Dr Michael Jones

Senior Staff Scientist:

Phone: +44 20 8722 4254

Email: [email protected]

Also on:  michael-jones-0640a550

Location: Sutton

Dr Michael Jones

Phone: +44 20 8722 4254

Email: [email protected]

Also on:  michael-jones-0640a550

Location: Sutton

Biography and research overview

Dr Michael Jones joined the Institute of Cancer Research in 2001. His main research interests are in the design and analysis of complex large epidemiological studies. He has been working on the Breast Cancer Now Generations Study, a large cohort study of over 100,000 women, since it started in 2003.

This study investigates the effects of reproductive, environmental and lifestyle factors, including modifiable factors such as physical activity and weight, as well as density of the breast, hormone levels, and genetic factors, on the risk of breast cancer overall and by sub-type. This information is essential for women’s understanding of their risks and for prevention, to give advice to policy makers and doctors, and to individual women and their children about their risk of breast cancer, and what they might be able to do to reduce that risk.

His work with the Generations Study has shown that risks associated with menopausal hormone therapy have been previously underestimated, risk of breast cancer is greater if women start smoking in adolescence, and (reassuringly) that breast cancer risk in Generations Study participants was not associated with night-shift work.

Dr Jones obtained a B.A. (Hons) in Natural Sciences (Physics) from the University of Cambridge, an MSc in Applied Statistics from the University of Oxford, and a PhD in Epidemiology from the London School of Hygiene & Tropical Medicine, University of London.

He  previously worked at the National Physical Laboratory and the Business Statistics Office (now part of the Office for National Statistics) in the UK. After completing his MSc he became a Research Fellow (Biostatistician) at the Menzies Centre for Population Health Research (University of Tasmania, Australia) working on the design and analysis of case-control and cohort studies.

He received a WHO Special Training Award held at the International Agency for Research on Cancer (Lyon, France) where he worked on world-wide incidence and mortality trends in malignant melanoma.  After he returned to the UK to complete a PhD he was appointed as a Lecturer in Medical Statistics at the London School of Hygiene & Tropical Medicine and was the inaugural Course Organiser for a distance learning MSc in Epidemiology: Principles and Practice. 

Types of Publications

Journal articles

Swerdlow, A.J. Jones, M.E. Schoemaker, M.J. Hemming, J. Thomas, D. Williamson, J. Ashworth, A (2011) The Breakthrough Generations Study: design of a long-term UK cohort study to investigate breast cancer aetiology.. Show Abstract full text

BACKGROUND: The rationale, design, recruitment and follow-up methods are described for the Breakthrough Generations Study, a UK cohort study started in 2003, targeted at investigation of breast cancer aetiology. METHODS: Cohort members have been recruited by a participant referral method intended to assemble economically a large general population cohort from whom detailed questionnaire information and blood samples can be obtained repeatedly over decades, with high completeness of follow-up and inclusion of large numbers of related individuals. 'First-generation' recruits were women contacted directly, or who volunteered directly, to join the study. They nominated female friends and family, whom we contacted, and those who joined ('second generation') nominated others, reiterated for up to 28 generations. RESULTS: The method has successfully been used during 2003-2011 to recruit 112,049 motivated participants with a broad geographic and socioeconomic distribution, aged 16-102 years, who have completed detailed questionnaires; 92% of the participants gave blood samples at recruitment. When eligible, 2½ years after recruitment, >98% completed the first follow-up questionnaire. Thirty percent are first-degree relatives of other study members. CONCLUSION: The 'generational' recruitment method has enabled recruitment of a large cohort who appear to have the commitment to enable long-term continuing data and sample collection, to investigate the effects of changing endogenous and exogenous factors on cancer risk.

Jones, M.E. Schoemaker, M.J. Wright, L. McFadden, E. Griffin, J. Thomas, D. Hemming, J. Wright, K. Ashworth, A. Swerdlow, A.J (2016) Menopausal hormone therapy and breast cancer: what is the true size of the increased risk?. Show Abstract full text

BACKGROUND: Menopausal hormone therapy (MHT) increases breast cancer risk; however, most cohort studies omit MHT use after enrolment and many infer menopausal age. METHODS: We used information from serial questionnaires from the UK Generations Study cohort to estimate hazard ratios (HRs) for breast cancer among post-menopausal women with known menopausal age, and examined biases induced when not updating data on MHT use and including women with inferred menopausal age. RESULTS: Among women recruited in 2003-2009, at 6 years of follow-up, 58 148 had reached menopause and 96% had completed a follow-up questionnaire. Among 39 183 women with known menopausal age, 775 developed breast cancer, and the HR in relation to current oestrogen plus progestogen MHT use (based on 52 current oestrogen plus progestogen MHT users in breast cancer cases) relative to those with no previous MHT use was 2.74 (95% confidence interval (CI): 2.05-3.65) for a median duration of 5.4 years of current use, reaching 3.27 (95% CI: 1.53-6.99) at 15+ years of use. The excess HR was underestimated by 53% if oestrogen plus progestogen MHT use was not updated after recruitment, 13% if women with uncertain menopausal age were included, and 59% if both applied. The HR for oestrogen-only MHT was not increased (HR=1.00; 95% CI: 0.66-1.54). CONCLUSIONS: Lack of updating MHT status through follow-up and inclusion of women with inferred menopausal age is likely to result in substantial underestimation of the excess relative risks for oestrogen plus progestogen MHT use in studies with long follow-up, limited updating of exposures, and changing or short durations of use.

Jones, M.E. Schoemaker, M.J. Wright, L.B. Ashworth, A. Swerdlow, A.J (2017) Smoking and risk of breast cancer in the Generations Study cohort.. Show Abstract full text

<h4>Background</h4>Plausible biological reasons exist regarding why smoking could affect breast cancer risk, but epidemiological evidence is inconsistent.<h4>Methods</h4>We used serial questionnaire information from the Generations Study cohort (United Kingdom) to estimate HRs for breast cancer in relation to smoking adjusted for potentially confounding factors, including alcohol intake.<h4>Results</h4>Among 102,927 women recruited 2003-2013, with an average of 7.7 years of follow-up, 1815 developed invasive breast cancer. The HR (reference group was never smokers) was 1.14 (95% CI 1.03-1.25; P = 0.010) for ever smokers, 1.24 (95% CI 1.08-1.43; P = 0.002) for starting smoking at ages < 17 years, and 1.23 (1.07-1.41; P = 0.004) for starting smoking 1-4 years after menarche. Breast cancer risk was not statistically associated with interval from initiation of smoking to first birth (P-trend = 0.97). Women with a family history of breast cancer (ever smoker vs never smoker HR 1.35; 95% CI 1.12-1.62; P = 0.002) had a significantly larger HR in relation to ever smokers (P for interaction = 0.039) than women without (ever smoker vs never smoker HR 1.07; 95% CI 0.96-1.20; P = 0.22). The interaction was prominent for age at starting smoking (P = 0.003) and starting smoking relative to age at menarche (P = 0.0001).<h4>Conclusions</h4>Smoking was associated with a modest but significantly increased risk of breast cancer, particularly among women who started smoking at adolescent or peri-menarcheal ages. The relative risk of breast cancer associated with smoking was greater for women with a family history of the disease.

Jones, M.E. Schoemaker, M.J. McFadden, E.C. Wright, L.B. Johns, L.E. Swerdlow, A.J (2019) Night shift work and risk of breast cancer in women: the Generations Study cohort.. Show Abstract full text

<h4>Background</h4>It is plausible that night shift work could affect breast cancer risk, possibly by melatonin suppression or circadian clock disruption, but epidemiological evidence is inconclusive.<h4>Methods</h4>Using serial questionnaires from the Generations Study cohort, we estimated hazard ratios (HR) and 95% confidence intervals (95%CI) for breast cancer in relation to being a night shift worker within the last 10 years, adjusted for potential confounders.<h4>Results</h4>Among 102,869 women recruited in 2003-2014, median follow-up 9.5 years, 2059 developed invasive breast cancer. The HR in relation to night shift work was 1.00 (95%CI: 0.86-1.15). There was a significant trend with average hours of night work per week (P = 0.035), but no significantly raised risks for hours worked per night, nights worked per week, average hours worked per week, cumulative years of employment, cumulative hours, time since cessation, type of occupation, age starting night shift work, or age starting in relation to first pregnancy.<h4>Conclusions</h4>The lack of overall association, and no association with all but one measure of dose, duration, and intensity in our data, does not support an increased risk of breast cancer from night shift work in women.

Jones, M.E. Folkerd, E.J. Doody, D.A. Iqbal, J. Dowsett, M. Ashworth, A. Swerdlow, A.J (2007) Effect of delays in processing blood samples on measured endogenous plasma sex hormone levels in women.. Show Abstract full text

Time spent in transit may affect the concentration of various constituents of collected blood samples and, consequently, results of sex hormone assays. Whole blood was collected from 46 women, and one third was processed immediately, one third was stored at ambient conditions (22 degrees C) for 1 day, and one third was stored for 2 days. Estradiol concentration increased by 7.1% [95% confidence interval (95% CI), 3.2-11.3%] after a delay in processing of 1 day and by 5.6% (95% CI, 0.2-11.4%) after a delay in processing of 2 days; the change was most apparent at lower than median concentrations. Progesterone concentrations showed no substantial change. Testosterone concentrations changed by 23.9% (95% CI, 17.8-30.3%) after a delay of 1 day but little thereafter. The sex hormone-binding globulin concentration decreased by 6.6% (95% CI, 4.6-8.6%) and 10.9% (95% CI, 8.1-13.6%), follicle-stimulating hormone increased by 7.4% (95% CI, 4.2-10.7%) and 13.9% (95% CI, 8.7-19.3%), and luteinizing hormone increased by 4.9% (95% CI, 1.3-8.5%) and 6.7% (95% CI, 2.2-11.5%) after a delay in processing of 1 and 2 days. Increases in calculated values for biologically available levels of estradiol and testosterone were greater than the increases seen in measured total hormone concentrations. Similar changes are likely when samples are delayed in transit, and evidence of etiology may be obscured unless study designs or analyses take into account processing delays.

Jones, M.E. Swerdlow, A.J (1998) Bias in the standardized mortality ratio when using general population rates to estimate expected number of deaths.. Show Abstract full text

Cohort studies often compare the observed number of cases arising in a group under investigation with the number expected to occur on the basis of general population rates. The general population is taken to represent unexposed persons, but it is almost inevitably biased in that it comprises all types of people including exposed ones. To identify circumstances when this bias matters, the authors modeled its effect in relation to the size of the observed standardized mortality ratio (SMR) and the prevalence of exposed individuals in the general population. The authors found that bias may be a major problem, causing substantial underestimation of the true relative risk, when either the prevalence of exposure in the general population or the SMR are large. The bias can cause an apparent trend in SMRs with age when none exists. It also places a limit on the maximum size of the observed SMR, no matter how large the true relative risk. A table is provided showing the extent of bias in different circumstances. Cohort studies of people with common diseases or exposures, or that find large SMRs, when using general population expectations, need to consider the extent of bias from this source.

Swerdlow, A.J. Jones, M.E. British Tamoxifen Second Cancer Study Group, (2005) Tamoxifen treatment for breast cancer and risk of endometrial cancer: a case-control study.. Show Abstract full text

BACKGROUND: Tamoxifen treatment of breast cancer is associated with an increased risk of endometrial cancer, but tamoxifen-related risks of endometrial cancer are unclear in premenopausal women, in long-term users of tamoxifen, and in women for whom several years have passed since ending treatment. We conducted a case-control study in Britain to investigate these risks. METHODS: We compared treatment information on 813 case patients who had endometrial cancer after their diagnosis for breast cancer and 1067 control patients who had breast cancer but not subsequent endometrial cancer. We assessed risk by conditional logistic regression analysis. All statistical tests were two-sided. RESULTS: Overall, tamoxifen treatment, compared with no treatment, was associated with an increased risk of endometrial cancer (odds ratio [OR] = 2.4; 95% confidence interval [CI] = 1.8 to 3.0). Risk increased statistically significantly (P(trend)<.001) with duration of treatment (for > or =5 years of treatment compared with no treatment, OR = 3.6, 95% CI = 2.6 to 4.8). As an indication of background levels of treatment, 16% of control patients received 5 years or more of treatment. Risk of endometrial cancer adjusted for treatment duration did not diminish in follow-up to at least 5 years after the last treatment ended. Risk of endometrial cancer was not associated with the daily dose of tamoxifen and was comparable in pre- and postmenopausal women. Ever treatment with tamoxifen was associated with a much greater risk of Mullerian and mesodermal mixed endometrial tumors (OR = 13.5, 95% CI = 4.1 to 44.5) than of adenocarcinoma (OR = 2.1, 95% CI = 1.6 to 2.7) or clear cell and papillary serous tumors (OR = 3.1, 95% CI = 0.8 to 17.9). CONCLUSIONS: There is an increasing risk of endometrial cancer associated with longer tamoxifen treatment, extending well beyond 5 years. The increased risk of endometrial cancer associated with tamoxifen treatment should be considered clinically for both premenopausal and postmenopausal women during treatment and for at least 5 years after the last treatment.

Swerdlow, A.J. Wright, L.B. Schoemaker, M.J. Jones, M.E (2018) Maternal breast cancer risk in relation to birthweight and gestation of her offspring.. Show Abstract full text

<h4>Background</h4>Parity and age at first pregnancy are well-established risk factors for breast cancer, but the effects of other characteristics of pregnancies are uncertain and the literature is inconsistent.<h4>Methods</h4>In a cohort of 83,451 parous women from the general population of the UK, which collected detailed information on each pregnancy and a wide range of potential confounders, we investigated the associations of length of gestation and birthweight of offspring in a woman's pregnancies with her breast cancer risk, adjusting for a full range of non-reproductive as well as reproductive risk factors unlike in previous large studies.<h4>Results</h4>Gestation of the first-born offspring was significantly inversely related to the risk of pre-menopausal breast cancer (p trend = 0.03; hazard ratio (HR) for 26-31 compared with 40-41 weeks, the baseline group, = 2.38, 95% confidence interval (CI) 1.26-4.49), and was borderline significantly related to risk of breast cancer overall (p trend = 0.05). Risk was significantly raised in mothers of high birthweight first-born (HR for breast cancer overall = 1.53, 95% CI 1.06-2.21 for ≥ 4500 g compared with 3000-3499 g, the baseline group). For gestation and birthweight of most recent birth, there were no clear effects. Analyses without adjustment for confounders (other than age) gave similar results.<h4>Conclusions</h4>Our data add to evidence that short gestation pregnancies may increase the risk of breast cancer, at least pre-menopausally, perhaps by hormonal stimulation and breast proliferation early in pregnancy without the opportunity for the differentiation that occurs in late pregnancy. High birthweight first pregnancies may increase breast cancer risk, possibly through the association of birthweight with oestrogen and insulin-like growth factor 1 levels.

Jones, M.E. Schoemaker, M.J. Rae, M. Folkerd, E.J. Dowsett, M. Ashworth, A. Swerdlow, A.J (2014) Reproducibility of estradiol and testosterone levels in postmenopausal women over 5 years: results from the breakthrough generations study.. Show Abstract full text

Prospective cohort studies examining sex hormones in relation to cancer risk have generally collected blood samples at 1 time point, with an assumption that hormone levels measured in these samples will be reliable markers of true levels at other times. In postmenopausal women, body fat is a major source of estradiol; therefore, changes in adiposity may affect the correlation of single measurements to more relevant long-term averages. To estimate the intraclass correlation coefficient (ICC) for estradiol and testosterone, we collected repeat blood samples from 119 postmenopausal women (average age = 59.4 (standard deviation, 4.7) years) from the United Kingdom during 2004-2005 and again during 2010-2011. The ICCs (adjusted for assay variation) were 0.73 (95% confidence interval: 0.63, 0.82) for total estradiol and 0.59 (95% confidence interval: 0.47, 0.72) for total testosterone. The ICCs were 3%-5% larger after adjustment for change in body mass index (weight (kg)/height (m)(2)) or leptin, which are 2 markers of change in adiposity. There was no increase in ICCs after adjustment for change in age, alcohol consumption, smoking, exercise, time between waking and blood collection, or season. The results suggest that other factors account for within-woman variation in these sex hormones.

Jones, M.E. Schoemaker, M. Rae, M. Folkerd, E.J. Dowsett, M. Ashworth, A. Swerdlow, A.J (2013) Changes in estradiol and testosterone levels in postmenopausal women after changes in body mass index.. Show Abstract full text

CONTEXT: Endogenous sex hormones are risk factors for postmenopausal breast cancer. A potential route for favorable hormonal modification is weight loss. OBJECTIVE: The objective of the study was to measure change in plasma estradiol and testosterone levels in postmenopausal women in relation to change in body mass index (BMI) and plasma leptin. SETTING: The setting was a cohort study of over 100,000 female volunteers from the general population, United Kingdom. PARTICIPANTS: The participants were a sample of 177 postmenopausal women aged over 45 years who provided blood samples during 2004-2005 and again during 2010-2011. MAIN OUTCOME MEASURE: Outcomes were percentage change in plasma estradiol and testosterone levels per 1 kg/m² change in BMI and per 1 ng/mL change in plasma leptin. RESULTS: Among women with reduction in BMI, estradiol decreased 12.7% (95% confidence interval: [6.4%, 19.5%]; P < .0001) per kg/m² and among women with increased BMI estradiol increased 6.4% [0.2%, 12.9%] (P = .042). The corresponding figures for testosterone were 10.7% [3.0%, 19.0%] (P = .006) and 1.9% [-5.4%, 9.7%] (P = .61) per kg/m². For women with decreases and increases in leptin, estradiol decreased by 3.6% [1.3%, 6.0%] (P = .003) per ng/mL and increased by 1.7% [-0.3%, 3.6%] (P = .094), respectively. The corresponding figures for testosterone were 4.8% [2.0%, 7.8%] (P = .009) and 0.3% [-2.0%, 2.6%] (P = .82) per ng/mL. CONCLUSIONS: In postmenopausal women, changes in BMI and plasma leptin occurring over several years are associated with changes in estradiol and testosterone levels. The results suggest that fat loss by an individual can result in substantial decreases in postmenopausal estradiol and testosterone levels and provides support for weight management to lessen breast cancer risk.

Seah, M.P. Jones, M.E (1984) Roughness Contributions to Resolution in Ion Sputter Depth Profiles of Polycrystalline Metal-Films.
Herlihy, E. Gies, P.H. Roy, C.R. Jones, M (1994) Personal dosimetry of solar UV radiation for different outdoor activities. Show Abstract full text

Quantifying individual exposure to ultraviolet radiation (UVR) is critical to understanding the etiology of a number of diseases including nonmelanotic and melanotic skin cancers. Measurements of personal exposure to solar UVR were made in Hobart, Tasmania in February (summer) 1991 for six different outdoor activities using UVR-sensitive polysulfone (PS) film attached at seven anatomical sites. Concurrent behavioral and environmental observations were also made. To date many studies have relied on subject recall to quantify past solar UVR exposures. To gain insight into the accuracy of subject recall the measured UVR exposures received by different subjects using the PS film were compared to those calculated from personal diaries and ambient solar UVB levels from a monitoring station. In general, when UVR exposure activities took place under close supervision, good correlations were obtained between the PS badges and the ambient measurements/diaries approach. Ultraviolet radiation exposures for the field study involving 94 subjects engaged in a number of outdoor activities are presented.

Bruce, J.C. Bond, S.T. Jones, M.E (2002) Teaching epidemiology and statistics by distance learning. full text
Sullivan, S.A. Marsden, K.A. Lowenthal, R.M. Jupe, D.M. Jones, M.E (1992) Circulating CD34+ cells: an adverse prognostic factor in the myelodysplastic syndromes. Show Abstract full text

As part of an epidemiological survey of myelodysplastic syndromes (MDS) in southern Tasmania, 62 MDS patients identified over a 2 year period were tested for the presence of CD34, the human progenitor cell antigen (HPCA), in their peripheral blood. The results were correlated with transformation to acute myeloid leukemia (AML) and patient survival, and CD34+ status was compared as a prognostic indicator with Bournemouth score, cytogenetics, and CFU-GM colony growth which were also assessed. Circulating CD34+ cells were found in 23 of the 62 MDS patients; 9 of the 23 patients with circulating CD34+ cells transformed to AML, as compared with none of the 39 CD34 negative patients (P less than 0.0001); and 11 of the 23 patients with circulating CD34+ cells were dead at the end of the 2 year period, as opposed to 6 of the 39 with no CD34+ cells (P less than 0.03). The Bournemouth score was also significantly associated with transformation to AML (P less than 0.0001) and poor survival (P less than 0.04). These were the only significant associations of the possible prognostic factors studied with either transformation or survival. In summary, the presence of circulating CD34+ cells was significantly associated with both progression to AML and poor survival and was found to be a better prognostic indicator than cytogenetics or CFU-GM colony growth.

Jones, M.E. Shugg, D. Dwyer, T. Young, B. Bonett, A (1992) Interstate differences in incidence and mortality from melanoma. A re-examination of the latitudinal gradient. Show Abstract full text

OBJECTIVE: To investigate the patterns of cutaneous malignant melanoma (CMM) mortality in Australia. DESIGN: A descriptive analysis of melanoma incidence and mortality in Australia supplemented by a case series analysis of melanoma survival. Melanoma mortality rates were based on tabulations supplied by the Australian Bureau of Statistics for the years 1969-1989. Melanoma incidence rates were based on State cancer registry records for the years 1977-1990. The case series survival analysis was based on detailed individual records from the population-based cancer registries in Tasmania and South Australia. MAIN OUTCOME MEASURES: The level of and rise in melanoma mortality rates during 1969-1989 in Australia; the five-year survival rates for Tasmanian and South Australian cases; and male:female incidence ratios related to latitude. RESULTS: We found annual increases in melanoma mortality rates of 2.5% in men (P < 0.0001) and 1.1% in women (P < 0.0001) for all Australia. The five-year survival rates (with 95% confidence intervals [CI]) were: 67% (59%-75%) for Tasmanian men; 79% (76%-83%) for South Australian men; 80% (74%-86%) for Tasmanian women and 88% (86%-91%) for South Australian women. A change in the male:female incidence ratio with latitude was also found--women have significantly higher incidence rates at higher latitudes, but similar rates to men at lower latitudes. CONCLUSIONS: The age standardised mortality from CMM for the period 1969 to 1989 shows little variation by State for women, despite a considerable range in latitude. CMM mortality in men is increasing at a faster rate than that in women. Between 1982 and 1987 the male:female incidence ratio in high latitudes in the Southern Hemisphere showed an excess of cases in women, a finding which we believe has not been reported before.

Saxena, S. Majeed, A. Jones, M (1999) Socioeconomic differences in childhood consultation rates in general practice in England and Wales: prospective cohort study. Show Abstract full text

OBJECTIVE: To establish how consultation rates in children for episodes of illness, preventive activities, and home visits vary by social class. DESIGN: Analysis of prospectively collected data from the fourth national survey of morbidity in general practice, carried out between September 1991 and August 1992. SETTING: 60 general practices in England and Wales. SUBJECTS: 106 102 children aged 0 to 15 years registered with the participating practices. MAIN OUTCOME MEASURES: Mean overall consultation rates for any reason, illness by severity of underlying disease, preventive episodes, home visits, and specific diagnostic category (infections, asthma, and injuries). RESULTS: Overall consultation rates increased from registrar general's social classes I-II to classes IV-V in a linear pattern (for IV-V v I-II rate ratio 1.18; 95% confidence interval 1.14 to 1. 22). Children from social classes IV-V consulted more frequently than children from classes I-II for illnesses (rate ratio 1.23; 1.15 to 1.30), including infections, asthma, and injuries and poisonings. They also had significantly higher consultation rates for minor, moderate, and serious illnesses and higher home visiting rates (rate ratio 2.00; 1.81 to 2.18). Consultations for preventive activities were lower in children from social classes IV-V than in children from social classes I-II (rate ratio 0.95; 0.86 to 1.05). CONCLUSIONS: Childhood consultation rates for episodes of illness increase from social classes I-II through to classes IV-V. The findings on severity of underlying illness suggest the health of children from lower social classes is worse than that of children from higher social classes. These results reinforce the need to identify and target children for preventive health care in their socioeconomic context.

Swerdlow, A.J. Jones, M.E (2007) Ovarian cancer risk in premenopausal and perimenopausal women treated with Tamoxifen: a case-control study.. Show Abstract full text

As tamoxifen stimulates ovarian steroidogenesis in premenopausal women, induces ovulation and increases the incidence of benign ovarian cysts, there has been concern that it might also increase ovarian cancer risk in women treated premenopausally. In a national case-control study in Britain, treatment histories were collected for 158 cases of ovarian cancer after breast cancer diagnosed at ages under 55 years and 464 controls who had breast cancer at these ages without subsequent ovarian cancer. Risk of ovarian cancer was not raised for women overall who had taken tamoxifen (odds ratio (OR)=0.9, 95% confidence interval (CI) 0.6-1.3) or for those treated when premenopausal (OR=1.0, 95% CI 0.6-1.6) or perimenopausal (OR=0.7, 95% CI 0.2-2.4). There was also no relation of risk to daily dose, duration or cumulative dose of tamoxifen, or time since last use. There was, however, a significantly raised risk in relation to non-hormonal chemotherapy. The results suggest that tamoxifen treatment of premenopausal or perimenopausal women does not materially affect ovarian cancer risk, but that non-hormonal chemotherapy might increase risk.

Swerdlow, A.J. Bruce, C. Cooke, R. Coulson, P. Jones, M.E (2022) Infertility and risk of breast cancer in men: a national case-control study in England and Wales.. Show Abstract full text

<h4>Purpose</h4>Breast cancer is uncommon in men and its aetiology is largely unknown, reflecting the limited size of studies thus far conducted. In general, number of children fathered has been found a risk factor inconsistently, and infertility not. We therefore investigated in a case-control study, the relation of risk of breast cancer in men to infertility and number of children.<h4>Patients and methods</h4>We conducted a national case-control study in England and Wales, interviewing 1998 cases incident 2005-17 and 1597 male controls, which included questions on infertility and offspring.<h4>Results</h4>Risk of breast cancer was statistically significantly associated with male-origin infertility (OR = 2.03 (95% confidence interval (CI) 1.18-3.49)) but not if a couple's infertility had been diagnosed as of origin from the female partner (OR = 0.86 (0.51-1.45)). Risk was statistically significantly raised for men who had not fathered any children (OR = 1.50 (95% CI 1.21-1.86)) compared with men who were fathers. These associations were statistically significantly present for invasive tumours but not statistically significant for in situ tumours.<h4>Conclusion</h4>Our data give strong evidence that risk of breast cancer is increased for men who are infertile. The reason is not clear and needs investigation.

Swerdlow, A.J. Bruce, C. Cooke, R. Coulson, P. Schoemaker, M.J. Jones, M.E (2022) Risk of breast cancer in men in relation to weight change: A national case-control study in England and Wales.. Show Abstract full text

Breast cancer is uncommon in men and knowledge about its causation limited. Obesity is a risk factor but there has been no investigation of whether weight change is an independent risk factor, as it is in women. In a national case-control study, 1998 men with breast cancer incident in England and Wales during 2005 to 2017 and 1597 male controls were interviewed about risk factors for breast cancer including anthropometric factors at several ages. Relative risks of breast cancer in relation to changes in body mass index (BMI) and waist/height ratios at these ages were obtained by logistic regression modelling. There were significant trends of increasing breast cancer risk with increase in BMI from age 20 to 40 (odds ratio [OR] 1.11 [95% confidence interval (CI) 1.05-1.17] per 2 kg/m<sup>2</sup> increase in BMI; P < .001), and from age 40 to 60 (OR 1.12 [1.04-1.20]; P = .003), and with increase in self-reported adiposity compared to peers at age 11 to BMI compared with peers at age 20 (OR 1.19 [1.09-1.30]; P < .001). Increase in waist/height ratio from age 20 to 5 years before diagnosis was also highly significantly associated with risk (OR 1.13 [1.08-1.19]; P < .001). The associations with increases in BMI and waist/height ratio were significant independently of each other and of BMI or waist/height ratio at the start of the period of change analysed, and effects were similar for invasive and in situ tumours separately. Increases in BMI and abdominal obesity are each risk factors for breast cancer in men, independently of obesity per se. These associations might relate to increasing oestrogen levels with weight gain, but this needs investigation.

Swerdlow, A.J. Bruce, C. Cooke, R. Coulson, P. Griffin, J. Butlin, A. Smith, B. Swerdlow, M.J. Jones, M.E (2021) Obesity and Breast Cancer Risk in Men: A National Case-Control Study in England and Wales.. Show Abstract full text

<h4>Background</h4>Breast cancer is rare in men, and information on its causes is very limited from studies that have generally been small. Adult obesity has been shown as a risk factor, but more detailed anthropometric relations have not been investigated.<h4>Methods</h4>We conducted an interview population-based case-control study of breast cancer in men in England and Wales including 1998 cases incident during 2005-2017 at ages younger than 80 years and 1597 male controls, with questions asked about a range of anthropometric variables at several ages. All tests of statistical significance were 2-sided.<h4>Results</h4>Risk of breast cancer statistically significantly increased with increasing body mass index (BMI) at ages 20 (odds ratio [OR] = 1.07, 95% confidence interval [CI] = 1.02 to 1.12 per 2-unit change in BMI), 40 (OR = 1.11, 95% CI = 1.07 to 1.16), and 60 (OR = 1.14, 95% CI = 1.09 to 1.19) years, but there was also an indication of raised risk for the lowest BMIs. Large waist circumference 5 years before interview was more strongly associated than was BMI with risk, and each showed independent associations. Associations were similar for invasive and in situ tumors separately and stronger for HER2-positive than HER2-negative tumors. Of the tumors, 99% were estrogen receptor positive.<h4>Conclusions</h4>Obesity at all adult ages, particularly recent abdominal obesity, is associated with raised risk of breast cancer in men, probably because of the conversion of testosterone to estrogen by aromatase in adipose tissue. The association is particularly strong for HER2-expressing tumors.

Types of Publications

Journal articles

Morris, D.H. Jones, M.E. Schoemaker, M.J. Ashworth, A. Swerdlow, A.J (2012) Familial concordance for height and its components: analyses from the Breakthrough Generations Study.. Show Abstract full text

OBJECTIVES: To assess familial resemblance for height, arm span, and components of these, and differences between concordance for short and tall heights. METHODS: We examined whether female relatives were similar for six anthropometric measurements (height, arm span, leg, trunk and arm length, and leg:trunk length ratio). Subjects were 31,622 related individuals aged 16-102 yr participating in the UK Breakthrough Generations Study. Height and arm span were self-reported, limb and trunk length were measured in a subset (N = 508) by study investigators, and paternal height was reported by the daughter. Data were analyzed using correlations and Poisson regression. RESULTS: Correlation coefficients within families were 0.4 for height, 0.3 for arm span, and 0.5 for leg length, trunk length, leg:trunk ratio, and arm length. Women had a relative risk (RR) of being short (i.e., in the lowest height quintile) of 2.3 (95% confidence interval [CI] = 2.1-2.5) if their mother was short, 2.1 (95% CI = 1.9-2.3) if their father was short, and 3.7 (95% CI = 3.4-4.0) if both parents were short. RRs of being tall (i.e., in the highest height quintile) were 2.3 (95% CI = 2.1-2.5), 2.4 (95% CI = 2.2-2.6), and 4.4 (95% CI = 4.1-4.8) if their mother, father or both were tall, respectively. CONCLUSIONS: We have shown, for the first time, that leg:trunk length ratio and arm length aggregate within families. Concordance seemed to be stronger for tall than short heights.

Morris, D.H. Jones, M.E. Schoemaker, M.J. McFadden, E. Ashworth, A. Swerdlow, A.J (2012) Body mass index, exercise, and other lifestyle factors in relation to age at natural menopause: analyses from the breakthrough generations study.. Show Abstract full text

The authors examined the effect of women's lifestyles on the timing of natural menopause using data from a cross-sectional questionnaire used in the United Kingdom-based Breakthrough Generations Study in 2003-2011. The analyses included 50,678 women (21,511 who had experienced a natural menopause) who were 40-98 years of age at study entry and did not have a history of breast cancer. Cox competing risks proportional hazards models were fitted to examine the relation of age at natural menopause to lifestyle and anthropometric factors. Results were adjusted for age at reporting, smoking status at menopause, parity, and body mass index at age 40 years, as appropriate. All P values were 2-sided. High adult weight (P(trend) < 0.001), high body mass index (P(trend) < 0.001), weight gain between the ages of 20 and 40 years (P(trend) = 0.01), not smoking (P < 0.001), increased alcohol consumption (P(trend) < 0.001), regular strenuous exercise (P < 0.01), and not being a vegetarian (P < 0.001) were associated with older age at menopause. Neither height nor history of an eating disorder was associated with menopausal age. These findings show the importance of lifestyle factors in determining menopausal age.

Cooke, R. Jones, M.E. Cunningham, D. Falk, S.J. Gilson, D. Hancock, B.W. Harris, S.J. Horwich, A. Hoskin, P.J. Illidge, T. Linch, D.C. Lister, T.A. Lucraft, H.H. Radford, J.A. Stevens, A.M. Syndikus, I. Williams, M.V. England and Wales Hodgkin Lymphoma Follow-up Group, . Swerdlow, A.J (2013) Breast cancer risk following Hodgkin lymphoma radiotherapy in relation to menstrual and reproductive factors.. Show Abstract full text

<h4>Background</h4>Women treated with supradiaphragmatic radiotherapy (sRT) for Hodgkin lymphoma (HL) at young ages have a substantially increased breast cancer risk. Little is known about how menarcheal and reproductive factors modify this risk.<h4>Methods</h4>We examined the effects of menarcheal age, pregnancy, and menopausal age on breast cancer risk following sRT in case-control data from questionnaires completed by 2497 women from a cohort of 5002 treated with sRT for HL at ages <36 during 1956-2003.<h4>Results</h4>Two-hundred and sixty women had been diagnosed with breast cancer. Breast cancer risk was significantly increased in patients treated within 6 months of menarche (odds ratio (OR) 5.52, 95% confidence interval (CI) (1.97-15.46)), and increased significantly with proximity of sRT to menarche (Ptrend<0.001). It was greatest when sRT was close to a late menarche, but based on small numbers and needing reexamination elsewhere. Risk was not significantly affected by full-term pregnancies before or after treatment. Risk was significantly reduced by early menopause (OR 0.55, 95% CI (0.35-0.85)), and increased with number of premenopausal years after treatment (Ptrend=0.003).<h4>Conclusion</h4>In summary, this paper shows for the first time that sRT close to menarche substantially increases breast cancer risk. Careful consideration should be given to follow-up of these women, and to measures that might reduce their future breast cancer risk.

Ong, J.-.S. Hwang, L.-.D. Cuellar-Partida, G. Martin, N.G. Chenevix-Trench, G. Quinn, M.C.J. Cornelis, M.C. Gharahkhani, P. Webb, P.M. MacGregor, S. Ovarian Cancer Association Consortium, (2018) Assessment of moderate coffee consumption and risk of epithelial ovarian cancer: a Mendelian randomization study.. Show Abstract full text

<h4>Background</h4>Coffee consumption has been shown to be associated with various health outcomes in observational studies. However, evidence for its association with epithelial ovarian cancer (EOC) is inconsistent and it is unclear whether these associations are causal.<h4>Methods</h4>We used single nucleotide polymorphisms associated with (i) coffee and (ii) caffeine consumption to perform Mendelian randomization (MR) on EOC risk. We conducted a two-sample MR using genetic data on 44 062 individuals of European ancestry from the Ovarian Cancer Association Consortium (OCAC), and combined instrumental variable estimates using a Wald-type ratio estimator.<h4>Results</h4>For all EOC cases, the causal odds ratio (COR) for genetically predicted consumption of one additional cup of coffee per day was 0.92 [95% confidence interval (CI): 0.79, 1.06]. The COR was 0.90 (95% CI: 0.73, 1.10) for high-grade serous EOC. The COR for genetically predicted consumption of an additional 80 mg caffeine was 1.01 (95% CI: 0.92, 1.11) for all EOC cases and 0.90 (95% CI: 0.73, 1.10) for high-grade serous cases.<h4>Conclusions</h4>We found no evidence indicative of a strong association between EOC risk and genetically predicted coffee or caffeine levels. However, our estimates were not statistically inconsistent with earlier observational studies and we were unable to rule out small protective associations.

Johns, L.E. Jones, M.E. Schoemaker, M.J. McFadden, E. Ashworth, A. Swerdlow, A.J (2018) Domestic light at night and breast cancer risk: a prospective analysis of 105 000 UK women in the Generations Study.. Show Abstract full text

<h4>Background</h4>Circadian disruption caused by exposure to light at night (LAN) has been proposed as a risk factor for breast cancer and a reason for secular increases in incidence. Studies to date have largely been ecological or case-control in design and findings have been mixed.<h4>Methods</h4>We investigated the relationship between LAN and breast cancer risk in the UK Generations Study. Bedroom light levels and sleeping patterns at age 20 and at study recruitment were obtained by questionnaire. Analyses were conducted on 105 866 participants with no prior history of breast cancer. During an average of 6.1 years of follow-up, 1775 cases of breast cancer were diagnosed. Cox proportional hazard models were used to calculate hazard ratios (HRs), adjusting for potential confounding factors.<h4>Results</h4>There was no association between LAN level and breast cancer risk overall (highest compared with lowest LAN level at recruitment: HR=1.01, 95% confidence interval (CI): 0.88-1.15), or for invasive (HR=0.98, 95% CI: 0.85-1.13) or in situ (HR=0.96, 95% CI: 0.83-1.11) breast cancer, or oestrogen-receptor (ER) positive (HR=0.98, 95% CI: 0.84-1.14); or negative (HR=1.16, 95% CI: 0.82-1.65) tumours separately. The findings did not differ by menopausal status. Adjusting for sleep duration, sleeping at unusual times (non-peak sleep) and history of night work did not affect the results. Night waking with exposure to light, occurring around age 20, was associated with a reduced risk of premenopausal breast cancer (HR for breast cancer overall=0.74, 95% CI: 0.55-0.99; HR for ER-positive breast cancer=0.69, 95% CI: 0.49-0.97).<h4>Conclusions</h4>In this prospective cohort analysis of LAN, there was no evidence that LAN exposure increased the risk of subsequent breast cancer, although the suggestion of a lower breast cancer risk in pre-menopausal women with a history of night waking in their twenties may warrant further investigation.

Abubakar, M. Chang-Claude, J. Ali, H.R. Chatterjee, N. Coulson, P. Daley, F. Blows, F. Benitez, J. Milne, R.L. Brenner, H. Stegmaier, C. Mannermaa, A. Rudolph, A. Sinn, P. Couch, F.J. Devilee, P. Tollenaar, R.A.E.M. Seynaeve, C. Figueroa, J. Lissowska, J. Hewitt, S. Hooning, M.J. Hollestelle, A. Foekens, R. Koppert, L.B. kConFab Investigators, . Bolla, M.K. Wang, Q. Jones, M.E. Schoemaker, M.J. Keeman, R. Easton, D.F. Swerdlow, A.J. Sherman, M.E. Schmidt, M.K. Pharoah, P.D. Garcia-Closas, M (2018) Etiology of hormone receptor positive breast cancer differs by levels of histologic grade and proliferation.. Show Abstract full text

Limited epidemiological evidence suggests that the etiology of hormone receptor positive (HR+) breast cancer may differ by levels of histologic grade and proliferation. We pooled risk factor and pathology data on 5,905 HR+ breast cancer cases and 26,281 controls from 11 epidemiological studies. Proliferation was determined by centralized automated measures of KI67 in tissue microarrays. Odds ratios (OR), 95% confidence intervals (CI) and p-values for case-case and case-control comparisons for risk factors in relation to levels of grade and quartiles (Q1-Q4) of KI67 were estimated using polytomous logistic regression models. Case-case comparisons showed associations between nulliparity and high KI67 [OR (95% CI) for Q4 vs. Q1 = 1.54 (1.22, 1.95)]; obesity and high grade [grade 3 vs. 1 = 1.68 (1.31, 2.16)] and current use of combined hormone therapy (HT) and low grade [grade 3 vs. 1 = 0.27 (0.16, 0.44)] tumors. In case-control comparisons, nulliparity was associated with elevated risk of tumors with high but not low levels of proliferation [1.43 (1.14, 1.81) for KI67 Q4 vs. 0.83 (0.60, 1.14) for KI67 Q1]; obesity among women ≥50 years with high but not low grade tumors [1.55 (1.17, 2.06) for grade 3 vs. 0.88 (0.66, 1.16) for grade 1] and HT with low but not high grade tumors [3.07 (2.22, 4.23) for grade 1 vs. 0.85 (0.55, 1.30) for grade 3]. Menarcheal age and family history were similarly associated with HR+ tumors of different grade or KI67 levels. These findings provide insights into the etiologic heterogeneity of HR+ tumors.

Phelan, C.M. Kuchenbaecker, K.B. Tyrer, J.P. Kar, S.P. Lawrenson, K. Winham, S.J. Dennis, J. Pirie, A. Riggan, M.J. Chornokur, G. Earp, M.A. Lyra, P.C. Lee, J.M. Coetzee, S. Beesley, J. McGuffog, L. Soucy, P. Dicks, E. Lee, A. Barrowdale, D. Lecarpentier, J. Leslie, G. Aalfs, C.M. Aben, K.K.H. Adams, M. Adlard, J. Andrulis, I.L. Anton-Culver, H. Antonenkova, N. AOCS study group, . Aravantinos, G. Arnold, N. Arun, B.K. Arver, B. Azzollini, J. Balmaña, J. Banerjee, S.N. Barjhoux, L. Barkardottir, R.B. Bean, Y. Beckmann, M.W. Beeghly-Fadiel, A. Benitez, J. Bermisheva, M. Bernardini, M.Q. Birrer, M.J. Bjorge, L. Black, A. Blankstein, K. Blok, M.J. Bodelon, C. Bogdanova, N. Bojesen, A. Bonanni, B. Borg, Å. Bradbury, A.R. Brenton, J.D. Brewer, C. Brinton, L. Broberg, P. Brooks-Wilson, A. Bruinsma, F. Brunet, J. Buecher, B. Butzow, R. Buys, S.S. Caldes, T. Caligo, M.A. Campbell, I. Cannioto, R. Carney, M.E. Cescon, T. Chan, S.B. Chang-Claude, J. Chanock, S. Chen, X.Q. Chiew, Y.-.E. Chiquette, J. Chung, W.K. Claes, K.B.M. Conner, T. Cook, L.S. Cook, J. Cramer, D.W. Cunningham, J.M. D'Aloisio, A.A. Daly, M.B. Damiola, F. Damirovna, S.D. Dansonka-Mieszkowska, A. Dao, F. Davidson, R. DeFazio, A. Delnatte, C. Doheny, K.F. Diez, O. Ding, Y.C. Doherty, J.A. Domchek, S.M. Dorfling, C.M. Dörk, T. Dossus, L. Duran, M. Dürst, M. Dworniczak, B. Eccles, D. Edwards, T. Eeles, R. Eilber, U. Ejlertsen, B. Ekici, A.B. Ellis, S. Elvira, M. EMBRACE Study, . Eng, K.H. Engel, C. Evans, D.G. Fasching, P.A. Ferguson, S. Ferrer, S.F. Flanagan, J.M. Fogarty, Z.C. Fortner, R.T. Fostira, F. Foulkes, W.D. Fountzilas, G. Fridley, B.L. Friebel, T.M. Friedman, E. Frost, D. Ganz, P.A. Garber, J. García, M.J. Garcia-Barberan, V. Gehrig, A. GEMO Study Collaborators, . Gentry-Maharaj, A. Gerdes, A.-.M. Giles, G.G. Glasspool, R. Glendon, G. Godwin, A.K. Goldgar, D.E. Goranova, T. Gore, M. Greene, M.H. Gronwald, J. Gruber, S. Hahnen, E. Haiman, C.A. Håkansson, N. Hamann, U. Hansen, T.V.O. Harrington, P.A. Harris, H.R. Hauke, J. HEBON Study, . Hein, A. Henderson, A. Hildebrandt, M.A.T. Hillemanns, P. Hodgson, S. Høgdall, C.K. Høgdall, E. Hogervorst, F.B.L. Holland, H. Hooning, M.J. Hosking, K. Huang, R.-.Y. Hulick, P.J. Hung, J. Hunter, D.J. Huntsman, D.G. Huzarski, T. Imyanitov, E.N. Isaacs, C. Iversen, E.S. Izatt, L. Izquierdo, A. Jakubowska, A. James, P. Janavicius, R. Jernetz, M. Jensen, A. Jensen, U.B. John, E.M. Johnatty, S. Jones, M.E. Kannisto, P. Karlan, B.Y. Karnezis, A. Kast, K. KConFab Investigators, . Kennedy, C.J. Khusnutdinova, E. Kiemeney, L.A. Kiiski, J.I. Kim, S.-.W. Kjaer, S.K. Köbel, M. Kopperud, R.K. Kruse, T.A. Kupryjanczyk, J. Kwong, A. Laitman, Y. Lambrechts, D. Larrañaga, N. Larson, M.C. Lazaro, C. Le, N.D. Le Marchand, L. Lee, J.W. Lele, S.B. Leminen, A. Leroux, D. Lester, J. Lesueur, F. Levine, D.A. Liang, D. Liebrich, C. Lilyquist, J. Lipworth, L. Lissowska, J. Lu, K.H. Lubinński, J. Luccarini, C. Lundvall, L. Mai, P.L. Mendoza-Fandiño, G. Manoukian, S. Massuger, L.F.A.G. May, T. Mazoyer, S. McAlpine, J.N. McGuire, V. McLaughlin, J.R. McNeish, I. Meijers-Heijboer, H. Meindl, A. Menon, U. Mensenkamp, A.R. Merritt, M.A. Milne, R.L. Mitchell, G. Modugno, F. Moes-Sosnowska, J. Moffitt, M. Montagna, M. Moysich, K.B. Mulligan, A.M. Musinsky, J. Nathanson, K.L. Nedergaard, L. Ness, R.B. Neuhausen, S.L. Nevanlinna, H. Niederacher, D. Nussbaum, R.L. Odunsi, K. Olah, E. Olopade, O.I. Olsson, H. Olswold, C. O'Malley, D.M. Ong, K.-.R. Onland-Moret, N.C. OPAL study group, . Orr, N. Orsulic, S. Osorio, A. Palli, D. Papi, L. Park-Simon, T.-.W. Paul, J. Pearce, C.L. Pedersen, I.S. Peeters, P.H.M. Peissel, B. Peixoto, A. Pejovic, T. Pelttari, L.M. Permuth, J.B. Peterlongo, P. Pezzani, L. Pfeiler, G. Phillips, K.-.A. Piedmonte, M. Pike, M.C. Piskorz, A.M. Poblete, S.R. Pocza, T. Poole, E.M. Poppe, B. Porteous, M.E. Prieur, F. Prokofyeva, D. Pugh, E. Pujana, M.A. Pujol, P. Radice, P. Rantala, J. Rappaport-Fuerhauser, C. Rennert, G. Rhiem, K. Rice, P. Richardson, A. Robson, M. Rodriguez, G.C. Rodríguez-Antona, C. Romm, J. Rookus, M.A. Rossing, M.A. Rothstein, J.H. Rudolph, A. Runnebaum, I.B. Salvesen, H.B. Sandler, D.P. Schoemaker, M.J. Senter, L. Setiawan, V.W. Severi, G. Sharma, P. Shelford, T. Siddiqui, N. Side, L.E. Sieh, W. Singer, C.F. Sobol, H. Song, H. Southey, M.C. Spurdle, A.B. Stadler, Z. Steinemann, D. Stoppa-Lyonnet, D. Sucheston-Campbell, L.E. Sukiennicki, G. Sutphen, R. Sutter, C. Swerdlow, A.J. Szabo, C.I. Szafron, L. Tan, Y.Y. Taylor, J.A. Tea, M.-.K. Teixeira, M.R. Teo, S.-.H. Terry, K.L. Thompson, P.J. Thomsen, L.C.V. Thull, D.L. Tihomirova, L. Tinker, A.V. Tischkowitz, M. Tognazzo, S. Toland, A.E. Tone, A. Trabert, B. Travis, R.C. Trichopoulou, A. Tung, N. Tworoger, S.S. van Altena, A.M. Van Den Berg, D. van der Hout, A.H. van der Luijt, R.B. Van Heetvelde, M. Van Nieuwenhuysen, E. van Rensburg, E.J. Vanderstichele, A. Varon-Mateeva, R. Vega, A. Edwards, D.V. Vergote, I. Vierkant, R.A. Vijai, J. Vratimos, A. Walker, L. Walsh, C. Wand, D. Wang-Gohrke, S. Wappenschmidt, B. Webb, P.M. Weinberg, C.R. Weitzel, J.N. Wentzensen, N. Whittemore, A.S. Wijnen, J.T. Wilkens, L.R. Wolk, A. Woo, M. Wu, X. Wu, A.H. Yang, H. Yannoukakos, D. Ziogas, A. Zorn, K.K. Narod, S.A. Easton, D.F. Amos, C.I. Schildkraut, J.M. Ramus, S.J. Ottini, L. Goodman, M.T. Park, S.K. Kelemen, L.E. Risch, H.A. Thomassen, M. Offit, K. Simard, J. Schmutzler, R.K. Hazelett, D. Monteiro, A.N. Couch, F.J. Berchuck, A. Chenevix-Trench, G. Goode, E.L. Sellers, T.A. Gayther, S.A. Antoniou, A.C. Pharoah, P.D.P (2017) Identification of 12 new susceptibility loci for different histotypes of epithelial ovarian cancer.. Show Abstract full text

To identify common alleles associated with different histotypes of epithelial ovarian cancer (EOC), we pooled data from multiple genome-wide genotyping projects totaling 25,509 EOC cases and 40,941 controls. We identified nine new susceptibility loci for different EOC histotypes: six for serous EOC histotypes (3q28, 4q32.3, 8q21.11, 10q24.33, 18q11.2 and 22q12.1), two for mucinous EOC (3q22.3 and 9q31.1) and one for endometrioid EOC (5q12.3). We then performed meta-analysis on the results for high-grade serous ovarian cancer with the results from analysis of 31,448 BRCA1 and BRCA2 mutation carriers, including 3,887 mutation carriers with EOC. This identified three additional susceptibility loci at 2q13, 8q24.1 and 12q24.31. Integrated analyses of genes and regulatory biofeatures at each locus predicted candidate susceptibility genes, including OBFC1, a new candidate susceptibility gene for low-grade and borderline serous EOC.

Nichols, H.B. Schoemaker, M.J. Wright, L.B. McGowan, C. Brook, M.N. McClain, K.M. Jones, M.E. Adami, H.-.O. Agnoli, C. Baglietto, L. Bernstein, L. Bertrand, K.A. Blot, W.J. Boutron-Ruault, M.-.C. Butler, L. Chen, Y. Doody, M.M. Dossus, L. Eliassen, A.H. Giles, G.G. Gram, I.T. Hankinson, S.E. Hoffman-Bolton, J. Kaaks, R. Key, T.J. Kirsh, V.A. Kitahara, C.M. Koh, W.-.P. Larsson, S.C. Lund, E. Ma, H. Merritt, M.A. Milne, R.L. Navarro, C. Overvad, K. Ozasa, K. Palmer, J.R. Peeters, P.H. Riboli, E. Rohan, T.E. Sadakane, A. Sund, M. Tamimi, R.M. Trichopoulou, A. Vatten, L. Visvanathan, K. Weiderpass, E. Willett, W.C. Wolk, A. Zeleniuch-Jacquotte, A. Zheng, W. Sandler, D.P. Swerdlow, A.J (2017) The Premenopausal Breast Cancer Collaboration: A Pooling Project of Studies Participating in the National Cancer Institute Cohort Consortium.. Show Abstract full text

Breast cancer is a leading cancer diagnosis among premenopausal women around the world. Unlike rates in postmenopausal women, incidence rates of advanced breast cancer have increased in recent decades for premenopausal women. Progress in identifying contributors to breast cancer risk among premenopausal women has been constrained by the limited numbers of premenopausal breast cancer cases in individual studies and resulting low statistical power to subcategorize exposures or to study specific subtypes. The Premenopausal Breast Cancer Collaborative Group was established to facilitate cohort-based analyses of risk factors for premenopausal breast cancer by pooling individual-level data from studies participating in the United States National Cancer Institute Cohort Consortium. This article describes the Group, including the rationale for its initial aims related to pregnancy, obesity, and physical activity. We also describe the 20 cohort studies with data submitted to the Group by June 2016. The infrastructure developed for this work can be leveraged to support additional investigations. <i>Cancer Epidemiol Biomarkers Prev; 26(9); 1360-9. ©2017 AACR</i>.

Brouckaert, O. Rudolph, A. Laenen, A. Keeman, R. Bolla, M.K. Wang, Q. Soubry, A. Wildiers, H. Andrulis, I.L. Arndt, V. Beckmann, M.W. Benitez, J. Blomqvist, C. Bojesen, S.E. Brauch, H. Brennan, P. Brenner, H. Chenevix-Trench, G. Choi, J.-.Y. Cornelissen, S. Couch, F.J. Cox, A. Cross, S.S. Czene, K. Eriksson, M. Fasching, P.A. Figueroa, J. Flyger, H. Giles, G.G. González-Neira, A. Guénel, P. Hall, P. Hollestelle, A. Hopper, J.L. Ito, H. Jones, M. Kang, D. kConFab, . Knight, J.A. Kosma, V.-.M. Li, J. Lindblom, A. Lilyquist, J. Lophatananon, A. Mannermaa, A. Manoukian, S. Margolin, S. Matsuo, K. Muir, K. Nevanlinna, H. Peterlongo, P. Pylkäs, K. Saajrang, S. Seynaeve, C. Shen, C.-.Y. Shu, X.-.O. Southey, M.C. Swerdlow, A. Teo, S.-.H. Tollenaar, R.A.E.M. Truong, T. Tseng, C.-.C. van den Broek, A.J. van Deurzen, C.H.M. Winqvist, R. Wu, A.H. Yip, C.H. Yu, J.-.C. Zheng, W. Milne, R.L. Pharoah, P.D.P. Easton, D.F. Schmidt, M.K. Garcia-Closas, M. Chang-Claude, J. Lambrechts, D. Neven, P (2017) Reproductive profiles and risk of breast cancer subtypes: a multi-center case-only study.. Show Abstract full text

<h4>Background</h4>Previous studies have shown that reproductive factors are differentially associated with breast cancer (BC) risk by subtypes. The aim of this study was to investigate associations between reproductive factors and BC subtypes, and whether these vary by age at diagnosis.<h4>Methods</h4>We used pooled data on tumor markers (estrogen and progesterone receptor, human epidermal growth factor receptor-2 (HER2)) and reproductive risk factors (parity, age at first full-time pregnancy (FFTP) and age at menarche) from 28,095 patients with invasive BC from 34 studies participating in the Breast Cancer Association Consortium (BCAC). In a case-only analysis, we used logistic regression to assess associations between reproductive factors and BC subtype compared to luminal A tumors as a reference. The interaction between age and parity in BC subtype risk was also tested, across all ages and, because age was modeled non-linearly, specifically at ages 35, 55 and 75 years.<h4>Results</h4>Parous women were more likely to be diagnosed with triple negative BC (TNBC) than with luminal A BC, irrespective of age (OR for parity = 1.38, 95% CI 1.16-1.65, p = 0.0004; p for interaction with age = 0.076). Parous women were also more likely to be diagnosed with luminal and non-luminal HER2-like BCs and this effect was slightly more pronounced at an early age (p for interaction with age = 0.037 and 0.030, respectively). For instance, women diagnosed at age 35 were 1.48 (CI 1.01-2.16) more likely to have luminal HER2-like BC than luminal A BC, while this association was not significant at age 75 (OR = 0.72, CI 0.45-1.14). While age at menarche was not significantly associated with BC subtype, increasing age at FFTP was non-linearly associated with TNBC relative to luminal A BC. An age at FFTP of 25 versus 20 years lowered the risk for TNBC (OR = 0.78, CI 0.70-0.88, p < 0.0001), but this effect was not apparent at a later FFTP.<h4>Conclusions</h4>Our main findings suggest that parity is associated with TNBC across all ages at BC diagnosis, whereas the association with luminal HER2-like BC was present only for early onset BC.

Wright, L.B. Schoemaker, M.J. Jones, M.E. Ashworth, A. Swerdlow, A.J (2018) Breast cancer risk in relation to history of preeclampsia and hyperemesis gravidarum: Prospective analysis in the Generations Study.. Show Abstract full text

Preeclampsia and hyperemesis gravidarum are pregnancy complications associated with altered sex hormone levels. Previous studies suggest preeclampsia may be associated with a decreased risk of subsequent breast cancer and hyperemesis with an increased risk, but the evidence remains unclear. We used data from the Generations Study, a large prospective study of women in the United Kingdom, to estimate relative risks of breast cancer in relation to a history of preeclampsia and hyperemesis using Cox regression adjusting for known breast cancer risk factors. During 7.5 years average follow-up of 82,053 parous women, 1,969 were diagnosed with invasive or in situ breast cancer. Women who had experienced preeclampsia during pregnancy had a significantly decreased risk of premenopausal breast cancer (hazard ratio (HR) =0.67, 95% confidence interval (CI): 0.49-0.90) and of HER2-enriched tumours (HR = 0.33, 95% CI: 0.12-0.91), but there was no association with overall (HR = 0.90, 95% CI: 0.80-1.02) or postmenopausal (HR = 0.97, 95% CI: 0.85-1.12) breast cancer risk. Risk reductions among premenopausal women were strongest within 20 years since the last pregnancy with preeclampsia. Hyperemesis was associated with a significantly increased risk of HER2-enriched tumours (HR = 1.76, 95% CI: 1.07-2.87), but not with other intrinsic subtypes or breast cancer risk overall. These results provide evidence that preeclampsia is associated with a decreased risk of premenopausal and HER2-enriched breast cancer and that hyperemesis, although not associated with breast cancer risk overall, may be associated with raised risk of HER2-enriched tumours.

Guo, Y. Warren Andersen, S. Shu, X.-.O. Michailidou, K. Bolla, M.K. Wang, Q. Garcia-Closas, M. Milne, R.L. Schmidt, M.K. Chang-Claude, J. Dunning, A. Bojesen, S.E. Ahsan, H. Aittomäki, K. Andrulis, I.L. Anton-Culver, H. Arndt, V. Beckmann, M.W. Beeghly-Fadiel, A. Benitez, J. Bogdanova, N.V. Bonanni, B. Børresen-Dale, A.-.L. Brand, J. Brauch, H. Brenner, H. Brüning, T. Burwinkel, B. Casey, G. Chenevix-Trench, G. Couch, F.J. Cox, A. Cross, S.S. Czene, K. Devilee, P. Dörk, T. Dumont, M. Fasching, P.A. Figueroa, J. Flesch-Janys, D. Fletcher, O. Flyger, H. Fostira, F. Gammon, M. Giles, G.G. Guénel, P. Haiman, C.A. Hamann, U. Hooning, M.J. Hopper, J.L. Jakubowska, A. Jasmine, F. Jenkins, M. John, E.M. Johnson, N. Jones, M.E. Kabisch, M. Kibriya, M. Knight, J.A. Koppert, L.B. Kosma, V.-.M. Kristensen, V. Le Marchand, L. Lee, E. Li, J. Lindblom, A. Luben, R. Lubinski, J. Malone, K.E. Mannermaa, A. Margolin, S. Marme, F. McLean, C. Meijers-Heijboer, H. Meindl, A. Neuhausen, S.L. Nevanlinna, H. Neven, P. Olson, J.E. Perez, J.I.A. Perkins, B. Peterlongo, P. Phillips, K.-.A. Pylkäs, K. Rudolph, A. Santella, R. Sawyer, E.J. Schmutzler, R.K. Seynaeve, C. Shah, M. Shrubsole, M.J. Southey, M.C. Swerdlow, A.J. Toland, A.E. Tomlinson, I. Torres, D. Truong, T. Ursin, G. Van Der Luijt, R.B. Verhoef, S. Whittemore, A.S. Winqvist, R. Zhao, H. Zhao, S. Hall, P. Simard, J. Kraft, P. Pharoah, P. Hunter, D. Easton, D.F. Zheng, W (2016) Genetically Predicted Body Mass Index and Breast Cancer Risk: Mendelian Randomization Analyses of Data from 145,000 Women of European Descent.. Show Abstract full text

<h4>Background</h4>Observational epidemiological studies have shown that high body mass index (BMI) is associated with a reduced risk of breast cancer in premenopausal women but an increased risk in postmenopausal women. It is unclear whether this association is mediated through shared genetic or environmental factors.<h4>Methods</h4>We applied Mendelian randomization to evaluate the association between BMI and risk of breast cancer occurrence using data from two large breast cancer consortia. We created a weighted BMI genetic score comprising 84 BMI-associated genetic variants to predicted BMI. We evaluated genetically predicted BMI in association with breast cancer risk using individual-level data from the Breast Cancer Association Consortium (BCAC) (cases  =  46,325, controls  =  42,482). We further evaluated the association between genetically predicted BMI and breast cancer risk using summary statistics from 16,003 cases and 41,335 controls from the Discovery, Biology, and Risk of Inherited Variants in Breast Cancer (DRIVE) Project. Because most studies measured BMI after cancer diagnosis, we could not conduct a parallel analysis to adequately evaluate the association of measured BMI with breast cancer risk prospectively.<h4>Results</h4>In the BCAC data, genetically predicted BMI was found to be inversely associated with breast cancer risk (odds ratio [OR]  =  0.65 per 5 kg/m2 increase, 95% confidence interval [CI]: 0.56-0.75, p = 3.32 × 10-10). The associations were similar for both premenopausal (OR   =   0.44, 95% CI:0.31-0.62, p  =  9.91 × 10-8) and postmenopausal breast cancer (OR  =  0.57, 95% CI: 0.46-0.71, p  =  1.88 × 10-8). This association was replicated in the data from the DRIVE consortium (OR  =  0.72, 95% CI: 0.60-0.84, p   =   1.64 × 10-7). Single marker analyses identified 17 of the 84 BMI-associated single nucleotide polymorphisms (SNPs) in association with breast cancer risk at p < 0.05; for 16 of them, the allele associated with elevated BMI was associated with reduced breast cancer risk.<h4>Conclusions</h4>BMI predicted by genome-wide association studies (GWAS)-identified variants is inversely associated with the risk of both pre- and postmenopausal breast cancer. The reduced risk of postmenopausal breast cancer associated with genetically predicted BMI observed in this study differs from the positive association reported from studies using measured adult BMI. Understanding the reasons for this discrepancy may reveal insights into the complex relationship of genetic determinants of body weight in the etiology of breast cancer.

Trabert, B. Poole, E.M. White, E. Visvanathan, K. Adami, H.-.O. Anderson, G.L. Brasky, T.M. Brinton, L.A. Fortner, R.T. Gaudet, M. Hartge, P. Hoffman-Bolton, J. Jones, M. Lacey, J.V. Larsson, S.C. Mackenzie, G.G. Schouten, L.J. Sandler, D.P. O'Brien, K. Patel, A.V. Peters, U. Prizment, A. Robien, K. Setiawan, V.W. Swerdlow, A. van den Brandt, P.A. Weiderpass, E. Wilkens, L.R. Wolk, A. Wentzensen, N. Tworoger, S.S. Ovarian Cancer Cohort Consortium (OC3), (2019) Analgesic Use and Ovarian Cancer Risk: An Analysis in the Ovarian Cancer Cohort Consortium.. Show Abstract full text

<h4>Background</h4>Aspirin use is associated with reduced risk of several cancers. A pooled analysis of 12 case-control studies showed a 10% decrease in ovarian cancer risk with regular aspirin use, which was stronger for daily and low-dose users. To prospectively investigate associations of analgesic use with ovarian cancer, we analyzed data from 13 studies in the Ovarian Cancer Cohort Consortium (OC3).<h4>Methods</h4>The current study included 758 829 women who at study enrollment self-reported analgesic use, among whom 3514 developed ovarian cancer. Using Cox regression, we assessed associations between frequent medication use and risk of ovarian cancer. Dose and duration were also evaluated. All statistical tests were two-sided.<h4>Results</h4>Women who used aspirin almost daily (≥6 days/wk) vs infrequent/nonuse experienced a 10% reduction in ovarian cancer risk (rate ratio [RR] = 0.90, 95% confidence interval [CI] = 0.82 to 1.00, P = .05). Frequent use (≥4 days/wk) of aspirin (RR = 0.95, 95% CI = 0.88 to 1.03), nonaspirin nonsteroidal anti-inflammatory drugs (NSAIDs; RR = 1.00, 95% CI = 0.90 to 1.11), or acetaminophen (RR = 1.05, 95% CI = 0.88 to 1.24) was not associated with risk. Daily acetaminophen use (RR = 1.28, 95% CI = 1.00 to 1.65, P = .05) was associated with elevated ovarian cancer risk. Risk estimates for frequent, long-term (10+ years) use of aspirin (RR = 1.15, 95% CI = 0.98 to 1.34) or nonaspirin NSAIDs (RR = 1.19, 95% CI = 0.84 to 1.68) were modestly elevated, although not statistically significantly so.<h4>Conclusions</h4>This large, prospective analysis suggests that women who use aspirin daily have a slightly lower risk of developing ovarian cancer (∼10% lower than infrequent/nonuse)-similar to the risk reduction observed in case-control analyses. The observed potential elevated risks for 10+ years of frequent aspirin and NSAID use require further study but could be due to confounding by medical indications for use or variation in drug dosing.

Premenopausal Breast Cancer Collaborative Group, . Schoemaker, M.J. Nichols, H.B. Wright, L.B. Brook, M.N. Jones, M.E. O'Brien, K.M. Adami, H.-.O. Baglietto, L. Bernstein, L. Bertrand, K.A. Boutron-Ruault, M.-.C. Braaten, T. Chen, Y. Connor, A.E. Dorronsoro, M. Dossus, L. Eliassen, A.H. Giles, G.G. Hankinson, S.E. Kaaks, R. Key, T.J. Kirsh, V.A. Kitahara, C.M. Koh, W.-.P. Larsson, S.C. Linet, M.S. Ma, H. Masala, G. Merritt, M.A. Milne, R.L. Overvad, K. Ozasa, K. Palmer, J.R. Peeters, P.H. Riboli, E. Rohan, T.E. Sadakane, A. Sund, M. Tamimi, R.M. Trichopoulou, A. Ursin, G. Vatten, L. Visvanathan, K. Weiderpass, E. Willett, W.C. Wolk, A. Yuan, J.-.M. Zeleniuch-Jacquotte, A. Sandler, D.P. Swerdlow, A.J (2018) Association of Body Mass Index and Age With Subsequent Breast Cancer Risk in Premenopausal Women.. Show Abstract full text

<h4>Importance</h4>The association between increasing body mass index (BMI; calculated as weight in kilograms divided by height in meters squared) and risk of breast cancer is unique in cancer epidemiology in that a crossover effect exists, with risk reduction before and risk increase after menopause. The inverse association with premenopausal breast cancer risk is poorly characterized but might be important in the understanding of breast cancer causation.<h4>Objective</h4>To investigate the association of BMI with premenopausal breast cancer risk, in particular by age at BMI, attained age, risk factors for breast cancer, and tumor characteristics.<h4>Design, setting, and participants</h4>This multicenter analysis used pooled individual-level data from 758 592 premenopausal women from 19 prospective cohorts to estimate hazard ratios (HRs) of premenopausal breast cancer in association with BMI from ages 18 through 54 years using Cox proportional hazards regression analysis. Median follow-up was 9.3 years (interquartile range, 4.9-13.5 years) per participant, with 13 082 incident cases of breast cancer. Participants were recruited from January 1, 1963, through December 31, 2013, and data were analyzed from September 1, 2013, through December 31, 2017.<h4>Exposures</h4>Body mass index at ages 18 to 24, 25 to 34, 35 to 44, and 45 to 54 years.<h4>Main outcomes and measures</h4>Invasive or in situ premenopausal breast cancer.<h4>Results</h4>Among the 758 592 premenopausal women (median age, 40.6 years; interquartile range, 35.2-45.5 years) included in the analysis, inverse linear associations of BMI with breast cancer risk were found that were stronger for BMI at ages 18 to 24 years (HR per 5 kg/m2 [5.0-U] difference, 0.77; 95% CI, 0.73-0.80) than for BMI at ages 45 to 54 years (HR per 5.0-U difference, 0.88; 95% CI, 0.86-0.91). The inverse associations were observed even among nonoverweight women. There was a 4.2-fold risk gradient between the highest and lowest BMI categories (BMI≥35.0 vs <17.0) at ages 18 to 24 years (HR, 0.24; 95% CI, 0.14-0.40). Hazard ratios did not appreciably vary by attained age or between strata of other breast cancer risk factors. Associations were stronger for estrogen receptor-positive and/or progesterone receptor-positive than for hormone receptor-negative breast cancer for BMI at every age group (eg, for BMI at age 18 to 24 years: HR per 5.0-U difference for estrogen receptor-positive and progesterone receptor-positive tumors, 0.76 [95% CI, 0.70-0.81] vs hormone receptor-negative tumors, 0.85 [95% CI: 0.76-0.95]); BMI at ages 25 to 54 years was not consistently associated with triple-negative or hormone receptor-negative breast cancer overall.<h4>Conclusions and relevance</h4>The results of this study suggest that increased adiposity is associated with a reduced risk of premenopausal breast cancer at a greater magnitude than previously shown and across the entire distribution of BMI. The strongest associations of risk were observed for BMI in early adulthood. Understanding the biological mechanisms underlying these associations could have important preventive potential.

Schoemaker, M.J. Jones, M.E. Higgins, C.D. Wright, A.F. UK Clinical Cytogenetics Group, . Swerdlow, A.J (2019) Mortality and cancer incidence in carriers of constitutional t(11;22)(q23;q11) translocations: A prospective study.. Show Abstract full text

The constitutional t(11;22)(q23;q11) translocation is the only recurrent non-Robertsonian translocation known in humans. Carriers are phenotypically normal and are usually referred for cytogenetic testing because of multiple miscarriages, infertility, or having aneuploidy in offspring. A breast cancer predisposition has been suggested, but previous studies have been small and had methodological shortcomings. We therefore conducted a long-term prospective study of cancer and mortality risk in carriers. We followed 65 male and 101 female carriers of t(11;22)(q23;q11) diagnosed in cytogenetic laboratories in Britain during 1976-2005 for cancer and deaths for an average of 21.4 years per subject. Standardised mortality (SMR) and incidence (SIR) ratios were calculated comparing the numbers of observed events with those expected from national age-, sex-, country- and calendar-period-specific population rates. Cancer incidence was borderline significantly raised for cancer overall (SIR = 1.56, 95% CI: 0.98-2.36, n = 22), and significantly raised for invasive breast cancer (SIR = 2.74, 95% CI: 1.18-5.40, n = 8) and in situ breast cancer (SIR = 13.0, 95% CI: 3.55-33.4, n = 4). Breast cancer risks were particularly increased at ages <50 (SIR = 4.37, 95% CI: 1.42-10.2 for invasive, SIR = 22.8, 95% CI: 2.76-82.5 for in situ). Mortality was borderline significantly raised for breast cancer (SMR = 4.82, 95% CI: 0.99-14.1) but not significantly raised for other cancers or causes. Individuals diagnosed with t(11;22)(q23;q11) appear to be at several-fold increased breast cancer risk, with the greatest risks at premenopausal ages. Further research is required to understand the genetic mechanism involving 11q23 and 22q11 and there may be a need for enhanced breast cancer surveillance among female carriers.

Schoemaker, M.J. Jones, M.E. Higgins, C.D. Wright, A.F. United Kingdom Clinical Cytogenetics Group, . Swerdlow, A.J (2019) Mortality and Cancer Incidence in Carriers of Balanced Robertsonian Translocations: A National Cohort Study.. Show Abstract full text

A balanced robertsonian translocation (rob) results from fusion of 2 acrocentric chromosomes. Carriers are phenotypically normal and are often diagnosed because of recurrent miscarriages, infertility, or aneuploid offspring. Mortality and site-specific cancer risks in carriers have not been prospectively investigated. We followed 1,987 carriers diagnosed in Great Britain for deaths and cancer risk, over an average of 24.1 years. Standardized mortality and incidence ratios were calculated comparing the number of observed events against population rates. Overall mortality was higher for carriers diagnosed before age 15 years (standardized mortality ratio (SMR) = 2.00, 95% confidence interval (CI): 1.09, 3.35), similar for those diagnosed aged 15-44 years (SMR = 1.06, 95% CI: 0.86-1.28), and lower for those diagnosed aged 45-84 years (SMR = 0.81, 95% CI: 0.68, 0.95). Cancer incidence was higher for non-Hodgkin lymphoma (standardized incidence ratio (SIR) = 1.90, 95% CI: 1.01, 3.24) and childhood leukemia (SIR = 14.5, 95% CI: 1.75, 52.2), the latter particularly in rob(15;21) carriers (SIR = 447.8, 95% CI: 11.3, 2,495). Rob(13;14) carriers had a higher breast cancer risk (SIR = 1.58, 95% CI: 1.12, 2.15). Mortality risks relative to the population in diagnosed carriers depend on age at cytogenetic diagnosis, possibly reflecting age-specific cytogenetic referral reasons. Carriers might be at greater risk of childhood leukemia and non-Hodgkin lymphoma and those diagnosed with rob(13;14) of breast cancer.

Ruth, K.S. Soares, A.L.G. Borges, M.-.C. Eliassen, A.H. Hankinson, S.E. Jones, M.E. Kraft, P. Nichols, H.B. Sandler, D.P. Schoemaker, M.J. Taylor, J.A. Zeleniuch-Jacquotte, A. Lawlor, D.A. Swerdlow, A.J. Murray, A (2019) Genome-wide association study of anti-Müllerian hormone levels in pre-menopausal women of late reproductive age and relationship with genetic determinants of reproductive lifespan.. Show Abstract full text

Anti-Müllerian hormone (AMH) is required for sexual differentiation in the fetus, and in adult females AMH is produced by growing ovarian follicles. Consequently, AMH levels are correlated with ovarian reserve, declining towards menopause when the oocyte pool is exhausted. A previous genome-wide association study identified three genetic variants in and around the AMH gene that explained 25% of variation in AMH levels in adolescent males but did not identify any genetic associations reaching genome-wide significance in adolescent females. To explore the role of genetic variation in determining AMH levels in women of late reproductive age, we carried out a genome-wide meta-analysis in 3344 pre-menopausal women from five cohorts (median age 44-48 years at blood draw). A single genetic variant, rs16991615, previously associated with age at menopause, reached genome-wide significance at P = 3.48 × 10-10, with a per allele difference in age-adjusted inverse normal AMH of 0.26 standard deviations (SD) (95% confidence interval (CI) [0.18,0.34]). We investigated whether genetic determinants of female reproductive lifespan were more generally associated with pre-menopausal AMH levels. Genetically-predicted age at menarche had no robust association but genetically-predicted age at menopause was associated with lower AMH levels by 0.18 SD (95% CI [0.14,0.21]) in age-adjusted inverse normal AMH per one-year earlier age at menopause. Our findings provide genetic support for the well-established use of AMH as a marker of ovarian reserve.

Mavaddat, N. Michailidou, K. Dennis, J. Lush, M. Fachal, L. Lee, A. Tyrer, J.P. Chen, T.-.H. Wang, Q. Bolla, M.K. Yang, X. Adank, M.A. Ahearn, T. Aittomäki, K. Allen, J. Andrulis, I.L. Anton-Culver, H. Antonenkova, N.N. Arndt, V. Aronson, K.J. Auer, P.L. Auvinen, P. Barrdahl, M. Beane Freeman, L.E. Beckmann, M.W. Behrens, S. Benitez, J. Bermisheva, M. Bernstein, L. Blomqvist, C. Bogdanova, N.V. Bojesen, S.E. Bonanni, B. Børresen-Dale, A.-.L. Brauch, H. Bremer, M. Brenner, H. Brentnall, A. Brock, I.W. Brooks-Wilson, A. Brucker, S.Y. Brüning, T. Burwinkel, B. Campa, D. Carter, B.D. Castelao, J.E. Chanock, S.J. Chlebowski, R. Christiansen, H. Clarke, C.L. Collée, J.M. Cordina-Duverger, E. Cornelissen, S. Couch, F.J. Cox, A. Cross, S.S. Czene, K. Daly, M.B. Devilee, P. Dörk, T. Dos-Santos-Silva, I. Dumont, M. Durcan, L. Dwek, M. Eccles, D.M. Ekici, A.B. Eliassen, A.H. Ellberg, C. Engel, C. Eriksson, M. Evans, D.G. Fasching, P.A. Figueroa, J. Fletcher, O. Flyger, H. Försti, A. Fritschi, L. Gabrielson, M. Gago-Dominguez, M. Gapstur, S.M. García-Sáenz, J.A. Gaudet, M.M. Georgoulias, V. Giles, G.G. Gilyazova, I.R. Glendon, G. Goldberg, M.S. Goldgar, D.E. González-Neira, A. Grenaker Alnæs, G.I. Grip, M. Gronwald, J. Grundy, A. Guénel, P. Haeberle, L. Hahnen, E. Haiman, C.A. Håkansson, N. Hamann, U. Hankinson, S.E. Harkness, E.F. Hart, S.N. He, W. Hein, A. Heyworth, J. Hillemanns, P. Hollestelle, A. Hooning, M.J. Hoover, R.N. Hopper, J.L. Howell, A. Huang, G. Humphreys, K. Hunter, D.J. Jakimovska, M. Jakubowska, A. Janni, W. John, E.M. Johnson, N. Jones, M.E. Jukkola-Vuorinen, A. Jung, A. Kaaks, R. Kaczmarek, K. Kataja, V. Keeman, R. Kerin, M.J. Khusnutdinova, E. Kiiski, J.I. Knight, J.A. Ko, Y.-.D. Kosma, V.-.M. Koutros, S. Kristensen, V.N. Krüger, U. Kühl, T. Lambrechts, D. Le Marchand, L. Lee, E. Lejbkowicz, F. Lilyquist, J. Lindblom, A. Lindström, S. Lissowska, J. Lo, W.-.Y. Loibl, S. Long, J. Lubiński, J. Lux, M.P. MacInnis, R.J. Maishman, T. Makalic, E. Maleva Kostovska, I. Mannermaa, A. Manoukian, S. Margolin, S. Martens, J.W.M. Martinez, M.E. Mavroudis, D. McLean, C. Meindl, A. Menon, U. Middha, P. Miller, N. Moreno, F. Mulligan, A.M. Mulot, C. Muñoz-Garzon, V.M. Neuhausen, S.L. Nevanlinna, H. Neven, P. Newman, W.G. Nielsen, S.F. Nordestgaard, B.G. Norman, A. Offit, K. Olson, J.E. Olsson, H. Orr, N. Pankratz, V.S. Park-Simon, T.-.W. Perez, J.I.A. Pérez-Barrios, C. Peterlongo, P. Peto, J. Pinchev, M. Plaseska-Karanfilska, D. Polley, E.C. Prentice, R. Presneau, N. Prokofyeva, D. Purrington, K. Pylkäs, K. Rack, B. Radice, P. Rau-Murthy, R. Rennert, G. Rennert, H.S. Rhenius, V. Robson, M. Romero, A. Ruddy, K.J. Ruebner, M. Saloustros, E. Sandler, D.P. Sawyer, E.J. Schmidt, D.F. Schmutzler, R.K. Schneeweiss, A. Schoemaker, M.J. Schumacher, F. Schürmann, P. Schwentner, L. Scott, C. Scott, R.J. Seynaeve, C. Shah, M. Sherman, M.E. Shrubsole, M.J. Shu, X.-.O. Slager, S. Smeets, A. Sohn, C. Soucy, P. Southey, M.C. Spinelli, J.J. Stegmaier, C. Stone, J. Swerdlow, A.J. Tamimi, R.M. Tapper, W.J. Taylor, J.A. Terry, M.B. Thöne, K. Tollenaar, R.A.E.M. Tomlinson, I. Truong, T. Tzardi, M. Ulmer, H.-.U. Untch, M. Vachon, C.M. van Veen, E.M. Vijai, J. Weinberg, C.R. Wendt, C. Whittemore, A.S. Wildiers, H. Willett, W. Winqvist, R. Wolk, A. Yang, X.R. Yannoukakos, D. Zhang, Y. Zheng, W. Ziogas, A. ABCTB Investigators, . kConFab/AOCS Investigators, . NBCS Collaborators, . Dunning, A.M. Thompson, D.J. Chenevix-Trench, G. Chang-Claude, J. Schmidt, M.K. Hall, P. Milne, R.L. Pharoah, P.D.P. Antoniou, A.C. Chatterjee, N. Kraft, P. García-Closas, M. Simard, J. Easton, D.F (2019) Polygenic Risk Scores for Prediction of Breast Cancer and Breast Cancer Subtypes.. Show Abstract full text

Stratification of women according to their risk of breast cancer based on polygenic risk scores (PRSs) could improve screening and prevention strategies. Our aim was to develop PRSs, optimized for prediction of estrogen receptor (ER)-specific disease, from the largest available genome-wide association dataset and to empirically validate the PRSs in prospective studies. The development dataset comprised 94,075 case subjects and 75,017 control subjects of European ancestry from 69 studies, divided into training and validation sets. Samples were genotyped using genome-wide arrays, and single-nucleotide polymorphisms (SNPs) were selected by stepwise regression or lasso penalized regression. The best performing PRSs were validated in an independent test set comprising 11,428 case subjects and 18,323 control subjects from 10 prospective studies and 190,040 women from UK Biobank (3,215 incident breast cancers). For the best PRSs (313 SNPs), the odds ratio for overall disease per 1 standard deviation in ten prospective studies was 1.61 (95%CI: 1.57-1.65) with area under receiver-operator curve (AUC) = 0.630 (95%CI: 0.628-0.651). The lifetime risk of overall breast cancer in the top centile of the PRSs was 32.6%. Compared with women in the middle quintile, those in the highest 1% of risk had 4.37- and 2.78-fold risks, and those in the lowest 1% of risk had 0.16- and 0.27-fold risks, of developing ER-positive and ER-negative disease, respectively. Goodness-of-fit tests indicated that this PRS was well calibrated and predicts disease risk accurately in the tails of the distribution. This PRS is a powerful and reliable predictor of breast cancer risk that may improve breast cancer prevention programs.

Escala-Garcia, M. Guo, Q. Dörk, T. Canisius, S. Keeman, R. Dennis, J. Beesley, J. Lecarpentier, J. Bolla, M.K. Wang, Q. Abraham, J. Andrulis, I.L. Anton-Culver, H. Arndt, V. Auer, P.L. Beckmann, M.W. Behrens, S. Benitez, J. Bermisheva, M. Bernstein, L. Blomqvist, C. Boeckx, B. Bojesen, S.E. Bonanni, B. Børresen-Dale, A.-.L. Brauch, H. Brenner, H. Brentnall, A. Brinton, L. Broberg, P. Brock, I.W. Brucker, S.Y. Burwinkel, B. Caldas, C. Caldés, T. Campa, D. Canzian, F. Carracedo, A. Carter, B.D. Castelao, J.E. Chang-Claude, J. Chanock, S.J. Chenevix-Trench, G. Cheng, T.-.Y.D. Chin, S.-.F. Clarke, C.L. NBCS Collaborators, . Cordina-Duverger, E. Couch, F.J. Cox, D.G. Cox, A. Cross, S.S. Czene, K. Daly, M.B. Devilee, P. Dunn, J.A. Dunning, A.M. Durcan, L. Dwek, M. Earl, H.M. Ekici, A.B. Eliassen, A.H. Ellberg, C. Engel, C. Eriksson, M. Evans, D.G. Figueroa, J. Flesch-Janys, D. Flyger, H. Gabrielson, M. Gago-Dominguez, M. Galle, E. Gapstur, S.M. García-Closas, M. García-Sáenz, J.A. Gaudet, M.M. George, A. Georgoulias, V. Giles, G.G. Glendon, G. Goldgar, D.E. González-Neira, A. Alnæs, G.I.G. Grip, M. Guénel, P. Haeberle, L. Hahnen, E. Haiman, C.A. Håkansson, N. Hall, P. Hamann, U. Hankinson, S. Harkness, E.F. Harrington, P.A. Hart, S.N. Hartikainen, J.M. Hein, A. Hillemanns, P. Hiller, L. Holleczek, B. Hollestelle, A. Hooning, M.J. Hoover, R.N. Hopper, J.L. Howell, A. Huang, G. Humphreys, K. Hunter, D.J. Janni, W. John, E.M. Jones, M.E. Jukkola-Vuorinen, A. Jung, A. Kaaks, R. Kabisch, M. Kaczmarek, K. Kerin, M.J. Khan, S. Khusnutdinova, E. Kiiski, J.I. Kitahara, C.M. Knight, J.A. Ko, Y.-.D. Koppert, L.B. Kosma, V.-.M. Kraft, P. Kristensen, V.N. Krüger, U. Kühl, T. Lambrechts, D. Le Marchand, L. Lee, E. Lejbkowicz, F. Li, L. Lindblom, A. Lindström, S. Linet, M. Lissowska, J. Lo, W.-.Y. Loibl, S. Lubiński, J. Lux, M.P. MacInnis, R.J. Maierthaler, M. Maishman, T. Makalic, E. Mannermaa, A. Manoochehri, M. Manoukian, S. Margolin, S. Martinez, M.E. Mavroudis, D. McLean, C. Meindl, A. Middha, P. Miller, N. Milne, R.L. Moreno, F. Mulligan, A.M. Mulot, C. Nassir, R. Neuhausen, S.L. Newman, W.T. Nielsen, S.F. Nordestgaard, B.G. Norman, A. Olsson, H. Orr, N. Pankratz, V.S. Park-Simon, T.-.W. Perez, J.I.A. Pérez-Barrios, C. Peterlongo, P. Petridis, C. Pinchev, M. Prajzendanc, K. Prentice, R. Presneau, N. Prokofieva, D. Pylkäs, K. Rack, B. Radice, P. Ramachandran, D. Rennert, G. Rennert, H.S. Rhenius, V. Romero, A. Roylance, R. Saloustros, E. Sawyer, E.J. Schmidt, D.F. Schmutzler, R.K. Schneeweiss, A. Schoemaker, M.J. Schumacher, F. Schwentner, L. Scott, R.J. Scott, C. Seynaeve, C. Shah, M. Simard, J. Smeets, A. Sohn, C. Southey, M.C. Swerdlow, A.J. Talhouk, A. Tamimi, R.M. Tapper, W.J. Teixeira, M.R. Tengström, M. Terry, M.B. Thöne, K. Tollenaar, R.A.E.M. Tomlinson, I. Torres, D. Truong, T. Turman, C. Turnbull, C. Ulmer, H.-.U. Untch, M. Vachon, C. van Asperen, C.J. van den Ouweland, A.M.W. van Veen, E.M. Wendt, C. Whittemore, A.S. Willett, W. Winqvist, R. Wolk, A. Yang, X.R. Zhang, Y. Easton, D.F. Fasching, P.A. Nevanlinna, H. Eccles, D.M. Pharoah, P.D.P. Schmidt, M.K (2019) Genome-wide association study of germline variants and breast cancer-specific mortality.. Show Abstract full text

<h4>Background</h4>We examined the associations between germline variants and breast cancer mortality using a large meta-analysis of women of European ancestry.<h4>Methods</h4>Meta-analyses included summary estimates based on Cox models of twelve datasets using ~10.4 million variants for 96,661 women with breast cancer and 7697 events (breast cancer-specific deaths). Oestrogen receptor (ER)-specific analyses were based on 64,171 ER-positive (4116) and 16,172 ER-negative (2125) patients. We evaluated the probability of a signal to be a true positive using the Bayesian false discovery probability (BFDP).<h4>Results</h4>We did not find any variant associated with breast cancer-specific mortality at P < 5 × 10<sup>-8</sup>. For ER-positive disease, the most significantly associated variant was chr7:rs4717568 (BFDP = 7%, P = 1.28 × 10<sup>-7</sup>, hazard ratio [HR] = 0.88, 95% confidence interval [CI] = 0.84-0.92); the closest gene is AUTS2. For ER-negative disease, the most significant variant was chr7:rs67918676 (BFDP = 11%, P = 1.38 × 10<sup>-7</sup>, HR = 1.27, 95% CI = 1.16-1.39); located within a long intergenic non-coding RNA gene (AC004009.3), close to the HOXA gene cluster.<h4>Conclusions</h4>We uncovered germline variants on chromosome 7 at BFDP < 15% close to genes for which there is biological evidence related to breast cancer outcome. However, the paucity of variants associated with mortality at genome-wide significance underpins the challenge in providing genetic-based individualised prognostic information for breast cancer patients.

Wu, L. Shi, W. Long, J. Guo, X. Michailidou, K. Beesley, J. Bolla, M.K. Shu, X.-.O. Lu, Y. Cai, Q. Al-Ejeh, F. Rozali, E. Wang, Q. Dennis, J. Li, B. Zeng, C. Feng, H. Gusev, A. Barfield, R.T. Andrulis, I.L. Anton-Culver, H. Arndt, V. Aronson, K.J. Auer, P.L. Barrdahl, M. Baynes, C. Beckmann, M.W. Benitez, J. Bermisheva, M. Blomqvist, C. Bogdanova, N.V. Bojesen, S.E. Brauch, H. Brenner, H. Brinton, L. Broberg, P. Brucker, S.Y. Burwinkel, B. Caldés, T. Canzian, F. Carter, B.D. Castelao, J.E. Chang-Claude, J. Chen, X. Cheng, T.-.Y.D. Christiansen, H. Clarke, C.L. NBCS Collaborators, . Collée, M. Cornelissen, S. Couch, F.J. Cox, D. Cox, A. Cross, S.S. Cunningham, J.M. Czene, K. Daly, M.B. Devilee, P. Doheny, K.F. Dörk, T. Dos-Santos-Silva, I. Dumont, M. Dwek, M. Eccles, D.M. Eilber, U. Eliassen, A.H. Engel, C. Eriksson, M. Fachal, L. Fasching, P.A. Figueroa, J. Flesch-Janys, D. Fletcher, O. Flyger, H. Fritschi, L. Gabrielson, M. Gago-Dominguez, M. Gapstur, S.M. García-Closas, M. Gaudet, M.M. Ghoussaini, M. Giles, G.G. Goldberg, M.S. Goldgar, D.E. González-Neira, A. Guénel, P. Hahnen, E. Haiman, C.A. Håkansson, N. Hall, P. Hallberg, E. Hamann, U. Harrington, P. Hein, A. Hicks, B. Hillemanns, P. Hollestelle, A. Hoover, R.N. Hopper, J.L. Huang, G. Humphreys, K. Hunter, D.J. Jakubowska, A. Janni, W. John, E.M. Johnson, N. Jones, K. Jones, M.E. Jung, A. Kaaks, R. Kerin, M.J. Khusnutdinova, E. Kosma, V.-.M. Kristensen, V.N. Lambrechts, D. Le Marchand, L. Li, J. Lindström, S. Lissowska, J. Lo, W.-.Y. Loibl, S. Lubinski, J. Luccarini, C. Lux, M.P. MacInnis, R.J. Maishman, T. Kostovska, I.M. Mannermaa, A. Manson, J.E. Margolin, S. Mavroudis, D. Meijers-Heijboer, H. Meindl, A. Menon, U. Meyer, J. Mulligan, A.M. Neuhausen, S.L. Nevanlinna, H. Neven, P. Nielsen, S.F. Nordestgaard, B.G. Olopade, O.I. Olson, J.E. Olsson, H. Peterlongo, P. Peto, J. Plaseska-Karanfilska, D. Prentice, R. Presneau, N. Pylkäs, K. Rack, B. Radice, P. Rahman, N. Rennert, G. Rennert, H.S. Rhenius, V. Romero, A. Romm, J. Rudolph, A. Saloustros, E. Sandler, D.P. Sawyer, E.J. Schmidt, M.K. Schmutzler, R.K. Schneeweiss, A. Scott, R.J. Scott, C.G. Seal, S. Shah, M. Shrubsole, M.J. Smeets, A. Southey, M.C. Spinelli, J.J. Stone, J. Surowy, H. Swerdlow, A.J. Tamimi, R.M. Tapper, W. Taylor, J.A. Terry, M.B. Tessier, D.C. Thomas, A. Thöne, K. Tollenaar, R.A.E.M. Torres, D. Truong, T. Untch, M. Vachon, C. Van Den Berg, D. Vincent, D. Waisfisz, Q. Weinberg, C.R. Wendt, C. Whittemore, A.S. Wildiers, H. Willett, W.C. Winqvist, R. Wolk, A. Xia, L. Yang, X.R. Ziogas, A. Ziv, E. kConFab/AOCS Investigators, . Dunning, A.M. Pharoah, P.D.P. Simard, J. Milne, R.L. Edwards, S.L. Kraft, P. Easton, D.F. Chenevix-Trench, G. Zheng, W (2018) A transcriptome-wide association study of 229,000 women identifies new candidate susceptibility genes for breast cancer.. Show Abstract full text

The breast cancer risk variants identified in genome-wide association studies explain only a small fraction of the familial relative risk, and the genes responsible for these associations remain largely unknown. To identify novel risk loci and likely causal genes, we performed a transcriptome-wide association study evaluating associations of genetically predicted gene expression with breast cancer risk in 122,977 cases and 105,974 controls of European ancestry. We used data from the Genotype-Tissue Expression Project to establish genetic models to predict gene expression in breast tissue and evaluated model performance using data from The Cancer Genome Atlas. Of the 8,597 genes evaluated, significant associations were identified for 48 at a Bonferroni-corrected threshold of P < 5.82 × 10<sup>-6</sup>, including 14 genes at loci not yet reported for breast cancer. We silenced 13 genes and showed an effect for 11 on cell proliferation and/or colony-forming efficiency. Our study provides new insights into breast cancer genetics and biology.

Williams, C.L. Jones, M.E. Swerdlow, A.J. Botting, B.J. Davies, M.C. Jacobs, I. Bunch, K.J. Murphy, M.F.G. Sutcliffe, A.G (2018) Risks of ovarian, breast, and corpus uteri cancer in women treated with assisted reproductive technology in Great Britain, 1991-2010: data linkage study including 2.2 million person years of observation.. Show Abstract full text

<h4>Objective</h4>To investigate the risks of ovarian, breast, and corpus uteri cancer in women who have had assisted reproduction.<h4>Design</h4>Large, population based, data linkage cohort study.<h4>Setting and participants</h4>All women who had assisted reproduction in Great Britain, 1991-2010, as recorded by the Human Fertilisation and Embryology Authority (HFEA).<h4>Interventions</h4>HFEA fertility records for cohort members were linked to national cancer registrations.<h4>Main outcome measures</h4>Observed first diagnosis of ovarian, breast, and corpus uteri cancer in cohort members were compared with age, sex, and period specific expectation. Standardised incidence ratios (SIRs) were calculated by use of age, sex, and period specific national incidence rates.<h4>Results</h4>255 786 women contributed 2 257 789 person years' follow-up. No significant increased risk of corpus uteri cancer (164 cancers observed <i>v</i> 146.9 cancers expected; SIR 1.12, 95% confidence interval 0.95 to 1.30) was found during an average of 8.8 years' follow-up. This study found no significantly increased risks of breast cancer overall (2578 <i>v</i> 2641.2; SIR 0.98, 0.94 to 1.01) or invasive breast cancer (2272 <i>v</i> 2371.4; SIR 0.96, 0.92 to 1.00). An increased risk of in situ breast cancer (291 <i>v</i> 253.5; SIR 1.15, 1.02 to 1.29; absolute excess risk (AER) 1.7 cases per 100 000 person years, 95% confidence interval 0.2 to 3.2) was detected, associated with an increasing number of treatment cycles (P=0.03). There was an increased risk of ovarian cancer (405 <i>v</i> 291.82; SIR 1.39, 1.26 to 1.53; AER 5.0 cases per 100 000 person years, 3.3 to 6.9), both invasive (264 <i>v</i> 188.1; SIR 1.40, 1.24 to 1.58; AER 3.4 cases per 100 000 person years, 2.0 to 4.9) and borderline (141 <i>v</i> 103.7; SIR 1.36, 1.15 to 1.60; AER 1.7 cases per 100 000 person years, 0.7 to 2.8). Increased risks of ovarian tumours were limited to women with endometriosis, low parity, or both. This study found no increased risk of any ovarian tumour in women treated because of only male factor or unexplained infertility.<h4>Conclusions</h4>No increased risk of corpus uteri or invasive breast cancer was detected in women who had had assisted reproduction, but increased risks of in situ breast cancer and invasive and borderline ovarian tumours were found in this study. Our results suggest that ovarian tumour risks could be due to patient characteristics, rather than assisted reproduction itself, although both surveillance bias and the effect of treatment are also possibilities. Ongoing monitoring of this population is essential.

McCullough, M.L. Zoltick, E.S. Weinstein, S.J. Fedirko, V. Wang, M. Cook, N.R. Eliassen, A.H. Zeleniuch-Jacquotte, A. Agnoli, C. Albanes, D. Barnett, M.J. Buring, J.E. Campbell, P.T. Clendenen, T.V. Freedman, N.D. Gapstur, S.M. Giovannucci, E.L. Goodman, G.G. Haiman, C.A. Ho, G.Y.F. Horst, R.L. Hou, T. Huang, W.-.Y. Jenab, M. Jones, M.E. Joshu, C.E. Krogh, V. Lee, I.-.M. Lee, J.E. Männistö, S. Le Marchand, L. Mondul, A.M. Neuhouser, M.L. Platz, E.A. Purdue, M.P. Riboli, E. Robsahm, T.E. Rohan, T.E. Sasazuki, S. Schoemaker, M.J. Sieri, S. Stampfer, M.J. Swerdlow, A.J. Thomson, C.A. Tretli, S. Tsugane, S. Ursin, G. Visvanathan, K. White, K.K. Wu, K. Yaun, S.-.S. Zhang, X. Willett, W.C. Gail, M.H. Ziegler, R.G. Smith-Warner, S.A (2019) Circulating Vitamin D and Colorectal Cancer Risk: An International Pooling Project of 17 Cohorts.. Show Abstract full text

<h4>Background</h4>Experimental and epidemiological studies suggest a protective role for vitamin D in colorectal carcinogenesis, but evidence is inconclusive. Circulating 25-hydroxyvitamin D (25(OH)D) concentrations that minimize risk are unknown. Current Institute of Medicine (IOM) vitamin D guidance is based solely on bone health.<h4>Methods</h4>We pooled participant-level data from 17 cohorts, comprising 5706 colorectal cancer case participants and 7107 control participants with a wide range of circulating 25(OH)D concentrations. For 30.1% of participants, 25(OH)D was newly measured. Previously measured 25(OH)D was calibrated to the same assay to permit estimating risk by absolute concentrations. Study-specific relative risks (RRs) for prediagnostic season-standardized 25(OH)D concentrations were calculated using conditional logistic regression and pooled using random effects models.<h4>Results</h4>Compared with the lower range of sufficiency for bone health (50-<62.5 nmol/L), deficient 25(OH)D (<30 nmol/L) was associated with 31% higher colorectal cancer risk (RR = 1.31, 95% confidence interval [CI] = 1.05 to 1.62); 25(OH)D above sufficiency (75-<87.5 and 87.5-<100 nmol/L) was associated with 19% (RR = 0.81, 95% CI = 0.67 to 0.99) and 27% (RR = 0.73, 95% CI = 0.59 to 0.91) lower risk, respectively. At 25(OH)D of 100 nmol/L or greater, risk did not continue to decline and was not statistically significantly reduced (RR = 0.91, 95% CI = 0.67 to 1.24, 3.5% of control participants). Associations were minimally affected when adjusting for body mass index, physical activity, or other risk factors. For each 25 nmol/L increment in circulating 25(OH)D, colorectal cancer risk was 19% lower in women (RR = 0.81, 95% CI = 0.75 to 0.87) and 7% lower in men (RR = 0.93, 95% CI = 0.86 to 1.00) (two-sided Pheterogeneity by sex = .008). Associations were inverse in all subgroups, including colorectal subsite, geographic region, and season of blood collection.<h4>Conclusions</h4>Higher circulating 25(OH)D was related to a statistically significant, substantially lower colorectal cancer risk in women and non-statistically significant lower risk in men. Optimal 25(OH)D concentrations for colorectal cancer risk reduction, 75-100 nmol/L, appear higher than current IOM recommendations.

Horne, H.N. Oh, H. Sherman, M.E. Palakal, M. Hewitt, S.M. Schmidt, M.K. Milne, R.L. Hardisson, D. Benitez, J. Blomqvist, C. Bolla, M.K. Brenner, H. Chang-Claude, J. Cora, R. Couch, F.J. Cuk, K. Devilee, P. Easton, D.F. Eccles, D.M. Eilber, U. Hartikainen, J.M. Heikkilä, P. Holleczek, B. Hooning, M.J. Jones, M. Keeman, R. Mannermaa, A. Martens, J.W.M. Muranen, T.A. Nevanlinna, H. Olson, J.E. Orr, N. Perez, J.I.A. Pharoah, P.D.P. Ruddy, K.J. Saum, K.-.U. Schoemaker, M.J. Seynaeve, C. Sironen, R. Smit, V.T.H.B.M. Swerdlow, A.J. Tengström, M. Thomas, A.S. Timmermans, A.M. Tollenaar, R.A.E.M. Troester, M.A. van Asperen, C.J. van Deurzen, C.H.M. Van Leeuwen, F.F. Van't Veer, L.J. García-Closas, M. Figueroa, J.D (2018) E-cadherin breast tumor expression, risk factors and survival: Pooled analysis of 5,933 cases from 12 studies in the Breast Cancer Association Consortium.. Show Abstract full text

E-cadherin (CDH1) is a putative tumor suppressor gene implicated in breast carcinogenesis. Yet, whether risk factors or survival differ by E-cadherin tumor expression is unclear. We evaluated E-cadherin tumor immunohistochemistry expression using tissue microarrays of 5,933 female invasive breast cancers from 12 studies from the Breast Cancer Consortium. H-scores were calculated and case-case odds ratios (OR) and 95% confidence intervals (CIs) were estimated using logistic regression. Survival analyses were performed using Cox regression models. All analyses were stratified by estrogen receptor (ER) status and histologic subtype. E-cadherin low cases (N = 1191, 20%) were more frequently of lobular histology, low grade, >2 cm, and HER2-negative. Loss of E-cadherin expression (score < 100) was associated with menopausal hormone use among ER-positive tumors (ever compared to never users, OR = 1.24, 95% CI = 0.97-1.59), which was stronger when we evaluated complete loss of E-cadherin (i.e. H-score = 0), OR = 1.57, 95% CI = 1.06-2.33. Breast cancer specific mortality was unrelated to E-cadherin expression in multivariable models. E-cadherin low expression is associated with lobular histology, tumor characteristics and menopausal hormone use, with no evidence of an association with breast cancer specific survival. These data support loss of E-cadherin expression as an important marker of tumor subtypes.

Schoemaker, M.J. Jones, M.E. Allen, S. Hoare, J. Ashworth, A. Dowsett, M. Swerdlow, A.J (2017) Childhood body size and pubertal timing in relation to adult mammographic density phenotype.. Show Abstract full text

<h4>Background</h4>An earlier age at onset of breast development and longer time between pubertal stages has been implicated in breast cancer risk. It is not clear whether associations of breast cancer risk with puberty or predictors of onset of puberty, such as weight and height, are mediated via mammographic density, an important risk factor for breast cancer.<h4>Methods</h4>We investigated whether childhood body size and pubertal timing and tempo, collected by questionnaire, are associated with percentage and absolute area mammographic density at ages 47-73 years in 1105 women recruited to a prospective study.<h4>Results</h4>After controlling for adult adiposity, weight at ages 7 and 11 years was strongly significantly inversely associated with percentage and absolute dense area (p trend <0.001), and positively associated with absolute non-dense area. Greater height at age 7, but not age 11, was associated with lower percentage density (p trend = 0.016). Later age at menarche and age at when regular periods were established was associated with increased density, but additional adjustment for childhood weight attenuated the association. A longer interval between thelarche and menarche, and between thelarche and regular periods, was associated with increased dense area, even after adjusting for childhood weight (p trend = 0.013 and 0.028, respectively), and was independent of age at pubertal onset.<h4>Conclusions</h4>Greater prepubertal weight and earlier pubertal onset are associated with lower adult breast density, but age at pubertal onset does not appear to have an independent effect on adult density after controlling for childhood adiposity. A possible effect of pubertal tempo on density needs further investigation.

Jiao, X. Aravidis, C. Marikkannu, R. Rantala, J. Picelli, S. Adamovic, T. Liu, T. Maguire, P. Kremeyer, B. Luo, L. von Holst, S. Kontham, V. Thutkawkorapin, J. Margolin, S. Du, Q. Lundin, J. Michailidou, K. Bolla, M.K. Wang, Q. Dennis, J. Lush, M. Ambrosone, C.B. Andrulis, I.L. Anton-Culver, H. Antonenkova, N.N. Arndt, V. Beckmann, M.W. Blomqvist, C. Blot, W. Boeckx, B. Bojesen, S.E. Bonanni, B. Brand, J.S. Brauch, H. Brenner, H. Broeks, A. Brüning, T. Burwinkel, B. Cai, Q. Chang-Claude, J. NBCS Collaborators, . Couch, F.J. Cox, A. Cross, S.S. Deming-Halverson, S.L. Devilee, P. Dos-Santos-Silva, I. Dörk, T. Eriksson, M. Fasching, P.A. Figueroa, J. Flesch-Janys, D. Flyger, H. Gabrielson, M. García-Closas, M. Giles, G.G. González-Neira, A. Guénel, P. Guo, Q. Gündert, M. Haiman, C.A. Hallberg, E. Hamann, U. Harrington, P. Hooning, M.J. Hopper, J.L. Huang, G. Jakubowska, A. Jones, M.E. Kerin, M.J. Kosma, V.-.M. Kristensen, V.N. Lambrechts, D. Le Marchand, L. Lubinski, J. Mannermaa, A. Martens, J.W.M. Meindl, A. Milne, R.L. Mulligan, A.M. Neuhausen, S.L. Nevanlinna, H. Peto, J. Pylkäs, K. Radice, P. Rhenius, V. Sawyer, E.J. Schmidt, M.K. Schmutzler, R.K. Seynaeve, C. Shah, M. Simard, J. Southey, M.C. Swerdlow, A.J. Truong, T. Wendt, C. Winqvist, R. Zheng, W. kConFab/AOCS Investigators, . Benitez, J. Dunning, A.M. Pharoah, P.D.P. Easton, D.F. Czene, K. Hall, P. Lindblom, A (2017) <i>PHIP</i> - a novel candidate breast cancer susceptibility locus on 6q14.1.. Show Abstract full text

Most non-<i>BRCA1/2</i> breast cancer families have no identified genetic cause. We used linkage and haplotype analyses in familial and sporadic breast cancer cases to identify a susceptibility locus on chromosome 6q. Two independent genome-wide linkage analysis studies suggested a 3 Mb locus on chromosome 6q and two unrelated Swedish families with a LOD >2 together seemed to share a haplotype in 6q14.1. We hypothesized that this region harbored a rare high-risk founder allele contributing to breast cancer in these two families. Sequencing of DNA and RNA from the two families did not detect any pathogenic mutations. Finally, 29 SNPs in the region were analyzed in 44,214 cases and 43,532 controls from BCAC, and the original haplotypes in the two families were suggested as low-risk alleles for European and Swedish women specifically. There was also some support for one additional independent moderate-risk allele in Swedish familial samples. The results were consistent with our previous findings in familial breast cancer and supported a breast cancer susceptibility locus at 6q14.1 around the <i>PHIP</i> gene.

Brewer, H.R. Jones, M.E. Schoemaker, M.J. Ashworth, A. Swerdlow, A.J (2017) Family history and risk of breast cancer: an analysis accounting for family structure.. Show Abstract full text

<h4>Purpose</h4>Family history is an important risk factor for breast cancer incidence, but the parameters conventionally used to categorize it are based solely on numbers and/or ages of breast cancer cases in the family and take no account of the size and age-structure of the woman's family.<h4>Methods</h4>Using data from the Generations Study, a cohort of over 113,000 women from the general UK population, we analyzed breast cancer risk in relation to first-degree family history using a family history score (FHS) that takes account of the expected number of family cases based on the family's age-structure and national cancer incidence rates.<h4>Results</h4>Breast cancer risk increased significantly (P <sub>trend</sub> < 0.0001) with greater FHS. There was a 3.5-fold (95% CI 2.56-4.79) range of risk between the lowest and highest FHS groups, whereas women who had two or more relatives with breast cancer, the strongest conventional familial risk factor, had a 2.5-fold (95% CI 1.83-3.47) increase in risk. Using likelihood ratio tests, the best model for determining breast cancer risk due to family history was that combining FHS and age of relative at diagnosis.<h4>Conclusions</h4>A family history score based on expected as well as observed breast cancers in a family can give greater risk discrimination on breast cancer incidence than conventional parameters based solely on cases in affected relatives. Our modeling suggests that a yet stronger predictor of risk might be a combination of this score and age at diagnosis in relatives.

Zeng, C. Guo, X. Long, J. Kuchenbaecker, K.B. Droit, A. Michailidou, K. Ghoussaini, M. Kar, S. Freeman, A. Hopper, J.L. Milne, R.L. Bolla, M.K. Wang, Q. Dennis, J. Agata, S. Ahmed, S. Aittomäki, K. Andrulis, I.L. Anton-Culver, H. Antonenkova, N.N. Arason, A. Arndt, V. Arun, B.K. Arver, B. Bacot, F. Barrowdale, D. Baynes, C. Beeghly-Fadiel, A. Benitez, J. Bermisheva, M. Blomqvist, C. Blot, W.J. Bogdanova, N.V. Bojesen, S.E. Bonanni, B. Borresen-Dale, A.-.L. Brand, J.S. Brauch, H. Brennan, P. Brenner, H. Broeks, A. Brüning, T. Burwinkel, B. Buys, S.S. Cai, Q. Caldes, T. Campbell, I. Carpenter, J. Chang-Claude, J. Choi, J.-.Y. Claes, K.B.M. Clarke, C. Cox, A. Cross, S.S. Czene, K. Daly, M.B. de la Hoya, M. De Leeneer, K. Devilee, P. Diez, O. Domchek, S.M. Doody, M. Dorfling, C.M. Dörk, T. Dos-Santos-Silva, I. Dumont, M. Dwek, M. Dworniczak, B. Egan, K. Eilber, U. Einbeigi, Z. Ejlertsen, B. Ellis, S. Frost, D. Lalloo, F. EMBRACE, . Fasching, P.A. Figueroa, J. Flyger, H. Friedlander, M. Friedman, E. Gambino, G. Gao, Y.-.T. Garber, J. García-Closas, M. Gehrig, A. Damiola, F. Lesueur, F. Mazoyer, S. Stoppa-Lyonnet, D. behalf of GEMO Study Collaborators, . Giles, G.G. Godwin, A.K. Goldgar, D.E. González-Neira, A. Greene, M.H. Guénel, P. Haeberle, L. Haiman, C.A. Hallberg, E. Hamann, U. Hansen, T.V.O. Hart, S. Hartikainen, J.M. Hartman, M. Hassan, N. Healey, S. Hogervorst, F.B.L. Verhoef, S. HEBON, . Hendricks, C.B. Hillemanns, P. Hollestelle, A. Hulick, P.J. Hunter, D.J. Imyanitov, E.N. Isaacs, C. Ito, H. Jakubowska, A. Janavicius, R. Jaworska-Bieniek, K. Jensen, U.B. John, E.M. Joly Beauparlant, C. Jones, M. Kabisch, M. Kang, D. Karlan, B.Y. Kauppila, S. Kerin, M.J. Khan, S. Khusnutdinova, E. Knight, J.A. Konstantopoulou, I. Kraft, P. Kwong, A. Laitman, Y. Lambrechts, D. Lazaro, C. Le Marchand, L. Lee, C.N. Lee, M.H. Lester, J. Li, J. Liljegren, A. Lindblom, A. Lophatananon, A. Lubinski, J. Mai, P.L. Mannermaa, A. Manoukian, S. Margolin, S. Marme, F. Matsuo, K. McGuffog, L. Meindl, A. Menegaux, F. Montagna, M. Muir, K. Mulligan, A.M. Nathanson, K.L. Neuhausen, S.L. Nevanlinna, H. Newcomb, P.A. Nord, S. Nussbaum, R.L. Offit, K. Olah, E. Olopade, O.I. Olswold, C. Osorio, A. Papi, L. Park-Simon, T.-.W. Paulsson-Karlsson, Y. Peeters, S. Peissel, B. Peterlongo, P. Peto, J. Pfeiler, G. Phelan, C.M. Presneau, N. Radice, P. Rahman, N. Ramus, S.J. Rashid, M.U. Rennert, G. Rhiem, K. Rudolph, A. Salani, R. Sangrajrang, S. Sawyer, E.J. Schmidt, M.K. Schmutzler, R.K. Schoemaker, M.J. Schürmann, P. Seynaeve, C. Shen, C.-.Y. Shrubsole, M.J. Shu, X.-.O. Sigurdson, A. Singer, C.F. Slager, S. Soucy, P. Southey, M. Steinemann, D. Swerdlow, A. Szabo, C.I. Tchatchou, S. Teixeira, M.R. Teo, S.H. Terry, M.B. Tessier, D.C. Teulé, A. Thomassen, M. Tihomirova, L. Tischkowitz, M. Toland, A.E. Tung, N. Turnbull, C. van den Ouweland, A.M.W. van Rensburg, E.J. Ven den Berg, D. Vijai, J. Wang-Gohrke, S. Weitzel, J.N. Whittemore, A.S. Winqvist, R. Wong, T.Y. Wu, A.H. Yannoukakos, D. Yu, J.-.C. Pharoah, P.D.P. Hall, P. Chenevix-Trench, G. KConFab, . AOCS Investigators, . Dunning, A.M. Simard, J. Couch, F.J. Antoniou, A.C. Easton, D.F. Zheng, W (2016) Identification of independent association signals and putative functional variants for breast cancer risk through fine-scale mapping of the 12p11 locus.. Show Abstract full text

<h4>Background</h4>Multiple recent genome-wide association studies (GWAS) have identified a single nucleotide polymorphism (SNP), rs10771399, at 12p11 that is associated with breast cancer risk.<h4>Method</h4>We performed a fine-scale mapping study of a 700 kb region including 441 genotyped and more than 1300 imputed genetic variants in 48,155 cases and 43,612 controls of European descent, 6269 cases and 6624 controls of East Asian descent and 1116 cases and 932 controls of African descent in the Breast Cancer Association Consortium (BCAC; http://bcac.ccge.medschl.cam.ac.uk/ ), and in 15,252 BRCA1 mutation carriers in the Consortium of Investigators of Modifiers of BRCA1/2 (CIMBA). Stepwise regression analyses were performed to identify independent association signals. Data from the Encyclopedia of DNA Elements project (ENCODE) and the Cancer Genome Atlas (TCGA) were used for functional annotation.<h4>Results</h4>Analysis of data from European descendants found evidence for four independent association signals at 12p11, represented by rs7297051 (odds ratio (OR) = 1.09, 95 % confidence interval (CI) = 1.06-1.12; P = 3 × 10(-9)), rs805510 (OR = 1.08, 95 % CI = 1.04-1.12, P = 2 × 10(-5)), and rs1871152 (OR = 1.04, 95 % CI = 1.02-1.06; P = 2 × 10(-4)) identified in the general populations, and rs113824616 (P = 7 × 10(-5)) identified in the meta-analysis of BCAC ER-negative cases and BRCA1 mutation carriers. SNPs rs7297051, rs805510 and rs113824616 were also associated with breast cancer risk at P < 0.05 in East Asians, but none of the associations were statistically significant in African descendants. Multiple candidate functional variants are located in putative enhancer sequences. Chromatin interaction data suggested that PTHLH was the likely target gene of these enhancers. Of the six variants with the strongest evidence of potential functionality, rs11049453 was statistically significantly associated with the expression of PTHLH and its nearby gene CCDC91 at P < 0.05.<h4>Conclusion</h4>This study identified four independent association signals at 12p11 and revealed potentially functional variants, providing additional insights into the underlying biological mechanism(s) for the association observed between variants at 12p11 and breast cancer risk.

Wentzensen, N. Poole, E.M. Trabert, B. White, E. Arslan, A.A. Patel, A.V. Setiawan, V.W. Visvanathan, K. Weiderpass, E. Adami, H.-.O. Black, A. Bernstein, L. Brinton, L.A. Buring, J. Butler, L.M. Chamosa, S. Clendenen, T.V. Dossus, L. Fortner, R. Gapstur, S.M. Gaudet, M.M. Gram, I.T. Hartge, P. Hoffman-Bolton, J. Idahl, A. Jones, M. Kaaks, R. Kirsh, V. Koh, W.-.P. Lacey, J.V. Lee, I.-.M. Lundin, E. Merritt, M.A. Onland-Moret, N.C. Peters, U. Poynter, J.N. Rinaldi, S. Robien, K. Rohan, T. Sandler, D.P. Schairer, C. Schouten, L.J. Sjöholm, L.K. Sieri, S. Swerdlow, A. Tjonneland, A. Travis, R. Trichopoulou, A. van den Brandt, P.A. Wilkens, L. Wolk, A. Yang, H.P. Zeleniuch-Jacquotte, A. Tworoger, S.S (2016) Ovarian Cancer Risk Factors by Histologic Subtype: An Analysis From the Ovarian Cancer Cohort Consortium.. Show Abstract full text

<h4>Purpose</h4>An understanding of the etiologic heterogeneity of ovarian cancer is important for improving prevention, early detection, and therapeutic approaches. We evaluated 14 hormonal, reproductive, and lifestyle factors by histologic subtype in the Ovarian Cancer Cohort Consortium (OC3).<h4>Patients and methods</h4>Among 1.3 million women from 21 studies, 5,584 invasive epithelial ovarian cancers were identified (3,378 serous, 606 endometrioid, 331 mucinous, 269 clear cell, 1,000 other). By using competing-risks Cox proportional hazards regression stratified by study and birth year and adjusted for age, parity, and oral contraceptive use, we assessed associations for all invasive cancers by histology. Heterogeneity was evaluated by likelihood ratio test.<h4>Results</h4>Most risk factors exhibited significant heterogeneity by histology. Higher parity was most strongly associated with endometrioid (relative risk [RR] per birth, 0.78; 95% CI, 0.74 to 0.83) and clear cell (RR, 0.68; 95% CI, 0.61 to 0.76) carcinomas (P value for heterogeneity [P-het] < .001). Similarly, age at menopause, endometriosis, and tubal ligation were only associated with endometrioid and clear cell tumors (P-het ≤ .01). Family history of breast cancer (P-het = .008) had modest heterogeneity. Smoking was associated with an increased risk of mucinous (RR per 20 pack-years, 1.26; 95% CI, 1.08 to 1.46) but a decreased risk of clear cell (RR, 0.72; 95% CI, 0.55 to 0.94) tumors (P-het = .004). Unsupervised clustering by risk factors separated endometrioid, clear cell, and low-grade serous carcinomas from high-grade serous and mucinous carcinomas.<h4>Conclusion</h4>The heterogeneous associations of risk factors with ovarian cancer subtypes emphasize the importance of conducting etiologic studies by ovarian cancer subtypes. Most established risk factors were more strongly associated with nonserous carcinomas, which demonstrate challenges for risk prediction of serous cancers, the most fatal subtype.

Southey, M.C. Goldgar, D.E. Winqvist, R. Pylkäs, K. Couch, F. Tischkowitz, M. Foulkes, W.D. Dennis, J. Michailidou, K. van Rensburg, E.J. Heikkinen, T. Nevanlinna, H. Hopper, J.L. Dörk, T. Claes, K.B. Reis-Filho, J. Teo, Z.L. Radice, P. Catucci, I. Peterlongo, P. Tsimiklis, H. Odefrey, F.A. Dowty, J.G. Schmidt, M.K. Broeks, A. Hogervorst, F.B. Verhoef, S. Carpenter, J. Clarke, C. Scott, R.J. Fasching, P.A. Haeberle, L. Ekici, A.B. Beckmann, M.W. Peto, J. Dos-Santos-Silva, I. Fletcher, O. Johnson, N. Bolla, M.K. Sawyer, E.J. Tomlinson, I. Kerin, M.J. Miller, N. Marme, F. Burwinkel, B. Yang, R. Guénel, P. Truong, T. Menegaux, F. Sanchez, M. Bojesen, S. Nielsen, S.F. Flyger, H. Benitez, J. Zamora, M.P. Perez, J.I.A. Menéndez, P. Anton-Culver, H. Neuhausen, S. Ziogas, A. Clarke, C.A. Brenner, H. Arndt, V. Stegmaier, C. Brauch, H. Brüning, T. Ko, Y.-.D. Muranen, T.A. Aittomäki, K. Blomqvist, C. Bogdanova, N.V. Antonenkova, N.N. Lindblom, A. Margolin, S. Mannermaa, A. Kataja, V. Kosma, V.-.M. Hartikainen, J.M. Spurdle, A.B. Investigators, K. Australian Ovarian Cancer Study Group, . Wauters, E. Smeets, D. Beuselinck, B. Floris, G. Chang-Claude, J. Rudolph, A. Seibold, P. Flesch-Janys, D. Olson, J.E. Vachon, C. Pankratz, V.S. McLean, C. Haiman, C.A. Henderson, B.E. Schumacher, F. Le Marchand, L. Kristensen, V. Alnæs, G.G. Zheng, W. Hunter, D.J. Lindstrom, S. Hankinson, S.E. Kraft, P. Andrulis, I. Knight, J.A. Glendon, G. Mulligan, A.M. Jukkola-Vuorinen, A. Grip, M. Kauppila, S. Devilee, P. Tollenaar, R.A.E.M. Seynaeve, C. Hollestelle, A. Garcia-Closas, M. Figueroa, J. Chanock, S.J. Lissowska, J. Czene, K. Darabi, H. Eriksson, M. Eccles, D.M. Rafiq, S. Tapper, W.J. Gerty, S.M. Hooning, M.J. Martens, J.W.M. Collée, J.M. Tilanus-Linthorst, M. Hall, P. Li, J. Brand, J.S. Humphreys, K. Cox, A. Reed, M.W.R. Luccarini, C. Baynes, C. Dunning, A.M. Hamann, U. Torres, D. Ulmer, H.U. Rüdiger, T. Jakubowska, A. Lubinski, J. Jaworska, K. Durda, K. Slager, S. Toland, A.E. Ambrosone, C.B. Yannoukakos, D. Swerdlow, A. Ashworth, A. Orr, N. Jones, M. González-Neira, A. Pita, G. Alonso, M.R. Álvarez, N. Herrero, D. Tessier, D.C. Vincent, D. Bacot, F. Simard, J. Dumont, M. Soucy, P. Eeles, R. Muir, K. Wiklund, F. Gronberg, H. Schleutker, J. Nordestgaard, B.G. Weischer, M. Travis, R.C. Neal, D. Donovan, J.L. Hamdy, F.C. Khaw, K.-.T. Stanford, J.L. Blot, W.J. Thibodeau, S. Schaid, D.J. Kelley, J.L. Maier, C. Kibel, A.S. Cybulski, C. Cannon-Albright, L. Butterbach, K. Park, J. Kaneva, R. Batra, J. Teixeira, M.R. Kote-Jarai, Z. Olama, A.A.A. Benlloch, S. Renner, S.P. Hartmann, A. Hein, A. Ruebner, M. Lambrechts, D. Van Nieuwenhuysen, E. Vergote, I. Lambretchs, S. Doherty, J.A. Rossing, M.A. Nickels, S. Eilber, U. Wang-Gohrke, S. Odunsi, K. Sucheston-Campbell, L.E. Friel, G. Lurie, G. Killeen, J.L. Wilkens, L.R. Goodman, M.T. Runnebaum, I. Hillemanns, P.A. Pelttari, L.M. Butzow, R. Modugno, F. Edwards, R.P. Ness, R.B. Moysich, K.B. du Bois, A. Heitz, F. Harter, P. Kommoss, S. Karlan, B.Y. Walsh, C. Lester, J. Jensen, A. Kjaer, S.K. Høgdall, E. Peissel, B. Bonanni, B. Bernard, L. Goode, E.L. Fridley, B.L. Vierkant, R.A. Cunningham, J.M. Larson, M.C. Fogarty, Z.C. Kalli, K.R. Liang, D. Lu, K.H. Hildebrandt, M.A.T. Wu, X. Levine, D.A. Dao, F. Bisogna, M. Berchuck, A. Iversen, E.S. Marks, J.R. Akushevich, L. Cramer, D.W. Schildkraut, J. Terry, K.L. Poole, E.M. Stampfer, M. Tworoger, S.S. Bandera, E.V. Orlow, I. Olson, S.H. Bjorge, L. Salvesen, H.B. van Altena, A.M. Aben, K.K.H. Kiemeney, L.A. Massuger, L.F.A.G. Pejovic, T. Bean, Y. Brooks-Wilson, A. Kelemen, L.E. Cook, L.S. Le, N.D. Górski, B. Gronwald, J. Menkiszak, J. Høgdall, C.K. Lundvall, L. Nedergaard, L. Engelholm, S.A. Dicks, E. Tyrer, J. Campbell, I. McNeish, I. Paul, J. Siddiqui, N. Glasspool, R. Whittemore, A.S. Rothstein, J.H. McGuire, V. Sieh, W. Cai, H. Shu, X.-.O. Teten, R.T. Sutphen, R. McLaughlin, J.R. Narod, S.A. Phelan, C.M. Monteiro, A.N. Fenstermacher, D. Lin, H.-.Y. Permuth, J.B. Sellers, T.A. Chen, Y.A. Tsai, Y.-.Y. Chen, Z. Gentry-Maharaj, A. Gayther, S.A. Ramus, S.J. Menon, U. Wu, A.H. Pearce, C.L. Van Den Berg, D. Pike, M.C. Dansonka-Mieszkowska, A. Plisiecka-Halasa, J. Moes-Sosnowska, J. Kupryjanczyk, J. Pharoah, P.D. Song, H. Winship, I. Chenevix-Trench, G. Giles, G.G. Tavtigian, S.V. Easton, D.F. Milne, R.L (2016) PALB2, CHEK2 and ATM rare variants and cancer risk: data from COGS.. Show Abstract full text

<h4>Background</h4>The rarity of mutations in PALB2, CHEK2 and ATM make it difficult to estimate precisely associated cancer risks. Population-based family studies have provided evidence that at least some of these mutations are associated with breast cancer risk as high as those associated with rare BRCA2 mutations. We aimed to estimate the relative risks associated with specific rare variants in PALB2, CHEK2 and ATM via a multicentre case-control study.<h4>Methods</h4>We genotyped 10 rare mutations using the custom iCOGS array: PALB2 c.1592delT, c.2816T>G and c.3113G>A, CHEK2 c.349A>G, c.538C>T, c.715G>A, c.1036C>T, c.1312G>T, and c.1343T>G and ATM c.7271T>G. We assessed associations with breast cancer risk (42 671 cases and 42 164 controls), as well as prostate (22 301 cases and 22 320 controls) and ovarian (14 542 cases and 23 491 controls) cancer risk, for each variant.<h4>Results</h4>For European women, strong evidence of association with breast cancer risk was observed for PALB2 c.1592delT OR 3.44 (95% CI 1.39 to 8.52, p=7.1×10<sup>-5</sup>), PALB2 c.3113G>A OR 4.21 (95% CI 1.84 to 9.60, p=6.9×10<sup>-8</sup>) and ATM c.7271T>G OR 11.0 (95% CI 1.42 to 85.7, p=0.0012). We also found evidence of association with breast cancer risk for three variants in CHEK2, c.349A>G OR 2.26 (95% CI 1.29 to 3.95), c.1036C>T OR 5.06 (95% CI 1.09 to 23.5) and c.538C>T OR 1.33 (95% CI 1.05 to 1.67) (p≤0.017). Evidence for prostate cancer risk was observed for CHEK2 c.1343T>G OR 3.03 (95% CI 1.53 to 6.03, p=0.0006) for African men and CHEK2 c.1312G>T OR 2.21 (95% CI 1.06 to 4.63, p=0.030) for European men. No evidence of association with ovarian cancer was found for any of these variants.<h4>Conclusions</h4>This report adds to accumulating evidence that at least some variants in these genes are associated with an increased risk of breast cancer that is clinically important.

Schoemaker, M.J. Jones, M.E. Wright, L.B. Griffin, J. McFadden, E. Ashworth, A. Swerdlow, A.J (2016) Psychological stress, adverse life events and breast cancer incidence: a cohort investigation in 106,000 women in the United Kingdom.. Show Abstract full text

<h4>Background</h4>Women diagnosed with breast cancer frequently attribute their cancer to psychological stress, but scientific evidence is inconclusive. We investigated whether experienced frequency of stress and adverse life events affect subsequent breast cancer risk.<h4>Methods</h4>Breast cancer incidence was analysed with respect to stress variables collected at enrolment in a prospective cohort study of 106,000 women in the United Kingdom, with 1783 incident breast cancer cases. Relative risks (RR) were obtained as hazard ratios using Cox proportional hazards models.<h4>Results</h4>There was no association of breast cancer risk overall with experienced frequency of stress. Risk was reduced for death of a close relative during the 5 years preceding study entry (RR = 0.87, 95 % confidence interval (CI): 0.78-0.97), but not for death of a spouse/partner or close friend, personal illness/injury, or divorce/separation. There was a positive association of divorce with oestrogen-receptor-negative (RR = 1.54, 95 % CI: 1.01-2.34), but not with oestrogen-receptor-positive breast cancer. Risk was raised in women who were under age 20 at the death of their mother (RR = 1.31, 95 % CI: 1.02-1.67), but not of their father, and the effect was attenuated after excluding mothers with breast or ovarian cancer (RR = 1.17, 95 % CI: 0.85-1.61).<h4>Conclusions</h4>This large prospective study did not show consistent evidence for an association of breast cancer risk with perceived stress levels or adverse life events in the preceding 5 years, or loss of parents during childhood and adolescence.

Meeks, H.D. Song, H. Michailidou, K. Bolla, M.K. Dennis, J. Wang, Q. Barrowdale, D. Frost, D. EMBRACE, . McGuffog, L. Ellis, S. Feng, B. Buys, S.S. Hopper, J.L. Southey, M.C. Tesoriero, A. kConFab Investigators, . James, P.A. Bruinsma, F. Campbell, I.G. Australia Ovarian Cancer Study Group, . Broeks, A. Schmidt, M.K. Hogervorst, F.B.L. HEBON, . Beckman, M.W. Fasching, P.A. Fletcher, O. Johnson, N. Sawyer, E.J. Riboli, E. Banerjee, S. Menon, U. Tomlinson, I. Burwinkel, B. Hamann, U. Marme, F. Rudolph, A. Janavicius, R. Tihomirova, L. Tung, N. Garber, J. Cramer, D. Terry, K.L. Poole, E.M. Tworoger, S.S. Dorfling, C.M. van Rensburg, E.J. Godwin, A.K. Guénel, P. Truong, T. GEMO Study Collaborators, . Stoppa-Lyonnet, D. Damiola, F. Mazoyer, S. Sinilnikova, O.M. Isaacs, C. Maugard, C. Bojesen, S.E. Flyger, H. Gerdes, A.-.M. Hansen, T.V.O. Jensen, A. Kjaer, S.K. Hogdall, C. Hogdall, E. Pedersen, I.S. Thomassen, M. Benitez, J. González-Neira, A. Osorio, A. Hoya, M.D.L. Segura, P.P. Diez, O. Lazaro, C. Brunet, J. Anton-Culver, H. Eunjung, L. John, E.M. Neuhausen, S.L. Ding, Y.C. Castillo, D. Weitzel, J.N. Ganz, P.A. Nussbaum, R.L. Chan, S.B. Karlan, B.Y. Lester, J. Wu, A. Gayther, S. Ramus, S.J. Sieh, W. Whittermore, A.S. Monteiro, A.N.A. Phelan, C.M. Terry, M.B. Piedmonte, M. Offit, K. Robson, M. Levine, D. Moysich, K.B. Cannioto, R. Olson, S.H. Daly, M.B. Nathanson, K.L. Domchek, S.M. Lu, K.H. Liang, D. Hildebrant, M.A.T. Ness, R. Modugno, F. Pearce, L. Goodman, M.T. Thompson, P.J. Brenner, H. Butterbach, K. Meindl, A. Hahnen, E. Wappenschmidt, B. Brauch, H. Brüning, T. Blomqvist, C. Khan, S. Nevanlinna, H. Pelttari, L.M. Aittomäki, K. Butzow, R. Bogdanova, N.V. Dörk, T. Lindblom, A. Margolin, S. Rantala, J. Kosma, V.-.M. Mannermaa, A. Lambrechts, D. Neven, P. Claes, K.B.M. Maerken, T.V. Chang-Claude, J. Flesch-Janys, D. Heitz, F. Varon-Mateeva, R. Peterlongo, P. Radice, P. Viel, A. Barile, M. Peissel, B. Manoukian, S. Montagna, M. Oliani, C. Peixoto, A. Teixeira, M.R. Collavoli, A. Hallberg, E. Olson, J.E. Goode, E.L. Hart, S.N. Shimelis, H. Cunningham, J.M. Giles, G.G. Milne, R.L. Healey, S. Tucker, K. Haiman, C.A. Henderson, B.E. Goldberg, M.S. Tischkowitz, M. Simard, J. Soucy, P. Eccles, D.M. Le, N. Borresen-Dale, A.-.L. Kristensen, V. Salvesen, H.B. Bjorge, L. Bandera, E.V. Risch, H. Zheng, W. Beeghly-Fadiel, A. Cai, H. Pylkäs, K. Tollenaar, R.A.E.M. Ouweland, A.M.W.V.D. Andrulis, I.L. Knight, J.A. OCGN, . Narod, S. Devilee, P. Winqvist, R. Figueroa, J. Greene, M.H. Mai, P.L. Loud, J.T. García-Closas, M. Schoemaker, M.J. Czene, K. Darabi, H. McNeish, I. Siddiquil, N. Glasspool, R. Kwong, A. Park, S.K. Teo, S.H. Yoon, S.-.Y. Matsuo, K. Hosono, S. Woo, Y.L. Gao, Y.-.T. Foretova, L. Singer, C.F. Rappaport-Feurhauser, C. Friedman, E. Laitman, Y. Rennert, G. Imyanitov, E.N. Hulick, P.J. Olopade, O.I. Senter, L. Olah, E. Doherty, J.A. Schildkraut, J. Koppert, L.B. Kiemeney, L.A. Massuger, L.F.A.G. Cook, L.S. Pejovic, T. Li, J. Borg, A. Öfverholm, A. Rossing, M.A. Wentzensen, N. Henriksson, K. Cox, A. Cross, S.S. Pasini, B.J. Shah, M. Kabisch, M. Torres, D. Jakubowska, A. Lubinski, J. Gronwald, J. Agnarsson, B.A. Kupryjanczyk, J. Moes-Sosnowska, J. Fostira, F. Konstantopoulou, I. Slager, S. Jones, M. PRostate cancer AssoCiation group To Investigate Cancer Associated aLterations in the genome, . Antoniou, A.C. Berchuck, A. Swerdlow, A. Chenevix-Trench, G. Dunning, A.M. Pharoah, P.D.P. Hall, P. Easton, D.F. Couch, F.J. Spurdle, A.B. Goldgar, D.E (2016) BRCA2 Polymorphic Stop Codon K3326X and the Risk of Breast, Prostate, and Ovarian Cancers.. Show Abstract full text

<h4>Background</h4>The K3326X variant in BRCA2 (BRCA2*c.9976A>T; p.Lys3326*; rs11571833) has been found to be associated with small increased risks of breast cancer. However, it is not clear to what extent linkage disequilibrium with fully pathogenic mutations might account for this association. There is scant information about the effect of K3326X in other hormone-related cancers.<h4>Methods</h4>Using weighted logistic regression, we analyzed data from the large iCOGS study including 76 637 cancer case patients and 83 796 control patients to estimate odds ratios (ORw) and 95% confidence intervals (CIs) for K3326X variant carriers in relation to breast, ovarian, and prostate cancer risks, with weights defined as probability of not having a pathogenic BRCA2 variant. Using Cox proportional hazards modeling, we also examined the associations of K3326X with breast and ovarian cancer risks among 7183 BRCA1 variant carriers. All statistical tests were two-sided.<h4>Results</h4>The K3326X variant was associated with breast (ORw = 1.28, 95% CI = 1.17 to 1.40, P = 5.9x10(-) (6)) and invasive ovarian cancer (ORw = 1.26, 95% CI = 1.10 to 1.43, P = 3.8x10(-3)). These associations were stronger for serous ovarian cancer and for estrogen receptor-negative breast cancer (ORw = 1.46, 95% CI = 1.2 to 1.70, P = 3.4x10(-5) and ORw = 1.50, 95% CI = 1.28 to 1.76, P = 4.1x10(-5), respectively). For BRCA1 mutation carriers, there was a statistically significant inverse association of the K3326X variant with risk of ovarian cancer (HR = 0.43, 95% CI = 0.22 to 0.84, P = .013) but no association with breast cancer. No association with prostate cancer was observed.<h4>Conclusions</h4>Our study provides evidence that the K3326X variant is associated with risk of developing breast and ovarian cancers independent of other pathogenic variants in BRCA2. Further studies are needed to determine the biological mechanism of action responsible for these associations.

Lei, J. Rudolph, A. Moysich, K.B. Behrens, S. Goode, E.L. Bolla, M.K. Dennis, J. Dunning, A.M. Easton, D.F. Wang, Q. Benitez, J. Hopper, J.L. Southey, M.C. Schmidt, M.K. Broeks, A. Fasching, P.A. Haeberle, L. Peto, J. Dos-Santos-Silva, I. Sawyer, E.J. Tomlinson, I. Burwinkel, B. Marmé, F. Guénel, P. Truong, T. Bojesen, S.E. Flyger, H. Nielsen, S.F. Nordestgaard, B.G. González-Neira, A. Menéndez, P. Anton-Culver, H. Neuhausen, S.L. Brenner, H. Arndt, V. Meindl, A. Schmutzler, R.K. Brauch, H. Hamann, U. Nevanlinna, H. Fagerholm, R. Dörk, T. Bogdanova, N.V. Mannermaa, A. Hartikainen, J.M. Australian Ovarian Study Group, . kConFab Investigators, . Van Dijck, L. Smeets, A. Flesch-Janys, D. Eilber, U. Radice, P. Peterlongo, P. Couch, F.J. Hallberg, E. Giles, G.G. Milne, R.L. Haiman, C.A. Schumacher, F. Simard, J. Goldberg, M.S. Kristensen, V. Borresen-Dale, A.-.L. Zheng, W. Beeghly-Fadiel, A. Winqvist, R. Grip, M. Andrulis, I.L. Glendon, G. García-Closas, M. Figueroa, J. Czene, K. Brand, J.S. Darabi, H. Eriksson, M. Hall, P. Li, J. Cox, A. Cross, S.S. Pharoah, P.D.P. Shah, M. Kabisch, M. Torres, D. Jakubowska, A. Lubinski, J. Ademuyiwa, F. Ambrosone, C.B. Swerdlow, A. Jones, M. Chang-Claude, J (2016) Genetic variation in the immunosuppression pathway genes and breast cancer susceptibility: a pooled analysis of 42,510 cases and 40,577 controls from the Breast Cancer Association Consortium.. Show Abstract full text

Immunosuppression plays a pivotal role in assisting tumors to evade immune destruction and promoting tumor development. We hypothesized that genetic variation in the immunosuppression pathway genes may be implicated in breast cancer tumorigenesis. We included 42,510 female breast cancer cases and 40,577 controls of European ancestry from 37 studies in the Breast Cancer Association Consortium (2015) with available genotype data for 3595 single nucleotide polymorphisms (SNPs) in 133 candidate genes. Associations between genotyped SNPs and overall breast cancer risk, and secondarily according to estrogen receptor (ER) status, were assessed using multiple logistic regression models. Gene-level associations were assessed based on principal component analysis. Gene expression analyses were conducted using RNA sequencing level 3 data from The Cancer Genome Atlas for 989 breast tumor samples and 113 matched normal tissue samples. SNP rs1905339 (A>G) in the STAT3 region was associated with an increased breast cancer risk (per allele odds ratio 1.05, 95 % confidence interval 1.03-1.08; p value = 1.4 × 10(-6)). The association did not differ significantly by ER status. On the gene level, in addition to TGFBR2 and CCND1, IL5 and GM-CSF showed the strongest associations with overall breast cancer risk (p value = 1.0 × 10(-3) and 7.0 × 10(-3), respectively). Furthermore, STAT3 and IL5 but not GM-CSF were differentially expressed between breast tumor tissue and normal tissue (p value = 2.5 × 10(-3), 4.5 × 10(-4) and 0.63, respectively). Our data provide evidence that the immunosuppression pathway genes STAT3, IL5, and GM-CSF may be novel susceptibility loci for breast cancer in women of European ancestry.

Kar, S.P. Beesley, J. Amin Al Olama, A. Michailidou, K. Tyrer, J. Kote-Jarai, Z. Lawrenson, K. Lindstrom, S. Ramus, S.J. Thompson, D.J. ABCTB Investigators, . Kibel, A.S. Dansonka-Mieszkowska, A. Michael, A. Dieffenbach, A.K. Gentry-Maharaj, A. Whittemore, A.S. Wolk, A. Monteiro, A. Peixoto, A. Kierzek, A. Cox, A. Rudolph, A. Gonzalez-Neira, A. Wu, A.H. Lindblom, A. Swerdlow, A. AOCS Study Group & Australian Cancer Study (Ovarian Cancer), . APCB BioResource, . Ziogas, A. Ekici, A.B. Burwinkel, B. Karlan, B.Y. Nordestgaard, B.G. Blomqvist, C. Phelan, C. McLean, C. Pearce, C.L. Vachon, C. Cybulski, C. Slavov, C. Stegmaier, C. Maier, C. Ambrosone, C.B. Høgdall, C.K. Teerlink, C.C. Kang, D. Tessier, D.C. Schaid, D.J. Stram, D.O. Cramer, D.W. Neal, D.E. Eccles, D. Flesch-Janys, D. Edwards, D.R.V. Wokozorczyk, D. Levine, D.A. Yannoukakos, D. Sawyer, E.J. Bandera, E.V. Poole, E.M. Goode, E.L. Khusnutdinova, E. Høgdall, E. Song, F. Bruinsma, F. Heitz, F. Modugno, F. Hamdy, F.C. Wiklund, F. Giles, G.G. Olsson, H. Wildiers, H. Ulmer, H.-.U. Pandha, H. Risch, H.A. Darabi, H. Salvesen, H.B. Nevanlinna, H. Gronberg, H. Brenner, H. Brauch, H. Anton-Culver, H. Song, H. Lim, H.-.Y. McNeish, I. Campbell, I. Vergote, I. Gronwald, J. Lubiński, J. Stanford, J.L. Benítez, J. Doherty, J.A. Permuth, J.B. Chang-Claude, J. Donovan, J.L. Dennis, J. Schildkraut, J.M. Schleutker, J. Hopper, J.L. Kupryjanczyk, J. Park, J.Y. Figueroa, J. Clements, J.A. Knight, J.A. Peto, J. Cunningham, J.M. Pow-Sang, J. Batra, J. Czene, K. Lu, K.H. Herkommer, K. Khaw, K.-.T. kConFab Investigators, . Matsuo, K. Muir, K. Offitt, K. Chen, K. Moysich, K.B. Aittomäki, K. Odunsi, K. Kiemeney, L.A. Massuger, L.F.A.G. Fitzgerald, L.M. Cook, L.S. Cannon-Albright, L. Hooning, M.J. Pike, M.C. Bolla, M.K. Luedeke, M. Teixeira, M.R. Goodman, M.T. Schmidt, M.K. Riggan, M. Aly, M. Rossing, M.A. Beckmann, M.W. Moisse, M. Sanderson, M. Southey, M.C. Jones, M. Lush, M. Hildebrandt, M.A.T. Hou, M.-.F. Schoemaker, M.J. Garcia-Closas, M. Bogdanova, N. Rahman, N. NBCS Investigators, . Le, N.D. Orr, N. Wentzensen, N. Pashayan, N. Peterlongo, P. Guénel, P. Brennan, P. Paulo, P. Webb, P.M. Broberg, P. Fasching, P.A. Devilee, P. Wang, Q. Cai, Q. Li, Q. Kaneva, R. Butzow, R. Kopperud, R.K. Schmutzler, R.K. Stephenson, R.A. MacInnis, R.J. Hoover, R.N. Winqvist, R. Ness, R. Milne, R.L. Travis, R.C. Benlloch, S. Olson, S.H. McDonnell, S.K. Tworoger, S.S. Maia, S. Berndt, S. Lee, S.C. Teo, S.-.H. Thibodeau, S.N. Bojesen, S.E. Gapstur, S.M. Kjær, S.K. Pejovic, T. Tammela, T.L.J. GENICA Network, . PRACTICAL consortium, . Dörk, T. Brüning, T. Wahlfors, T. Key, T.J. Edwards, T.L. Menon, U. Hamann, U. Mitev, V. Kosma, V.-.M. Setiawan, V.W. Kristensen, V. Arndt, V. Vogel, W. Zheng, W. Sieh, W. Blot, W.J. Kluzniak, W. Shu, X.-.O. Gao, Y.-.T. Schumacher, F. Freedman, M.L. Berchuck, A. Dunning, A.M. Simard, J. Haiman, C.A. Spurdle, A. Sellers, T.A. Hunter, D.J. Henderson, B.E. Kraft, P. Chanock, S.J. Couch, F.J. Hall, P. Gayther, S.A. Easton, D.F. Chenevix-Trench, G. Eeles, R. Pharoah, P.D.P. Lambrechts, D (2016) Genome-Wide Meta-Analyses of Breast, Ovarian, and Prostate Cancer Association Studies Identify Multiple New Susceptibility Loci Shared by at Least Two Cancer Types.. Show Abstract full text

<h4>Unlabelled</h4>Breast, ovarian, and prostate cancers are hormone-related and may have a shared genetic basis, but this has not been investigated systematically by genome-wide association (GWA) studies. Meta-analyses combining the largest GWA meta-analysis data sets for these cancers totaling 112,349 cases and 116,421 controls of European ancestry, all together and in pairs, identified at P < 10(-8) seven new cross-cancer loci: three associated with susceptibility to all three cancers (rs17041869/2q13/BCL2L11; rs7937840/11q12/INCENP; rs1469713/19p13/GATAD2A), two breast and ovarian cancer risk loci (rs200182588/9q31/SMC2; rs8037137/15q26/RCCD1), and two breast and prostate cancer risk loci (rs5013329/1p34/NSUN4; rs9375701/6q23/L3MBTL3). Index variants in five additional regions previously associated with only one cancer also showed clear association with a second cancer type. Cell-type-specific expression quantitative trait locus and enhancer-gene interaction annotations suggested target genes with potential cross-cancer roles at the new loci. Pathway analysis revealed significant enrichment of death receptor signaling genes near loci with P < 10(-5) in the three-cancer meta-analysis.<h4>Significance</h4>We demonstrate that combining large-scale GWA meta-analysis findings across cancer types can identify completely new risk loci common to breast, ovarian, and prostate cancers. We show that the identification of such cross-cancer risk loci has the potential to shed new light on the shared biology underlying these hormone-related cancers. Cancer Discov; 6(9); 1052-67. ©2016 AACR.This article is highlighted in the In This Issue feature, p. 932.

Hamdi, Y. Soucy, P. Adoue, V. Michailidou, K. Canisius, S. Lemaçon, A. Droit, A. Andrulis, I.L. Anton-Culver, H. Arndt, V. Baynes, C. Blomqvist, C. Bogdanova, N.V. Bojesen, S.E. Bolla, M.K. Bonanni, B. Borresen-Dale, A.-.L. Brand, J.S. Brauch, H. Brenner, H. Broeks, A. Burwinkel, B. Chang-Claude, J. NBCS Collaborators, . Couch, F.J. Cox, A. Cross, S.S. Czene, K. Darabi, H. Dennis, J. Devilee, P. Dörk, T. Dos-Santos-Silva, I. Eriksson, M. Fasching, P.A. Figueroa, J. Flyger, H. García-Closas, M. Giles, G.G. Goldberg, M.S. González-Neira, A. Grenaker-Alnæs, G. Guénel, P. Haeberle, L. Haiman, C.A. Hamann, U. Hallberg, E. Hooning, M.J. Hopper, J.L. Jakubowska, A. Jones, M. Kabisch, M. Kataja, V. Lambrechts, D. Le Marchand, L. Lindblom, A. Lubinski, J. Mannermaa, A. Maranian, M. Margolin, S. Marme, F. Milne, R.L. Neuhausen, S.L. Nevanlinna, H. Neven, P. Olswold, C. Peto, J. Plaseska-Karanfilska, D. Pylkäs, K. Radice, P. Rudolph, A. Sawyer, E.J. Schmidt, M.K. Shu, X.-.O. Southey, M.C. Swerdlow, A. Tollenaar, R.A.E.M. Tomlinson, I. Torres, D. Truong, T. Vachon, C. Van Den Ouweland, A.M.W. Wang, Q. Winqvist, R. kConFab/AOCS Investigators, . Zheng, W. Benitez, J. Chenevix-Trench, G. Dunning, A.M. Pharoah, P.D.P. Kristensen, V. Hall, P. Easton, D.F. Pastinen, T. Nord, S. Simard, J (2016) Association of breast cancer risk with genetic variants showing differential allelic expression: Identification of a novel breast cancer susceptibility locus at 4q21.. Show Abstract full text

There are significant inter-individual differences in the levels of gene expression. Through modulation of gene expression, cis-acting variants represent an important source of phenotypic variation. Consequently, cis-regulatory SNPs associated with differential allelic expression are functional candidates for further investigation as disease-causing variants. To investigate whether common variants associated with differential allelic expression were involved in breast cancer susceptibility, a list of genes was established on the basis of their involvement in cancer related pathways and/or mechanisms. Thereafter, using data from a genome-wide map of allelic expression associated SNPs, 313 genetic variants were selected and their association with breast cancer risk was then evaluated in 46,451 breast cancer cases and 42,599 controls of European ancestry ascertained from 41 studies participating in the Breast Cancer Association Consortium. The associations were evaluated with overall breast cancer risk and with estrogen receptor negative and positive disease. One novel breast cancer susceptibility locus on 4q21 (rs11099601) was identified (OR = 1.05, P = 5.6x10-6). rs11099601 lies in a 135 kb linkage disequilibrium block containing several genes, including, HELQ, encoding the protein HEL308 a DNA dependant ATPase and DNA Helicase involved in DNA repair, MRPS18C encoding the Mitochondrial Ribosomal Protein S18C and FAM175A (ABRAXAS), encoding a BRCA1 BRCT domain-interacting protein involved in DNA damage response and double-strand break (DSB) repair. Expression QTL analysis in breast cancer tissue showed rs11099601 to be associated with HELQ (P = 8.28x10-14), MRPS18C (P = 1.94x10-27) and FAM175A (P = 3.83x10-3), explaining about 20%, 14% and 1%, respectively of the variance inexpression of these genes in breast carcinomas.

Easton, D.F. Lesueur, F. Decker, B. Michailidou, K. Li, J. Allen, J. Luccarini, C. Pooley, K.A. Shah, M. Bolla, M.K. Wang, Q. Dennis, J. Ahmad, J. Thompson, E.R. Damiola, F. Pertesi, M. Voegele, C. Mebirouk, N. Robinot, N. Durand, G. Forey, N. Luben, R.N. Ahmed, S. Aittomäki, K. Anton-Culver, H. Arndt, V. Australian Ovarian Cancer Study Group, . Baynes, C. Beckman, M.W. Benitez, J. Van Den Berg, D. Blot, W.J. Bogdanova, N.V. Bojesen, S.E. Brenner, H. Chang-Claude, J. Chia, K.S. Choi, J.-.Y. Conroy, D.M. Cox, A. Cross, S.S. Czene, K. Darabi, H. Devilee, P. Eriksson, M. Fasching, P.A. Figueroa, J. Flyger, H. Fostira, F. García-Closas, M. Giles, G.G. Glendon, G. González-Neira, A. Guénel, P. Haiman, C.A. Hall, P. Hart, S.N. Hartman, M. Hooning, M.J. Hsiung, C.-.N. Ito, H. Jakubowska, A. James, P.A. John, E.M. Johnson, N. Jones, M. Kabisch, M. Kang, D. kConFab Investigators, . Kosma, V.-.M. Kristensen, V. Lambrechts, D. Li, N. Lifepool Investigators, . Lindblom, A. Long, J. Lophatananon, A. Lubinski, J. Mannermaa, A. Manoukian, S. Margolin, S. Matsuo, K. Meindl, A. Mitchell, G. Muir, K. NBCS Investigators, . Nevelsteen, I. van den Ouweland, A. Peterlongo, P. Phuah, S.Y. Pylkäs, K. Rowley, S.M. Sangrajrang, S. Schmutzler, R.K. Shen, C.-.Y. Shu, X.-.O. Southey, M.C. Surowy, H. Swerdlow, A. Teo, S.H. Tollenaar, R.A.E.M. Tomlinson, I. Torres, D. Truong, T. Vachon, C. Verhoef, S. Wong-Brown, M. Zheng, W. Zheng, Y. Nevanlinna, H. Scott, R.J. Andrulis, I.L. Wu, A.H. Hopper, J.L. Couch, F.J. Winqvist, R. Burwinkel, B. Sawyer, E.J. Schmidt, M.K. Rudolph, A. Dörk, T. Brauch, H. Hamann, U. Neuhausen, S.L. Milne, R.L. Fletcher, O. Pharoah, P.D.P. Campbell, I.G. Dunning, A.M. Le Calvez-Kelm, F. Goldgar, D.E. Tavtigian, S.V. Chenevix-Trench, G (2016) No evidence that protein truncating variants in BRIP1 are associated with breast cancer risk: implications for gene panel testing.. Show Abstract full text

<h4>Background</h4>BRCA1 interacting protein C-terminal helicase 1 (BRIP1) is one of the Fanconi Anaemia Complementation (FANC) group family of DNA repair proteins. Biallelic mutations in BRIP1 are responsible for FANC group J, and previous studies have also suggested that rare protein truncating variants in BRIP1 are associated with an increased risk of breast cancer. These studies have led to inclusion of BRIP1 on targeted sequencing panels for breast cancer risk prediction.<h4>Methods</h4>We evaluated a truncating variant, p.Arg798Ter (rs137852986), and 10 missense variants of BRIP1, in 48 144 cases and 43 607 controls of European origin, drawn from 41 studies participating in the Breast Cancer Association Consortium (BCAC). Additionally, we sequenced the coding regions of BRIP1 in 13 213 cases and 5242 controls from the UK, 1313 cases and 1123 controls from three population-based studies as part of the Breast Cancer Family Registry, and 1853 familial cases and 2001 controls from Australia.<h4>Results</h4>The rare truncating allele of rs137852986 was observed in 23 cases and 18 controls in Europeans in BCAC (OR 1.09, 95% CI 0.58 to 2.03, p=0.79). Truncating variants were found in the sequencing studies in 34 cases (0.21%) and 19 controls (0.23%) (combined OR 0.90, 95% CI 0.48 to 1.70, p=0.75).<h4>Conclusions</h4>These results suggest that truncating variants in BRIP1, and in particular p.Arg798Ter, are not associated with a substantial increase in breast cancer risk. Such observations have important implications for the reporting of results from breast cancer screening panels.

Couch, F.J. Kuchenbaecker, K.B. Michailidou, K. Mendoza-Fandino, G.A. Nord, S. Lilyquist, J. Olswold, C. Hallberg, E. Agata, S. Ahsan, H. Aittomäki, K. Ambrosone, C. Andrulis, I.L. Anton-Culver, H. Arndt, V. Arun, B.K. Arver, B. Barile, M. Barkardottir, R.B. Barrowdale, D. Beckmann, L. Beckmann, M.W. Benitez, J. Blank, S.V. Blomqvist, C. Bogdanova, N.V. Bojesen, S.E. Bolla, M.K. Bonanni, B. Brauch, H. Brenner, H. Burwinkel, B. Buys, S.S. Caldes, T. Caligo, M.A. Canzian, F. Carpenter, J. Chang-Claude, J. Chanock, S.J. Chung, W.K. Claes, K.B.M. Cox, A. Cross, S.S. Cunningham, J.M. Czene, K. Daly, M.B. Damiola, F. Darabi, H. de la Hoya, M. Devilee, P. Diez, O. Ding, Y.C. Dolcetti, R. Domchek, S.M. Dorfling, C.M. Dos-Santos-Silva, I. Dumont, M. Dunning, A.M. Eccles, D.M. Ehrencrona, H. Ekici, A.B. Eliassen, H. Ellis, S. Fasching, P.A. Figueroa, J. Flesch-Janys, D. Försti, A. Fostira, F. Foulkes, W.D. Friebel, T. Friedman, E. Frost, D. Gabrielson, M. Gammon, M.D. Ganz, P.A. Gapstur, S.M. Garber, J. Gaudet, M.M. Gayther, S.A. Gerdes, A.-.M. Ghoussaini, M. Giles, G.G. Glendon, G. Godwin, A.K. Goldberg, M.S. Goldgar, D.E. González-Neira, A. Greene, M.H. Gronwald, J. Guénel, P. Gunter, M. Haeberle, L. Haiman, C.A. Hamann, U. Hansen, T.V.O. Hart, S. Healey, S. Heikkinen, T. Henderson, B.E. Herzog, J. Hogervorst, F.B.L. Hollestelle, A. Hooning, M.J. Hoover, R.N. Hopper, J.L. Humphreys, K. Hunter, D.J. Huzarski, T. Imyanitov, E.N. Isaacs, C. Jakubowska, A. James, P. Janavicius, R. Jensen, U.B. John, E.M. Jones, M. Kabisch, M. Kar, S. Karlan, B.Y. Khan, S. Khaw, K.-.T. Kibriya, M.G. Knight, J.A. Ko, Y.-.D. Konstantopoulou, I. Kosma, V.-.M. Kristensen, V. Kwong, A. Laitman, Y. Lambrechts, D. Lazaro, C. Lee, E. Le Marchand, L. Lester, J. Lindblom, A. Lindor, N. Lindstrom, S. Liu, J. Long, J. Lubinski, J. Mai, P.L. Makalic, E. Malone, K.E. Mannermaa, A. Manoukian, S. Margolin, S. Marme, F. Martens, J.W.M. McGuffog, L. Meindl, A. Miller, A. Milne, R.L. Miron, P. Montagna, M. Mazoyer, S. Mulligan, A.M. Muranen, T.A. Nathanson, K.L. Neuhausen, S.L. Nevanlinna, H. Nordestgaard, B.G. Nussbaum, R.L. Offit, K. Olah, E. Olopade, O.I. Olson, J.E. Osorio, A. Park, S.K. Peeters, P.H. Peissel, B. Peterlongo, P. Peto, J. Phelan, C.M. Pilarski, R. Poppe, B. Pylkäs, K. Radice, P. Rahman, N. Rantala, J. Rappaport, C. Rennert, G. Richardson, A. Robson, M. Romieu, I. Rudolph, A. Rutgers, E.J. Sanchez, M.-.J. Santella, R.M. Sawyer, E.J. Schmidt, D.F. Schmidt, M.K. Schmutzler, R.K. Schumacher, F. Scott, R. Senter, L. Sharma, P. Simard, J. Singer, C.F. Sinilnikova, O.M. Soucy, P. Southey, M. Steinemann, D. Stenmark-Askmalm, M. Stoppa-Lyonnet, D. Swerdlow, A. Szabo, C.I. Tamimi, R. Tapper, W. Teixeira, M.R. Teo, S.-.H. Terry, M.B. Thomassen, M. Thompson, D. Tihomirova, L. Toland, A.E. Tollenaar, R.A.E.M. Tomlinson, I. Truong, T. Tsimiklis, H. Teulé, A. Tumino, R. Tung, N. Turnbull, C. Ursin, G. van Deurzen, C.H.M. van Rensburg, E.J. Varon-Mateeva, R. Wang, Z. Wang-Gohrke, S. Weiderpass, E. Weitzel, J.N. Whittemore, A. Wildiers, H. Winqvist, R. Yang, X.R. Yannoukakos, D. Yao, S. Zamora, M.P. Zheng, W. Hall, P. Kraft, P. Vachon, C. Slager, S. Chenevix-Trench, G. Pharoah, P.D.P. Monteiro, A.A.N. García-Closas, M. Easton, D.F. Antoniou, A.C (2016) Identification of four novel susceptibility loci for oestrogen receptor negative breast cancer.. Show Abstract full text

Common variants in 94 loci have been associated with breast cancer including 15 loci with genome-wide significant associations (P<5 × 10(-8)) with oestrogen receptor (ER)-negative breast cancer and BRCA1-associated breast cancer risk. In this study, to identify new ER-negative susceptibility loci, we performed a meta-analysis of 11 genome-wide association studies (GWAS) consisting of 4,939 ER-negative cases and 14,352 controls, combined with 7,333 ER-negative cases and 42,468 controls and 15,252 BRCA1 mutation carriers genotyped on the iCOGS array. We identify four previously unidentified loci including two loci at 13q22 near KLF5, a 2p23.2 locus near WDR43 and a 2q33 locus near PPIL3 that display genome-wide significant associations with ER-negative breast cancer. In addition, 19 known breast cancer risk loci have genome-wide significant associations and 40 had moderate associations (P<0.05) with ER-negative disease. Using functional and eQTL studies we implicate TRMT61B and WDR43 at 2p23.2 and PPIL3 at 2q33 in ER-negative breast cancer aetiology. All ER-negative loci combined account for ∼11% of familial relative risk for ER-negative disease and may contribute to improved ER-negative and BRCA1 breast cancer risk prediction.

Abubakar, M. Orr, N. Daley, F. Coulson, P. Ali, H.R. Blows, F. Benitez, J. Milne, R. Brenner, H. Stegmaier, C. Mannermaa, A. Chang-Claude, J. Rudolph, A. Sinn, P. Couch, F.J. Devilee, P. Tollenaar, R.A.E.M. Seynaeve, C. Figueroa, J. Sherman, M.E. Lissowska, J. Hewitt, S. Eccles, D. Hooning, M.J. Hollestelle, A. Martens, J.W.M. van Deurzen, C.H.M. kConFab Investigators, . Bolla, M.K. Wang, Q. Jones, M. Schoemaker, M. Wesseling, J. van Leeuwen, F.E. Van 't Veer, L. Easton, D. Swerdlow, A.J. Dowsett, M. Pharoah, P.D. Schmidt, M.K. Garcia-Closas, M (2016) Prognostic value of automated KI67 scoring in breast cancer: a centralised evaluation of 8088 patients from 10 study groups.. Show Abstract full text

<h4>Background</h4>The value of KI67 in breast cancer prognostication has been questioned due to concerns on the analytical validity of visual KI67 assessment and methodological limitations of published studies. Here, we investigate the prognostic value of automated KI67 scoring in a large, multicentre study, and compare this with pathologists' visual scores available in a subset of patients.<h4>Methods</h4>We utilised 143 tissue microarrays containing 15,313 tumour tissue cores from 8088 breast cancer patients in 10 collaborating studies. A total of 1401 deaths occurred during a median follow-up of 7.5 years. Centralised KI67 assessment was performed using an automated scoring protocol. The relationship of KI67 levels with 10-year breast cancer specific survival (BCSS) was investigated using Kaplan-Meier survival curves and Cox proportional hazard regression models adjusted for known prognostic factors.<h4>Results</h4>Patients in the highest quartile of KI67 (>12 % positive KI67 cells) had a worse 10-year BCSS than patients in the lower three quartiles. This association was statistically significant for ER-positive patients (hazard ratio (HR) (95 % CI) at baseline = 1.96 (1.31-2.93); P = 0.001) but not for ER-negative patients (1.23 (0.86-1.77); P = 0.248) (P-heterogeneity = 0.064). In spite of differences in characteristics of the study populations, the estimates of HR were consistent across all studies (P-heterogeneity = 0.941 for ER-positive and P-heterogeneity = 0.866 for ER-negative). Among ER-positive cancers, KI67 was associated with worse prognosis in both node-negative (2.47 (1.16-5.27)) and node-positive (1.74 (1.05-2.86)) tumours (P-heterogeneity = 0.671). Further classification according to ER, PR and HER2 showed statistically significant associations with prognosis among hormone receptor-positive patients regardless of HER2 status (P-heterogeneity = 0.270) and among triple-negative patients (1.70 (1.02-2.84)). Model fit parameters were similar for visual and automated measures of KI67 in a subset of 2440 patients with information from both sources.<h4>Conclusions</h4>Findings from this large-scale multicentre analysis with centrally generated automated KI67 scores show strong evidence in support of a prognostic value for automated KI67 scoring in breast cancer. Given the advantages of automated scoring in terms of its potential for standardisation, reproducibility and throughput, automated methods appear to be promising alternatives to visual scoring for KI67 assessment.

Abubakar, M. Howat, W.J. Daley, F. Zabaglo, L. McDuffus, L.-.A. Blows, F. Coulson, P. Raza Ali, H. Benitez, J. Milne, R. Brenner, H. Stegmaier, C. Mannermaa, A. Chang-Claude, J. Rudolph, A. Sinn, P. Couch, F.J. Tollenaar, R.A.E.M. Devilee, P. Figueroa, J. Sherman, M.E. Lissowska, J. Hewitt, S. Eccles, D. Hooning, M.J. Hollestelle, A. Wm Martens, J. Hm van Deurzen, C. kConFab Investigators, . Bolla, M.K. Wang, Q. Jones, M. Schoemaker, M. Broeks, A. van Leeuwen, F.E. Van't Veer, L. Swerdlow, A.J. Orr, N. Dowsett, M. Easton, D. Schmidt, M.K. Pharoah, P.D. Garcia-Closas, M (2016) High-throughput automated scoring of Ki67 in breast cancer tissue microarrays from the Breast Cancer Association Consortium.. Show Abstract full text

Automated methods are needed to facilitate high-throughput and reproducible scoring of Ki67 and other markers in breast cancer tissue microarrays (TMAs) in large-scale studies. To address this need, we developed an automated protocol for Ki67 scoring and evaluated its performance in studies from the Breast Cancer Association Consortium. We utilized 166 TMAs containing 16,953 tumour cores representing 9,059 breast cancer cases, from 13 studies, with information on other clinical and pathological characteristics. TMAs were stained for Ki67 using standard immunohistochemical procedures, and scanned and digitized using the Ariol system. An automated algorithm was developed for the scoring of Ki67, and scores were compared to computer assisted visual (CAV) scores in a subset of 15 TMAs in a training set. We also assessed the correlation between automated Ki67 scores and other clinical and pathological characteristics. Overall, we observed good discriminatory accuracy (AUC = 85%) and good agreement (kappa = 0.64) between the automated and CAV scoring methods in the training set. The performance of the automated method varied by TMA (kappa range= 0.37-0.87) and study (kappa range = 0.39-0.69). The automated method performed better in satisfactory cores (kappa = 0.68) than suboptimal (kappa = 0.51) cores (p-value for comparison = 0.005); and among cores with higher total nuclei counted by the machine (4,000-4,500 cells: kappa = 0.78) than those with lower counts (50-500 cells: kappa = 0.41; p-value = 0.010). Among the 9,059 cases in this study, the correlations between automated Ki67 and clinical and pathological characteristics were found to be in the expected directions. Our findings indicate that automated scoring of Ki67 can be an efficient method to obtain good quality data across large numbers of TMAs from multicentre studies. However, robust algorithm development and rigorous pre- and post-analytical quality control procedures are necessary in order to ensure satisfactory performance.

Zhang, B. Shu, X.-.O. Delahanty, R.J. Zeng, C. Michailidou, K. Bolla, M.K. Wang, Q. Dennis, J. Wen, W. Long, J. Li, C. Dunning, A.M. Chang-Claude, J. Shah, M. Perkins, B.J. Czene, K. Darabi, H. Eriksson, M. Bojesen, S.E. Nordestgaard, B.G. Nielsen, S.F. Flyger, H. Lambrechts, D. Neven, P. Wildiers, H. Floris, G. Schmidt, M.K. Rookus, M.A. van den Hurk, K. de Kort, W.L.A.M. Couch, F.J. Olson, J.E. Hallberg, E. Vachon, C. Rudolph, A. Seibold, P. Flesch-Janys, D. Peto, J. Dos-Santos-Silva, I. Fletcher, O. Johnson, N. Nevanlinna, H. Muranen, T.A. Aittomäki, K. Blomqvist, C. Li, J. Humphreys, K. Brand, J. Guénel, P. Truong, T. Cordina-Duverger, E. Menegaux, F. Burwinkel, B. Marme, F. Yang, R. Surowy, H. Benitez, J. Zamora, M.P. Perez, J.I.A. Cox, A. Cross, S.S. Reed, M.W.R. Andrulis, I.L. Knight, J.A. Glendon, G. Tchatchou, S. Sawyer, E.J. Tomlinson, I. Kerin, M.J. Miller, N. Chenevix-Trench, G. kConFab Investigators, Australian Ovarian Study Group, . Haiman, C.A. Henderson, B.E. Schumacher, F. Marchand, L.L. Lindblom, A. Margolin, S. Hooning, M.J. Martens, J.W.M. Tilanus-Linthorst, M.M.A. Collée, J.M. Hopper, J.L. Southey, M.C. Tsimiklis, H. Apicella, C. Slager, S. Toland, A.E. Ambrosone, C.B. Yannoukakos, D. Giles, G.G. Milne, R.L. McLean, C. Fasching, P.A. Haeberle, L. Ekici, A.B. Beckmann, M.W. Brenner, H. Dieffenbach, A.K. Arndt, V. Stegmaier, C. Swerdlow, A.J. Ashworth, A. Orr, N. Jones, M. Figueroa, J. Garcia-Closas, M. Brinton, L. Lissowska, J. Dumont, M. Winqvist, R. Pylkäs, K. Jukkola-Vuorinen, A. Grip, M. Brauch, H. Brüning, T. Ko, Y.-.D. Peterlongo, P. Manoukian, S. Bonanni, B. Radice, P. Bogdanova, N. Antonenkova, N. Dörk, T. Mannermaa, A. Kataja, V. Kosma, V.-.M. Hartikainen, J.M. Devilee, P. Seynaeve, C. Van Asperen, C.J. Jakubowska, A. Lubiński, J. Jaworska-Bieniek, K. Durda, K. Hamann, U. Torres, D. Schmutzler, R.K. Neuhausen, S.L. Anton-Culver, H. Kristensen, V.N. Grenaker Alnæs, G.I. DRIVE Project, . Pierce, B.L. Kraft, P. Peters, U. Lindstrom, S. Seminara, D. Burgess, S. Ahsan, H. Whittemore, A.S. John, E.M. Gammon, M.D. Malone, K.E. Tessier, D.C. Vincent, D. Bacot, F. Luccarini, C. Baynes, C. Ahmed, S. Maranian, M. Healey, C.S. González-Neira, A. Pita, G. Alonso, M.R. Álvarez, N. Herrero, D. Pharoah, P.D.P. Simard, J. Hall, P. Hunter, D.J. Easton, D.F. Zheng, W (2015) Height and Breast Cancer Risk: Evidence From Prospective Studies and Mendelian Randomization.. Show Abstract full text

<h4>Background</h4>Epidemiological studies have linked adult height with breast cancer risk in women. However, the magnitude of the association, particularly by subtypes of breast cancer, has not been established. Furthermore, the mechanisms of the association remain unclear.<h4>Methods</h4>We performed a meta-analysis to investigate associations between height and breast cancer risk using data from 159 prospective cohorts totaling 5216302 women, including 113178 events. In a consortium with individual-level data from 46325 case patients and 42482 control patients, we conducted a Mendelian randomization analysis using a genetic score that comprised 168 height-associated variants as an instrument. This association was further evaluated in a second consortium using summary statistics data from 16003 case patients and 41335 control patients.<h4>Results</h4>The pooled relative risk of breast cancer was 1.17 (95% confidence interval [CI] = 1.15 to 1.19) per 10cm increase in height in the meta-analysis of prospective studies. In Mendelian randomization analysis, the odds ratio of breast cancer per 10cm increase in genetically predicted height was 1.22 (95% CI = 1.13 to 1.32) in the first consortium and 1.21 (95% CI = 1.05 to 1.39) in the second consortium. The association was found in both premenopausal and postmenopausal women but restricted to hormone receptor-positive breast cancer. Analyses of height-associated variants identified eight new loci associated with breast cancer risk after adjusting for multiple comparisons, including three loci at 1q21.2, DNAJC27, and CCDC91 at genome-wide significance level P < 5×10(-8).<h4>Conclusions</h4>Our study provides strong evidence that adult height is a risk factor for breast cancer in women and certain genetic factors and biological pathways affecting adult height have an important role in the etiology of breast cancer.

van Veldhoven, K. Polidoro, S. Baglietto, L. Severi, G. Sacerdote, C. Panico, S. Mattiello, A. Palli, D. Masala, G. Krogh, V. Agnoli, C. Tumino, R. Frasca, G. Flower, K. Curry, E. Orr, N. Tomczyk, K. Jones, M.E. Ashworth, A. Swerdlow, A. Chadeau-Hyam, M. Lund, E. Garcia-Closas, M. Sandanger, T.M. Flanagan, J.M. Vineis, P (2015) Epigenome-wide association study reveals decreased average methylation levels years before breast cancer diagnosis.. Show Abstract full text

<h4>Background</h4>Interest in the potential of DNA methylation in peripheral blood as a biomarker of cancer risk is increasing. We aimed to assess whether epigenome-wide DNA methylation measured in peripheral blood samples obtained before onset of the disease is associated with increased risk of breast cancer. We report on three independent prospective nested case-control studies from the European Prospective Investigation into Cancer and Nutrition (EPIC-Italy; n = 162 matched case-control pairs), the Norwegian Women and Cancer study (NOWAC; n = 168 matched pairs), and the Breakthrough Generations Study (BGS; n = 548 matched pairs). We used the Illumina 450k array to measure methylation in the EPIC and NOWAC cohorts. Whole-genome bisulphite sequencing (WGBS) was performed on the BGS cohort using pooled DNA samples, combined to reach 50× coverage across ~16 million CpG sites in the genome including 450k array CpG sites. Mean β values over all probes were calculated as a measurement for epigenome-wide methylation.<h4>Results</h4>In EPIC, we found that high epigenome-wide methylation was associated with lower risk of breast cancer (odds ratio (OR) per 1 SD = 0.61, 95 % confidence interval (CI) 0.47-0.80; -0.2 % average difference in epigenome-wide methylation for cases and controls). Specifically, this was observed in gene bodies (OR = 0.51, 95 % CI 0.38-0.69) but not in gene promoters (OR = 0.92, 95 % CI 0.64-1.32). The association was not replicated in NOWAC (OR = 1.03 95 % CI 0.81-1.30). The reasons for heterogeneity across studies are unclear. However, data from the BGS cohort was consistent with epigenome-wide hypomethylation in breast cancer cases across the overlapping 450k probe sites (difference in average epigenome-wide methylation in case and control DNA pools = -0.2 %).<h4>Conclusions</h4>We conclude that epigenome-wide hypomethylation of DNA from pre-diagnostic blood samples may be predictive of breast cancer risk and may thus be useful as a clinical biomarker.

Orr, N. Dudbridge, F. Dryden, N. Maguire, S. Novo, D. Perrakis, E. Johnson, N. Ghoussaini, M. Hopper, J.L. Southey, M.C. Apicella, C. Stone, J. Schmidt, M.K. Broeks, A. Van't Veer, L.J. Hogervorst, F.B. Fasching, P.A. Haeberle, L. Ekici, A.B. Beckmann, M.W. Gibson, L. Aitken, Z. Warren, H. Sawyer, E. Tomlinson, I. Kerin, M.J. Miller, N. Burwinkel, B. Marme, F. Schneeweiss, A. Sohn, C. Guénel, P. Truong, T. Cordina-Duverger, E. Sanchez, M. Bojesen, S.E. Nordestgaard, B.G. Nielsen, S.F. Flyger, H. Benitez, J. Zamora, M.P. Arias Perez, J.I. Menéndez, P. Anton-Culver, H. Neuhausen, S.L. Brenner, H. Dieffenbach, A.K. Arndt, V. Stegmaier, C. Hamann, U. Brauch, H. Justenhoven, C. Brüning, T. Ko, Y.-.D. Nevanlinna, H. Aittomäki, K. Blomqvist, C. Khan, S. Bogdanova, N. Dörk, T. Lindblom, A. Margolin, S. Mannermaa, A. Kataja, V. Kosma, V.-.M. Hartikainen, J.M. Chenevix-Trench, G. Beesley, J. Lambrechts, D. Moisse, M. Floris, G. Beuselinck, B. Chang-Claude, J. Rudolph, A. Seibold, P. Flesch-Janys, D. Radice, P. Peterlongo, P. Peissel, B. Pensotti, V. Couch, F.J. Olson, J.E. Slettedahl, S. Vachon, C. Giles, G.G. Milne, R.L. McLean, C. Haiman, C.A. Henderson, B.E. Schumacher, F. Le Marchand, L. Simard, J. Goldberg, M.S. Labrèche, F. Dumont, M. Kristensen, V. Alnæs, G.G. Nord, S. Borresen-Dale, A.-.L. Zheng, W. Deming-Halverson, S. Shrubsole, M. Long, J. Winqvist, R. Pylkäs, K. Jukkola-Vuorinen, A. Grip, M. Andrulis, I.L. Knight, J.A. Glendon, G. Tchatchou, S. Devilee, P. Tollenaar, R.A.E.M. Seynaeve, C.M. Van Asperen, C.J. Garcia-Closas, M. Figueroa, J. Chanock, S.J. Lissowska, J. Czene, K. Darabi, H. Eriksson, M. Klevebring, D. Hooning, M.J. Hollestelle, A. van Deurzen, C.H.M. Kriege, M. Hall, P. Li, J. Liu, J. Humphreys, K. Cox, A. Cross, S.S. Reed, M.W.R. Pharoah, P.D.P. Dunning, A.M. Shah, M. Perkins, B.J. Jakubowska, A. Lubinski, J. Jaworska-Bieniek, K. Durda, K. Ashworth, A. Swerdlow, A. Jones, M. Schoemaker, M.J. Meindl, A. Schmutzler, R.K. Olswold, C. Slager, S. Toland, A.E. Yannoukakos, D. Muir, K. Lophatananon, A. Stewart-Brown, S. Siriwanarangsan, P. Matsuo, K. Ito, H. Iwata, H. Ishiguro, J. Wu, A.H. Tseng, C.-.C. Van Den Berg, D. Stram, D.O. Teo, S.H. Yip, C.H. Kang, P. Ikram, M.K. Shu, X.-.O. Lu, W. Gao, Y.-.T. Cai, H. Kang, D. Choi, J.-.Y. Park, S.K. Noh, D.-.Y. Hartman, M. Miao, H. Lim, W.Y. Lee, S.C. Sangrajrang, S. Gaborieau, V. Brennan, P. Mckay, J. Wu, P.-.E. Hou, M.-.F. Yu, J.-.C. Shen, C.-.Y. Blot, W. Cai, Q. Signorello, L.B. Luccarini, C. Bayes, C. Ahmed, S. Maranian, M. Healey, C.S. González-Neira, A. Pita, G. Alonso, M.R. Álvarez, N. Herrero, D. Tessier, D.C. Vincent, D. Bacot, F. Hunter, D.J. Lindstrom, S. Dennis, J. Michailidou, K. Bolla, M.K. Easton, D.F. dos Santos Silva, I. Fletcher, O. Peto, J. GENICA Network, . kConFab Investigators, . Australian Ovarian Cancer Study Group, (2015) Fine-mapping identifies two additional breast cancer susceptibility loci at 9q31.2.. Show Abstract full text

We recently identified a novel susceptibility variant, rs865686, for estrogen-receptor positive breast cancer at 9q31.2. Here, we report a fine-mapping analysis of the 9q31.2 susceptibility locus using 43 160 cases and 42 600 controls of European ancestry ascertained from 52 studies and a further 5795 cases and 6624 controls of Asian ancestry from nine studies. Single nucleotide polymorphism (SNP) rs676256 was most strongly associated with risk in Europeans (odds ratios [OR] = 0.90 [0.88-0.92]; P-value = 1.58 × 10(-25)). This SNP is one of a cluster of highly correlated variants, including rs865686, that spans ∼14.5 kb. We identified two additional independent association signals demarcated by SNPs rs10816625 (OR = 1.12 [1.08-1.17]; P-value = 7.89 × 10(-09)) and rs13294895 (OR = 1.09 [1.06-1.12]; P-value = 2.97 × 10(-11)). SNP rs10816625, but not rs13294895, was also associated with risk of breast cancer in Asian individuals (OR = 1.12 [1.06-1.18]; P-value = 2.77 × 10(-05)). Functional genomic annotation using data derived from breast cancer cell-line models indicates that these SNPs localise to putative enhancer elements that bind known drivers of hormone-dependent breast cancer, including ER-α, FOXA1 and GATA-3. In vitro analyses indicate that rs10816625 and rs13294895 have allele-specific effects on enhancer activity and suggest chromatin interactions with the KLF4 gene locus. These results demonstrate the power of dense genotyping in large studies to identify independent susceptibility variants. Analysis of associations using subjects with different ancestry, combined with bioinformatic and genomic characterisation, can provide strong evidence for the likely causative alleles and their functional basis.

Lin, W.-.Y. Camp, N.J. Ghoussaini, M. Beesley, J. Michailidou, K. Hopper, J.L. Apicella, C. Southey, M.C. Stone, J. Schmidt, M.K. Broeks, A. Van't Veer, L.J. Th Rutgers, E.J. Muir, K. Lophatananon, A. Stewart-Brown, S. Siriwanarangsan, P. Fasching, P.A. Haeberle, L. Ekici, A.B. Beckmann, M.W. Peto, J. Dos-Santos-Silva, I. Fletcher, O. Johnson, N. Bolla, M.K. Wang, Q. Dennis, J. Sawyer, E.J. Cheng, T. Tomlinson, I. Kerin, M.J. Miller, N. Marmé, F. Surowy, H.M. Burwinkel, B. Guénel, P. Truong, T. Menegaux, F. Mulot, C. Bojesen, S.E. Nordestgaard, B.G. Nielsen, S.F. Flyger, H. Benitez, J. Zamora, M.P. Arias Perez, J.I. Menéndez, P. González-Neira, A. Pita, G. Alonso, M.R. Alvarez, N. Herrero, D. Anton-Culver, H. Brenner, H. Dieffenbach, A.K. Arndt, V. Stegmaier, C. Meindl, A. Lichtner, P. Schmutzler, R.K. Müller-Myhsok, B. Brauch, H. Brüning, T. Ko, Y.-.D. GENICA Network, . Tessier, D.C. Vincent, D. Bacot, F. Nevanlinna, H. Aittomäki, K. Blomqvist, C. Khan, S. Matsuo, K. Ito, H. Iwata, H. Horio, A. Bogdanova, N.V. Antonenkova, N.N. Dörk, T. Lindblom, A. Margolin, S. Mannermaa, A. Kataja, V. Kosma, V.-.M. Hartikainen, J.M. kConFab Investigators, . Australian Ovarian Cancer Study Group, . Wu, A.H. Tseng, C.-.C. Van Den Berg, D. Stram, D.O. Neven, P. Wauters, E. Wildiers, H. Lambrechts, D. Chang-Claude, J. Rudolph, A. Seibold, P. Flesch-Janys, D. Radice, P. Peterlongo, P. Manoukian, S. Bonanni, B. Couch, F.J. Wang, X. Vachon, C. Purrington, K. Giles, G.G. Milne, R.L. Mclean, C. Haiman, C.A. Henderson, B.E. Schumacher, F. Le Marchand, L. Simard, J. Goldberg, M.S. Labrèche, F. Dumont, M. Teo, S.H. Yip, C.H. Hassan, N. Vithana, E.N. Kristensen, V. Zheng, W. Deming-Halverson, S. Shrubsole, M.J. Long, J. Winqvist, R. Pylkäs, K. Jukkola-Vuorinen, A. Kauppila, S. Andrulis, I.L. Knight, J.A. Glendon, G. Tchatchou, S. Devilee, P. Tollenaar, R.A.E.M. Seynaeve, C. Van Asperen, C.J. García-Closas, M. Figueroa, J. Lissowska, J. Brinton, L. Czene, K. Darabi, H. Eriksson, M. Brand, J.S. Hooning, M.J. Hollestelle, A. Van Den Ouweland, A.M.W. Jager, A. Li, J. Liu, J. Humphreys, K. Shu, X.-.O. Lu, W. Gao, Y.-.T. Cai, H. Cross, S.S. Reed, M.W.R. Blot, W. Signorello, L.B. Cai, Q. Pharoah, P.D.P. Perkins, B. Shah, M. Blows, F.M. Kang, D. Yoo, K.-.Y. Noh, D.-.Y. Hartman, M. Miao, H. Chia, K.S. Putti, T.C. Hamann, U. Luccarini, C. Baynes, C. Ahmed, S. Maranian, M. Healey, C.S. Jakubowska, A. Lubinski, J. Jaworska-Bieniek, K. Durda, K. Sangrajrang, S. Gaborieau, V. Brennan, P. Mckay, J. Slager, S. Toland, A.E. Yannoukakos, D. Shen, C.-.Y. Hsiung, C.-.N. Wu, P.-.E. Ding, S.-.L. Ashworth, A. Jones, M. Orr, N. Swerdlow, A.J. Tsimiklis, H. Makalic, E. Schmidt, D.F. Bui, Q.M. Chanock, S.J. Hunter, D.J. Hein, R. Dahmen, N. Beckmann, L. Aaltonen, K. Muranen, T.A. Heikkinen, T. Irwanto, A. Rahman, N. Turnbull, C.A. Breast and Ovarian Cancer Susceptibility (BOCS) Study, . Waisfisz, Q. Meijers-Heijboer, H.E.J. Adank, M.A. Van Der Luijt, R.B. Hall, P. Chenevix-Trench, G. Dunning, A. Easton, D.F. Cox, A (2015) Identification and characterization of novel associations in the CASP8/ALS2CR12 region on chromosome 2 with breast cancer risk.. Show Abstract full text

Previous studies have suggested that polymorphisms in CASP8 on chromosome 2 are associated with breast cancer risk. To clarify the role of CASP8 in breast cancer susceptibility, we carried out dense genotyping of this region in the Breast Cancer Association Consortium (BCAC). Single-nucleotide polymorphisms (SNPs) spanning a 1 Mb region around CASP8 were genotyped in 46 450 breast cancer cases and 42 600 controls of European origin from 41 studies participating in the BCAC as part of a custom genotyping array experiment (iCOGS). Missing genotypes and SNPs were imputed and, after quality exclusions, 501 typed and 1232 imputed SNPs were included in logistic regression models adjusting for study and ancestry principal components. The SNPs retained in the final model were investigated further in data from nine genome-wide association studies (GWAS) comprising in total 10 052 case and 12 575 control subjects. The most significant association signal observed in European subjects was for the imputed intronic SNP rs1830298 in ALS2CR12 (telomeric to CASP8), with per allele odds ratio and 95% confidence interval [OR (95% confidence interval, CI)] for the minor allele of 1.05 (1.03-1.07), P = 1 × 10(-5). Three additional independent signals from intronic SNPs were identified, in CASP8 (rs36043647), ALS2CR11 (rs59278883) and CFLAR (rs7558475). The association with rs1830298 was replicated in the imputed results from the combined GWAS (P = 3 × 10(-6)), yielding a combined OR (95% CI) of 1.06 (1.04-1.08), P = 1 × 10(-9). Analyses of gene expression associations in peripheral blood and normal breast tissue indicate that CASP8 might be the target gene, suggesting a mechanism involving apoptosis.

Lemnrau, A. Brook, M.N. Fletcher, O. Coulson, P. Tomczyk, K. Jones, M. Ashworth, A. Swerdlow, A. Orr, N. Garcia-Closas, M (2015) Mitochondrial DNA Copy Number in Peripheral Blood Cells and Risk of Developing Breast Cancer.. Show Abstract full text

Increased mitochondrial DNA (mtDNA) copy number in peripheral blood cells (PBC) has been associated with the risk of developing several tumor types. Here we evaluate sources of variation of this biomarker and its association with breast cancer risk in a prospective cohort study. mtDNA copy number was measured using quantitative real-time PCR on PBC DNA samples from participants in the UK-based Breakthrough Generations Study. Temporal and assay variation was evaluated in a serial study of 91 women, with two blood samples collected approximately 6-years apart. Then, associations with breast cancer risk factors and risk were evaluated in 1,108 cases and 1,099 controls using a nested case-control design. In the serial study, mtDNA copy number showed low assay variation but large temporal variation [assay intraclass correlation coefficient (ICC), 79.3%-87.9%; temporal ICC, 38.3%). Higher mtDNA copy number was significantly associated with younger age at blood collection, being premenopausal, having an older age at menopause, and never taking HRT, both in cases and controls. Based on measurements in a single blood sample taken on average 6 years before diagnosis, higher mtDNA copy number was associated with increased breast cancer risk [OR (95% CI) for highest versus lowest quartile, 1.37 (1.02-1.83); P trend = 0.007]. In conclusion, mtDNA copy number is associated with breast cancer risk and represents a promising biomarker for risk assessment. The relatively large temporal variation should be taken into account in future analyses.

Kote-Jarai, Z. Mikropoulos, C. Leongamornlert, D.A. Dadaev, T. Tymrakiewicz, M. Saunders, E.J. Jones, M. Jugurnauth-Little, S. Govindasami, K. Guy, M. Hamdy, F.C. Donovan, J.L. Neal, D.E. Lane, J.A. Dearnaley, D. Wilkinson, R.A. Sawyer, E.J. Morgan, A. Antoniou, A.C. Eeles, R.A. UK Genetic Prostate Cancer Study Collaborators, and the ProtecT Study Group, (2015) Prevalence of the HOXB13 G84E germline mutation in British men and correlation with prostate cancer risk, tumour characteristics and clinical outcomes.. Show Abstract full text

<h4>Background</h4>A rare recurrent missense variant in HOXB13 (rs138213197/G84E) was recently reported to be associated with hereditary prostate cancer. Population-based studies have established that, since the frequency of this single-nucleotide polymorphism (SNP) varies between geographic regions, the associated proportion of prostate cancer (PrCa) risk contribution is also highly variable by country.<h4>Patients and methods</h4>This is the largest comprehensive case-control study assessing the prevalence of the HOXB13 G84E variant to date and is the first in the UK population. We genotyped 8652 men diagnosed with PrCa within the UK Genetic Prostate Cancer Study (UKGPCS) and 5252 healthy men from the UK ProtecT study.<h4>Results</h4>HOXB13 G84E was identified in 0.5% of the healthy controls and 1.5% of the PrCa cases, and it was associated with a 2.93-fold increased risk of PrCa [95% confidence interval (CI) 1.94-4.59; P = 6.27 × 10(-8)]. The risk was even higher among men with family history of PrCa [odds ratio (OR) = 4.53, 95% CI 2.86-7.34; P = 3.1 × 10(-8)] and in young-onset PrCa (diagnosed up to the age of 55 years; OR = 3.11, 95% CI 1.98-5.00; P = 6.1 × 10(-7)). There was no significant association between Gleason Score, presenting prostate specific antigen, tumour-node-metastasis (TNM) stage or NCCN risk group and carrier status. HOXB13 G84E was not associated with overall or cancer-specific survival. We found that the polygenic PrCa risk score (PR score), calculated using the 71 known single-nucleotide polymorphisms (SNPs) associated with PrCa and the HOXB13 G84E variant act multiplicatively on PrCa risk. Based on the estimated prevalence and risk, this rare variant explains ∼1% of the familial risk of PrCa in the UK population.<h4>Conclusions</h4>The clinical importance of HOXB13 G84E in PrCa management has not been established. This variant was found to have no effect on prognostic implications but could be used for stratifying screening, by identifying men at high risk.<h4>Clinical trials numbers</h4>Prostate Testing for Cancer and Treatment (ProtecT): NCT02044172.<h4>Uk genetic prostate cancer study</h4>Epidemiology and Molecular Genetics Studies (UKGPCS): NCT01737242.

Glubb, D.M. Maranian, M.J. Michailidou, K. Pooley, K.A. Meyer, K.B. Kar, S. Carlebur, S. O'Reilly, M. Betts, J.A. Hillman, K.M. Kaufmann, S. Beesley, J. Canisius, S. Hopper, J.L. Southey, M.C. Tsimiklis, H. Apicella, C. Schmidt, M.K. Broeks, A. Hogervorst, F.B. van der Schoot, C.E. Muir, K. Lophatananon, A. Stewart-Brown, S. Siriwanarangsan, P. Fasching, P.A. Ruebner, M. Ekici, A.B. Beckmann, M.W. Peto, J. dos-Santos-Silva, I. Fletcher, O. Johnson, N. Pharoah, P.D.P. Bolla, M.K. Wang, Q. Dennis, J. Sawyer, E.J. Tomlinson, I. Kerin, M.J. Miller, N. Burwinkel, B. Marme, F. Yang, R. Surowy, H. Guénel, P. Truong, T. Menegaux, F. Sanchez, M. Bojesen, S.E. Nordestgaard, B.G. Nielsen, S.F. Flyger, H. González-Neira, A. Benitez, J. Zamora, M.P. Arias Perez, J.I. Anton-Culver, H. Neuhausen, S.L. Brenner, H. Dieffenbach, A.K. Arndt, V. Stegmaier, C. Meindl, A. Schmutzler, R.K. Brauch, H. Ko, Y.-.D. Brüning, T. GENICA Network, . Nevanlinna, H. Muranen, T.A. Aittomäki, K. Blomqvist, C. Matsuo, K. Ito, H. Iwata, H. Tanaka, H. Dörk, T. Bogdanova, N.V. Helbig, S. Lindblom, A. Margolin, S. Mannermaa, A. Kataja, V. Kosma, V.-.M. Hartikainen, J.M. kConFab Investigators, . Wu, A.H. Tseng, C.-.C. Van Den Berg, D. Stram, D.O. Lambrechts, D. Zhao, H. Weltens, C. van Limbergen, E. Chang-Claude, J. Flesch-Janys, D. Rudolph, A. Seibold, P. Radice, P. Peterlongo, P. Barile, M. Capra, F. Couch, F.J. Olson, J.E. Hallberg, E. Vachon, C. Giles, G.G. Milne, R.L. McLean, C. Haiman, C.A. Henderson, B.E. Schumacher, F. Le Marchand, L. Simard, J. Goldberg, M.S. Labrèche, F. Dumont, M. Teo, S.H. Yip, C.H. See, M.-.H. Cornes, B. Cheng, C.-.Y. Ikram, M.K. Kristensen, V. Norwegian Breast Cancer Study, . Zheng, W. Halverson, S.L. Shrubsole, M. Long, J. Winqvist, R. Pylkäs, K. Jukkola-Vuorinen, A. Kauppila, S. Andrulis, I.L. Knight, J.A. Glendon, G. Tchatchou, S. Devilee, P. Tollenaar, R.A.E.M. Seynaeve, C. Van Asperen, C.J. García-Closas, M. Figueroa, J. Chanock, S.J. Lissowska, J. Czene, K. Klevebring, D. Darabi, H. Eriksson, M. Hooning, M.J. Hollestelle, A. Martens, J.W.M. Collée, J.M. Hall, P. Li, J. Humphreys, K. Shu, X.-.O. Lu, W. Gao, Y.-.T. Cai, H. Cox, A. Cross, S.S. Reed, M.W.R. Blot, W. Signorello, L.B. Cai, Q. Shah, M. Ghoussaini, M. Kang, D. Choi, J.-.Y. Park, S.K. Noh, D.-.Y. Hartman, M. Miao, H. Lim, W.Y. Tang, A. Hamann, U. Torres, D. Jakubowska, A. Lubinski, J. Jaworska, K. Durda, K. Sangrajrang, S. Gaborieau, V. Brennan, P. McKay, J. Olswold, C. Slager, S. Toland, A.E. Yannoukakos, D. Shen, C.-.Y. Wu, P.-.E. Yu, J.-.C. Hou, M.-.F. Swerdlow, A. Ashworth, A. Orr, N. Jones, M. Pita, G. Alonso, M.R. Álvarez, N. Herrero, D. Tessier, D.C. Vincent, D. Bacot, F. Luccarini, C. Baynes, C. Ahmed, S. Healey, C.S. Brown, M.A. Ponder, B.A.J. Chenevix-Trench, G. Thompson, D.J. Edwards, S.L. Easton, D.F. Dunning, A.M. French, J.D (2015) Fine-scale mapping of the 5q11.2 breast cancer locus reveals at least three independent risk variants regulating MAP3K1.. Show Abstract full text

Genome-wide association studies (GWASs) have revealed SNP rs889312 on 5q11.2 to be associated with breast cancer risk in women of European ancestry. In an attempt to identify the biologically relevant variants, we analyzed 909 genetic variants across 5q11.2 in 103,991 breast cancer individuals and control individuals from 52 studies in the Breast Cancer Association Consortium. Multiple logistic regression analyses identified three independent risk signals: the strongest associations were with 15 correlated variants (iCHAV1), where the minor allele of the best candidate, rs62355902, associated with significantly increased risks of both estrogen-receptor-positive (ER(+): odds ratio [OR] = 1.24, 95% confidence interval [CI] = 1.21-1.27, ptrend = 5.7 × 10(-44)) and estrogen-receptor-negative (ER(-): OR = 1.10, 95% CI = 1.05-1.15, ptrend = 3.0 × 10(-4)) tumors. After adjustment for rs62355902, we found evidence of association of a further 173 variants (iCHAV2) containing three subsets with a range of effects (the strongest was rs113317823 [pcond = 1.61 × 10(-5)]) and five variants composing iCHAV3 (lead rs11949391; ER(+): OR = 0.90, 95% CI = 0.87-0.93, pcond = 1.4 × 10(-4)). Twenty-six percent of the prioritized candidate variants coincided with four putative regulatory elements that interact with the MAP3K1 promoter through chromatin looping and affect MAP3K1 promoter activity. Functional analysis indicated that the cancer risk alleles of four candidates (rs74345699 and rs62355900 [iCHAV1], rs16886397 [iCHAV2a], and rs17432750 [iCHAV3]) increased MAP3K1 transcriptional activity. Chromatin immunoprecipitation analysis revealed diminished GATA3 binding to the minor (cancer-protective) allele of rs17432750, indicating a mechanism for its action. We propose that the cancer risk alleles act to increase MAP3K1 expression in vivo and might promote breast cancer cell survival.

Flanagan, J.M. Brook, M.N. Orr, N. Tomczyk, K. Coulson, P. Fletcher, O. Jones, M.E. Schoemaker, M.J. Ashworth, A. Swerdlow, A. Brown, R. Garcia-Closas, M (2015) Temporal stability and determinants of white blood cell DNA methylation in the breakthrough generations study.. Show Abstract full text

<h4>Background</h4>Epigenome-wide association studies (EWAS) using measurements of blood DNA methylation are performed to identify associations of methylation changes with environmental and lifestyle exposures and disease risk. However, little is known about the variation of methylation markers in the population and their stability over time, both important factors in the design and interpretation of EWAS. We aimed to identify stable variable methylated probes (VMP), i.e., markers that are variable in the population, yet stable over time.<h4>Methods</h4>We estimated the intraclass correlation coefficient (ICC) for each probe on the Illumina 450K methylation array in paired samples collected approximately 6 years apart from 92 participants in the Breakthrough Generations Study. We also evaluated relationships with age, reproductive and hormonal history, weight, alcohol intake, and smoking.<h4>Results</h4>Approximately 17% of probes had an ICC > 0.50 and were considered stable VMPs (stable-VMPs). Stable-VMPs were enriched for probes located in "shores" bordering CpG islands, and at approximately 1.3 kb downstream from the transcription start site in the transition between the unmethylated promoter and methylated gene body. Both cross-sectional and longitudinal data analyses provided strong evidence for associations between changes in methylation levels and aging. Smoking-related probes at 2q37.1 and AHRR were stable-VMPs and related to time since quitting. We also observed associations between methylation and weight changes.<h4>Conclusion</h4>Our results provide support for the use of white blood cell DNA methylation as a biomarker of exposure in EWAS.<h4>Impact</h4>Larger studies, preferably with repeated measures over time, will be required to establish associations between specific probes and exposures.

Schoeps, A. Rudolph, A. Seibold, P. Dunning, A.M. Milne, R.L. Bojesen, S.E. Swerdlow, A. Andrulis, I. Brenner, H. Behrens, S. Orr, N. Jones, M. Ashworth, A. Li, J. Cramp, H. Connley, D. Czene, K. Darabi, H. Chanock, S.J. Lissowska, J. Figueroa, J.D. Knight, J. Glendon, G. Mulligan, A.M. Dumont, M. Severi, G. Baglietto, L. Olson, J. Vachon, C. Purrington, K. Moisse, M. Neven, P. Wildiers, H. Spurdle, A. Kosma, V.-.M. Kataja, V. Hartikainen, J.M. Hamann, U. Ko, Y.-.D. Dieffenbach, A.K. Arndt, V. Stegmaier, C. Malats, N. Arias Perez, J.I. Benítez, J. Flyger, H. Nordestgaard, B.G. Truong, T. Cordina-Duverger, E. Menegaux, F. dos Santos Silva, I. Fletcher, O. Johnson, N. Häberle, L. Beckmann, M.W. Ekici, A.B. Braaf, L. Atsma, F. van den Broek, A.J. Makalic, E. Schmidt, D.F. Southey, M.C. Cox, A. Simard, J. Giles, G.G. Lambrechts, D. Mannermaa, A. Brauch, H. Guénel, P. Peto, J. Fasching, P.A. Hopper, J. Flesch-Janys, D. Couch, F. Chenevix-Trench, G. Pharoah, P.D.P. Garcia-Closas, M. Schmidt, M.K. Hall, P. Easton, D.F. Chang-Claude, J (2014) Identification of new genetic susceptibility loci for breast cancer through consideration of gene-environment interactions.. Show Abstract full text

Genes that alter disease risk only in combination with certain environmental exposures may not be detected in genetic association analysis. By using methods accounting for gene-environment (G × E) interaction, we aimed to identify novel genetic loci associated with breast cancer risk. Up to 34,475 cases and 34,786 controls of European ancestry from up to 23 studies in the Breast Cancer Association Consortium were included. Overall, 71,527 single nucleotide polymorphisms (SNPs), enriched for association with breast cancer, were tested for interaction with 10 environmental risk factors using three recently proposed hybrid methods and a joint test of association and interaction. Analyses were adjusted for age, study, population stratification, and confounding factors as applicable. Three SNPs in two independent loci showed statistically significant association: SNPs rs10483028 and rs2242714 in perfect linkage disequilibrium on chromosome 21 and rs12197388 in ARID1B on chromosome 6. While rs12197388 was identified using the joint test with parity and with age at menarche (P-values = 3 × 10(-07)), the variants on chromosome 21 q22.12, which showed interaction with adult body mass index (BMI) in 8,891 postmenopausal women, were identified by all methods applied. SNP rs10483028 was associated with breast cancer in women with a BMI below 25 kg/m(2) (OR = 1.26, 95% CI 1.15-1.38) but not in women with a BMI of 30 kg/m(2) or higher (OR = 0.89, 95% CI 0.72-1.11, P for interaction = 3.2 × 10(-05)). Our findings confirm comparable power of the recent methods for detecting G × E interaction and the utility of using G × E interaction analyses to identify new susceptibility loci.

Schoemaker, M.J. Folkerd, E.J. Jones, M.E. Rae, M. Allen, S. Ashworth, A. Dowsett, M. Swerdlow, A.J (2014) Combined effects of endogenous sex hormone levels and mammographic density on postmenopausal breast cancer risk: results from the Breakthrough Generations Study.. Show Abstract full text

<h4>Background</h4>Mammographic density and sex hormone levels are strong risk factors for breast cancer, but it is unclear whether they represent the same aetiological entity or are independent risk factors.<h4>Methods</h4>Within the Breakthrough Generations Study cohort, we conducted a case-control study of 265 postmenopausal breast cancer cases and 343 controls with prediagnostic mammograms and blood samples. Plasma was assayed for oestradiol, testosterone and sex hormone-binding globulin (SHBG) concentrations and mammographic density assessed by Cumulus.<h4>Results</h4>Oestradiol and testosterone were negatively and SHBG positively associated with percentage density and absolute dense area, but after adjusting for body mass index the associations remained significant only for SHBG. Breast cancer risk was independently and significantly positively associated with percentage density (P=0.002), oestradiol (P=0.002) and testosterone (P=0.007) levels. Women in the highest tertile of both density and sex hormone level were at greatest risk, with an odds ratio of 7.81 (95% confidence interval (CI): 2.89-21.1) for oestradiol and 4.57 (95% CI: 1.75-11.9) for testosterone and high density compared with those who were in the lowest tertiles. The cumulative risk of breast cancer in the highest oestradiol and density tertiles, representing 8% of controls, was estimated as 12.8% at ages 50-69 years and 19.4% at ages 20-79 years, and in the lowest tertiles was 1.7% and 4.3%, respectively. Associations of breast cancer risk with tertiles of mammographic dense area were less strong than for percentage density.<h4>Conclusions</h4>Endogenous sex hormone levels and mammographic density are independent risk factors for postmenopausal breast cancer, which in combination can identify women who might benefit from increased frequency of screening and chemoprophylaxis.

Purrington, K.S. Slettedahl, S. Bolla, M.K. Michailidou, K. Czene, K. Nevanlinna, H. Bojesen, S.E. Andrulis, I.L. Cox, A. Hall, P. Carpenter, J. Yannoukakos, D. Haiman, C.A. Fasching, P.A. Mannermaa, A. Winqvist, R. Brenner, H. Lindblom, A. Chenevix-Trench, G. Benitez, J. Swerdlow, A. Kristensen, V. Guénel, P. Meindl, A. Darabi, H. Eriksson, M. Fagerholm, R. Aittomäki, K. Blomqvist, C. Nordestgaard, B.G. Nielsen, S.F. Flyger, H. Wang, X. Olswold, C. Olson, J.E. Mulligan, A.M. Knight, J.A. Tchatchou, S. Reed, M.W.R. Cross, S.S. Liu, J. Li, J. Humphreys, K. Clarke, C. Scott, R. ABCTB Investigators, . Fostira, F. Fountzilas, G. Konstantopoulou, I. Henderson, B.E. Schumacher, F. Le Marchand, L. Ekici, A.B. Hartmann, A. Beckmann, M.W. Hartikainen, J.M. Kosma, V.-.M. Kataja, V. Jukkola-Vuorinen, A. Pylkäs, K. Kauppila, S. Dieffenbach, A.K. Stegmaier, C. Arndt, V. Margolin, S. Australian Ovarian Cancer Study Group, . kConFab Investigators, . Balleine, R. Arias Perez, J.I. Pilar Zamora, M. Menéndez, P. Ashworth, A. Jones, M. Orr, N. Arveux, P. Kerbrat, P. Truong, T. Bugert, P. Toland, A.E. Ambrosone, C.B. Labrèche, F. Goldberg, M.S. Dumont, M. Ziogas, A. Lee, E. Dite, G.S. Apicella, C. Southey, M.C. Long, J. Shrubsole, M. Deming-Halverson, S. Ficarazzi, F. Barile, M. Peterlongo, P. Durda, K. Jaworska-Bieniek, K. Tollenaar, R.A.E.M. Seynaeve, C. GENICA Network, . Brüning, T. Ko, Y.-.D. Van Deurzen, C.H.M. Martens, J.W.M. Kriege, M. Figueroa, J.D. Chanock, S.J. Lissowska, J. Tomlinson, I. Kerin, M.J. Miller, N. Schneeweiss, A. Tapper, W.J. Gerty, S.M. Durcan, L. Mclean, C. Milne, R.L. Baglietto, L. dos Santos Silva, I. Fletcher, O. Johnson, N. Van'T Veer, L.J. Cornelissen, S. Försti, A. Torres, D. Rüdiger, T. Rudolph, A. Flesch-Janys, D. Nickels, S. Weltens, C. Floris, G. Moisse, M. Dennis, J. Wang, Q. Dunning, A.M. Shah, M. Brown, J. Simard, J. Anton-Culver, H. Neuhausen, S.L. Hopper, J.L. Bogdanova, N. Dörk, T. Zheng, W. Radice, P. Jakubowska, A. Lubinski, J. Devillee, P. Brauch, H. Hooning, M. García-Closas, M. Sawyer, E. Burwinkel, B. Marmee, F. Eccles, D.M. Giles, G.G. Peto, J. Schmidt, M. Broeks, A. Hamann, U. Chang-Claude, J. Lambrechts, D. Pharoah, P.D.P. Easton, D. Pankratz, V.S. Slager, S. Vachon, C.M. Couch, F.J (2014) Genetic variation in mitotic regulatory pathway genes is associated with breast tumor grade.. Show Abstract full text

Mitotic index is an important component of histologic grade and has an etiologic role in breast tumorigenesis. Several small candidate gene studies have reported associations between variation in mitotic genes and breast cancer risk. We measured associations between 2156 single nucleotide polymorphisms (SNPs) from 194 mitotic genes and breast cancer risk, overall and by histologic grade, in the Breast Cancer Association Consortium (BCAC) iCOGS study (n = 39 067 cases; n = 42 106 controls). SNPs in TACC2 [rs17550038: odds ratio (OR) = 1.24, 95% confidence interval (CI) 1.16-1.33, P = 4.2 × 10(-10)) and EIF3H (rs799890: OR = 1.07, 95% CI 1.04-1.11, P = 8.7 × 10(-6)) were significantly associated with risk of low-grade breast cancer. The TACC2 signal was retained (rs17550038: OR = 1.15, 95% CI 1.07-1.23, P = 7.9 × 10(-5)) after adjustment for breast cancer risk SNPs in the nearby FGFR2 gene, suggesting that TACC2 is a novel, independent genome-wide significant genetic risk locus for low-grade breast cancer. While no SNPs were individually associated with high-grade disease, a pathway-level gene set analysis showed that variation across the 194 mitotic genes was associated with high-grade breast cancer risk (P = 2.1 × 10(-3)). These observations will provide insight into the contribution of mitotic defects to histological grade and the etiology of breast cancer.

Murray, A. Schoemaker, M.J. Bennett, C.E. Ennis, S. Macpherson, J.N. Jones, M. Morris, D.H. Orr, N. Ashworth, A. Jacobs, P.A. Swerdlow, A.J (2014) Population-based estimates of the prevalence of FMR1 expansion mutations in women with early menopause and primary ovarian insufficiency.. Show Abstract full text

<h4>Purpose</h4>Primary ovarian insufficiency before the age of 40 years affects 1% of the female population and is characterized by permanent cessation of menstruation. Genetic causes include FMR1 expansion mutations. Previous studies have estimated mutation prevalence in clinical referrals for primary ovarian insufficiency, but these are likely to be biased as compared with cases in the general population. The prevalence of FMR1 expansion mutations in early menopause (between the ages of 40 and 45 years) has not been published.<h4>Methods</h4>We studied FMR1 CGG repeat number in more than 2,000 women from the Breakthrough Generations Study who underwent menopause before the age of 46 years. We determined the prevalence of premutation (55-200 CGG repeats) and intermediate (45-54 CGG repeats) alleles in women with primary ovarian insufficiency (n = 254) and early menopause (n = 1,881).<h4>Results</h4>The prevalence of the premutation was 2.0% in primary ovarian insufficiency, 0.7% in early menopause, and 0.4% in controls, corresponding to odds ratios of 5.4 (95% confidence interval = 1.7-17.4; P = 0.004) for primary ovarian insufficiency and 2.0 (95% confidence interval = 0.8-5.1; P = 0.12) for early menopause. Combining primary ovarian insufficiency and early menopause gave an odds ratio of 2.4 (95% confidence interval = 1.02-5.8; P = 0.04). Intermediate alleles were not significant risk factors for either early menopause or primary ovarian insufficiency.<h4>Conclusion</h4>FMR1 premutations are not as prevalent in women with ovarian insufficiency as previous estimates have suggested, but they still represent a substantial cause of primary ovarian insufficiency and early menopause.

McFadden, E. Jones, M.E. Schoemaker, M.J. Ashworth, A. Swerdlow, A.J (2014) The relationship between obesity and exposure to light at night: cross-sectional analyses of over 100,000 women in the Breakthrough Generations Study.. Show Abstract full text

There has been a worldwide epidemic of obesity in recent decades. In animal studies, there is convincing evidence that light exposure causes weight gain, even when calorie intake and physical activity are held constant. Disruption of sleep and circadian rhythms by exposure to light at night (LAN) might be one mechanism contributing to the rise in obesity, but it has not been well-investigated in humans. Using multinomial logistic regression, we examined the association between exposure to LAN and obesity in questionnaire data from over 100,000 women in the Breakthrough Generations Study, a cohort study of women aged 16 years or older who were living in the United Kingdom and recruited during 2003-2012. The odds of obesity, measured using body mass index, waist:hip ratio, waist:height ratio, and waist circumference, increased with increasing levels of LAN exposure (P < 0.001), even after adjustment for potential confounders such as sleep duration, alcohol intake, physical activity, and current smoking. We found a significant association between LAN exposure and obesity which was not explained by potential confounders we could measure. While the possibility of residual confounding cannot be excluded, the pattern is intriguing, accords with the results of animal experiments, and warrants further investigation.

Khan, S. Greco, D. Michailidou, K. Milne, R.L. Muranen, T.A. Heikkinen, T. Aaltonen, K. Dennis, J. Bolla, M.K. Liu, J. Hall, P. Irwanto, A. Humphreys, K. Li, J. Czene, K. Chang-Claude, J. Hein, R. Rudolph, A. Seibold, P. Flesch-Janys, D. Fletcher, O. Peto, J. dos Santos Silva, I. Johnson, N. Gibson, L. Aitken, Z. Hopper, J.L. Tsimiklis, H. Bui, M. Makalic, E. Schmidt, D.F. Southey, M.C. Apicella, C. Stone, J. Waisfisz, Q. Meijers-Heijboer, H. Adank, M.A. van der Luijt, R.B. Meindl, A. Schmutzler, R.K. Müller-Myhsok, B. Lichtner, P. Turnbull, C. Rahman, N. Chanock, S.J. Hunter, D.J. Cox, A. Cross, S.S. Reed, M.W.R. Schmidt, M.K. Broeks, A. Van't Veer, L.J. Hogervorst, F.B. Fasching, P.A. Schrauder, M.G. Ekici, A.B. Beckmann, M.W. Bojesen, S.E. Nordestgaard, B.G. Nielsen, S.F. Flyger, H. Benitez, J. Zamora, P.M. Perez, J.I.A. Haiman, C.A. Henderson, B.E. Schumacher, F. Le Marchand, L. Pharoah, P.D.P. Dunning, A.M. Shah, M. Luben, R. Brown, J. Couch, F.J. Wang, X. Vachon, C. Olson, J.E. Lambrechts, D. Moisse, M. Paridaens, R. Christiaens, M.-.R. Guénel, P. Truong, T. Laurent-Puig, P. Mulot, C. Marme, F. Burwinkel, B. Schneeweiss, A. Sohn, C. Sawyer, E.J. Tomlinson, I. Kerin, M.J. Miller, N. Andrulis, I.L. Knight, J.A. Tchatchou, S. Mulligan, A.M. Dörk, T. Bogdanova, N.V. Antonenkova, N.N. Anton-Culver, H. Darabi, H. Eriksson, M. Garcia-Closas, M. Figueroa, J. Lissowska, J. Brinton, L. Devilee, P. Tollenaar, R.A.E.M. Seynaeve, C. van Asperen, C.J. Kristensen, V.N. kConFab Investigators, . Australian Ovarian Cancer Study Group, . Slager, S. Toland, A.E. Ambrosone, C.B. Yannoukakos, D. Lindblom, A. Margolin, S. Radice, P. Peterlongo, P. Barile, M. Mariani, P. Hooning, M.J. Martens, J.W.M. Collée, J.M. Jager, A. Jakubowska, A. Lubinski, J. Jaworska-Bieniek, K. Durda, K. Giles, G.G. McLean, C. Brauch, H. Brüning, T. Ko, Y.-.D. GENICA Network, . Brenner, H. Dieffenbach, A.K. Arndt, V. Stegmaier, C. Swerdlow, A. Ashworth, A. Orr, N. Jones, M. Simard, J. Goldberg, M.S. Labrèche, F. Dumont, M. Winqvist, R. Pylkäs, K. Jukkola-Vuorinen, A. Grip, M. Kataja, V. Kosma, V.-.M. Hartikainen, J.M. Mannermaa, A. Hamann, U. Chenevix-Trench, G. Blomqvist, C. Aittomäki, K. Easton, D.F. Nevanlinna, H (2014) MicroRNA related polymorphisms and breast cancer risk.. Show Abstract full text

Genetic variations, such as single nucleotide polymorphisms (SNPs) in microRNAs (miRNA) or in the miRNA binding sites may affect the miRNA dependent gene expression regulation, which has been implicated in various cancers, including breast cancer, and may alter individual susceptibility to cancer. We investigated associations between miRNA related SNPs and breast cancer risk. First we evaluated 2,196 SNPs in a case-control study combining nine genome wide association studies (GWAS). Second, we further investigated 42 SNPs with suggestive evidence for association using 41,785 cases and 41,880 controls from 41 studies included in the Breast Cancer Association Consortium (BCAC). Combining the GWAS and BCAC data within a meta-analysis, we estimated main effects on breast cancer risk as well as risks for estrogen receptor (ER) and age defined subgroups. Five miRNA binding site SNPs associated significantly with breast cancer risk: rs1045494 (odds ratio (OR) 0.92; 95% confidence interval (CI): 0.88-0.96), rs1052532 (OR 0.97; 95% CI: 0.95-0.99), rs10719 (OR 0.97; 95% CI: 0.94-0.99), rs4687554 (OR 0.97; 95% CI: 0.95-0.99, and rs3134615 (OR 1.03; 95% CI: 1.01-1.05) located in the 3' UTR of CASP8, HDDC3, DROSHA, MUSTN1, and MYCL1, respectively. DROSHA belongs to miRNA machinery genes and has a central role in initial miRNA processing. The remaining genes are involved in different molecular functions, including apoptosis and gene expression regulation. Further studies are warranted to elucidate whether the miRNA binding site SNPs are the causative variants for the observed risk effects.

Johnson, N. Dudbridge, F. Orr, N. Gibson, L. Jones, M.E. Schoemaker, M.J. Folkerd, E.J. Haynes, B.P. Hopper, J.L. Southey, M.C. Dite, G.S. Apicella, C. Schmidt, M.K. Broeks, A. Van't Veer, L.J. Atsma, F. Muir, K. Lophatananon, A. Fasching, P.A. Beckmann, M.W. Ekici, A.B. Renner, S.P. Sawyer, E. Tomlinson, I. Kerin, M. Miller, N. Burwinkel, B. Marme, F. Schneeweiss, A. Sohn, C. Guenel, P. Truong, T. Cordina, E. Menegaux, F. Bojesen, S.E. Nordestgaard, B.G. Flyger, H. Milne, R. Zamora, M.P. Arias Perez, J.I. Benitez, J. Bernstein, L. Anton-Culver, H. Ziogas, A. Clarke Dur, C. Brenner, H. Muller, H. Arndt, V. Dieffenbach, A.K. Meindl, A. Heil, J. Bartram, C.R. Schmutzler, R.K. Brauch, H. Justenhoven, C. Ko, Y.D. Network, G. Nevanlinna, H. Muranen, T.A. Aittomaki, K. Blomqvist, C. Matsuo, K. Dork, T. Bogdanova, N.V. Antonenkova, N.N. Lindblom, A. Mannermaa, A. Kataja, V. Kosma, V.M. Hartikainen, J.M. Chenevix-Trench, G. Beesley, J. kConFab, I. Australian Ovarian Cancer Study, G. Wu, A.H. Van den Berg, D. Tseng, C.C. Lambrechts, D. Smeets, D. Neven, P. Wildiers, H. Chang-Claude, J. Rudolph, A. Nickels, S. Flesch-Janys, D. Radice, P. Peterlongo, P. Bonanni, B. Pensotti, V. Couch, F.J. Olson, J.E. Wang, X. Fredericksen, Z. Pankratz, V.S. Giles, G.G. Severi, G. Baglietto, L. Haiman, C. Simard, J. Goldberg, M.S. Labreche, F. Dumont, M. Soucy, P. Teo, S. Yip, C.H. Phuah, S.Y. Cornes, B.K. Kristensen, V.N. Grenaker Alnaes, G. Borresen-Dale, A.L. Zheng, W. Winqvist, R. Pylkas, K. Jukkola-Vuorinen, A. Grip, M. Andrulis, I.L. Knight, J.A. Glendon, G. Mulligan, A.M. Devillee, P. Figueroa, J. Chanock, S.J. Lissowska, J. Sherman, M.E. Hall, P. Schoof, N. Hooning, M. Hollestelle, A. Oldenburg, R.A. Tilanus-Linthorst, M. Liu, J. Cox, A. Brock, I.W. Reed, M.W. Cross, S.S. Blot, W. Signorello, L.B. Pharoah, P.D. Dunning, A.M. Shah, M. Kang, D. Noh, D.Y. Park, S.K. Choi, J.Y. Hartman, M. Miao, H. Lim, W.Y. Tang, A. Hamann, U. Forsti, A. Rudiger, T. Ulmer, H.U. Jakubowska, A. Lubinski, J. Jaworska-Bieniek, K. Durda, K. Sangrajrang, S. Gaborieau, V. Brennan, P. McKay, J. Slager, S. Toland, A.E. Vachon, C. Yannoukakos, D. Shen, C.Y. Yu, J.C. Huang, C.S. Hou, M.F. Gonzalez-Neira, A. Tessier, D.C. Vincent, D. Bacot, F. Luccarini, C. Dennis, J. Michailidou, K. Bolla, M.K. Wang, J. Easton, D.F. Garcia-Closas, M. Dowsett, M. Ashworth, A. Swerdlow, A.J. Peto, J. dos Santos Silva, I. Fletcher, O (2014) Genetic variation at CYP3A is associated with age at menarche and breast cancer risk: a case-control study. Show Abstract full text

INTRODUCTION: We have previously shown that a tag single nucleotide polymorphism (rs10235235), which maps to the CYP3A locus (7q22.1), was associated with a reduction in premenopausal urinary estrone glucuronide levels and a modest reduction in risk of breast cancer in women age </=50 years. METHODS: We further investigated the association of rs10235235 with breast cancer risk in a large case control study of 47,346 cases and 47,570 controls from 52 studies participating in the Breast Cancer Association Consortium. Genotyping of rs10235235 was conducted using a custom Illumina Infinium array. Stratified analyses were conducted to determine whether this association was modified by age at diagnosis, ethnicity, age at menarche or tumor characteristics. RESULTS: We confirmed the association of rs10235235 with breast cancer risk for women of European ancestry but found no evidence that this association differed with age at diagnosis. Heterozygote and homozygote odds ratios (ORs) were OR = 0.98 (95% CI 0.94, 1.01; P = 0.2) and OR = 0.80 (95% CI 0.69, 0.93; P = 0.004), respectively (P(trend) = 0.02). There was no evidence of effect modification by tumor characteristics. rs10235235 was, however, associated with age at menarche in controls (P(trend) = 0.005) but not cases (P(trend) = 0.97). Consequently the association between rs10235235 and breast cancer risk differed according to age at menarche (P(het) = 0.02); the rare allele of rs10235235 was associated with a reduction in breast cancer risk for women who had their menarche age >/=15 years (OR(het) = 0.84, 95% CI 0.75, 0.94; OR(hom) = 0.81, 95% CI 0.51, 1.30; P(trend) = 0.002) but not for those who had their menarche age </=11 years (OR(het) = 1.06, 95% CI 0.95, 1.19, OR(hom) = 1.07, 95% CI 0.67, 1.72; P(trend) = 0.29). CONCLUSIONS: To our knowledge rs10235235 is the first single nucleotide polymorphism to be associated with both breast cancer risk and age at menarche consistent with the well-documented association between later age at menarche and a reduction in breast cancer risk. These associations are likely mediated via an effect on circulating hormone levels.

Bodicoat, D.H. Schoemaker, M.J. Jones, M.E. McFadden, E. Griffin, J. Ashworth, A. Swerdlow, A.J (2014) Timing of pubertal stages and breast cancer risk: the Breakthrough Generations Study.. Show Abstract full text

<h4>Introduction</h4>Breast development and hormonal changes at puberty might affect breast cancer risk, but epidemiological analyses have focussed largely on age at menarche and not at other pubertal stages.<h4>Methods</h4>We investigated associations between the timing of pubertal stages and breast cancer risk using data from a cohort study of 104,931 women (Breakthrough Generations Study, UK, 2003-2013). Pubertal variables were reported retrospectively at baseline. Breast cancer risk was analysed using Cox regression models with breast cancer diagnosis as the outcome of interest, attained age as the underlying time variable, and adjustment for potentially confounding variables.<h4>Results</h4>During follow-up (mean = 4.1 years), 1094 breast cancers (including ductal carcinoma in situ) occurred. An increased breast cancer risk was associated with earlier thelarche (age when breast growth begins; HR [95% CI] = 1.23 [1.02, 1.48], 1 [referent] and 0.80 [0.69, 0.93] for ≤10, 11-12 and ≥13 years respectively), menarche (initiation of menses; 1.06 [0.93, 1.21], 1 [referent] and 0.78 [0.62, 0.99] for ≤12, 13-14 and ≥15 years), regular periods (0.99 [0.83, 1.18], 1 [referent] and 0.74 [0.59, 0.92] for ≤12, 13-14 and ≥15 years) and age reached adult height (1.25 [1.03, 1.52], 1 [referent] and 1.07 [0.87, 1.32] for ≤14, 15-16 and ≥17 years), and with increased time between thelarche and menarche (0.87 [0.65, 1.15], 1 [referent], 1.14 [0.96, 1.34] and 1.27 [1.04, 1.55] for <0, 0, 1 and ≥2 years), and shorter time between menarche and regular periods (1 [referent], 0.87 [0.73, 1.04] and 0.66 [0.50, 0.88] for 0, 1 and ≥2 years). These associations were generally similar when considered separately for premenopausal and postmenopausal breast cancer.<h4>Conclusions</h4>Breast duct development may be a time of heightened susceptibility to risk of carcinogenesis, and greater attention needs to be given to the relation of breast cancer risk to the different stages of puberty.

Agarwal, D. Pineda, S. Michailidou, K. Herranz, J. Pita, G. Moreno, L.T. Alonso, M.R. Dennis, J. Wang, Q. Bolla, M.K. Meyer, K.B. Menéndez-Rodríguez, P. Hardisson, D. Mendiola, M. González-Neira, A. Lindblom, A. Margolin, S. Swerdlow, A. Ashworth, A. Orr, N. Jones, M. Matsuo, K. Ito, H. Iwata, H. Kondo, N. kConFab Investigators, . Australian Ovarian Cancer Study Group, . Hartman, M. Hui, M. Lim, W.Y. Iau, P.T.-.C. Sawyer, E. Tomlinson, I. Kerin, M. Miller, N. Kang, D. Choi, J.-.Y. Park, S.K. Noh, D.-.Y. Hopper, J.L. Schmidt, D.F. Makalic, E. Southey, M.C. Teo, S.H. Yip, C.H. Sivanandan, K. Tay, W.-.T. Brauch, H. Brüning, T. Hamann, U. GENICA Network, . Dunning, A.M. Shah, M. Andrulis, I.L. Knight, J.A. Glendon, G. Tchatchou, S. Schmidt, M.K. Broeks, A. Rosenberg, E.H. van't Veer, L.J. Fasching, P.A. Renner, S.P. Ekici, A.B. Beckmann, M.W. Shen, C.-.Y. Hsiung, C.-.N. Yu, J.-.C. Hou, M.-.F. Blot, W. Cai, Q. Wu, A.H. Tseng, C.-.C. Van Den Berg, D. Stram, D.O. Cox, A. Brock, I.W. Reed, M.W.R. Muir, K. Lophatananon, A. Stewart-Brown, S. Siriwanarangsan, P. Zheng, W. Deming-Halverson, S. Shrubsole, M.J. Long, J. Shu, X.-.O. Lu, W. Gao, Y.-.T. Zhang, B. Radice, P. Peterlongo, P. Manoukian, S. Mariette, F. Sangrajrang, S. McKay, J. Couch, F.J. Toland, A.E. TNBCC, . Yannoukakos, D. Fletcher, O. Johnson, N. dos Santos Silva, I. Peto, J. Marme, F. Burwinkel, B. Guénel, P. Truong, T. Sanchez, M. Mulot, C. Bojesen, S.E. Nordestgaard, B.G. Flyer, H. Brenner, H. Dieffenbach, A.K. Arndt, V. Stegmaier, C. Mannermaa, A. Kataja, V. Kosma, V.-.M. Hartikainen, J.M. Lambrechts, D. Yesilyurt, B.T. Floris, G. Leunen, K. Chang-Claude, J. Rudolph, A. Seibold, P. Flesch-Janys, D. Wang, X. Olson, J.E. Vachon, C. Purrington, K. Giles, G.G. Severi, G. Baglietto, L. Haiman, C.A. Henderson, B.E. Schumacher, F. Marchand, L.L. Simard, J. Dumont, M. Goldberg, M.S. Labréche, F. Winqvist, R. Pylkäs, K. Jukkola-Vuorinen, A. Grip, M. Devilee, P. Tollenaar, R.A.E.M. Seynaeve, C. García-Closas, M. Chanock, S.J. Lissowska, J. Figueroa, J.D. Czene, K. Eriksson, M. Humphreys, K. Darabi, H. Hooning, M.J. Kriege, M. Collée, J.M. Tilanus-Linthorst, M. Li, J. Jakubowska, A. Lubinski, J. Jaworska-Bieniek, K. Durda, K. Nevanlinna, H. Muranen, T.A. Aittomäki, K. Blomqvist, C. Bogdanova, N. Dörk, T. Hall, P. Chenevix-Trench, G. Easton, D.F. Pharroah, P.D.P. Arias-Perez, J.I. Zamora, P. Benítez, J. Milne, R.L (2014) FGF receptor genes and breast cancer susceptibility: results from the Breast Cancer Association Consortium.. Show Abstract full text

<h4>Background</h4>Breast cancer is one of the most common malignancies in women. Genome-wide association studies have identified FGFR2 as a breast cancer susceptibility gene. Common variation in other fibroblast growth factor (FGF) receptors might also modify risk. We tested this hypothesis by studying genotyped single-nucleotide polymorphisms (SNPs) and imputed SNPs in FGFR1, FGFR3, FGFR4 and FGFRL1 in the Breast Cancer Association Consortium.<h4>Methods</h4>Data were combined from 49 studies, including 53 835 cases and 50 156 controls, of which 89 050 (46 450 cases and 42 600 controls) were of European ancestry, 12 893 (6269 cases and 6624 controls) of Asian and 2048 (1116 cases and 932 controls) of African ancestry. Associations with risk of breast cancer, overall and by disease sub-type, were assessed using unconditional logistic regression.<h4>Results</h4>Little evidence of association with breast cancer risk was observed for SNPs in the FGF receptor genes. The strongest evidence in European women was for rs743682 in FGFR3; the estimated per-allele odds ratio was 1.05 (95% confidence interval=1.02-1.09, P=0.0020), which is substantially lower than that observed for SNPs in FGFR2.<h4>Conclusion</h4>Our results suggest that common variants in the other FGF receptors are not associated with risk of breast cancer to the degree observed for FGFR2.

Perry, J.R.B. Corre, T. Esko, T. Chasman, D.I. Fischer, K. Franceschini, N. He, C. Kutalik, Z. Mangino, M. Rose, L.M. Vernon Smith, A. Stolk, L. Sulem, P. Weedon, M.N. Zhuang, W.V. Arnold, A. Ashworth, A. Bergmann, S. Buring, J.E. Burri, A. Chen, C. Cornelis, M.C. Couper, D.J. Goodarzi, M.O. Gudnason, V. Harris, T. Hofman, A. Jones, M. Kraft, P. Launer, L. Laven, J.S.E. Li, G. McKnight, B. Masciullo, C. Milani, L. Orr, N. Psaty, B.M. ReproGen Consortium, . Ridker, P.M. Rivadeneira, F. Sala, C. Salumets, A. Schoemaker, M. Traglia, M. Waeber, G. Chanock, S.J. Demerath, E.W. Garcia, M. Hankinson, S.E. Hu, F.B. Hunter, D.J. Lunetta, K.L. Metspalu, A. Montgomery, G.W. Murabito, J.M. Newman, A.B. Ong, K.K. Spector, T.D. Stefansson, K. Swerdlow, A.J. Thorsteinsdottir, U. Van Dam, R.M. Uitterlinden, A.G. Visser, J.A. Vollenweider, P. Toniolo, D. Murray, A (2013) A genome-wide association study of early menopause and the combined impact of identified variants.. Show Abstract full text

Early menopause (EM) affects up to 10% of the female population, reducing reproductive lifespan considerably. Currently, it constitutes the leading cause of infertility in the western world, affecting mainly those women who postpone their first pregnancy beyond the age of 30 years. The genetic aetiology of EM is largely unknown in the majority of cases. We have undertaken a meta-analysis of genome-wide association studies (GWASs) in 3493 EM cases and 13 598 controls from 10 independent studies. No novel genetic variants were discovered, but the 17 variants previously associated with normal age at natural menopause as a quantitative trait (QT) were also associated with EM and primary ovarian insufficiency (POI). Thus, EM has a genetic aetiology which overlaps variation in normal age at menopause and is at least partly explained by the additive effects of the same polygenic variants. The combined effect of the common variants captured by the single nucleotide polymorphism arrays was estimated to account for ∼30% of the variance in EM. The association between the combined 17 variants and the risk of EM was greater than the best validated non-genetic risk factor, smoking.

Garcia-Closas, M. Couch, F.J. Lindstrom, S. Michailidou, K. Schmidt, M.K. Brook, M.N. Orr, N. Rhie, S.K. Riboli, E. Feigelson, H.S. Le Marchand, L. Buring, J.E. Eccles, D. Miron, P. Fasching, P.A. Brauch, H. Chang-Claude, J. Carpenter, J. Godwin, A.K. Nevanlinna, H. Giles, G.G. Cox, A. Hopper, J.L. Bolla, M.K. Wang, Q. Dennis, J. Dicks, E. Howat, W.J. Schoof, N. Bojesen, S.E. Lambrechts, D. Broeks, A. Andrulis, I.L. Guénel, P. Burwinkel, B. Sawyer, E.J. Hollestelle, A. Fletcher, O. Winqvist, R. Brenner, H. Mannermaa, A. Hamann, U. Meindl, A. Lindblom, A. Zheng, W. Devillee, P. Goldberg, M.S. Lubinski, J. Kristensen, V. Swerdlow, A. Anton-Culver, H. Dörk, T. Muir, K. Matsuo, K. Wu, A.H. Radice, P. Teo, S.H. Shu, X.-.O. Blot, W. Kang, D. Hartman, M. Sangrajrang, S. Shen, C.-.Y. Southey, M.C. Park, D.J. Hammet, F. Stone, J. Veer, L.J.V. Rutgers, E.J. Lophatananon, A. Stewart-Brown, S. Siriwanarangsan, P. Peto, J. Schrauder, M.G. Ekici, A.B. Beckmann, M.W. Dos Santos Silva, I. Johnson, N. Warren, H. Tomlinson, I. Kerin, M.J. Miller, N. Marme, F. Schneeweiss, A. Sohn, C. Truong, T. Laurent-Puig, P. Kerbrat, P. Nordestgaard, B.G. Nielsen, S.F. Flyger, H. Milne, R.L. Perez, J.I.A. Menéndez, P. Müller, H. Arndt, V. Stegmaier, C. Lichtner, P. Lochmann, M. Justenhoven, C. Ko, Y.-.D. Gene ENvironmental Interaction and breast CAncer (GENICA) Network, . Muranen, T.A. Aittomäki, K. Blomqvist, C. Greco, D. Heikkinen, T. Ito, H. Iwata, H. Yatabe, Y. Antonenkova, N.N. Margolin, S. Kataja, V. Kosma, V.-.M. Hartikainen, J.M. Balleine, R. kConFab Investigators, . Tseng, C.-.C. Berg, D.V.D. Stram, D.O. Neven, P. Dieudonné, A.-.S. Leunen, K. Rudolph, A. Nickels, S. Flesch-Janys, D. Peterlongo, P. Peissel, B. Bernard, L. Olson, J.E. Wang, X. Stevens, K. Severi, G. Baglietto, L. McLean, C. Coetzee, G.A. Feng, Y. Henderson, B.E. Schumacher, F. Bogdanova, N.V. Labrèche, F. Dumont, M. Yip, C.H. Taib, N.A.M. Cheng, C.-.Y. Shrubsole, M. Long, J. Pylkäs, K. Jukkola-Vuorinen, A. Kauppila, S. Knight, J.A. Glendon, G. Mulligan, A.M. Tollenaar, R.A.E.M. Seynaeve, C.M. Kriege, M. Hooning, M.J. van den Ouweland, A.M.W. van Deurzen, C.H.M. Lu, W. Gao, Y.-.T. Cai, H. Balasubramanian, S.P. Cross, S.S. Reed, M.W.R. Signorello, L. Cai, Q. Shah, M. Miao, H. Chan, C.W. Chia, K.S. Jakubowska, A. Jaworska, K. Durda, K. Hsiung, C.-.N. Wu, P.-.E. Yu, J.-.C. Ashworth, A. Jones, M. Tessier, D.C. González-Neira, A. Pita, G. Alonso, M.R. Vincent, D. Bacot, F. Ambrosone, C.B. Bandera, E.V. John, E.M. Chen, G.K. Hu, J.J. Rodriguez-Gil, J.L. Bernstein, L. Press, M.F. Ziegler, R.G. Millikan, R.M. Deming-Halverson, S.L. Nyante, S. Ingles, S.A. Waisfisz, Q. Tsimiklis, H. Makalic, E. Schmidt, D. Bui, M. Gibson, L. Müller-Myhsok, B. Schmutzler, R.K. Hein, R. Dahmen, N. Beckmann, L. Aaltonen, K. Czene, K. Irwanto, A. Liu, J. Turnbull, C. Familial Breast Cancer Study (FBCS), . Rahman, N. Meijers-Heijboer, H. Uitterlinden, A.G. Rivadeneira, F. Australian Breast Cancer Tissue Bank (ABCTB) Investigators, . Olswold, C. Slager, S. Pilarski, R. Ademuyiwa, F. Konstantopoulou, I. Martin, N.G. Montgomery, G.W. Slamon, D.J. Rauh, C. Lux, M.P. Jud, S.M. Bruning, T. Weaver, J. Sharma, P. Pathak, H. Tapper, W. Gerty, S. Durcan, L. Trichopoulos, D. Tumino, R. Peeters, P.H. Kaaks, R. Campa, D. Canzian, F. Weiderpass, E. Johansson, M. Khaw, K.-.T. Travis, R. Clavel-Chapelon, F. Kolonel, L.N. Chen, C. Beck, A. Hankinson, S.E. Berg, C.D. Hoover, R.N. Lissowska, J. Figueroa, J.D. Chasman, D.I. Gaudet, M.M. Diver, W.R. Willett, W.C. Hunter, D.J. Simard, J. Benitez, J. Dunning, A.M. Sherman, M.E. Chenevix-Trench, G. Chanock, S.J. Hall, P. Pharoah, P.D.P. Vachon, C. Easton, D.F. Haiman, C.A. Kraft, P (2013) Genome-wide association studies identify four ER negative-specific breast cancer risk loci.. Show Abstract full text

Estrogen receptor (ER)-negative tumors represent 20-30% of all breast cancers, with a higher proportion occurring in younger women and women of African ancestry. The etiology and clinical behavior of ER-negative tumors are different from those of tumors expressing ER (ER positive), including differences in genetic predisposition. To identify susceptibility loci specific to ER-negative disease, we combined in a meta-analysis 3 genome-wide association studies of 4,193 ER-negative breast cancer cases and 35,194 controls with a series of 40 follow-up studies (6,514 cases and 41,455 controls), genotyped using a custom Illumina array, iCOGS, developed by the Collaborative Oncological Gene-environment Study (COGS). SNPs at four loci, 1q32.1 (MDM4, P = 2.1 × 10(-12) and LGR6, P = 1.4 × 10(-8)), 2p24.1 (P = 4.6 × 10(-8)) and 16q12.2 (FTO, P = 4.0 × 10(-8)), were associated with ER-negative but not ER-positive breast cancer (P > 0.05). These findings provide further evidence for distinct etiological pathways associated with invasive ER-positive and ER-negative breast cancers.

Bojesen, S.E. Pooley, K.A. Johnatty, S.E. Beesley, J. Michailidou, K. Tyrer, J.P. Edwards, S.L. Pickett, H.A. Shen, H.C. Smart, C.E. Hillman, K.M. Mai, P.L. Lawrenson, K. Stutz, M.D. Lu, Y. Karevan, R. Woods, N. Johnston, R.L. French, J.D. Chen, X. Weischer, M. Nielsen, S.F. Maranian, M.J. Ghoussaini, M. Ahmed, S. Baynes, C. Bolla, M.K. Wang, Q. Dennis, J. McGuffog, L. Barrowdale, D. Lee, A. Healey, S. Lush, M. Tessier, D.C. Vincent, D. Bacot, F. Australian Cancer Study, . Australian Ovarian Cancer Study, . Kathleen Cuningham Foundation Consortium for Research into Familial Breast Cancer (kConFab), . Gene Environment Interaction and Breast Cancer (GENICA), . Swedish Breast Cancer Study (SWE-BRCA), . Hereditary Breast and Ovarian Cancer Research Group Netherlands (HEBON), . Epidemiological study of BRCA1 & BRCA2 Mutation Carriers (EMBRACE), . Genetic Modifiers of Cancer Risk in BRCA1/2 Mutation Carriers (GEMO), . Vergote, I. Lambrechts, S. Despierre, E. Risch, H.A. González-Neira, A. Rossing, M.A. Pita, G. Doherty, J.A. Alvarez, N. Larson, M.C. Fridley, B.L. Schoof, N. Chang-Claude, J. Cicek, M.S. Peto, J. Kalli, K.R. Broeks, A. Armasu, S.M. Schmidt, M.K. Braaf, L.M. Winterhoff, B. Nevanlinna, H. Konecny, G.E. Lambrechts, D. Rogmann, L. Guénel, P. Teoman, A. Milne, R.L. Garcia, J.J. Cox, A. Shridhar, V. Burwinkel, B. Marme, F. Hein, R. Sawyer, E.J. Haiman, C.A. Wang-Gohrke, S. Andrulis, I.L. Moysich, K.B. Hopper, J.L. Odunsi, K. Lindblom, A. Giles, G.G. Brenner, H. Simard, J. Lurie, G. Fasching, P.A. Carney, M.E. Radice, P. Wilkens, L.R. Swerdlow, A. Goodman, M.T. Brauch, H. Garcia-Closas, M. Hillemanns, P. Winqvist, R. Dürst, M. Devilee, P. Runnebaum, I. Jakubowska, A. Lubinski, J. Mannermaa, A. Butzow, R. Bogdanova, N.V. Dörk, T. Pelttari, L.M. Zheng, W. Leminen, A. Anton-Culver, H. Bunker, C.H. Kristensen, V. Ness, R.B. Muir, K. Edwards, R. Meindl, A. Heitz, F. Matsuo, K. du Bois, A. Wu, A.H. Harter, P. Teo, S.-.H. Schwaab, I. Shu, X.-.O. Blot, W. Hosono, S. Kang, D. Nakanishi, T. Hartman, M. Yatabe, Y. Hamann, U. Karlan, B.Y. Sangrajrang, S. Kjaer, S.K. Gaborieau, V. Jensen, A. Eccles, D. Høgdall, E. Shen, C.-.Y. Brown, J. Woo, Y.L. Shah, M. Azmi, M.A.N. Luben, R. Omar, S.Z. Czene, K. Vierkant, R.A. Nordestgaard, B.G. Flyger, H. Vachon, C. Olson, J.E. Wang, X. Levine, D.A. Rudolph, A. Weber, R.P. Flesch-Janys, D. Iversen, E. Nickels, S. Schildkraut, J.M. Silva, I.D.S. Cramer, D.W. Gibson, L. Terry, K.L. Fletcher, O. Vitonis, A.F. van der Schoot, C.E. Poole, E.M. Hogervorst, F.B.L. Tworoger, S.S. Liu, J. Bandera, E.V. Li, J. Olson, S.H. Humphreys, K. Orlow, I. Blomqvist, C. Rodriguez-Rodriguez, L. Aittomäki, K. Salvesen, H.B. Muranen, T.A. Wik, E. Brouwers, B. Krakstad, C. Wauters, E. Halle, M.K. Wildiers, H. Kiemeney, L.A. Mulot, C. Aben, K.K. Laurent-Puig, P. Altena, A.M. Truong, T. Massuger, L.F.A.G. Benitez, J. Pejovic, T. Perez, J.I.A. Hoatlin, M. Zamora, M.P. Cook, L.S. Balasubramanian, S.P. Kelemen, L.E. Schneeweiss, A. Le, N.D. Sohn, C. Brooks-Wilson, A. Tomlinson, I. Kerin, M.J. Miller, N. Cybulski, C. Henderson, B.E. Menkiszak, J. Schumacher, F. Wentzensen, N. Le Marchand, L. Yang, H.P. Mulligan, A.M. Glendon, G. Engelholm, S.A. Knight, J.A. Høgdall, C.K. Apicella, C. Gore, M. Tsimiklis, H. Song, H. Southey, M.C. Jager, A. den Ouweland, A.M.W. Brown, R. Martens, J.W.M. Flanagan, J.M. Kriege, M. Paul, J. Margolin, S. Siddiqui, N. Severi, G. Whittemore, A.S. Baglietto, L. McGuire, V. Stegmaier, C. Sieh, W. Müller, H. Arndt, V. Labrèche, F. Gao, Y.-.T. Goldberg, M.S. Yang, G. Dumont, M. McLaughlin, J.R. Hartmann, A. Ekici, A.B. Beckmann, M.W. Phelan, C.M. Lux, M.P. Permuth-Wey, J. Peissel, B. Sellers, T.A. Ficarazzi, F. Barile, M. Ziogas, A. Ashworth, A. Gentry-Maharaj, A. Jones, M. Ramus, S.J. Orr, N. Menon, U. Pearce, C.L. Brüning, T. Pike, M.C. Ko, Y.-.D. Lissowska, J. Figueroa, J. Kupryjanczyk, J. Chanock, S.J. Dansonka-Mieszkowska, A. Jukkola-Vuorinen, A. Rzepecka, I.K. Pylkäs, K. Bidzinski, M. Kauppila, S. Hollestelle, A. Seynaeve, C. Tollenaar, R.A.E.M. Durda, K. Jaworska, K. Hartikainen, J.M. Kosma, V.-.M. Kataja, V. Antonenkova, N.N. Long, J. Shrubsole, M. Deming-Halverson, S. Lophatananon, A. Siriwanarangsan, P. Stewart-Brown, S. Ditsch, N. Lichtner, P. Schmutzler, R.K. Ito, H. Iwata, H. Tajima, K. Tseng, C.-.C. Stram, D.O. van den Berg, D. Yip, C.H. Ikram, M.K. Teh, Y.-.C. Cai, H. Lu, W. Signorello, L.B. Cai, Q. Noh, D.-.Y. Yoo, K.-.Y. Miao, H. Iau, P.T.-.C. Teo, Y.Y. McKay, J. Shapiro, C. Ademuyiwa, F. Fountzilas, G. Hsiung, C.-.N. Yu, J.-.C. Hou, M.-.F. Healey, C.S. Luccarini, C. Peock, S. Stoppa-Lyonnet, D. Peterlongo, P. Rebbeck, T.R. Piedmonte, M. Singer, C.F. Friedman, E. Thomassen, M. Offit, K. Hansen, T.V.O. Neuhausen, S.L. Szabo, C.I. Blanco, I. Garber, J. Narod, S.A. Weitzel, J.N. Montagna, M. Olah, E. Godwin, A.K. Yannoukakos, D. Goldgar, D.E. Caldes, T. Imyanitov, E.N. Tihomirova, L. Arun, B.K. Campbell, I. Mensenkamp, A.R. van Asperen, C.J. van Roozendaal, K.E.P. Meijers-Heijboer, H. Collée, J.M. Oosterwijk, J.C. Hooning, M.J. Rookus, M.A. van der Luijt, R.B. Os, T.A.M. Evans, D.G. Frost, D. Fineberg, E. Barwell, J. Walker, L. Kennedy, M.J. Platte, R. Davidson, R. Ellis, S.D. Cole, T. Bressac-de Paillerets, B. Buecher, B. Damiola, F. Faivre, L. Frenay, M. Sinilnikova, O.M. Caron, O. Giraud, S. Mazoyer, S. Bonadona, V. Caux-Moncoutier, V. Toloczko-Grabarek, A. Gronwald, J. Byrski, T. Spurdle, A.B. Bonanni, B. Zaffaroni, D. Giannini, G. Bernard, L. Dolcetti, R. Manoukian, S. Arnold, N. Engel, C. Deissler, H. Rhiem, K. Niederacher, D. Plendl, H. Sutter, C. Wappenschmidt, B. Borg, A. Melin, B. Rantala, J. Soller, M. Nathanson, K.L. Domchek, S.M. Rodriguez, G.C. Salani, R. Kaulich, D.G. Tea, M.-.K. Paluch, S.S. Laitman, Y. Skytte, A.-.B. Kruse, T.A. Jensen, U.B. Robson, M. Gerdes, A.-.M. Ejlertsen, B. Foretova, L. Savage, S.A. Lester, J. Soucy, P. Kuchenbaecker, K.B. Olswold, C. Cunningham, J.M. Slager, S. Pankratz, V.S. Dicks, E. Lakhani, S.R. Couch, F.J. Hall, P. Monteiro, A.N.A. Gayther, S.A. Pharoah, P.D.P. Reddel, R.R. Goode, E.L. Greene, M.H. Easton, D.F. Berchuck, A. Antoniou, A.C. Chenevix-Trench, G. Dunning, A.M (2013) Multiple independent variants at the TERT locus are associated with telomere length and risks of breast and ovarian cancer.. Show Abstract full text

TERT-locus SNPs and leukocyte telomere measures are reportedly associated with risks of multiple cancers. Using the Illumina custom genotyping array iCOGs, we analyzed ∼480 SNPs at the TERT locus in breast (n = 103,991), ovarian (n = 39,774) and BRCA1 mutation carrier (n = 11,705) cancer cases and controls. Leukocyte telomere measurements were also available for 53,724 participants. Most associations cluster into three independent peaks. The minor allele at the peak 1 SNP rs2736108 associates with longer telomeres (P = 5.8 × 10(-7)), lower risks for estrogen receptor (ER)-negative (P = 1.0 × 10(-8)) and BRCA1 mutation carrier (P = 1.1 × 10(-5)) breast cancers and altered promoter assay signal. The minor allele at the peak 2 SNP rs7705526 associates with longer telomeres (P = 2.3 × 10(-14)), higher risk of low-malignant-potential ovarian cancer (P = 1.3 × 10(-15)) and greater promoter activity. The minor alleles at the peak 3 SNPs rs10069690 and rs2242652 increase ER-negative (P = 1.2 × 10(-12)) and BRCA1 mutation carrier (P = 1.6 × 10(-14)) breast and invasive ovarian (P = 1.3 × 10(-11)) cancer risks but not via altered telomere length. The cancer risk alleles of rs2242652 and rs10069690, respectively, increase silencing and generate a truncated TERT splice variant.

Warren, H. Dudbridge, F. Fletcher, O. Orr, N. Johnson, N. Hopper, J.L. Apicella, C. Southey, M.C. Mahmoodi, M. Schmidt, M.K. Broeks, A. Cornelissen, S. Braaf, L.M. Muir, K.R. Lophatananon, A. Chaiwerawattana, A. Wiangnon, S. Fasching, P.A. Beckmann, M.W. Ekici, A.B. Schulz-Wendtland, R. Sawyer, E.J. Tomlinson, I. Kerin, M. Burwinkel, B. Marme, F. Schneeweiss, A. Sohn, C. Guénel, P. Truong, T. Laurent-Puig, P. Mulot, C. Bojesen, S.E. Nielsen, S.F. Flyger, H. Nordestgaard, B.G. Milne, R.L. Benítez, J. Arias-Pérez, J.-.I. Zamora, M.P. Anton-Culver, H. Ziogas, A. Bernstein, L. Dur, C.C. Brenner, H. Müller, H. Arndt, V. Langheinz, A. Meindl, A. Golatta, M. Bartram, C.R. Schmutzler, R.K. Brauch, H. Justenhoven, C. Brüning, T. GENICA Network, . Chang-Claude, J. Wang-Gohrke, S. Eilber, U. Dörk, T. Schürmann, P. Bremer, M. Hillemanns, P. Nevanlinna, H. Muranen, T.A. Aittomäki, K. Blomqvist, C. Bogdanova, N. Antonenkova, N. Rogov, Y. Bermisheva, M. Prokofyeva, D. Zinnatullina, G. Khusnutdinova, E. Lindblom, A. Margolin, S. Mannermaa, A. Kosma, V.-.M. Hartikainen, J.M. Kataja, V. Chenevix-Trench, G. Beesley, J. Chen, X. kConFab Investigators, . Australian Ovarian Cancer Study Group, . Lambrechts, D. Smeets, A. Paridaens, R. Weltens, C. Flesch-Janys, D. Buck, K. Behrens, S. Peterlongo, P. Bernard, L. Manoukian, S. Radice, P. Couch, F.J. Vachon, C. Wang, X. Olson, J. Giles, G. Baglietto, L. McLean, C.A. Severi, G. John, E.M. Miron, A. Winqvist, R. Pylkäs, K. Jukkola-Vuorinen, A. Grip, M. Andrulis, I.L. Knight, J.A. Mulligan, A.M. Weerasooriya, N. Devilee, P. Tollenaar, R.A.E.M. Martens, J.W.M. Seynaeve, C.M. Hooning, M.J. Hollestelle, A. Jager, A. Tilanus-Linthorst, M.M.A. Hall, P. Czene, K. Liu, J. Li, J. Cox, A. Cross, S.S. Brock, I.W. Reed, M.W.R. Pharoah, P. Blows, F.M. Dunning, A.M. Ghoussaini, M. Ashworth, A. Swerdlow, A. Jones, M. Schoemaker, M. Easton, D.F. Humphreys, M. Wang, Q. Peto, J. dos-Santos-Silva, I (2012) 9q31.2-rs865686 as a susceptibility locus for estrogen receptor-positive breast cancer: evidence from the Breast Cancer Association Consortium.. Show Abstract full text

BACKGROUND: Our recent genome-wide association study identified a novel breast cancer susceptibility locus at 9q31.2 (rs865686). METHODS: To further investigate the rs865686-breast cancer association, we conducted a replication study within the Breast Cancer Association Consortium, which comprises 37 case-control studies (48,394 cases, 50,836 controls). RESULTS: This replication study provides additional strong evidence of an inverse association between rs865686 and breast cancer risk [study-adjusted per G-allele OR, 0.90; 95% confidence interval (CI), 0.88; 0.91, P = 2.01 × 10(-29)] among women of European ancestry. There were ethnic differences in the estimated minor (G)-allele frequency among controls [0.09, 0.30, and 0.38 among, respectively, Asians, Eastern Europeans, and other Europeans; P for heterogeneity (P(het)) = 1.3 × 10(-143)], but no evidence of ethnic differences in per allele OR (P(het) = 0.43). rs865686 was associated with estrogen receptor-positive (ER(+)) disease (per G-allele OR, 0.89; 95% CI, 0.86-0.91; P = 3.13 × 10(-22)) but less strongly, if at all, with ER-negative (ER(-)) disease (OR, 0.98; 95% CI, 0.94-1.02; P = 0.26; P(het) = 1.16 × 10(-6)), with no evidence of independent heterogeneity by progesterone receptor or HER2 status. The strength of the breast cancer association decreased with increasing age at diagnosis, with case-only analysis showing a trend in the number of copies of the G allele with increasing age at diagnosis (P for linear trend = 0.0095), but only among women with ER(+) tumors. CONCLUSIONS: This study is the first to show that rs865686 is a susceptibility marker for ER(+) breast cancer. IMPACT: The findings further support the view that genetic susceptibility varies according to tumor subtype.

Orr, N. Lemnrau, A. Cooke, R. Fletcher, O. Tomczyk, K. Jones, M. Johnson, N. Lord, C.J. Mitsopoulos, C. Zvelebil, M. McDade, S.S. Buck, G. Blancher, C. KConFab Consortium, . Trainer, A.H. James, P.A. Bojesen, S.E. Bokmand, S. Nevanlinna, H. Mattson, J. Friedman, E. Laitman, Y. Palli, D. Masala, G. Zanna, I. Ottini, L. Giannini, G. Hollestelle, A. Ouweland, A.M.W.V.D. Novaković, S. Krajc, M. Gago-Dominguez, M. Castelao, J.E. Olsson, H. Hedenfalk, I. Easton, D.F. Pharoah, P.D.P. Dunning, A.M. Bishop, D.T. Neuhausen, S.L. Steele, L. Houlston, R.S. Garcia-Closas, M. Ashworth, A. Swerdlow, A.J (2012) Genome-wide association study identifies a common variant in RAD51B associated with male breast cancer risk.. Show Abstract full text

We conducted a genome-wide association study of male breast cancer comprising 823 cases and 2,795 controls of European ancestry, with validation in independent sample sets totaling 438 cases and 474 controls. A SNP in RAD51B at 14q24.1 was significantly associated with male breast cancer risk (P = 3.02 × 10(-13); odds ratio (OR) = 1.57). We also refine association at 16q12.1 to a SNP within TOX3 (P = 3.87 × 10(-15); OR = 1.50).

Johnson, N. Walker, K. Gibson, L.J. Orr, N. Folkerd, E. Haynes, B. Palles, C. Coupland, B. Schoemaker, M. Jones, M. Broderick, P. Sawyer, E. Kerin, M. Tomlinson, I.P. Zvelebil, M. Chilcott-Burns, S. Tomczyk, K. Simpson, G. Williamson, J. Hillier, S.G. Ross, G. Houlston, R.S. Swerdlow, A. Ashworth, A. Dowsett, M. Peto, J. Dos Santos Silva, I. Fletcher, O (2012) CYP3A variation, premenopausal estrone levels, and breast cancer risk.. Show Abstract full text

BACKGROUND: Epidemiological studies have provided strong evidence for a role of endogenous sex steroids in the etiology of breast cancer. Our aim was to identify common variants in genes involved in sex steroid synthesis or metabolism that are associated with hormone levels and the risk of breast cancer in premenopausal women. METHODS: We measured urinary levels of estrone glucuronide (E1G) using a protocol specifically developed to account for cyclic variation in hormone levels during the menstrual cycle in 729 healthy premenopausal women. We genotyped 642 single-nucleotide polymorphisms (SNPs) in these women; a single SNP, rs10273424, was further tested for association with the risk of breast cancer using data from 10 551 breast cancer case patients and 17 535 control subjects. All statistical tests were two-sided. RESULTS: rs10273424, which maps approximately 50 kb centromeric to the cytochrome P450 3A (CYP3A) gene cluster at chromosome 7q22.1, was associated with a 21.8% reduction in E1G levels (95% confidence interval [CI] = 27.8% to 15.3% reduction; P = 2.7 × 10(-9)) and a modest reduction in the risk of breast cancer in case patients who were diagnosed at or before age 50 years (odds ratio [OR] = 0.91, 95% CI = 0.83 to 0.99; P = .03) but not in those diagnosed after age 50 years (OR = 1.01, 95% CI = 0.93 to 1.10; P = .82). CONCLUSIONS: Genetic variation in noncoding sequences flanking the CYP3A locus contributes to variance in premenopausal E1G levels and is associated with the risk of breast cancer in younger patients. This association may have wider implications given that the most predominantly expressed CYP3A gene, CYP3A4, is responsible for metabolism of endogenous and exogenous hormones and hormonal agents used in the treatment of breast cancer.

Ghoussaini, M. Fletcher, O. Michailidou, K. Turnbull, C. Schmidt, M.K. Dicks, E. Dennis, J. Wang, Q. Humphreys, M.K. Luccarini, C. Baynes, C. Conroy, D. Maranian, M. Ahmed, S. Driver, K. Johnson, N. Orr, N. dos Santos Silva, I. Waisfisz, Q. Meijers-Heijboer, H. Uitterlinden, A.G. Rivadeneira, F. Netherlands Collaborative Group on Hereditary Breast and Ovarian Cancer (HEBON), . Hall, P. Czene, K. Irwanto, A. Liu, J. Nevanlinna, H. Aittomäki, K. Blomqvist, C. Meindl, A. Schmutzler, R.K. Müller-Myhsok, B. Lichtner, P. Chang-Claude, J. Hein, R. Nickels, S. Flesch-Janys, D. Tsimiklis, H. Makalic, E. Schmidt, D. Bui, M. Hopper, J.L. Apicella, C. Park, D.J. Southey, M. Hunter, D.J. Chanock, S.J. Broeks, A. Verhoef, S. Hogervorst, F.B.L. Fasching, P.A. Lux, M.P. Beckmann, M.W. Ekici, A.B. Sawyer, E. Tomlinson, I. Kerin, M. Marme, F. Schneeweiss, A. Sohn, C. Burwinkel, B. Guénel, P. Truong, T. Cordina-Duverger, E. Menegaux, F. Bojesen, S.E. Nordestgaard, B.G. Nielsen, S.F. Flyger, H. Milne, R.L. Alonso, M.R. González-Neira, A. Benítez, J. Anton-Culver, H. Ziogas, A. Bernstein, L. Dur, C.C. Brenner, H. Müller, H. Arndt, V. Stegmaier, C. Familial Breast Cancer Study (FBCS), . Justenhoven, C. Brauch, H. Brüning, T. Gene Environment Interaction of Breast Cancer in Germany (GENICA) Network, . Wang-Gohrke, S. Eilber, U. Dörk, T. Schürmann, P. Bremer, M. Hillemanns, P. Bogdanova, N.V. Antonenkova, N.N. Rogov, Y.I. Karstens, J.H. Bermisheva, M. Prokofieva, D. Khusnutdinova, E. Lindblom, A. Margolin, S. Mannermaa, A. Kataja, V. Kosma, V.-.M. Hartikainen, J.M. Lambrechts, D. Yesilyurt, B.T. Floris, G. Leunen, K. Manoukian, S. Bonanni, B. Fortuzzi, S. Peterlongo, P. Couch, F.J. Wang, X. Stevens, K. Lee, A. Giles, G.G. Baglietto, L. Severi, G. McLean, C. Alnaes, G.G. Kristensen, V. Børrensen-Dale, A.-.L. John, E.M. Miron, A. Winqvist, R. Pylkäs, K. Jukkola-Vuorinen, A. Kauppila, S. Andrulis, I.L. Glendon, G. Mulligan, A.M. Devilee, P. van Asperen, C.J. Tollenaar, R.A.E.M. Seynaeve, C. Figueroa, J.D. Garcia-Closas, M. Brinton, L. Lissowska, J. Hooning, M.J. Hollestelle, A. Oldenburg, R.A. van den Ouweland, A.M.W. Cox, A. Reed, M.W.R. Shah, M. Jakubowska, A. Lubinski, J. Jaworska, K. Durda, K. Jones, M. Schoemaker, M. Ashworth, A. Swerdlow, A. Beesley, J. Chen, X. kConFab Investigators, . Australian Ovarian Cancer Study Group, . Muir, K.R. Lophatananon, A. Rattanamongkongul, S. Chaiwerawattana, A. Kang, D. Yoo, K.-.Y. Noh, D.-.Y. Shen, C.-.Y. Yu, J.-.C. Wu, P.-.E. Hsiung, C.-.N. Perkins, A. Swann, R. Velentzis, L. Eccles, D.M. Tapper, W.J. Gerty, S.M. Graham, N.J. Ponder, B.A.J. Chenevix-Trench, G. Pharoah, P.D.P. Lathrop, M. Dunning, A.M. Rahman, N. Peto, J. Easton, D.F (2012) Genome-wide association analysis identifies three new breast cancer susceptibility loci.. Show Abstract full text

Breast cancer is the most common cancer among women. To date, 22 common breast cancer susceptibility loci have been identified accounting for ∼8% of the heritability of the disease. We attempted to replicate 72 promising associations from two independent genome-wide association studies (GWAS) in ∼70,000 cases and ∼68,000 controls from 41 case-control studies and 9 breast cancer GWAS. We identified three new breast cancer risk loci at 12p11 (rs10771399; P = 2.7 × 10(-35)), 12q24 (rs1292011; P = 4.3 × 10(-19)) and 21q21 (rs2823093; P = 1.1 × 10(-12)). rs10771399 was associated with similar relative risks for both estrogen receptor (ER)-negative and ER-positive breast cancer, whereas the other two loci were associated only with ER-positive disease. Two of the loci lie in regions that contain strong plausible candidate genes: PTHLH (12p11) has a crucial role in mammary gland development and the establishment of bone metastasis in breast cancer, and NRIP1 (21q21) encodes an ER cofactor and has a role in the regulation of breast cancer cell growth.

Brennan, K. Garcia-Closas, M. Orr, N. Fletcher, O. Jones, M. Ashworth, A. Swerdlow, A. Thorne, H. KConFab Investigators, . Riboli, E. Vineis, P. Dorronsoro, M. Clavel-Chapelon, F. Panico, S. Onland-Moret, N.C. Trichopoulos, D. Kaaks, R. Khaw, K.-.T. Brown, R. Flanagan, J.M (2012) Intragenic ATM methylation in peripheral blood DNA as a biomarker of breast cancer risk.. Show Abstract full text

Few studies have evaluated the association between DNA methylation in white blood cells (WBC) and the risk of breast cancer. The evaluation of WBC DNA methylation as a biomarker of cancer risk is of particular importance as peripheral blood is often available in prospective cohorts and easier to obtain than tumor or normal tissues. Here, we used prediagnostic blood samples from three studies to analyze WBC DNA methylation of two ATM intragenic loci (ATMmvp2a and ATMmvp2b) and genome-wide DNA methylation in long interspersed nuclear element-1 (LINE1) repetitive elements. Samples were from a case-control study derived from a cohort of high-risk breast cancer families (KConFab) and nested case-control studies in two prospective cohorts: Breakthrough Generations Study (BGS) and European Prospective Investigation into Cancer and Nutrition (EPIC). Bisulfite pyrosequencing was used to quantify methylation from 640 incident cases of invasive breast cancer and 741 controls. Quintile analyses for ATMmvp2a showed an increased risk of breast cancer limited to women in the highest quintile [OR, 1.89; 95% confidence interval (CI), 1.36-2.64; P = 1.64 × 10(-4)]. We found no significant differences in estimates across studies or in analyses stratified by family history or menopausal status. However, a more consistent association was observed in younger than in older women and individually significant in KConFab and BGS, but not EPIC. We observed no differences in LINE1 or ATMmvp2b methylation between cases and controls. Together, our findings indicate that WBC DNA methylation levels at ATM could be a marker of breast cancer risk and further support the pursuit of epigenome-wide association studies of peripheral blood DNA methylation.

Swerdlow, A.J. Jones, M.E. Schoemaker, M.J. Hemming, J. Thomas, D. Williamson, J. Ashworth, A (2011) The Breakthrough Generations Study: design of a long-term UK cohort study to investigate breast cancer aetiology.. Show Abstract full text

BACKGROUND: The rationale, design, recruitment and follow-up methods are described for the Breakthrough Generations Study, a UK cohort study started in 2003, targeted at investigation of breast cancer aetiology. METHODS: Cohort members have been recruited by a participant referral method intended to assemble economically a large general population cohort from whom detailed questionnaire information and blood samples can be obtained repeatedly over decades, with high completeness of follow-up and inclusion of large numbers of related individuals. 'First-generation' recruits were women contacted directly, or who volunteered directly, to join the study. They nominated female friends and family, whom we contacted, and those who joined ('second generation') nominated others, reiterated for up to 28 generations. RESULTS: The method has successfully been used during 2003-2011 to recruit 112,049 motivated participants with a broad geographic and socioeconomic distribution, aged 16-102 years, who have completed detailed questionnaires; 92% of the participants gave blood samples at recruitment. When eligible, 2½ years after recruitment, >98% completed the first follow-up questionnaire. Thirty percent are first-degree relatives of other study members. CONCLUSION: The 'generational' recruitment method has enabled recruitment of a large cohort who appear to have the commitment to enable long-term continuing data and sample collection, to investigate the effects of changing endogenous and exogenous factors on cancer risk.

Orr, N. Cooke, R. Jones, M. Fletcher, O. Dudbridge, F. Chilcott-Burns, S. Tomczyk, K. Broderick, P. Houlston, R. Ashworth, A. Swerdlow, A (2011) Genetic variants at chromosomes 2q35, 5p12, 6q25.1, 10q26.13, and 16q12.1 influence the risk of breast cancer in men.. Show Abstract full text

Male breast cancer accounts for approximately 1% of all breast cancer. To date, risk factors for male breast cancer are poorly defined, but certain risk factors and genetic features appear common to both male and female breast cancer. Genome-wide association studies (GWAS) have recently identified common single nucleotide polymorphisms (SNPs) that influence female breast cancer risk; 12 of these have been independently replicated. To examine if these variants contribute to male breast cancer risk, we genotyped 433 male breast cancer cases and 1,569 controls. Five SNPs showed a statistically significant association with male breast cancer: rs13387042 (2q35) (odds ratio (OR)  = 1.30, p = 7.98×10⁻⁴), rs10941679 (5p12) (OR = 1.26, p = 0.007), rs9383938 (6q25.1) (OR = 1.39, p = 0.004), rs2981579 (FGFR2) (OR = 1.18, p = 0.03), and rs3803662 (TOX3) (OR = 1.48, p = 4.04×10⁻⁶). Comparing the ORs for male breast cancer with the published ORs for female breast cancer, three SNPs--rs13387042 (2q35), rs3803662 (TOX3), and rs6504950 (COX11)--showed significant differences in ORs (p<0.05) between sexes. Breast cancer is a heterogeneous disease; the relative risks associated with loci identified to date show subtype and, based on these data, gender specificity. Additional studies of well-defined patient subgroups could provide further insight into the biological basis of breast cancer development.

Murray, A. Bennett, C.E. Perry, J.R.B. Weedon, M.N. Jacobs, P.A. Morris, D.H. Orr, N. Schoemaker, M.J. Jones, M. Ashworth, A. Swerdlow, A.J. ReproGen Consortium, (2011) Common genetic variants are significant risk factors for early menopause: results from the Breakthrough Generations Study.. Show Abstract full text

Women become infertile approximately 10 years before menopause, and as more women delay childbirth into their 30s, the number of women who experience infertility is likely to increase. Tests that predict the timing of menopause would allow women to make informed reproductive decisions. Current predictors are only effective just prior to menopause, and there are no long-range indicators. Age at menopause and early menopause (EM) are highly heritable, suggesting a genetic aetiology. Recent genome-wide scans have identified four loci associated with variation in the age of normal menopause (40-60 years). We aimed to determine whether theses loci are also risk factors for EM. We tested the four menopause-associated genetic variants in a cohort of approximately 2000 women with menopause≤45 years from the Breakthrough Generations Study (BGS). All four variants significantly increased the odds of having EM. Comparing the 4.5% of individuals with the lowest number of risk alleles (two or three) with the 3.0% with the highest number (eight risk alleles), the odds ratio was 4.1 (95% CI 2.4-7.1, P=4.0×10(-7)). In combination, the four variants discriminated EM cases with a receiver operator characteristic area under the curve of 0.6. Four common genetic variants identified by genome-wide association studies, had a significant impact on the odds of having EM in an independent cohort from the BGS. The discriminative power is still limited, but as more variants are discovered they may be useful for predicting reproductive lifespan.

Morris, D.H. Jones, M.E. Schoemaker, M.J. Ashworth, A. Swerdlow, A.J (2011) Familial concordance for age at menarche: analyses from the Breakthrough Generations Study.. Show Abstract full text

Age at menarche is correlated within families, but estimates of the heritability of menarcheal age have a wide range (0.45-0.95). We examined the familial resemblance for age at menarche and the extent to which this is due to genetic and shared environmental factors. Between 2003 and 2010 data were retrospectively collected by questionnaire from participants within the UK-based Breakthrough Generations Study. These analyses included 25,970 female participants aged 16-98 with at least one female relative who was also a study participant. A woman's menarche was significantly delayed for each yearly increase in the menarcheal age of her monozygotic twin (average increase = 7.2 months, P < 0.001), dizygotic twin (average increase = 3.0 months, P = 0.03), older sister (average increase = 3.3 months, P < 0.001), mother (average increase = 3.4 months, P < 0.001), maternal grandmother (average increase = 1.5 months, P = 0.04), maternal aunt (average increase = 1.4 months, P < 0.001) and paternal aunt (average increase = 3.0 months, P < 0.001). There was not a significant association between the menarcheal ages of half-sister pairs or of paternal grandmother-granddaughter pairs, based on small numbers. Heritability was estimated as 0.57 [95% confidence interval 0.53, 0.61]. Shared environmental factors did not have an effect in the model. In conclusion, approximately half of the variation in age at menarche was attributable to additive genetic effects with the remainder attributable to non-shared environmental effects.

Morris, D.H. Jones, M.E. Schoemaker, M.J. Ashworth, A. Swerdlow, A.J (2011) Secular trends in age at menarche in women in the UK born 1908-93: results from the Breakthrough Generations Study.. Show Abstract full text

Menarcheal age decreased over time in Western countries until cohorts born in the mid-20th century. It then stabilised, but limited data are available for recent cohorts. Menarche data were collected retrospectively by questionnaire in 2003-10 from 94,170 women who were participating in the Breakthrough Generations Study, aged 16-98 years, born 1908-93 and resident in the UK. Average menarcheal age declined from women born in 1908-19 (mean=13.5 years) to those born in 1945-49 (mean=12.6 years). It was then stable for several birth cohorts, but resumed its downward trend in recent cohorts (mean=12.3 years in 1990-93 cohort). Trends differed between socio-economic groups, but the recent decline was present in each group. In conclusion, menarcheal age appears to have decreased again in recent cohorts after a period of stabilisation.

Morris, D.H. Jones, M.E. Schoemaker, M.J. Ashworth, A. Swerdlow, A.J (2011) Familial concordance for age at natural menopause: results from the Breakthrough Generations Study.. Show Abstract full text

OBJECTIVE: Existing estimates of the heritability of menopause age have a wide range. Furthermore, few studies have analyzed to what extent familial similarities might reflect shared environment, rather than shared genes. We therefore analyzed familial concordance for age at natural menopause and the effects of shared genetic and environmental factors on this concordance. METHODS: Participants were 2,060 individuals comprising first-degree relatives, aged 31 to 90 years, and participating in the UK Breakthrough Generations Study. Menopause data were collected using a self-administered questionnaire and analyzed using logistic regression and variance-components models. RESULTS: Women were at an increased risk of early menopause (≤45 y) if their mother (odds ratio, 6.2; P < 0.001) or non-twin sister (odds ratio, 5.5; P < 0.001) had had an early menopause. Likewise, women had an increased risk of late menopause (≥54 y) if their relative had had a late menopause (mother: odds ratio, 6.1; P < 0.01; non-twin sister: odds ratio, 2.3; P < 0.001). Estimated heritability was 41.6% (P < 0.01), with an additional 13.6% (P = 0.02) of the variation in menopause age attributed to environmental factors shared by sisters. CONCLUSIONS: We confirm that early menopause aggregates within families and show, for the first time, that there is also strong familial concordance for late menopause. Both genes and shared environment were the source of variation in menopause age. Past heritability estimates have not accounted for shared environment, and thus, the effect of genetic variants on menopause age may previously have been overestimated.

Fletcher, O. Johnson, N. Orr, N. Hosking, F.J. Gibson, L.J. Walker, K. Zelenika, D. Gut, I. Heath, S. Palles, C. Coupland, B. Broderick, P. Schoemaker, M. Jones, M. Williamson, J. Chilcott-Burns, S. Tomczyk, K. Simpson, G. Jacobs, K.B. Chanock, S.J. Hunter, D.J. Tomlinson, I.P. Swerdlow, A. Ashworth, A. Ross, G. dos Santos Silva, I. Lathrop, M. Houlston, R.S. Peto, J (2011) Novel breast cancer susceptibility locus at 9q31.2: results of a genome-wide association study.. Show Abstract full text

BACKGROUND: Genome-wide association studies have identified several common genetic variants associated with breast cancer risk. It is likely, however, that a substantial proportion of such loci have not yet been discovered. METHODS: We compared 296,114 tagging single-nucleotide polymorphisms in 1694 breast cancer case subjects (92% with two primary cancers or at least two affected first-degree relatives) and 2365 control subjects, with validation in three independent series totaling 11,880 case subjects and 12,487 control subjects. Odds ratios (ORs) and associated 95% confidence intervals (CIs) in each stage and all stages combined were calculated using unconditional logistic regression. Heterogeneity was evaluated with Cochran Q and I(2) statistics. All statistical tests were two-sided. RESULTS: We identified a novel risk locus for breast cancer at 9q31.2 (rs865686: OR = 0.89, 95% CI = 0.85 to 0.92, P = 1.75 × 10(-10)). This single-nucleotide polymorphism maps to a gene desert, the nearest genes being Kruppel-like factor 4 (KLF4, 636 kb centromeric), RAD23 homolog B (RAD23B, 794 kb centromeric), and actin-like 7A (ACTL7A, 736 kb telomeric). We also identified two variants (rs3734805 and rs9383938) mapping to 6q25.1 estrogen receptor 1 (ESR1), which were associated with breast cancer in subjects of northern European ancestry (rs3734805: OR = 1.19, 95% CI = 1.11 to 1.27, P = 1.35 × 10(-7); rs9383938: OR = 1.18, 95% CI = 1.11 to 1.26, P = 1.41 × 10(-7)). A variant mapping to 10q26.13, approximately 300 kb telomeric to the established risk locus within the second intron of FGFR2, was also associated with breast cancer risk, although not at genome-wide statistical significance (rs10510102: OR = 1.12, 95% CI = 1.07 to 1.17, P = 1.58 × 10(-6)). CONCLUSIONS: These findings provide further evidence on the role of genetic variation in the etiology of breast cancer. Fine mapping will be needed to identify causal variants and to determine their functional effects.

Figueroa, J.D. Garcia-Closas, M. Humphreys, M. Platte, R. Hopper, J.L. Southey, M.C. Apicella, C. Hammet, F. Schmidt, M.K. Broeks, A. Tollenaar, R.A.E.M. Van't Veer, L.J. Fasching, P.A. Beckmann, M.W. Ekici, A.B. Strick, R. Peto, J. dos Santos Silva, I. Fletcher, O. Johnson, N. Sawyer, E. Tomlinson, I. Kerin, M. Burwinkel, B. Marme, F. Schneeweiss, A. Sohn, C. Bojesen, S. Flyger, H. Nordestgaard, B.G. Benítez, J. Milne, R.L. Ignacio Arias, J. Zamora, M.P. Brenner, H. Müller, H. Arndt, V. Rahman, N. Turnbull, C. Seal, S. Renwick, A. Brauch, H. Justenhoven, C. Brüning, T. GENICA Network, . Chang-Claude, J. Hein, R. Wang-Gohrke, S. Dörk, T. Schürmann, P. Bremer, M. Hillemanns, P. Nevanlinna, H. Heikkinen, T. Aittomäki, K. Blomqvist, C. Bogdanova, N. Antonenkova, N. Rogov, Y.I. Karstens, J.H. Bermisheva, M. Prokofieva, D. Gantcev, S.H. Khusnutdinova, E. Lindblom, A. Margolin, S. Chenevix-Trench, G. Beesley, J. Chen, X. kConFab AOCS Management Group, . Mannermaa, A. Kosma, V.-.M. Soini, Y. Kataja, V. Lambrechts, D. Yesilyurt, B.T. Chrisiaens, M.-.R. Peeters, S. Radice, P. Peterlongo, P. Manoukian, S. Barile, M. Couch, F. Lee, A.M. Diasio, R. Wang, X. Giles, G.G. Severi, G. Baglietto, L. Maclean, C. Offit, K. Robson, M. Joseph, V. Gaudet, M. John, E.M. Winqvist, R. Pylkäs, K. Jukkola-Vuorinen, A. Grip, M. Andrulis, I. Knight, J.A. Mulligan, A.M. O'Malley, F.P. Brinton, L.A. Sherman, M.E. Lissowska, J. Chanock, S.J. Hooning, M. Martens, J.W.M. van den Ouweland, A.M.W. Collée, J.M. Hall, P. Czene, K. Cox, A. Brock, I.W. Reed, M.W.R. Cross, S.S. Pharoah, P. Dunning, A.M. Kang, D. Yoo, K.-.Y. Noh, D.-.Y. Ahn, S.-.H. Jakubowska, A. Lubinski, J. Jaworska, K. Durda, K. Sangrajrang, S. Gaborieau, V. Brennan, P. McKay, J. Shen, C.-.Y. Ding, S.-.L. Hsu, H.-.M. Yu, J.-.C. Anton-Culver, H. Ziogas, A. Ashworth, A. Swerdlow, A. Jones, M. Orr, N. Trentham-Dietz, A. Egan, K. Newcomb, P. Titus-Ernstoff, L. Easton, D. Spurdle, A.B (2011) Associations of common variants at 1p11.2 and 14q24.1 (RAD51L1) with breast cancer risk and heterogeneity by tumor subtype: findings from the Breast Cancer Association Consortium.. Show Abstract full text

A genome-wide association study (GWAS) identified single-nucleotide polymorphisms (SNPs) at 1p11.2 and 14q24.1 (RAD51L1) as breast cancer susceptibility loci. The initial GWAS suggested stronger effects for both loci for estrogen receptor (ER)-positive tumors. Using data from the Breast Cancer Association Consortium (BCAC), we sought to determine whether risks differ by ER, progesterone receptor (PR), human epidermal growth factor receptor 2 (HER2), grade, node status, tumor size, and ductal or lobular morphology. We genotyped rs11249433 at 1p.11.2, and two highly correlated SNPs rs999737 and rs10483813 (r(2)= 0.98) at 14q24.1 (RAD51L1), for up to 46 036 invasive breast cancer cases and 46 930 controls from 39 studies. Analyses by tumor characteristics focused on subjects reporting to be white women of European ancestry and were based on 25 458 cases, of which 87% had ER data. The SNP at 1p11.2 showed significantly stronger associations with ER-positive tumors [per-allele odds ratio (OR) for ER-positive tumors was 1.13, 95% CI = 1.10-1.16 and, for ER-negative tumors, OR was 1.03, 95% CI = 0.98-1.07, case-only P-heterogeneity = 7.6 × 10(-5)]. The association with ER-positive tumors was stronger for tumors of lower grade (case-only P= 6.7 × 10(-3)) and lobular histology (case-only P= 0.01). SNPs at 14q24.1 were associated with risk for most tumor subtypes evaluated, including triple-negative breast cancers, which has not been described previously. Our results underscore the need for large pooling efforts with tumor pathology data to help refine risk estimates for SNP associations with susceptibility to different subtypes of breast cancer.

Schoemaker, M.J. Robertson, L. Wigertz, A. Jones, M.E. Hosking, F.J. Feychting, M. Lönn, S. McKinney, P.A. Hepworth, S.J. Muir, K.R. Auvinen, A. Salminen, T. Kiuru, A. Johansen, C. Houlston, R.S. Swerdlow, A.J (2010) Interaction between 5 genetic variants and allergy in glioma risk.. Show Abstract full text

The etiology of glioma is barely known. Epidemiologic studies have provided evidence for an inverse relation between glioma risk and allergic disease. Genome-wide association data have identified common genetic variants at 5p15.33 (rs2736100, TERT), 8q24.21 (rs4295627, CCDC26), 9p21.3 (rs4977756, CDKN2A-CDKN2B), 11q23.3 (rs498872, PHLDB1), and 20q13.33 (rs6010620, RTEL1) as determinants of glioma risk. The authors investigated whether there is interaction between the effects of allergy and these 5 variants on glioma risk. Data from 5 case-control studies carried out in Denmark, Finland, Sweden, and the United Kingdom (2000-2004) were used, totaling 1,029 cases and 1,668 controls. Risk was inversely associated with asthma, hay fever, eczema, and "any allergy," significantly for each factor except asthma, and was significantly positively associated with number of risk alleles for each of the 5 single nucleotide polymorphisms. There was interaction between asthma and rs498872 (greater protective effect of asthma with increasing number of risk alleles; per-allele interaction odds ratio (OR) = 0.65, P = 0.041), between "any allergy" and rs4977756 (smaller protective effect; interaction OR = 1.27, P = 0.047), and between "any allergy" and rs6010620 (greater protective effect; interaction OR = 0.70, P = 0.017). Case-only analyses provided further support for atopy interactions for rs4977756 and rs498872. This study provides evidence for possible gene-environment interactions in glioma development.

Morris, D.H. Jones, M.E. Schoemaker, M.J. Ashworth, A. Swerdlow, A.J (2010) Determinants of age at menarche in the UK: analyses from the Breakthrough Generations Study.. Show Abstract full text

BACKGROUND: Early menarche increases breast cancer risk but, aside from weight, information on its determinants is limited. METHODS: Age at menarche data were collected retrospectively by questionnaire from 81,606 women aged 16-98, resident in the UK and participating in the Breakthrough Generations Study. RESULTS: Menarche occurred earlier in women who had a low birthweight (P(trend)<0.001), were singletons (P<0.001), had prenatal exposure to pre-eclampsia (P<0.001) or maternal smoking (P=0.01), were not breastfed (P(trend)=0.03), were non-white (P<0.001), were heavy (P(trend)<0.001) or tall (P(trend)<0.001) compared with their peers at age 7 and exercised little as a child (P(trend)<0.001). Menarcheal age increased with number of siblings (P<0.001) independently of birth order, and had an inverse association with birth order after adjustment for sibship size (P<0.001). In a multivariate model, birthweight, ethnicity, weight, height, exercise, sibship size and birth order remained significant, and maternal age at birth became significant (positive association, P<0.001). CONCLUSION: Age at menarche was influenced by both pre- and post-natal factors, and these factors may affect breast cancer risk through this route.

Hosseini, M. Taslimi, S.H. Dinarvand, P. Jones, M.E. Mohammad, K (2010) Trends in weights, heights, BMI and comparison of their differences in urban and rural areas for Iranian children and adolescents 2-18-year-old between 1990-1991 and 1999.
Laing, S.P. Jones, M.E. Swerdlow, A.J. Burden, A.C. Gatling, W (2005) Psychosocial and socioeconomic risk factors for premature death in young people with type 1 diabetes.. Show Abstract full text

OBJECTIVE: Mortality from acute diabetes-related events is greatly raised in young adults with type 1 diabetes. Psychosocial and socioeconomic risk factors are examined for deaths from acute events separately from deaths due to other causes. RESEARCH DESIGN AND METHODS: This study had a nested case-control design. The cases were patients from the Diabetes UK cohort who died before age 40 years. Deaths were categorized as acute events or chronic conditions related to diabetes. Where possible, two matched control subjects were selected for each case. Data relating to psychosocial and socioeconomic factors and variables related to diabetes complications were extracted from the case notes. Risks of death were estimated by calculation of odds ratios (ORs). RESULTS: Case notes were obtained for 98 case and 137 control subjects. Fifty-one deaths were attributed to acute causes, 34 to chronic conditions related to diabetes, and the remaining 13 were unrelated to diabetes. Living alone (OR 4.4), past drug abuse (5.7), and previous psychiatric referral (4.6) were all significantly associated with death from acute events but not death from chronic conditions. There was no association between deaths from acute events and nephropathy, hypertension, neuropathy, or retinopathy, although all of these were associated with deaths from chronic conditions. CONCLUSIONS: The results indicate that psychosocial factors are powerful risk factors for mortality from acute events in patients with type 1 diabetes, although not for mortality from chronic conditions. The data enable the identification of a high-risk group suitable for targeting with preventive measures to reduce acute event mortality.

Graham, A. Fuller, A. Murphy, M. Jones, M. Forman, D. Swerdlow, A.J (1999) Maternal and child constitutional factors and the frequency of melanocytic naevi in children.. Show Abstract full text

This study examined the association between numbers of benign melanocytic naevi in 7-year-old children in Oxfordshire born in 1988-9 with their mother's arm naevus count, and maternal and child pigmentation factors. We believe this is the first time that the relationship between child and maternal naevus counts has been reported. A high naevus count in the child was associated with male sex (P = 0.009), freckling (P = 0.001) and propensity of the child to burn in the sun (P = 0.05). A low naevus count was observed in red-haired children (P = 0.02). The strongest association of child's naevus count was with a high maternal arm naevus count, independent of the child's pigmentation factors (trend P < 0.0001). Maternal pigmentation factors were not associated with child's naevus count independent of the child's own pigmentation factors. Maternal arm naevus counts may be a better predictor of child naevus count than the child's own pigmentation factors and children. There has not been examination, however, of the relationship between naevus counts in children and those in their parents. We therefore conducted a study of the occurrence of naevi in children aged 7-8 years in Oxfordshire, examining, in addition to sex and pigmentation factors in the child, the relationship of maternal pigmentation factors and maternal naevus counts with naevi in their offspring.

dos Santos Silva, I. Jones, M. Malveiro, F. Swerdlow, A (1999) Mortality in the Portuguese thorotrast study. Show Abstract full text

The Portuguese Thorotrast study cohort consists of a group of patients who received Thorotrast for diagnostic reasons between 1929 and 1956, and a group of similar patients who were given nonradioactive contrast agents. The cohort members were identified from medical records that contained information on reasons for the radiological investigation, type of procedure employed, and name and dose of the contrast medium used. This cohort was assembled in 1961, but its follow-up was interrupted in 1976. We have now reactivated this cohort and extended its follow-up through the end of 1996. Similar methods were used to follow up and ascertain cause of death for the Thorotrast-exposed and unexposed subjects. A total of 1931 patients who received Thorotrast systemically and 2258 unexposed subjects were initially identified from medical records. We were able to successfully follow up 58.6% (1131) of the Thorotrast patients and 45.7% (1032) of the unexposed patients. By the end of 1996, 92.2% of the Thorotrast patients and 75.2% of the unexposed patients were dead. Mortality from all causes was increased in the Thorotrast patients compared to those who were not exposed. This excess in mortality increased with time since exposure, peaking 30 years after administration of Thorotrast. The rise in overall mortality was essentially due to neoplasms [relative risk (RR) adjusted for sex, age and period = 6.04, 95% CI = 4.42-8.26], nonmalignant liver disorders (RR = 5.67, 95% CI 3.13-10.3) and nonmalignant hematological conditions (RR = 14.2, 95% CI = 2.54-79.3). The increase in mortality from neoplasms was explained mainly by increases in the risk of liver cancer (RR = 70.8, 95% CI = 19.9-251.3) and, to a much lesser extent, leukemia (RR = 15.2, 95% CI = 1. 28-181.7).

McLean, S. Wood, L.J. Montgomery, I.M. Davidson, J. Jones, M.E (1995) Trends in hotel patronage and drink driving in Hobart, Tasmania. Show Abstract full text

From 1990 to 1991 in the Hobart region there was a marked fall in both hotel patronage and the proportion of patrons subsequently driving with their blood alcohol concentration above the legal limit. This was associated with smaller falls in the number of drink drivers charged and alcohol-related road accidents, which continued in the following year. It appears that the pattern of drinking and driving is changing, presumably in response to random breath testing and tougher penalties for offences.

McLean, S. Wood, L.J. Montgomery, I.M. Davidson, J. Jones, M.E (1994) Promotion of responsible drinking in hotels. Show Abstract full text

This study reports on an intervention programme to promote responsible drinking in hotels. The licensees of eight hotels agreed to participate in a trial of measures designed to assist patrons to avoid drink-driving, and seven other hotels were used as controls. The interventions acceptable to licensees comprised commercial-quality promotional material with the theme "0.05 Know Your Limits", and a breath analysis machine and poster on its use. Patrons leaving the hotels on Thursday, Friday and Saturday nights were interviewed and breath-tested. Although the intervention material had been seen by one-third of patrons in the intervention hotels, there was no significant difference between them and control hotel patrons in either median BAC or the proportion who were going to drive with BAC over the legal limit. There was poor compliance by hotels with the intervention procedures, indicating that a major impediment to the implementation and evaluation of programmes to promote responsible drinking is a lack of motivation by many licensees, despite support by some licensees and the Australian Hotels Association.

McLean, S. Wood, L.J. Davidson, J. Montgomery, I.M. Jones, M.E (1993) Alcohol consumption and driving intentions amongst hotel patrons. Show Abstract full text

To examine the extent to which hotel patrons drink in excess of current health recommendations, and to identify risk factors for excessive drinking, hotel patrons were invited to participate in a survey of social drinking, which included a free breath test. Patrons were approached at 15 min. intervals, and 1000 subjects were studied. Amongst this group of hotel patrons interviewed, 1 in 2 had consumed alcohol in excess of the daily limit recommended by the National Health & Medical Research Council. One in 10 intending drivers had a BAC over the legal limit. Excessive drinking and drink driving appear to be prevalent amongst hotel patrons, and hotels should be targets for interventions designed to reduce these problems. For example, the National Guidelines for the Responsible Serving of Alcohol should be more widely practised.

Ponsonby, A.L. Jones, M.E. Lumley, J. Dwyer, T. Gilbert, N (1992) Climatic temperature and variation in the incidence of sudden infant death syndrome between the Australian states. Show Abstract full text

OBJECTIVE: To describe the relationship between climatic temperature and the incidence of sudden infant death syndrome (SIDS) for the Australian States and examine the extent to which differences in climatic temperature might explain the regional variation of SIDS in Australia. DESIGN: Case series study. A generalised linear model was used to model the association between monthly average temperature and the incidence of SIDS. SETTING: The report is population based. Data are available from all Australian States. SUBJECTS: Cases of SIDS from birth to less than 12 months of age occurring in Queensland (1981-1987), New South Wales (1981-1987), Victoria (1984-1987), Tasmania (1975-1989), South Australia (1980-1989), and Western Australia (1980-1988). RESULTS: Every one degree Celsius decrease in average monthly temperature within the range 9 degrees C to 25 degrees C is associated with a 10.6% (95% confidence interval, 9.6%-11.7%) increase in the incidence of SIDS. Climatic temperature accounts for 84% of the interstate variation in the rate of SIDS. After controlling for the effect of temperature, a significant overall difference in SIDS incidence remains (P less than 0.0001) for the Australian States. CONCLUSION: Climatic temperature accounts for most but not all of the regional variation of SIDS incidence in the Australian States. The remaining variation may reflect differences in the maternal and infant characteristics of the State populations.

Ponsonby, A.L. Jones, M.E. Lumley, J. Dwyer, T. Gilbert, N (1992) Sudden infant death syndrome: factors contributing to the difference in incidence between Victoria and Tasmania. Show Abstract full text

OBJECTIVE: To examine how much of the difference in incidence of sudden infant death syndrome (SIDS) between Tasmania and Victoria could be accounted for by the effect of differing climatic temperature and the effect of the differing prevalence of maternal and infant characteristics in the two State populations. DESIGN: A two population ecological comparison. Two previously published predictive models were applied to quantify the contribution of several factors to the higher incidence of SIDS in Tasmania compared with Victoria. SETTING: A population based study involving the two Australian States of Tasmania and Victoria. PATIENTS: The characteristics of the 1985 to 1987 live birth cohorts of Tasmania and Victoria were examined. Cases were defined as all infants dying in 1985 to 1987 whose cause of death was stated as SIDS. RESULTS: The rate of SIDS for Tasmania and Victoria 1985 to 1987 was 3.76 per 1000 live births and 2.18 per 1000 live births respectively. Adjustment of the Tasmanian rate for the effect of the interstate difference in climatic temperature resulted in a lower Tasmanian rate of 2.92 per 1000 live births. Adjustment for the effect of interstate differences in maternal age, birthweight, infant sex, month of birth and intention to breast-feed at hospital discharge decreased the Tasmanian rate to 2.47 per 1000 live births. CONCLUSION: Approximately 82% of the interstate difference in SIDS incidence between Tasmania and Victoria from 1985 to 1987 can be accounted for by differences in climatic temperature, maternal age, birth-weight, infant sex, month of birth and feeding intention at hospital discharge.

Beard, T.C. Eickhoff, R. Mejglo, Z.A. Jones, M. Bennett, S.A. Dwyer, T (1992) Population-based survey of human sodium and potassium excretion. Show Abstract full text

1. During the 1989 National Heart Foundation Risk Factor Prevalence Survey a subsample in Hobart collected 24 h urine samples to measure electrolyte excretion. 2. The ranges were 30-344 mmol/24 h for Na+ excretion (mean 160 mmol/24 h for men, 124 mmol/24 h for women), and 25-119 mmol/24 h for K+ excretion (mean 77 mmol/24 h for men, 68 mmol/24 h for women). 3. As in other surveys, women excreted about 20-25% less Na+ and K+ than men, although there was no significant sex difference in the ratio of Na+/K+ excretion. 4. The recommended dietary intake (RDI) for Na+ and K+ was followed simultaneously by 19% of subjects, and 13% had a 24 h urinary Na+/K+ ratio less than or equal to 1.0. 5. Observance of the RDI limited the value of iodized salt for goitre prophylaxis. 6. Sodium excretion rates were outside the therapeutic range of thiazide diuretics in 22% of subjects. 7. Diet groups for long-term prospective cohort studies to test the prophylactic value of avoiding salt could apparently be recruited from existing subsamples of the population.

Pal Choudhury, P. Wilcox, A.N. Brook, M.N. Zhang, Y. Ahearn, T. Orr, N. Coulson, P. Schoemaker, M.J. Jones, M.E. Gail, M.H. Swerdlow, A.J. Chatterjee, N. Garcia-Closas, M (2020) Comparative Validation of Breast Cancer Risk Prediction Models and Projections for Future Risk Stratification.. Show Abstract full text

<h4>Background</h4>External validation of risk models is critical for risk-stratified breast cancer prevention. We used the Individualized Coherent Absolute Risk Estimation (iCARE) as a flexible tool for risk model development and comparative model validation and to make projections for population risk stratification.<h4>Methods</h4>Performance of two recently developed models, one based on the Breast and Prostate Cancer Cohort Consortium analysis (iCARE-BPC3) and another based on a literature review (iCARE-Lit), were compared with two established models (Breast Cancer Risk Assessment Tool and International Breast Cancer Intervention Study Model) based on classical risk factors in a UK-based cohort of 64 874 white non-Hispanic women (863 patients) age 35-74 years. Risk projections in a target population of US white non-Hispanic women age 50-70 years assessed potential improvements in risk stratification by adding mammographic breast density (MD) and polygenic risk score (PRS).<h4>Results</h4>The best calibrated models were iCARE-Lit (expected to observed number of cases [E/O] = 0.98, 95% confidence interval [CI] = 0.87 to 1.11) for women younger than 50 years, and iCARE-BPC3 (E/O = 1.00, 95% CI = 0.93 to 1.09) for women 50 years or older. Risk projections using iCARE-BPC3 indicated classical risk factors can identify approximately 500 000 women at moderate to high risk (>3% 5-year risk) in the target population. Addition of MD and a 313-variant PRS is expected to increase this number to approximately 3.5 million women, and among them, approximately 153 000 are expected to develop invasive breast cancer within 5 years.<h4>Conclusions</h4>iCARE models based on classical risk factors perform similarly to or better than BCRAT or IBIS in white non-Hispanic women. Addition of MD and PRS can lead to substantial improvements in risk stratification. However, these integrated models require independent prospective validation before broad clinical applications.

Johansson, A. Palli, D. Masala, G. Grioni, S. Agnoli, C. Tumino, R. Giurdanella, M.C. Fasanelli, F. Sacerdote, C. Panico, S. Mattiello, A. Polidoro, S. Jones, M.E. Schoemaker, M.J. Orr, N. Tomczyk, K. Johnson, N. Fletcher, O. Perduca, V. Baglietto, L. Dugué, P.-.A. Southey, M.C. Giles, G.G. English, D.R. Milne, R.L. Severi, G. Ambatipudi, S. Cuenin, C. Chajès, V. Romieu, I. Herceg, Z. Swerdlow, A.J. Vineis, P. Flanagan, J.M (2019) Epigenome-wide association study for lifetime estrogen exposure identifies an epigenetic signature associated with breast cancer risk.. Show Abstract full text

<h4>Background</h4>It is well established that estrogens and other hormonal factors influence breast cancer susceptibility. We hypothesized that a woman's total lifetime estrogen exposure accumulates changes in DNA methylation, detectable in the blood, which could be used in risk assessment for breast cancer.<h4>Methods</h4>An estimated lifetime estrogen exposure (ELEE) model was defined using epidemiological data from EPIC-Italy (n = 31,864). An epigenome-wide association study (EWAS) of ELEE was performed using existing Illumina HumanMethylation450K Beadchip (HM450K) methylation data obtained from EPIC-Italy blood DNA samples (n = 216). A methylation index (MI) of ELEE based on 31 CpG sites was developed using HM450K data from EPIC-Italy and the Generations Study and evaluated for association with breast cancer risk in an independent dataset from the Generations Study (n = 440 incident breast cancer cases matched to 440 healthy controls) using targeted bisulfite sequencing. Lastly, a meta-analysis was conducted including three additional cohorts, consisting of 1187 case-control pairs.<h4>Results</h4>We observed an estimated 5% increase in breast cancer risk per 1-year longer ELEE (OR = 1.05, 95% CI 1.04-1.07, P = 3 × 10<sup>-12</sup>) in EPIC-Italy. The EWAS identified 694 CpG sites associated with ELEE (FDR Q < 0.05). We report a DNA methylation index (MI) associated with breast cancer risk that is validated in the Generations Study targeted bisulfite sequencing data (OR<sub>Q4_vs_Q1</sub> = 1.77, 95% CI 1.07-2.93, P = 0.027) and in the meta-analysis (OR<sub>Q4_vs_Q1</sub> = 1.43, 95% CI 1.05-2.00, P = 0.024); however, the correlation between the MI and ELEE was not validated across study cohorts.<h4>Conclusion</h4>We have identified a blood DNA methylation signature associated with breast cancer risk in this study. Further investigation is required to confirm the interaction between estrogen exposure and DNA methylation in the blood.

Ferreira, M.A. Gamazon, E.R. Al-Ejeh, F. Aittomäki, K. Andrulis, I.L. Anton-Culver, H. Arason, A. Arndt, V. Aronson, K.J. Arun, B.K. Asseryanis, E. Azzollini, J. Balmaña, J. Barnes, D.R. Barrowdale, D. Beckmann, M.W. Behrens, S. Benitez, J. Bermisheva, M. Białkowska, K. Blomqvist, C. Bogdanova, N.V. Bojesen, S.E. Bolla, M.K. Borg, A. Brauch, H. Brenner, H. Broeks, A. Burwinkel, B. Caldés, T. Caligo, M.A. Campa, D. Campbell, I. Canzian, F. Carter, J. Carter, B.D. Castelao, J.E. Chang-Claude, J. Chanock, S.J. Christiansen, H. Chung, W.K. Claes, K.B.M. Clarke, C.L. EMBRACE Collaborators, . GC-HBOC Study Collaborators, . GEMO Study Collaborators, . Couch, F.J. Cox, A. Cross, S.S. Czene, K. Daly, M.B. de la Hoya, M. Dennis, J. Devilee, P. Diez, O. Dörk, T. Dunning, A.M. Dwek, M. Eccles, D.M. Ejlertsen, B. Ellberg, C. Engel, C. Eriksson, M. Fasching, P.A. Fletcher, O. Flyger, H. Friedman, E. Frost, D. Gabrielson, M. Gago-Dominguez, M. Ganz, P.A. Gapstur, S.M. Garber, J. García-Closas, M. García-Sáenz, J.A. Gaudet, M.M. Giles, G.G. Glendon, G. Godwin, A.K. Goldberg, M.S. Goldgar, D.E. González-Neira, A. Greene, M.H. Gronwald, J. Guénel, P. Haiman, C.A. Hall, P. Hamann, U. He, W. Heyworth, J. Hogervorst, F.B.L. Hollestelle, A. Hoover, R.N. Hopper, J.L. Hulick, P.J. Humphreys, K. Imyanitov, E.N. ABCTB Investigators, . HEBON Investigators, . BCFR Investigators, . Isaacs, C. Jakimovska, M. Jakubowska, A. James, P.A. Janavicius, R. Jankowitz, R.C. John, E.M. Johnson, N. Joseph, V. Karlan, B.Y. Khusnutdinova, E. Kiiski, J.I. Ko, Y.-.D. Jones, M.E. Konstantopoulou, I. Kristensen, V.N. Laitman, Y. Lambrechts, D. Lazaro, C. Leslie, G. Lester, J. Lesueur, F. Lindström, S. Long, J. Loud, J.T. Lubiński, J. Makalic, E. Mannermaa, A. Manoochehri, M. Margolin, S. Maurer, T. Mavroudis, D. McGuffog, L. Meindl, A. Menon, U. Michailidou, K. Miller, A. Montagna, M. Moreno, F. Moserle, L. Mulligan, A.M. Nathanson, K.L. Neuhausen, S.L. Nevanlinna, H. Nevelsteen, I. Nielsen, F.C. Nikitina-Zake, L. Nussbaum, R.L. Offit, K. Olah, E. Olopade, O.I. Olsson, H. Osorio, A. Papp, J. Park-Simon, T.-.W. Parsons, M.T. Pedersen, I.S. Peixoto, A. Peterlongo, P. Pharoah, P.D.P. Plaseska-Karanfilska, D. Poppe, B. Presneau, N. Radice, P. Rantala, J. Rennert, G. Risch, H.A. Saloustros, E. Sanden, K. Sawyer, E.J. Schmidt, M.K. Schmutzler, R.K. Sharma, P. Shu, X.-.O. Simard, J. Singer, C.F. Soucy, P. Southey, M.C. Spinelli, J.J. Spurdle, A.B. Stone, J. Swerdlow, A.J. Tapper, W.J. Taylor, J.A. Teixeira, M.R. Terry, M.B. Teulé, A. Thomassen, M. Thöne, K. Thull, D.L. Tischkowitz, M. Toland, A.E. Torres, D. Truong, T. Tung, N. Vachon, C.M. van Asperen, C.J. van den Ouweland, A.M.W. van Rensburg, E.J. Vega, A. Viel, A. Wang, Q. Wappenschmidt, B. Weitzel, J.N. Wendt, C. Winqvist, R. Yang, X.R. Yannoukakos, D. Ziogas, A. Kraft, P. Antoniou, A.C. Zheng, W. Easton, D.F. Milne, R.L. Beesley, J. Chenevix-Trench, G (2019) Genome-wide association and transcriptome studies identify target genes and risk loci for breast cancer.. Show Abstract full text

Genome-wide association studies (GWAS) have identified more than 170 breast cancer susceptibility loci. Here we hypothesize that some risk-associated variants might act in non-breast tissues, specifically adipose tissue and immune cells from blood and spleen. Using expression quantitative trait loci (eQTL) reported in these tissues, we identify 26 previously unreported, likely target genes of overall breast cancer risk variants, and 17 for estrogen receptor (ER)-negative breast cancer, several with a known immune function. We determine the directional effect of gene expression on disease risk measured based on single and multiple eQTL. In addition, using a gene-based test of association that considers eQTL from multiple tissues, we identify seven (and four) regions with variants associated with overall (and ER-negative) breast cancer risk, which were not reported in previous GWAS. Further investigation of the function of the implicated genes in breast and immune cells may provide insights into the etiology of breast cancer.

Nichols, H.B. Schoemaker, M.J. Cai, J. Xu, J. Wright, L.B. Brook, M.N. Jones, M.E. Adami, H.-.O. Baglietto, L. Bertrand, K.A. Blot, W.J. Boutron-Ruault, M.-.C. Dorronsoro, M. Dossus, L. Eliassen, A.H. Giles, G.G. Gram, I.T. Hankinson, S.E. Hoffman-Bolton, J. Kaaks, R. Key, T.J. Kitahara, C.M. Larsson, S.C. Linet, M. Merritt, M.A. Milne, R.L. Pala, V. Palmer, J.R. Peeters, P.H. Riboli, E. Sund, M. Tamimi, R.M. Tjønneland, A. Trichopoulou, A. Ursin, G. Vatten, L. Visvanathan, K. Weiderpass, E. Wolk, A. Zheng, W. Weinberg, C.R. Swerdlow, A.J. Sandler, D.P (2019) Breast Cancer Risk After Recent Childbirth: A Pooled Analysis of 15 Prospective Studies.. Show Abstract full text

<h4>Background</h4>Parity is widely recognized as protective for breast cancer, but breast cancer risk may be increased shortly after childbirth. Whether this risk varies with breastfeeding, family history of breast cancer, or specific tumor subtype has rarely been evaluated.<h4>Objective</h4>To characterize breast cancer risk in relation to recent childbirth.<h4>Design</h4>Pooled analysis of individual-level data from 15 prospective cohort studies.<h4>Setting</h4>The international Premenopausal Breast Cancer Collaborative Group.<h4>Participants</h4>Women younger than 55 years.<h4>Measurements</h4>During 9.6 million person-years of follow-up, 18 826 incident cases of breast cancer were diagnosed. Hazard ratios (HRs) and 95% CIs for breast cancer were calculated using Cox proportional hazards regression.<h4>Results</h4>Compared with nulliparous women, parous women had an HR for breast cancer that peaked about 5 years after birth (HR, 1.80 [95% CI, 1.63 to 1.99]) before decreasing to 0.77 (CI, 0.67 to 0.88) after 34 years. The association crossed over from positive to negative about 24 years after birth. The overall pattern was driven by estrogen receptor (ER)-positive breast cancer; no crossover was seen for ER-negative cancer. Increases in breast cancer risk after childbirth were pronounced when combined with a family history of breast cancer and were greater for women who were older at first birth or who had more births. Breastfeeding did not modify overall risk patterns.<h4>Limitations</h4>Breast cancer diagnoses during pregnancy were not uniformly distinguishable from early postpartum diagnoses. Data on human epidermal growth factor receptor 2 (HER2) oncogene overexpression were limited.<h4>Conclusion</h4>Compared with nulliparous women, parous women have an increased risk for breast cancer for more than 20 years after childbirth. Health care providers should consider recent childbirth a risk factor for breast cancer in young women.<h4>Primary funding source</h4>The Avon Foundation, the National Institute of Environmental Health Sciences, Breast Cancer Now and the UK National Health Service, and the Institute of Cancer Research.

Rambau, P.F. Vierkant, R.A. Intermaggio, M.P. Kelemen, L.E. Goodman, M.T. Herpel, E. Pharoah, P.D. Kommoss, S. Jimenez-Linan, M. Karlan, B.Y. Gentry-Maharaj, A. Menon, U. Polo, S.H. Candido Dos Reis, F.J. Doherty, J.A. Gayther, S.A. Sharma, R. Larson, M.C. Harnett, P.R. Hatfield, E. de Andrade, J.M. Nelson, G.S. Steed, H. Schildkraut, J.M. Carney, M.E. Høgdall, E. Whittemore, A.S. Widschwendter, M. Kennedy, C.J. Wang, F. Wang, Q. Wang, C. Armasu, S.M. Daley, F. Coulson, P. Jones, M.E. Anglesio, M.S. Chow, C. de Fazio, A. García-Closas, M. Brucker, S.Y. Cybulski, C. Harris, H.R. Hartkopf, A.D. Huzarski, T. Jensen, A. Lubiński, J. Oszurek, O. Benitez, J. Mina, F. Staebler, A. Taran, F.A. Pasternak, J. Talhouk, A. Rossing, M.A. Hendley, J. AOCS Group, . Edwards, R.P. Fereday, S. Modugno, F. Ness, R.B. Sieh, W. El-Bahrawy, M.A. Winham, S.J. Lester, J. Kjaer, S.K. Gronwald, J. Sinn, P. Fasching, P.A. Chang-Claude, J. Moysich, K.B. Bowtell, D.D. Hernandez, B.Y. Luk, H. Behrens, S. Shah, M. Jung, A. Ghatage, P. Alsop, J. Alsop, K. García-Donas, J. Thompson, P.J. Swerdlow, A.J. Karpinskyj, C. Cazorla-Jiménez, A. García, M.J. Deen, S. Wilkens, L.R. Palacios, J. Berchuck, A. Koziak, J.M. Brenton, J.D. Cook, L.S. Goode, E.L. Huntsman, D.G. Ramus, S.J. Köbel, M (2018) Association of p16 expression with prognosis varies across ovarian carcinoma histotypes: an Ovarian Tumor Tissue Analysis consortium study.. Show Abstract full text

We aimed to validate the prognostic association of p16 expression in ovarian high-grade serous carcinomas (HGSC) and to explore it in other ovarian carcinoma histotypes. p16 protein expression was assessed by clinical-grade immunohistochemistry in 6525 ovarian carcinomas including 4334 HGSC using tissue microarrays from 24 studies participating in the Ovarian Tumor Tissue Analysis consortium. p16 expression patterns were interpreted as abnormal (either overexpression referred to as block expression or absence) or normal (heterogeneous). CDKN2A (which encodes p16) mRNA expression was also analyzed in a subset (n = 2280) mostly representing HGSC (n = 2010). Association of p16 expression with overall survival (OS) was determined within histotypes as was CDKN2A expression for HGSC only. p16 block expression was most frequent in HGSC (56%) but neither protein nor mRNA expression was associated with OS. However, relative to heterogeneous expression, block expression was associated with shorter OS in endometriosis-associated carcinomas, clear cell [hazard ratio (HR): 2.02, 95% confidence (CI) 1.47-2.77, p < 0.001] and endometrioid (HR: 1.88, 95% CI 1.30-2.75, p = 0.004), while absence was associated with shorter OS in low-grade serous carcinomas (HR: 2.95, 95% CI 1.61-5.38, p = 0.001). Absence was most frequent in mucinous carcinoma (50%), and was not associated with OS in this histotype. The prognostic value of p16 expression is histotype-specific and pattern dependent. We provide definitive evidence against an association of p16 expression with survival in ovarian HGSC as previously suggested. Block expression of p16 in clear cell and endometrioid carcinoma should be further validated as a prognostic marker, and absence in low-grade serous carcinoma justifies CDK4 inhibition.

Sud, A. Torr, B. Jones, M.E. Broggio, J. Scott, S. Loveday, C. Garrett, A. Gronthoud, F. Nicol, D.L. Jhanji, S. Boyce, S.A. Williams, M. Riboli, E. Muller, D.C. Kipps, E. Larkin, J. Navani, N. Swanton, C. Lyratzopoulos, G. McFerran, E. Lawler, M. Houlston, R. Turnbull, C (2020) Effect of delays in the 2-week-wait cancer referral pathway during the COVID-19 pandemic on cancer survival in the UK: a modelling study.. Show Abstract full text

<h4>Background</h4>During the COVID-19 lockdown, referrals via the 2-week-wait urgent pathway for suspected cancer in England, UK, are reported to have decreased by up to 84%. We aimed to examine the impact of different scenarios of lockdown-accumulated backlog in cancer referrals on cancer survival, and the impact on survival per referred patient due to delayed referral versus risk of death from nosocomial infection with severe acute respiratory syndrome coronavirus 2.<h4>Methods</h4>In this modelling study, we used age-stratified and stage-stratified 10-year cancer survival estimates for patients in England, UK, for 20 common tumour types diagnosed in 2008-17 at age 30 years and older from Public Health England. We also used data for cancer diagnoses made via the 2-week-wait referral pathway in 2013-16 from the Cancer Waiting Times system from NHS Digital. We applied per-day hazard ratios (HRs) for cancer progression that we generated from observational studies of delay to treatment. We quantified the annual numbers of cancers at stage I-III diagnosed via the 2-week-wait pathway using 2-week-wait age-specific and stage-specific breakdowns. From these numbers, we estimated the aggregate number of lives and life-years lost in England for per-patient delays of 1-6 months in presentation, diagnosis, or cancer treatment, or a combination of these. We assessed three scenarios of a 3-month period of lockdown during which 25%, 50%, and 75% of the normal monthly volumes of symptomatic patients delayed their presentation until after lockdown. Using referral-to-diagnosis conversion rates and COVID-19 case-fatality rates, we also estimated the survival increment per patient referred.<h4>Findings</h4>Across England in 2013-16, an average of 6281 patients with stage I-III cancer were diagnosed via the 2-week-wait pathway per month, of whom 1691 (27%) would be predicted to die within 10 years from their disease. Delays in presentation via the 2-week-wait pathway over a 3-month lockdown period (with an average presentational delay of 2 months per patient) would result in 181 additional lives and 3316 life-years lost as a result of a backlog of referrals of 25%, 361 additional lives and 6632 life-years lost for a 50% backlog of referrals, and 542 additional lives and 9948 life-years lost for a 75% backlog in referrals. Compared with all diagnostics for the backlog being done in month 1 after lockdown, additional capacity across months 1-3 would result in 90 additional lives and 1662 live-years lost due to diagnostic delays for the 25% backlog scenario, 183 additional lives and 3362 life-years lost under the 50% backlog scenario, and 276 additional lives and 5075 life-years lost under the 75% backlog scenario. However, a delay in additional diagnostic capacity with provision spread across months 3-8 after lockdown would result in 401 additional lives and 7332 life-years lost due to diagnostic delays under the 25% backlog scenario, 811 additional lives and 14 873 life-years lost under the 50% backlog scenario, and 1231 additional lives and 22 635 life-years lost under the 75% backlog scenario. A 2-month delay in 2-week-wait investigatory referrals results in an estimated loss of between 0·0 and 0·7 life-years per referred patient, depending on age and tumour type.<h4>Interpretation</h4>Prompt provision of additional capacity to address the backlog of diagnostics will minimise deaths as a result of diagnostic delays that could add to those predicted due to expected presentational delays. Prioritisation of patient groups for whom delay would result in most life-years lost warrants consideration as an option for mitigating the aggregate burden of mortality in patients with cancer.<h4>Funding</h4>None.

Sud, A. Jones, M.E. Broggio, J. Loveday, C. Torr, B. Garrett, A. Nicol, D.L. Jhanji, S. Boyce, S.A. Gronthoud, F. Ward, P. Handy, J.M. Yousaf, N. Larkin, J. Suh, Y.-.E. Scott, S. Pharoah, P.D.P. Swanton, C. Abbosh, C. Williams, M. Lyratzopoulos, G. Houlston, R. Turnbull, C (2020) Collateral damage: the impact on outcomes from cancer surgery of the COVID-19 pandemic.. Show Abstract full text

<h4>Background</h4>Cancer diagnostics and surgery have been disrupted by the response of health care services to the coronavirus disease 2019 (COVID-19) pandemic. Progression of cancers during delay will impact on patients' long-term survival.<h4>Patients and methods</h4>We generated per-day hazard ratios of cancer progression from observational studies and applied these to age-specific, stage-specific cancer survival for England 2013-2017. We modelled per-patient delay of 3 and 6 months and periods of disruption of 1 and 2 years. Using health care resource costing, we contextualise attributable lives saved and life-years gained (LYGs) from cancer surgery to equivalent volumes of COVID-19 hospitalisations.<h4>Results</h4>Per year, 94 912 resections for major cancers result in 80 406 long-term survivors and 1 717 051 LYGs. Per-patient delay of 3/6 months would cause attributable death of 4755/10 760 of these individuals with loss of 92 214/208 275 life-years, respectively. For cancer surgery, average LYGs per patient are 18.1 under standard conditions and 17.1/15.9 with a delay of 3/6 months (an average loss of 0.97/2.19 LYGs per patient), respectively. Taking into account health care resource units (HCRUs), surgery results on average per patient in 2.25 resource-adjusted life-years gained (RALYGs) under standard conditions and 2.12/1.97 RALYGs following delay of 3/6 months. For 94 912 hospital COVID-19 admissions, there are 482 022 LYGs requiring 1 052 949 HCRUs. Hospitalisation of community-acquired COVID-19 patients yields on average per patient 5.08 LYG and 0.46 RALYGs.<h4>Conclusions</h4>Modest delays in surgery for cancer incur significant impact on survival. Delay of 3/6 months in surgery for incident cancers would mitigate 19%/43% of LYGs, respectively, by hospitalisation of an equivalent volume of admissions for community-acquired COVID-19. This rises to 26%/59%, respectively, when considering RALYGs. To avoid a downstream public health crisis of avoidable cancer deaths, cancer diagnostic and surgical pathways must be maintained at normal throughput, with rapid attention to any backlog already accrued.

Kapoor, P.M. Mavaddat, N. Choudhury, P.P. Wilcox, A.N. Lindström, S. Behrens, S. Michailidou, K. Dennis, J. Bolla, M.K. Wang, Q. Jung, A. Abu-Ful, Z. Ahearn, T. Andrulis, I.L. Anton-Culver, H. Arndt, V. Aronson, K.J. Auer, P.L. Freeman, L.E.B. Becher, H. Beckmann, M.W. Beeghly-Fadiel, A. Benitez, J. Bernstein, L. Bojesen, S.E. Brauch, H. Brenner, H. Brüning, T. Cai, Q. Campa, D. Canzian, F. Carracedo, A. Carter, B.D. Castelao, J.E. Chanock, S.J. Chatterjee, N. Chenevix-Trench, G. Clarke, C.L. Couch, F.J. Cox, A. Cross, S.S. Czene, K. Dai, J.Y. Earp, H.S. Ekici, A.B. Eliassen, A.H. Eriksson, M. Evans, D.G. Fasching, P.A. Figueroa, J. Fritschi, L. Gabrielson, M. Gago-Dominguez, M. Gao, C. Gapstur, S.M. Gaudet, M.M. Giles, G.G. González-Neira, A. Guénel, P. Haeberle, L. Haiman, C.A. Håkansson, N. Hall, P. Hamann, U. Hatse, S. Heyworth, J. Holleczek, B. Hoover, R.N. Hopper, J.L. Howell, A. Hunter, D.J. ABCTB Investigators, . kConFab/AOCS Investigators, . John, E.M. Jones, M.E. Kaaks, R. Keeman, R. Kitahara, C.M. Ko, Y.-.D. Koutros, S. Kurian, A.W. Lambrechts, D. Le Marchand, L. Lee, E. Lejbkowicz, F. Linet, M. Lissowska, J. Llaneza, A. MacInnis, R.J. Martinez, M.E. Maurer, T. McLean, C. Neuhausen, S.L. Newman, W.G. Norman, A. O'Brien, K.M. Olshan, A.F. Olson, J.E. Olsson, H. Orr, N. Perou, C.M. Pita, G. Polley, E.C. Prentice, R.L. Rennert, G. Rennert, H.S. Ruddy, K.J. Sandler, D.P. Saunders, C. Schoemaker, M.J. Schöttker, B. Schumacher, F. Scott, C. Scott, R.J. Shu, X.-.O. Smeets, A. Southey, M.C. Spinelli, J.J. Stone, J. Swerdlow, A.J. Tamimi, R.M. Taylor, J.A. Troester, M.A. Vachon, C.M. van Veen, E.M. Wang, X. Weinberg, C.R. Weltens, C. Willett, W. Winham, S.J. Wolk, A. Yang, X.R. Zheng, W. Ziogas, A. Dunning, A.M. Pharoah, P.D.P. Schmidt, M.K. Kraft, P. Easton, D.F. Milne, R.L. García-Closas, M. Chang-Claude, J (2021) Combined Associations of a Polygenic Risk Score and Classical Risk Factors With Breast Cancer Risk.. Show Abstract full text

We evaluated the joint associations between a new 313-variant PRS (PRS313) and questionnaire-based breast cancer risk factors for women of European ancestry, using 72 284 cases and 80 354 controls from the Breast Cancer Association Consortium. Interactions were evaluated using standard logistic regression and a newly developed case-only method for breast cancer risk overall and by estrogen receptor status. After accounting for multiple testing, we did not find evidence that per-standard deviation PRS313 odds ratio differed across strata defined by individual risk factors. Goodness-of-fit tests did not reject the assumption of a multiplicative model between PRS313 and each risk factor. Variation in projected absolute lifetime risk of breast cancer associated with classical risk factors was greater for women with higher genetic risk (PRS313 and family history) and, on average, 17.5% higher in the highest vs lowest deciles of genetic risk. These findings have implications for risk prevention for women at increased risk of breast cancer.

Schoemaker, M.J. Nichols, H.B. Wright, L.B. Brook, M.N. Jones, M.E. O'Brien, K.M. Adami, H.-.O. Baglietto, L. Bernstein, L. Bertrand, K.A. Boutron-Ruault, M.-.C. Chen, Y. Connor, A.E. Dossus, L. Eliassen, A.H. Giles, G.G. Gram, I.T. Hankinson, S.E. Kaaks, R. Key, T.J. Kirsh, V.A. Kitahara, C.M. Larsson, S.C. Linet, M. Ma, H. Milne, R.L. Ozasa, K. Palmer, J.R. Riboli, E. Rohan, T.E. Sacerdote, C. Sadakane, A. Sund, M. Tamimi, R.M. Trichopoulou, A. Ursin, G. Visvanathan, K. Weiderpass, E. Willett, W.C. Wolk, A. Zeleniuch-Jacquotte, A. Sandler, D.P. Swerdlow, A.J (2020) Adult weight change and premenopausal breast cancer risk: A prospective pooled analysis of data from 628,463 women.. Show Abstract full text

Early-adulthood body size is strongly inversely associated with risk of premenopausal breast cancer. It is unclear whether subsequent changes in weight affect risk. We pooled individual-level data from 17 prospective studies to investigate the association of weight change with premenopausal breast cancer risk, considering strata of initial weight, timing of weight change, other breast cancer risk factors and breast cancer subtype. Hazard ratios (HR) and 95% confidence intervals (CI) were obtained using Cox regression. Among 628,463 women, 10,886 were diagnosed with breast cancer before menopause. Models adjusted for initial weight at ages 18-24 years and other breast cancer risk factors showed that weight gain from ages 18-24 to 35-44 or to 45-54 years was inversely associated with breast cancer overall (e.g., HR per 5 kg to ages 45-54: 0.96, 95% CI: 0.95-0.98) and with oestrogen-receptor(ER)-positive breast cancer (HR per 5 kg to ages 45-54: 0.96, 95% CI: 0.94-0.98). Weight gain from ages 25-34 was inversely associated with ER-positive breast cancer only and weight gain from ages 35-44 was not associated with risk. None of these weight gains were associated with ER-negative breast cancer. Weight loss was not consistently associated with overall or ER-specific risk after adjusting for initial weight. Weight increase from early-adulthood to ages 45-54 years is associated with a reduced premenopausal breast cancer risk independently of early-adulthood weight. Biological explanations are needed to account for these two separate factors.

Manoochehri, M. Jones, M. Tomczyk, K. Fletcher, O. Schoemaker, M.J. Swerdlow, A.J. Borhani, N. Hamann, U (2020) DNA methylation of the long intergenic noncoding RNA 299 gene in triple-negative breast cancer: results from a prospective study.. Show Abstract full text

Triple-negative breast cancer (TNBC) is an aggressive breast cancer subtype associated with a high rate of recurrence and poor prognosis. Recently we identified a hypermethylation in the long noncoding RNA 299 (LINC00299) gene in blood-derived DNA from TNBC patients compared with healthy controls implying that LINC00299 hypermethylation may serve as a circulating biomarker for TNBC. In the present study, we investigated whether LINC00299 methylation is associated with TNBC in a prospective nested breast cancer case-control study within the Generations Study. Methylation at cg06588802 in LINC00299 was measured in 154 TNBC cases and 159 breast cancer-free matched controls using MethyLight droplet digital PCR. To assess the association between methylation level and TNBC risk, logistic regression was used to calculate odd ratios and 95% confidence intervals, adjusted for smoking status. We found no evidence for association between methylation levels and TNBC overall (P = 0.062). Subgroup analysis according to age at diagnosis and age at blood draw revealed increased methylation levels in TNBC cases compared with controls in the young age groups [age 26-52 (P = 0.0025) and age 22-46 (P = 0.001), respectively]. Our results suggest a potential association of LINC00299 hypermethylation with TNBC in young women.

Maguire, S. Perraki, E. Tomczyk, K. Jones, M.E. Fletcher, O. Pugh, M. Winter, T. Thompson, K. Cooke, R. kConFab Consortium, . Trainer, A. James, P. Bojesen, S. Flyger, H. Nevanlinna, H. Mattson, J. Friedman, E. Laitman, Y. Palli, D. Masala, G. Zanna, I. Ottini, L. Silvestri, V. Hollestelle, A. Hooning, M.J. Novaković, S. Krajc, M. Gago-Dominguez, M. Castelao, J.E. Olsson, H. Hedenfalk, I. Saloustros, E. Georgoulias, V. Easton, D.F. Pharoah, P. Dunning, A.M. Bishop, D.T. Neuhausen, S.L. Steele, L. Ashworth, A. Garcia Closas, M. Houlston, R. Swerdlow, A. Orr, N (2021) Common Susceptibility Loci for Male Breast Cancer.. Show Abstract full text

<h4>Background</h4>The etiology of male breast cancer (MBC) is poorly understood. In particular, the extent to which the genetic basis of MBC differs from female breast cancer (FBC) is unknown. A previous genome-wide association study of MBC identified 2 predisposition loci for the disease, both of which were also associated with risk of FBC.<h4>Methods</h4>We performed genome-wide single nucleotide polymorphism genotyping of European ancestry MBC case subjects and controls in 3 stages. Associations between directly genotyped and imputed single nucleotide polymorphisms with MBC were assessed using fixed-effects meta-analysis of 1380 cases and 3620 controls. Replication genotyping of 810 cases and 1026 controls was used to validate variants with P values less than 1 × 10-06. Genetic correlation with FBC was evaluated using linkage disequilibrium score regression, by comprehensively examining the associations of published FBC risk loci with risk of MBC and by assessing associations between a FBC polygenic risk score and MBC. All statistical tests were 2-sided.<h4>Results</h4>The genome-wide association study identified 3 novel MBC susceptibility loci that attained genome-wide statistical significance (P < 5 × 10-08). Genetic correlation analysis revealed a strong shared genetic basis with estrogen receptor-positive FBC. Men in the top quintile of genetic risk had a fourfold increased risk of breast cancer relative to those in the bottom quintile (odds ratio = 3.86, 95% confidence interval = 3.07 to 4.87, P = 2.08 × 10-30).<h4>Conclusions</h4>These findings advance our understanding of the genetic basis of MBC, providing support for an overlapping genetic etiology with FBC and identifying a fourfold high-risk group of susceptible men.

Loveday, C. Sud, A. Jones, M.E. Broggio, J. Scott, S. Gronthound, F. Torr, B. Garrett, A. Nicol, D.L. Jhanji, S. Boyce, S.A. Williams, M. Barry, C. Riboli, E. Kipps, E. McFerran, E. Muller, D.C. Lyratzopoulos, G. Lawler, M. Abulafi, M. Houlston, R.S. Turnbull, C (2021) Prioritisation by FIT to mitigate the impact of delays in the 2-week wait colorectal cancer referral pathway during the COVID-19 pandemic: a UK modelling study.. Show Abstract full text

<h4>Objective</h4>To evaluate the impact of faecal immunochemical testing (FIT) prioritisation to mitigate the impact of delays in the colorectal cancer (CRC) urgent diagnostic (2-week-wait (2WW)) pathway consequent from the COVID-19 pandemic.<h4>Design</h4>We modelled the reduction in CRC survival and life years lost resultant from per-patient delays of 2-6 months in the 2WW pathway. We stratified by age group, individual-level benefit in CRC survival versus age-specific nosocomial COVID-19-related fatality per referred patient undergoing colonoscopy. We modelled mitigation strategies using thresholds of FIT triage of 2, 10 and 150 µg Hb/g to prioritise 2WW referrals for colonoscopy. To construct the underlying models, we employed 10-year net CRC survival for England 2008-2017, 2WW pathway CRC case and referral volumes and per-day-delay HRs generated from observational studies of diagnosis-to-treatment interval.<h4>Results</h4>Delay of 2/4/6 months across all 11 266 patients with CRC diagnosed per typical year via the 2WW pathway were estimated to result in 653/1419/2250 attributable deaths and loss of 9214/20 315/32 799 life years. Risk-benefit from urgent investigatory referral is particularly sensitive to nosocomial COVID-19 rates for patients aged >60. Prioritisation out of delay for the 18% of symptomatic referrals with FIT >10 µg Hb/g would avoid 89% of these deaths attributable to presentational/diagnostic delay while reducing immediate requirement for colonoscopy by >80%.<h4>Conclusions</h4>Delays in the pathway to CRC diagnosis and treatment have potential to cause significant mortality and loss of life years. FIT triage of symptomatic patients in primary care could streamline access to colonoscopy, reduce delays for true-positive CRC cases and reduce nosocomial COVID-19 mortality in older true-negative 2WW referrals. However, this strategy offers benefit only in short-term rationalisation of limited endoscopy services: the appreciable false-negative rate of FIT in symptomatic patients means most colonoscopies will still be required.

Johnson, N. Maguire, S. Morra, A. Kapoor, P.M. Tomczyk, K. Jones, M.E. Schoemaker, M.J. Gilham, C. Bolla, M.K. Wang, Q. Dennis, J. Ahearn, T.U. Andrulis, I.L. Anton-Culver, H. Antonenkova, N.N. Arndt, V. Aronson, K.J. Augustinsson, A. Baynes, C. Freeman, L.E.B. Beckmann, M.W. Benitez, J. Bermisheva, M. Blomqvist, C. Boeckx, B. Bogdanova, N.V. Bojesen, S.E. Brauch, H. Brenner, H. Burwinkel, B. Campa, D. Canzian, F. Castelao, J.E. Chanock, S.J. Chenevix-Trench, G. Clarke, C.L. NBCS Collaborators, . Conroy, D.M. Couch, F.J. Cox, A. Cross, S.S. Czene, K. Dörk, T. Eliassen, A.H. Engel, C. Evans, D.G. Fasching, P.A. Figueroa, J. Floris, G. Flyger, H. Gago-Dominguez, M. Gapstur, S.M. García-Closas, M. Gaudet, M.M. Giles, G.G. Goldberg, M.S. González-Neira, A. AOCS Group, . Guénel, P. Hahnen, E. Haiman, C.A. Håkansson, N. Hall, P. Hamann, U. Harrington, P.A. Hart, S.N. Hooning, M.J. Hopper, J.L. Howell, A. Hunter, D.J. ABCTB Investigators, . kConFab Investigators, . Jager, A. Jakubowska, A. John, E.M. Kaaks, R. Keeman, R. Khusnutdinova, E. Kitahara, C.M. Kosma, V.-.M. Koutros, S. Kraft, P. Kristensen, V.N. Kurian, A.W. Lambrechts, D. Le Marchand, L. Linet, M. Lubiński, J. Mannermaa, A. Manoukian, S. Margolin, S. Martens, J.W.M. Mavroudis, D. Mayes, R. Meindl, A. Milne, R.L. Neuhausen, S.L. Nevanlinna, H. Newman, W.G. Nielsen, S.F. Nordestgaard, B.G. Obi, N. Olshan, A.F. Olson, J.E. Olsson, H. Orban, E. Park-Simon, T.-.W. Peterlongo, P. Plaseska-Karanfilska, D. Pylkäs, K. Rennert, G. Rennert, H.S. Ruddy, K.J. Saloustros, E. Sandler, D.P. Sawyer, E.J. Schmutzler, R.K. Scott, C. Shu, X.-.O. Simard, J. Smichkoska, S. Sohn, C. Southey, M.C. Spinelli, J.J. Stone, J. Tamimi, R.M. Taylor, J.A. Tollenaar, R.A.E.M. Tomlinson, I. Troester, M.A. Truong, T. Vachon, C.M. van Veen, E.M. Wang, S.S. Weinberg, C.R. Wendt, C. Wildiers, H. Winqvist, R. Wolk, A. Zheng, W. Ziogas, A. Dunning, A.M. Pharoah, P.D.P. Easton, D.F. Howie, A.F. Peto, J. Dos-Santos-Silva, I. Swerdlow, A.J. Chang-Claude, J. Schmidt, M.K. Orr, N. Fletcher, O (2021) CYP3A7*1C allele: linking premenopausal oestrone and progesterone levels with risk of hormone receptor-positive breast cancers.. Show Abstract full text

<h4>Background</h4>Epidemiological studies provide strong evidence for a role of endogenous sex hormones in the aetiology of breast cancer. The aim of this analysis was to identify genetic variants that are associated with urinary sex-hormone levels and breast cancer risk.<h4>Methods</h4>We carried out a genome-wide association study of urinary oestrone-3-glucuronide and pregnanediol-3-glucuronide levels in 560 premenopausal women, with additional analysis of progesterone levels in 298 premenopausal women. To test for the association with breast cancer risk, we carried out follow-up genotyping in 90,916 cases and 89,893 controls from the Breast Cancer Association Consortium. All women were of European ancestry.<h4>Results</h4>For pregnanediol-3-glucuronide, there were no genome-wide significant associations; for oestrone-3-glucuronide, we identified a single peak mapping to the CYP3A locus, annotated by rs45446698. The minor rs45446698-C allele was associated with lower oestrone-3-glucuronide (-49.2%, 95% CI -56.1% to -41.1%, P = 3.1 × 10<sup>-18</sup>); in follow-up analyses, rs45446698-C was also associated with lower progesterone (-26.7%, 95% CI -39.4% to -11.6%, P = 0.001) and reduced risk of oestrogen and progesterone receptor-positive breast cancer (OR = 0.86, 95% CI 0.82-0.91, P = 6.9 × 10<sup>-8</sup>).<h4>Conclusions</h4>The CYP3A7*1C allele is associated with reduced risk of hormone receptor-positive breast cancer possibly mediated via an effect on the metabolism of endogenous sex hormones in premenopausal women.

Jones, M.E. Schoemaker, M.J. Wright, L. McFadden, E. Griffin, J. Thomas, D. Hemming, J. Wright, K. Ashworth, A. Swerdlow, A.J (2016) Menopausal hormone therapy and breast cancer: what is the true size of the increased risk?. Show Abstract full text

BACKGROUND: Menopausal hormone therapy (MHT) increases breast cancer risk; however, most cohort studies omit MHT use after enrolment and many infer menopausal age. METHODS: We used information from serial questionnaires from the UK Generations Study cohort to estimate hazard ratios (HRs) for breast cancer among post-menopausal women with known menopausal age, and examined biases induced when not updating data on MHT use and including women with inferred menopausal age. RESULTS: Among women recruited in 2003-2009, at 6 years of follow-up, 58 148 had reached menopause and 96% had completed a follow-up questionnaire. Among 39 183 women with known menopausal age, 775 developed breast cancer, and the HR in relation to current oestrogen plus progestogen MHT use (based on 52 current oestrogen plus progestogen MHT users in breast cancer cases) relative to those with no previous MHT use was 2.74 (95% confidence interval (CI): 2.05-3.65) for a median duration of 5.4 years of current use, reaching 3.27 (95% CI: 1.53-6.99) at 15+ years of use. The excess HR was underestimated by 53% if oestrogen plus progestogen MHT use was not updated after recruitment, 13% if women with uncertain menopausal age were included, and 59% if both applied. The HR for oestrogen-only MHT was not increased (HR=1.00; 95% CI: 0.66-1.54). CONCLUSIONS: Lack of updating MHT status through follow-up and inclusion of women with inferred menopausal age is likely to result in substantial underestimation of the excess relative risks for oestrogen plus progestogen MHT use in studies with long follow-up, limited updating of exposures, and changing or short durations of use.

Jones, M.E. Schoemaker, M.J. Wright, L.B. Ashworth, A. Swerdlow, A.J (2017) Smoking and risk of breast cancer in the Generations Study cohort.. Show Abstract full text

<h4>Background</h4>Plausible biological reasons exist regarding why smoking could affect breast cancer risk, but epidemiological evidence is inconsistent.<h4>Methods</h4>We used serial questionnaire information from the Generations Study cohort (United Kingdom) to estimate HRs for breast cancer in relation to smoking adjusted for potentially confounding factors, including alcohol intake.<h4>Results</h4>Among 102,927 women recruited 2003-2013, with an average of 7.7 years of follow-up, 1815 developed invasive breast cancer. The HR (reference group was never smokers) was 1.14 (95% CI 1.03-1.25; P = 0.010) for ever smokers, 1.24 (95% CI 1.08-1.43; P = 0.002) for starting smoking at ages < 17 years, and 1.23 (1.07-1.41; P = 0.004) for starting smoking 1-4 years after menarche. Breast cancer risk was not statistically associated with interval from initiation of smoking to first birth (P-trend = 0.97). Women with a family history of breast cancer (ever smoker vs never smoker HR 1.35; 95% CI 1.12-1.62; P = 0.002) had a significantly larger HR in relation to ever smokers (P for interaction = 0.039) than women without (ever smoker vs never smoker HR 1.07; 95% CI 0.96-1.20; P = 0.22). The interaction was prominent for age at starting smoking (P = 0.003) and starting smoking relative to age at menarche (P = 0.0001).<h4>Conclusions</h4>Smoking was associated with a modest but significantly increased risk of breast cancer, particularly among women who started smoking at adolescent or peri-menarcheal ages. The relative risk of breast cancer associated with smoking was greater for women with a family history of the disease.

Swerdlow, A.J. Higgins, C.D. Adlard, P. Jones, M.E. Preece, M.A (2003) Creutzfeldt-Jakob disease in United Kingdom patients treated with human pituitary growth hormone.. Show Abstract full text

OBJECTIVE: To investigate risk factors for Creutzfeldt-Jakob disease (CJD) in patients in the United Kingdom treated with human pituitary growth hormone (hGH). METHODS: Incidence rates of CJD, based on person-year denominators, were assessed in a cohort of 1,848 patients treated with hGH in the United Kingdom from 1959 through 1985 and followed to the end of 2000. RESULTS: CJD developed in 38 patients. Risk of CJD was significantly increased by treatment with hGH prepared by the Wilhelmi method of extraction from human pituitaries. Risk was further raised if this treatment was administered at ages 8 to 10 years. The peak risk of CJD was estimated to occur 20 years after first exposure, and the estimated lifetime cumulative risk of CJD in Wilhelmi-treated patients was 4.5%. CONCLUSIONS: Size-exclusion chromatography, used in non-Wilhelmi preparation methods, may prevent CJD infection. Susceptibility to CJD may vary with age, and susceptibility may be present in only a few percent of the population.

Crook, P.D. Jones, M.E. Hall, A.J (2003) Mortality of hepatitis B surface antigen-positive blood donors in England and Wales. full text
dos Santos Silva, I. Malveiro, F. Jones, M.E. Swerdlow, A.J (2003) Mortality after radiological investigation with radioactive Thorotrast: a follow-up study of up to fifty years in Portugal.. Show Abstract full text

Cerebral angiography using a radioactive radiological contrast medium, Thorotrast, was pioneered by Moniz in Portugal in the 1920s. Thorotrast is retained by the reticuloendothelial system, with a biological half-life of several hundred years, so that such patients suffer lifetime exposure to internal radiation. We studied mortality in Portuguese patients who were administered Thorotrast during the period 1928-1959 and in a comparison group of patients who received nonradioactive contrast agents. There were 1096 systemically exposed, 1014 unexposed, and, unique to the Portuguese study, 240 locally exposed Thorotrast patients who were successfully traced and followed up to the end of 1996. Mortality was significantly raised among systemically exposed Thorotrast patients relative to those unexposed for all causes [relative risk (RR) = 2.63], all neoplasms (RR = 6.72), liver cancer (RR = 42.4), chronic liver disease (RR = 5.12), other non-neoplastic diseases of the digestive system (RR = 4.87), neoplastic (RR = 21.9) and non-neoplastic hematological disorders (RR = 6.00), and non-neoplastic diseases of the respiratory system (RR = 4.31). Risks for most of these conditions increased significantly with time since first administration of the contrast medium and with cumulative alpha-particle radiation dose. Mortality was also significantly raised for non-neoplastic disorders of the nervous system (RR = 12.7) and ill-defined conditions (RR = 3.74), but these associations are likely to reflect the initial diagnosis, not Thorotrast exposure, because risks declined significantly with time and/or dose. There were no significant excess deaths from oropharyngeal or nasal cancers, or from any other cause, among patients exposed to Thorotrast locally for visualization of the perinasal sinuses, and no clear trend in risk with time since exposure. This study shows an association between systemic, but not local, exposure to Thorotrast and mortality from liver cancer, chronic liver disease, and neoplastic and non-neoplastic hematological disorders, with risks for these conditions remaining high for over 40 years after administration. Liver conditions, but not hematological disorders, showed a strong and consistent gradient with cumulative alpha-particle radiation dose.

Ponsonby, A.L. Dwyer, T. Gibbons, L.E. Cochrane, J.A. Jones, M.E. McCall, M.J (1992) Thermal environment and sudden infant death syndrome: case-control study. Show Abstract full text

OBJECTIVE: To compare the thermal environment of infants who died of the sudden infant death syndrome with that of age matched control infants. DESIGN: Case-control study. Infants who died were matched with two controls, one for age and one for age and birth weight. Thermal measurements were conducted at the death scene for cases and at the scene of last sleep for control infants, who were visited unexpectedly within four weeks of the index infant's death on a day of similar climatic conditions. A follow up questionnaire was administered to parents of cases and controls. SETTING: The geographical area served by the professional Tasmanian state ambulance service, which includes 94% of the Tasmanian population. SUBJECTS: 41 infants died of the sudden infant death syndrome at home; thermal observations at death scene were available for 28 (68%), parental questionnaire data were available for 40 (96%). 38 controls matched for age and 41 matched for age and birth weight. RESULTS: Cases had more excess thermal insulation for their given room temperature (2.3 togs) than matched controls (0.6 togs) (p = 0.009). For every excess thermal insulation unit (tog) the relative risk of the sudden infant death syndrome was 1.26 (95% confidence interval 1.05 to 1.52). The average thermal bedding value calculated from parental recall was similar to that observed by attendant ambulance officers (mean difference = 0.4 tog, p = 0.39). Cases were more likely to have been found prone (odds ratio 4.58; 1.48 to 14.11). Prone sleeping position was not a confounder or effect modifier of the relation between excess thermal insulation and the syndrome. CONCLUSIONS: Overheating and the prone sleeping position are independently associated with an increased risk of the sudden infant death syndrome. Further work on infant thermal balance and sudden infant death is required and guidelines for appropriate infant thermal care need to be developed.

Jones, M.E. Schoemaker, M.J. McFadden, E.C. Wright, L.B. Johns, L.E. Swerdlow, A.J (2019) Night shift work and risk of breast cancer in women: the Generations Study cohort.. Show Abstract full text

<h4>Background</h4>It is plausible that night shift work could affect breast cancer risk, possibly by melatonin suppression or circadian clock disruption, but epidemiological evidence is inconclusive.<h4>Methods</h4>Using serial questionnaires from the Generations Study cohort, we estimated hazard ratios (HR) and 95% confidence intervals (95%CI) for breast cancer in relation to being a night shift worker within the last 10 years, adjusted for potential confounders.<h4>Results</h4>Among 102,869 women recruited in 2003-2014, median follow-up 9.5 years, 2059 developed invasive breast cancer. The HR in relation to night shift work was 1.00 (95%CI: 0.86-1.15). There was a significant trend with average hours of night work per week (P = 0.035), but no significantly raised risks for hours worked per night, nights worked per week, average hours worked per week, cumulative years of employment, cumulative hours, time since cessation, type of occupation, age starting night shift work, or age starting in relation to first pregnancy.<h4>Conclusions</h4>The lack of overall association, and no association with all but one measure of dose, duration, and intensity in our data, does not support an increased risk of breast cancer from night shift work in women.

Jones, M.E. Folkerd, E.J. Doody, D.A. Iqbal, J. Dowsett, M. Ashworth, A. Swerdlow, A.J (2007) Effect of delays in processing blood samples on measured endogenous plasma sex hormone levels in women.. Show Abstract full text

Time spent in transit may affect the concentration of various constituents of collected blood samples and, consequently, results of sex hormone assays. Whole blood was collected from 46 women, and one third was processed immediately, one third was stored at ambient conditions (22 degrees C) for 1 day, and one third was stored for 2 days. Estradiol concentration increased by 7.1% [95% confidence interval (95% CI), 3.2-11.3%] after a delay in processing of 1 day and by 5.6% (95% CI, 0.2-11.4%) after a delay in processing of 2 days; the change was most apparent at lower than median concentrations. Progesterone concentrations showed no substantial change. Testosterone concentrations changed by 23.9% (95% CI, 17.8-30.3%) after a delay of 1 day but little thereafter. The sex hormone-binding globulin concentration decreased by 6.6% (95% CI, 4.6-8.6%) and 10.9% (95% CI, 8.1-13.6%), follicle-stimulating hormone increased by 7.4% (95% CI, 4.2-10.7%) and 13.9% (95% CI, 8.7-19.3%), and luteinizing hormone increased by 4.9% (95% CI, 1.3-8.5%) and 6.7% (95% CI, 2.2-11.5%) after a delay in processing of 1 and 2 days. Increases in calculated values for biologically available levels of estradiol and testosterone were greater than the increases seen in measured total hormone concentrations. Similar changes are likely when samples are delayed in transit, and evidence of etiology may be obscured unless study designs or analyses take into account processing delays.

Jones, M.E. Swerdlow, A.J (1998) Bias in the standardized mortality ratio when using general population rates to estimate expected number of deaths.. Show Abstract full text

Cohort studies often compare the observed number of cases arising in a group under investigation with the number expected to occur on the basis of general population rates. The general population is taken to represent unexposed persons, but it is almost inevitably biased in that it comprises all types of people including exposed ones. To identify circumstances when this bias matters, the authors modeled its effect in relation to the size of the observed standardized mortality ratio (SMR) and the prevalence of exposed individuals in the general population. The authors found that bias may be a major problem, causing substantial underestimation of the true relative risk, when either the prevalence of exposure in the general population or the SMR are large. The bias can cause an apparent trend in SMRs with age when none exists. It also places a limit on the maximum size of the observed SMR, no matter how large the true relative risk. A table is provided showing the extent of bias in different circumstances. Cohort studies of people with common diseases or exposures, or that find large SMRs, when using general population expectations, need to consider the extent of bias from this source.

Seah, M.P. Jones, M.E. Anthony, M.T (1984) Quantitative Xps - the Calibration of Spectrometer Intensity Energy Response Functions .2. Results of Interlaboratory Measurements for Commerical Instruments.
Swerdlow, A.J. Jones, M.E. British Tamoxifen Second Cancer Study Group, (2005) Tamoxifen treatment for breast cancer and risk of endometrial cancer: a case-control study.. Show Abstract full text

BACKGROUND: Tamoxifen treatment of breast cancer is associated with an increased risk of endometrial cancer, but tamoxifen-related risks of endometrial cancer are unclear in premenopausal women, in long-term users of tamoxifen, and in women for whom several years have passed since ending treatment. We conducted a case-control study in Britain to investigate these risks. METHODS: We compared treatment information on 813 case patients who had endometrial cancer after their diagnosis for breast cancer and 1067 control patients who had breast cancer but not subsequent endometrial cancer. We assessed risk by conditional logistic regression analysis. All statistical tests were two-sided. RESULTS: Overall, tamoxifen treatment, compared with no treatment, was associated with an increased risk of endometrial cancer (odds ratio [OR] = 2.4; 95% confidence interval [CI] = 1.8 to 3.0). Risk increased statistically significantly (P(trend)<.001) with duration of treatment (for > or =5 years of treatment compared with no treatment, OR = 3.6, 95% CI = 2.6 to 4.8). As an indication of background levels of treatment, 16% of control patients received 5 years or more of treatment. Risk of endometrial cancer adjusted for treatment duration did not diminish in follow-up to at least 5 years after the last treatment ended. Risk of endometrial cancer was not associated with the daily dose of tamoxifen and was comparable in pre- and postmenopausal women. Ever treatment with tamoxifen was associated with a much greater risk of Mullerian and mesodermal mixed endometrial tumors (OR = 13.5, 95% CI = 4.1 to 44.5) than of adenocarcinoma (OR = 2.1, 95% CI = 1.6 to 2.7) or clear cell and papillary serous tumors (OR = 3.1, 95% CI = 0.8 to 17.9). CONCLUSIONS: There is an increasing risk of endometrial cancer associated with longer tamoxifen treatment, extending well beyond 5 years. The increased risk of endometrial cancer associated with tamoxifen treatment should be considered clinically for both premenopausal and postmenopausal women during treatment and for at least 5 years after the last treatment.

Swerdlow, A.J. Wright, L.B. Schoemaker, M.J. Jones, M.E (2018) Maternal breast cancer risk in relation to birthweight and gestation of her offspring.. Show Abstract full text

<h4>Background</h4>Parity and age at first pregnancy are well-established risk factors for breast cancer, but the effects of other characteristics of pregnancies are uncertain and the literature is inconsistent.<h4>Methods</h4>In a cohort of 83,451 parous women from the general population of the UK, which collected detailed information on each pregnancy and a wide range of potential confounders, we investigated the associations of length of gestation and birthweight of offspring in a woman's pregnancies with her breast cancer risk, adjusting for a full range of non-reproductive as well as reproductive risk factors unlike in previous large studies.<h4>Results</h4>Gestation of the first-born offspring was significantly inversely related to the risk of pre-menopausal breast cancer (p trend = 0.03; hazard ratio (HR) for 26-31 compared with 40-41 weeks, the baseline group, = 2.38, 95% confidence interval (CI) 1.26-4.49), and was borderline significantly related to risk of breast cancer overall (p trend = 0.05). Risk was significantly raised in mothers of high birthweight first-born (HR for breast cancer overall = 1.53, 95% CI 1.06-2.21 for ≥ 4500 g compared with 3000-3499 g, the baseline group). For gestation and birthweight of most recent birth, there were no clear effects. Analyses without adjustment for confounders (other than age) gave similar results.<h4>Conclusions</h4>Our data add to evidence that short gestation pregnancies may increase the risk of breast cancer, at least pre-menopausally, perhaps by hormonal stimulation and breast proliferation early in pregnancy without the opportunity for the differentiation that occurs in late pregnancy. High birthweight first pregnancies may increase breast cancer risk, possibly through the association of birthweight with oestrogen and insulin-like growth factor 1 levels.

Dwyer, T. Viney, R. Jones, M (1991) Assessing school health education programs. Show Abstract full text

This review focuses on the component of health education directed at achieving changes in health behavior. Much of the work in this field has centered on health behavior that has a role in preventing future disease. Because the evidence is strongest in relation to coronary heart disease (17), considerable effort has been devoted to this area. Walter et al. (34) indicated that the most relevant forms of health behavior to be considered in school-based programs on heart disease are those relating to diet, physical activity, and smoking. Programs relating to each of these behaviors are addressed here.

Michailidou, K. Hall, P. Gonzalez-Neira, A. Ghoussaini, M. Dennis, J. Milne, R.L. Schmidt, M.K. Chang-Claude, J. Bojesen, S.E. Bolla, M.K. Wang, Q. Dicks, E. Lee, A. Turnbull, C. Rahman, N. Fletcher, O. Peto, J. Gibson, L. Silva, I.D.S. Nevanlinna, H. Muranen, T.A. Aittomaki, K. Blomqvist, C. Czene, K. Irwanto, A. Liu, J. Waisfisz, Q. Meijers-Heijboer, H. Adank, M. van der Luijt, R.B. Hein, R. Dahmen, N. Beckman, L. Meindl, A. Schmutzler, R.K. Mueller-Myhsok, B. Lichtner, P. Hopper, J.L. Southey, M.C. Makalic, E. Schmidt, D.F. Uitterlinden, A.G. Hofman, A. Hunter, D.J. Chanock, S.J. Vincent, D. Bacot, F. Tessier, D.C. Canisius, S. Wessels, L.F.A. Haiman, C.A. Shah, M. Luben, R. Brown, J. Luccarini, C. Schoof, N. Humphreys, K. Li, J. Nordestgaard, B.G. Nielsen, S.F. Flyger, H. Couch, F.J. Wang, X. Vachon, C. Stevens, K.N. Lambrechts, D. Moisse, M. Paridaens, R. Christiaens, M.-.R. Rudolph, A. Nickels, S. Flesch-Janys, D. Johnson, N. Aitken, Z. Aaltonen, K. Heikkinen, T. Broeks, A. Van't Veer, L.J. van der Schoot, C.E. Guenel, P. Truong, T. Laurent-Puig, P. Menegaux, F. Marme, F. Schneeweiss, A. Sohn, C. Burwinke, B. Pilar Zamora, M. Arias Perez, J.I. Pita, G. Rosario Alonso, M. Cox, A. Brock, I.W. Cross, S.S. Reed, M.W.R. Sawyer, E.J. Tomlinson, I. Kerin, M.J. Miller, N. Henderson, B.E. Schumacher, F. Le Marchand, L. Andrulis, I.L. Knight, J.A. Glendon, G. Mulligan, A.M. Lindblom, A. Margolin, S. Hooning, M.J. Hollestelle, A. van den Ouweland, A.M.W. Jager, A. Bui, Q.M. Stone, J. Dite, G.S. Apicella, C. Tsimiklis, H. Giles, G.G. Severi, G. Baglietto, L. Fasching, P.A. Haeberle, L. Ekici, A.B. Beckmann, M.W. Brenner, H. Mueller, H. Arndt, V. Stegmaier, C. Swerdlown, A. Ashworth, A. Orr, N. Jones, M. Figueroa, J. Lissowska, J. Brinton, L. Goldberg, M.S. Labreche, F. Dumont, M. Winqvist, R. Pylkas, K. Jukkola-Vuorinen, A. Grip, M. Brauch, H. Hamann, U. Bruening, T. Radice, P. Peterlongo, P. Manouldan, S. Bonanni, B. Devilee, P. Tollenaar, R.A.E.M. Seynaeve, C. van Asperen, C.J. Jakubowska, A. Lubinski, J. Jaworska, K. Durda, K. Mannermaa, A. Kataja, V. Kosma, V.-.M. Hartikainen, J.M. Bogdanova, N.V. Antonenkova, N.N. Doerk, T. Kristensen, V.N. Anton-Culver, H. Slager, S. Toland, A.E. Edge, S. Fostira, F. Kang, D. Yoo, K.-.Y. Noh, D.-.Y. Matsuo, K. Ito, H. Iwata, H. Sueta, A. Wu, A.H. Tseng, C.-.C. Van den Berg, D. Stram, D.O. Shu, X.-.O. Lu, W. Gao, Y.-.T. Cai, H. Teo, S.H. Yip, C.H. Phuah, S.Y. Cornes, B.K. Hartman, M. Miao, H. Lim, W.Y. Sng, J.-.H. Muir, K. Lophatananon, A. Stewart-Brown, S. Siriwanarangsan, P. Shen, C.-.Y. Hsiung, C.-.N. Wu, P.-.E. Ding, S.-.L. Sangrajrang, S. Gaborieau, V. Brennan, P. McKay, J. Blot, W.J. Signorello, L.B. Cai, Q. Zheng, W. Deming-Halverson, S. Shrubsole, M. Long, J. Simard, J. Garcia-Closas, M. Pharoah, P.D.P. Chenevix-Trench, G. Dunning, A.M. Benitez, J. Easton, D.F. Susceptibility, B.O.C. Res, H.B.O.C. Investigators, K. Grp, A.O.C.S. GENICA, G.E.I.B (2013) Large-scale genotyping identifies 41 new loci associated with breast cancer risk. full text
Lambrechts, D. Truong, T. Justenhoven, C. Humphreys, M.K. Wang, J. Hopper, J.L. Dite, G.S. Apicella, C. Southey, M.C. Schmidt, M.K. Broeks, A. Cornelissen, S. van Hien, R. Sawyer, E. Tomlinson, I. Kerin, M. Miller, N. Milne, R.L. Pilar Zamora, M. Arias Perez, J.I. Benitez, J. Hamann, U. Ko, Y.-.D. Bruening, T. Chang-Claude, J. Eilber, U. Hein, R. Nickels, S. Flesch-Janys, D. Wang-Gohrke, S. John, E.M. Miron, A. Winqvist, R. Pylkas, K. Jukkola-Vuorinen, A. Grip, M. Chenevix-Trench, G. Beesley, J. Chen, X. Menegaux, F. Cordina-Duverger, E. Shen, C.-.Y. Yu, J.-.C. Wu, P.-.E. Hou, M.-.F. Andrulis, I.L. Selander, T. Glendon, G. Mulligan, A.M. Anton-Culver, H. Ziogas, A. Muir, K.R. Lophatananon, A. Rattanamongkongul, S. Puttawibul, P. Jones, M. Orr, N. Ashworth, A. Swerdlow, A. Severi, G. Baglietto, L. Giles, G. Southey, M. Marme, F. Schneeweiss, A. Sohn, C. Burwinkel, B. Yesilyurt, B.T. Neven, P. Paridaens, R. Wildiers, H. Brenner, H. Mueller, H. Arndt, V. Stegmaier, C. Meindl, A. Schott, S. Bartram, C.R. Schmutzler, R.K. Cox, A. Brock, I.W. Elliott, G. Cross, S.S. Fasching, P.A. Schulz-Wendtland, R. Ekici, A.B. Beckmann, M.W. Fletcher, O. Johnson, N. Silva, I.D.S. Peto, J. Nevanlinna, H. Muranen, T.A. Aittomaki, K. Blomqvist, C. Doerk, T. Schuermann, P. Bremer, M. Hillemanns, P. Bogdanova, N.V. Antonenkova, N.N. Rogov, Y.I. Karstens, J.H. Khusnutdinova, E. Bermisheva, M. Prokofieva, D. Gancev, S. Jakubowska, A. Lubinski, J. Jaworska, K. Durda, K. Nordestgaard, B.G. Bojesen, S.E. Lanng, C. Mannermaa, A. Kataja, V. Kosma, V.-.M. Hartikainen, J.M. Radice, P. Peterlongo, P. Manoukian, S. Bernard, L. Couch, F.J. Olson, J.E. Wang, X. Fredericksen, Z. Alnaes, G.G. Kristensen, V. Borresen-Dale, A.-.L. Devilee, P. Tollenaar, R.A.E.M. Seynaeve, C.M. Hooning, M.J. Garcia-Closas, M. Chanock, S.J. Lissowska, J. Sherman, M.E. Hall, P. Liu, J. Czene, K. Kang, D. Yoo, K.-.Y. Noh, D.-.Y. Lindblom, A. Margolin, S. Dunning, A.M. Pharoah, P.D.P. Easton, D.F. Guenel, P. Brauch, H. Network, G.E.N.I.C.A. Investigators, K. Grp, A.O.C.S (2012) 11q13 is a susceptibility locus for hormone receptor positive breast cancer. full text
Ponsonby, A.L. Dwyer, T. Cochrane, J.A. Gibbons, L.E. Jones, M.E (1992) Characteristics of the infant thermal environment in the control population of a case-control study of SIDS. Show Abstract full text

This report examines the thermal environment during last sleep of a control population to investigate how the thermal environment of the infant's bedroom varies by season, external temperature and by certain maternal and infant characteristics. Two age-matched control infants were chosen for each case, one of which was also matched on birthweight. The home visits were not pre-arranged and were matched on climatic conditions, time of year and time period of day for the index case. The initial response rate for controls (n = 108) was 86%. Although there was a large amount of variation in the infant thermal environment, thermal insulation correlated with room temperature (r = -0.44, P = 0.0001) and external temperature (r = -0.30, P = 0.002). The thermal environment of the infant, as defined by excess thermal insulation for room temperature, did not vary by indoor or outdoor temperature, but higher average values were observed in teenage mothers (mean difference = 2.7 tog [95% Cl = 0.3, 5.2]), infants who slept in an adult bed (mean difference = 2.6 tog [-0.1, 5.4]) and infants with an illness (mean difference = 0.8 tog [-0.3, 1.9]). There was a tendency for the thermal environment of infants to be higher and more variable during winter, supporting previous hypotheses that paradoxical overheating may occur in some infants during winter. Further work is required to provide a set of recommendations on the optimal thermal conditions for post-neonatal infants.

Ponsonby, A.L. Dwyer, T. Jones, M.E (1992) Sudden infant death syndrome: seasonality and a biphasic model of pathogenesis. Show Abstract full text

STUDY OBJECTIVE: This paper examines the relationship between season, age, and the sudden infant death syndrome (SIDS). It provides a theoretical model for the pathogenesis of SIDS and uses it as a framework to consider risk factor mechanism. DESIGN: A case series analysis was used to examine season and age in relation to SIDS and seasonal pattern and age at death distribution of perinatal risk factors. SETTING: The source population for the SIDS cases in this study was all live births in the state of Tasmania, Australia, 1975 to 1987 inclusive. SUBJECTS: Cases were all infants born 1975 to 1987 who died of SIDS on whom birth notification information was available (n = 348). The live birth cohort 1980-87 (n = 55,944) was used as the control population for risk factor identification. MEASUREMENTS AND MAIN RESULTS: The median ages of death for spring, summer, autumn, and winter born infants were 115, 103.5, 91 and 78 days. Spring and summer born infants died at a significantly older median age than winter born infants. The month of birth distribution of SIDS cases did not alter significantly from a uniform, nonseasonal distribution (p greater than 0.25) but month of death was seasonally distributed (p less than 0.01). Premature and low birthweight infants died at an older median age (p less than 0.05) than term and non-low-birthweight infants. An excess of male infant deaths and infant deaths to older mothers occurred during winter (p less than 0.05). CONCLUSIONS: The pathogenesis of SIDS can be represented as a biphasic model with three pathways of risk factor operation. In this study, season influenced the age at death of SIDS infants. We propose that risk factors with a strong seasonal distribution are likely to be operating in the postnatal period.

Jones, M.E. Ponsonby, A.L. Dwyer, T. Gilbert, N (1994) The relation between climatic temperature and sudden infant death syndrome differs among communities: results from an ecologic analysis. Show Abstract full text

We examined the negative relation between temperature and the sudden infant death syndrome (SIDS) in 22 communities in seven countries. We estimated the percentage increase in SIDS rate for a 1 degree C drop in climatic temperature. The relation differed substantially among communities. In New Zealand and Australia (10 communities), the association was consistently strong; in Europe (seven communities), it varied from strong to weak; and in the USA (five communities), it was moderate or weak. We postulate that low climatic temperature indirectly increases the incidence of SIDS, particularly in countries where outdoor climatic temperature modifies the indoor temperature and clothing habits.

Hosseini, M. Carpenter, R.G. Mohammad, K. Jones, M.E (1999) Standardized percentile curves of body mass index of Iranian children compared to the US population reference. Show Abstract full text

OBJECTIVE: To present standardized percentile curves of body mass index (BMI) for Iranian children, and compare these to the US population reference. SUBJECTS: 1599 boys and 1702 girls aged 2-18 y living in urban Tehran as a part of a random cluster sample survey of 1 in 1000 families throughout Iran. MEASUREMENTS: Heights (cm) and weights (kg) were collected by trained health staff. RESULTS: Standardized BMI reference curves for Iranian boys and girls were constructed. The curves are shown to fit the data well. The development pattern of BMI for boys and girls are compared. CONCLUSIONS: The major differences observed between Iranian and the US BMI charts underline the need for population-specific reference data. For children over six years the 5th and 95th percentiles of our data may be used provisionally as cut-off points for defining thinness and obesity for Iranian children and adolescents.

Jones, M.E. Swerdlow, A.J. Griffith, M. Goldacre, M.J (1998) Prenatal risk factors for cryptorchidism: a record linkage study.. Show Abstract full text

Using data from the Oxford Record Linkage Study (ORLS), we conducted a case-control study to examine prenatal risk factors for cryptorchidism. We identified 1449 boys born during 1970-86 for whom there was a record of an orchidopexy during 1970-87. Up to eight controls were matched to each case on year of birth and hospital or place of delivery. For each boy and his mother we extracted abstracts of maternity and general hospital records from the ORLS. Low birthweight (trend P < 0.001), low social class (trend P < 0.001), breech presentation (relative risk 1.67; 95% confidence interval [CI] 1.16, 2.41), pre-eclampsia (1.17 [1.00, 1.37]), artificial feeding (1.22 [1.04, 1.45]) and episiotomy (1.13 [1.00, 1.27]) were identified as independent risk factors for cryptorchidism. Gestational age was not independently associated with cryptorchidism after adjusting for birthweight (P = 0.33), and this observation suggested that some cryptorchid boys may have suffered from intrauterine growth retardation. Low birthweight, breech presentation and pre-eclampsia may have in common poor placental function and impaired fetal growth, which may be causes of cryptorchidism.

Jones, M.E. Swerdlow, A.J. Griffith, M. Goldacre, M.J (1998) Risk of congenital inguinal hernia in siblings: a record linkage study.. Show Abstract full text

Using data from the Oxford Record Linkage Study (ORLS), we conducted a case-control study to estimate the sex-specific risks of inguinal hernia in siblings of children with this condition. There were 1921 male and 347 female cases born during 1970-86 who were operated on for inguinal hernia at ages 0-5 years during 1970-87, with 12,886 male and 2534 female control subjects. The relative risk of inguinal hernia was 5.8 [95% confidence interval 4.0-8.4] for brothers of male cases and 4.3 [2.1-8.7] for brothers of female cases (both relative to brothers of control subjects). The relative risk was 3.7 [1.8-7.9] for sisters of male cases and 17.8 [6.9-46.3] for sisters of female cases (both relative to sisters of control subjects). The pattern of sex-dependent risks, particularly the large risk for sisters of female cases, suggests a multifactorial threshold model for the disease. Girls have much lower rates of inguinal hernia than boys, and if these rates are low because of a low susceptibility to disease due to the absence of a sex-related risk factor, then those girls who develop disease might have a potentially large contribution to susceptibility from genetic or intrauterine risk factors unrelated to sex.

Jones, M.E. Swerdlow, A.J. Gill, L.E. Goldacre, M.J (1998) Pre-natal and early life risk factors for childhood onset diabetes mellitus: a record linkage study.. Show Abstract full text

BACKGROUND: Using data from the Oxford Record Linkage Study (ORLS) we conducted a case-control study to examine pre-natal and early life risk factors for childhood and adolescent onset diabetes mellitus. METHODS: We identified 160 boys and 155 girls born 1965-1986 and admitted to hospital with a diagnosis of diabetes during 1965-1987 in the ORLS area. Up to eight controls were matched to each case on sex, year of birth and hospital or place of birth. We linked the hospital records for each child to all of that child's hospital records and to his or her mother's maternity record. RESULTS: There were no significant associations between subsequent diabetes and birthweight, gestational age, birthweight for gestational age, maternal age and parity. There were increased risks with not breastfeeding (relative risk [RR] = 1.33; 95% CI: 0.76-2.34), and with diabetes recorded in the mother during pregnancy (RR = 5.87; 95% CI : 0.90-38.3), but these were not statistically significant. There was a significantly raised risk with pre-eclampsia or eclampsia during pregnancy (RR = 1.48; 95% CI: 1.05-2.10). CONCLUSIONS: Pre-eclampsia may be the result of an immunogenetic incompatibility between mother and fetus, and this early immunological disturbance might be related to incidence of diabetes in later life.

Swerdlow, A.J. Jones, M.E (1996) Mortality during 25 years of follow-up of a cohort with diabetes.. Show Abstract full text

BACKGROUND: Diabetes is one of the most common chronic diseases in Western populations. There have been few large published cohort studies of people with diabetes that have had more than 10 years of follow-up, and none other than the present one are in the UK. Such studies are important to understand the long-term fatal consequences of diabetes and their variation over time and between countries. METHODS: Cause-specific mortality was analysed in follow-up from 1966-1970 to December 1992 of 5783 members of the British Diabetic Association living in England and Wales during 1966-1970. Comparison was made with age-, sex- and calendar year-specific mortality by cause in the general population of England and Wales. RESULTS: During the follow-up 3399 (58.8%) subjects died. The relative risk of all-cause mortality in the cohort compared to the general population was 2.31 in women and 1.58 in men (both P < 0.001).Relative risks were greater for women than men at almost all ages and for each major diabetes-related cause of death. Absolute excess ('attributable') mortality rates were also greater in women than in men, except at ages < 50. Half the deaths in each sex were from circulatory diseases and only 3.4% were from renal disease. The relative risks of mortality for all-causes and circulatory diseases were particularly great at younger ages, but changed little with duration of follow-up. At ages < 40 the relative risks for all-causes were 3.75 in men and 5.51 in women and for ischaemic heart disease were 10.44 and 25.25 respectively (all P < 0.001). At these ages one-third of deaths were due to acute complications of diabetes, suicides and accidents, whereas at older ages these accounted for only 4% of deaths. CONCLUSIONS: The mortality rates at young ages in the cohort were around twice those in Sweden, Norway and Israel, suggesting that many of the deaths in England and Wales are preventable. The results also indicate a particular need for investigation and amelioration of cardiovascular risk factors in English and Welsh patients, especially women.

Jones, M.E. Swerdlow, A.J (1996) Bias caused by migration in case-control studies of prenatal risk factors for childhood and adult diseases.. Show Abstract full text

Case-control studies of prenatal risk factors for disease in later life often ascertain cases from within a defined area, trace the birth records of those cases born within the area, and select controls from birth records within the same area. Bias can occur in these studies if the disease risk factors are related to migration from the area. The effects of this bias were examined in a study in Oxfordshire, England. Cases (n = 218) of diabetes in children and young adults born during 1965-1986 were identified from hospital discharges during 1965-1986; controls (n = 753) were selected from livebirths during 1965-1986. By 1987, 219 controls (29.1%) had migrated from Oxfordshire or died. Low maternal parity and high social class were strongly related to migration, more than the other perinatal factors studied. Migration, therefore, could lead to apparent associations of diabetes risk with parity or social class. For a general instance, the authors show how much bias is caused by different degrees of migration and of association between migration and a perinatal risk factor. Examples are given of how migration can produce apparent trends in risk as well as increased or decreased individual relative risks. If more than 25% of controls migrate, bias may be appreciable.

Jones, M.E. Schoemaker, M.J. Rae, M. Folkerd, E.J. Dowsett, M. Ashworth, A. Swerdlow, A.J (2014) Reproducibility of estradiol and testosterone levels in postmenopausal women over 5 years: results from the breakthrough generations study.. Show Abstract full text

Prospective cohort studies examining sex hormones in relation to cancer risk have generally collected blood samples at 1 time point, with an assumption that hormone levels measured in these samples will be reliable markers of true levels at other times. In postmenopausal women, body fat is a major source of estradiol; therefore, changes in adiposity may affect the correlation of single measurements to more relevant long-term averages. To estimate the intraclass correlation coefficient (ICC) for estradiol and testosterone, we collected repeat blood samples from 119 postmenopausal women (average age = 59.4 (standard deviation, 4.7) years) from the United Kingdom during 2004-2005 and again during 2010-2011. The ICCs (adjusted for assay variation) were 0.73 (95% confidence interval: 0.63, 0.82) for total estradiol and 0.59 (95% confidence interval: 0.47, 0.72) for total testosterone. The ICCs were 3%-5% larger after adjustment for change in body mass index (weight (kg)/height (m)(2)) or leptin, which are 2 markers of change in adiposity. There was no increase in ICCs after adjustment for change in age, alcohol consumption, smoking, exercise, time between waking and blood collection, or season. The results suggest that other factors account for within-woman variation in these sex hormones.

Jones, M.E. Schoemaker, M. Rae, M. Folkerd, E.J. Dowsett, M. Ashworth, A. Swerdlow, A.J (2013) Changes in estradiol and testosterone levels in postmenopausal women after changes in body mass index.. Show Abstract full text

CONTEXT: Endogenous sex hormones are risk factors for postmenopausal breast cancer. A potential route for favorable hormonal modification is weight loss. OBJECTIVE: The objective of the study was to measure change in plasma estradiol and testosterone levels in postmenopausal women in relation to change in body mass index (BMI) and plasma leptin. SETTING: The setting was a cohort study of over 100,000 female volunteers from the general population, United Kingdom. PARTICIPANTS: The participants were a sample of 177 postmenopausal women aged over 45 years who provided blood samples during 2004-2005 and again during 2010-2011. MAIN OUTCOME MEASURE: Outcomes were percentage change in plasma estradiol and testosterone levels per 1 kg/m² change in BMI and per 1 ng/mL change in plasma leptin. RESULTS: Among women with reduction in BMI, estradiol decreased 12.7% (95% confidence interval: [6.4%, 19.5%]; P < .0001) per kg/m² and among women with increased BMI estradiol increased 6.4% [0.2%, 12.9%] (P = .042). The corresponding figures for testosterone were 10.7% [3.0%, 19.0%] (P = .006) and 1.9% [-5.4%, 9.7%] (P = .61) per kg/m². For women with decreases and increases in leptin, estradiol decreased by 3.6% [1.3%, 6.0%] (P = .003) per ng/mL and increased by 1.7% [-0.3%, 3.6%] (P = .094), respectively. The corresponding figures for testosterone were 4.8% [2.0%, 7.8%] (P = .009) and 0.3% [-2.0%, 2.6%] (P = .82) per ng/mL. CONCLUSIONS: In postmenopausal women, changes in BMI and plasma leptin occurring over several years are associated with changes in estradiol and testosterone levels. The results suggest that fat loss by an individual can result in substantial decreases in postmenopausal estradiol and testosterone levels and provides support for weight management to lessen breast cancer risk.

Jones, M.E. van Leeuwen, F.E. Hoogendoorn, W.E. Mourits, M.J. Hollema, H. van Boven, H. Press, M.F. Bernstein, L. Swerdlow, A.J (2012) Endometrial cancer survival after breast cancer in relation to tamoxifen treatment: pooled results from three countries.. Show Abstract full text

INTRODUCTION: Tamoxifen is an effective treatment for breast cancer but an undesirable side-effect is an increased risk of endometrial cancer, particularly rare tumor types associated with poor prognosis. We investigated whether tamoxifen therapy increases mortality among breast cancer patients subsequently diagnosed with endometrial cancer. METHODS: We pooled case-patient data from the three largest case-control studies of tamoxifen in relation to endometrial cancer after breast cancer (1,875 patients: Netherlands, 765; United Kingdom, 786; United States, 324) and collected follow-up information on vital status. Breast cancers were diagnosed in 1972 to 2005 with endometrial cancers diagnosed in 1978 to 2006. We used Cox proportional hazards survival analysis to estimate hazard ratios (HRs) and 95% confidence intervals (CI). RESULTS: A total of 1,104 deaths occurred during, on average, 5.8 years following endometrial cancer (32% attributed to breast cancer, 25% to endometrial cancer). Mortality from endometrial cancer increased significantly with unfavorable non-endometrioid morphologies (P < 0.0001), International Federation of Gynaecology and Obstetrics staging system for gynecological malignancy (FIGO) stage (P < 0.0001) and age (P < 0.0001). No overall association was observed between tamoxifen treatment and endometrial cancer mortality (HR = 1.17 (95% CI: (0.89 to 1.55)). Tamoxifen use for at least five years was associated with increased endometrial cancer mortality (HR = 1.59 (1.13 to 2.25)). This association appeared to be due primarily to the excess of unfavorable histologies and advanced stage in women using tamoxifen for five or more years since the association with mortality was no longer significant after adjustment for morphological type and FIGO stage (HR = 1.37 (0.97 to 1.93)). Those patients with endometrioid tumors, who stopped tamoxifen use at least five years before their endometrial cancer diagnosis, had a greater mortality risk from endometrial cancer than endometrioid patients with no tamoxifen exposure (HR = 2.11 (1.13 to 3.94)). The explanation for this latter observation is not apparent. CONCLUSIONS: Patients with endometrial cancer after breast cancer who received tamoxifen treatment for five years for breast cancer have greater endometrial cancer mortality risk than those who did not receive tamoxifen. This can be attributed to non-endometrioid histological subtypes with poorer prognosis among long term tamoxifen users.

Seah, M.P. Jones, M.E (1984) Roughness Contributions to Resolution in Ion Sputter Depth Profiles of Polycrystalline Metal-Films.
Herlihy, E. Gies, P.H. Roy, C.R. Jones, M (1994) Personal dosimetry of solar UV radiation for different outdoor activities. Show Abstract full text

Quantifying individual exposure to ultraviolet radiation (UVR) is critical to understanding the etiology of a number of diseases including nonmelanotic and melanotic skin cancers. Measurements of personal exposure to solar UVR were made in Hobart, Tasmania in February (summer) 1991 for six different outdoor activities using UVR-sensitive polysulfone (PS) film attached at seven anatomical sites. Concurrent behavioral and environmental observations were also made. To date many studies have relied on subject recall to quantify past solar UVR exposures. To gain insight into the accuracy of subject recall the measured UVR exposures received by different subjects using the PS film were compared to those calculated from personal diaries and ambient solar UVB levels from a monitoring station. In general, when UVR exposure activities took place under close supervision, good correlations were obtained between the PS badges and the ambient measurements/diaries approach. Ultraviolet radiation exposures for the field study involving 94 subjects engaged in a number of outdoor activities are presented.

Bruce, J.C. Bond, S.T. Jones, M.E (2002) Teaching epidemiology and statistics by distance learning. full text
Sullivan, S.A. Marsden, K.A. Lowenthal, R.M. Jupe, D.M. Jones, M.E (1992) Circulating CD34+ cells: an adverse prognostic factor in the myelodysplastic syndromes. Show Abstract full text

As part of an epidemiological survey of myelodysplastic syndromes (MDS) in southern Tasmania, 62 MDS patients identified over a 2 year period were tested for the presence of CD34, the human progenitor cell antigen (HPCA), in their peripheral blood. The results were correlated with transformation to acute myeloid leukemia (AML) and patient survival, and CD34+ status was compared as a prognostic indicator with Bournemouth score, cytogenetics, and CFU-GM colony growth which were also assessed. Circulating CD34+ cells were found in 23 of the 62 MDS patients; 9 of the 23 patients with circulating CD34+ cells transformed to AML, as compared with none of the 39 CD34 negative patients (P less than 0.0001); and 11 of the 23 patients with circulating CD34+ cells were dead at the end of the 2 year period, as opposed to 6 of the 39 with no CD34+ cells (P less than 0.03). The Bournemouth score was also significantly associated with transformation to AML (P less than 0.0001) and poor survival (P less than 0.04). These were the only significant associations of the possible prognostic factors studied with either transformation or survival. In summary, the presence of circulating CD34+ cells was significantly associated with both progression to AML and poor survival and was found to be a better prognostic indicator than cytogenetics or CFU-GM colony growth.

Jones, M.E. Shugg, D. Dwyer, T. Young, B. Bonett, A (1992) Interstate differences in incidence and mortality from melanoma. A re-examination of the latitudinal gradient. Show Abstract full text

OBJECTIVE: To investigate the patterns of cutaneous malignant melanoma (CMM) mortality in Australia. DESIGN: A descriptive analysis of melanoma incidence and mortality in Australia supplemented by a case series analysis of melanoma survival. Melanoma mortality rates were based on tabulations supplied by the Australian Bureau of Statistics for the years 1969-1989. Melanoma incidence rates were based on State cancer registry records for the years 1977-1990. The case series survival analysis was based on detailed individual records from the population-based cancer registries in Tasmania and South Australia. MAIN OUTCOME MEASURES: The level of and rise in melanoma mortality rates during 1969-1989 in Australia; the five-year survival rates for Tasmanian and South Australian cases; and male:female incidence ratios related to latitude. RESULTS: We found annual increases in melanoma mortality rates of 2.5% in men (P < 0.0001) and 1.1% in women (P < 0.0001) for all Australia. The five-year survival rates (with 95% confidence intervals [CI]) were: 67% (59%-75%) for Tasmanian men; 79% (76%-83%) for South Australian men; 80% (74%-86%) for Tasmanian women and 88% (86%-91%) for South Australian women. A change in the male:female incidence ratio with latitude was also found--women have significantly higher incidence rates at higher latitudes, but similar rates to men at lower latitudes. CONCLUSIONS: The age standardised mortality from CMM for the period 1969 to 1989 shows little variation by State for women, despite a considerable range in latitude. CMM mortality in men is increasing at a faster rate than that in women. Between 1982 and 1987 the male:female incidence ratio in high latitudes in the Southern Hemisphere showed an excess of cases in women, a finding which we believe has not been reported before.

Saxena, S. Majeed, A. Jones, M (1999) Socioeconomic differences in childhood consultation rates in general practice in England and Wales: prospective cohort study. Show Abstract full text

OBJECTIVE: To establish how consultation rates in children for episodes of illness, preventive activities, and home visits vary by social class. DESIGN: Analysis of prospectively collected data from the fourth national survey of morbidity in general practice, carried out between September 1991 and August 1992. SETTING: 60 general practices in England and Wales. SUBJECTS: 106 102 children aged 0 to 15 years registered with the participating practices. MAIN OUTCOME MEASURES: Mean overall consultation rates for any reason, illness by severity of underlying disease, preventive episodes, home visits, and specific diagnostic category (infections, asthma, and injuries). RESULTS: Overall consultation rates increased from registrar general's social classes I-II to classes IV-V in a linear pattern (for IV-V v I-II rate ratio 1.18; 95% confidence interval 1.14 to 1. 22). Children from social classes IV-V consulted more frequently than children from classes I-II for illnesses (rate ratio 1.23; 1.15 to 1.30), including infections, asthma, and injuries and poisonings. They also had significantly higher consultation rates for minor, moderate, and serious illnesses and higher home visiting rates (rate ratio 2.00; 1.81 to 2.18). Consultations for preventive activities were lower in children from social classes IV-V than in children from social classes I-II (rate ratio 0.95; 0.86 to 1.05). CONCLUSIONS: Childhood consultation rates for episodes of illness increase from social classes I-II through to classes IV-V. The findings on severity of underlying illness suggest the health of children from lower social classes is worse than that of children from higher social classes. These results reinforce the need to identify and target children for preventive health care in their socioeconomic context.

Swerdlow, A.J. Jones, M.E (2007) Ovarian cancer risk in premenopausal and perimenopausal women treated with Tamoxifen: a case-control study.. Show Abstract full text

As tamoxifen stimulates ovarian steroidogenesis in premenopausal women, induces ovulation and increases the incidence of benign ovarian cysts, there has been concern that it might also increase ovarian cancer risk in women treated premenopausally. In a national case-control study in Britain, treatment histories were collected for 158 cases of ovarian cancer after breast cancer diagnosed at ages under 55 years and 464 controls who had breast cancer at these ages without subsequent ovarian cancer. Risk of ovarian cancer was not raised for women overall who had taken tamoxifen (odds ratio (OR)=0.9, 95% confidence interval (CI) 0.6-1.3) or for those treated when premenopausal (OR=1.0, 95% CI 0.6-1.6) or perimenopausal (OR=0.7, 95% CI 0.2-2.4). There was also no relation of risk to daily dose, duration or cumulative dose of tamoxifen, or time since last use. There was, however, a significantly raised risk in relation to non-hormonal chemotherapy. The results suggest that tamoxifen treatment of premenopausal or perimenopausal women does not materially affect ovarian cancer risk, but that non-hormonal chemotherapy might increase risk.

Park, J. Choi, J.-.Y. Choi, J. Chung, S. Song, N. Park, S.K. Han, W. Noh, D.-.Y. Ahn, S.-.H. Lee, J.W. Kim, M.K. Jee, S.H. Wen, W. Bolla, M.K. Wang, Q. Dennis, J. Michailidou, K. Shah, M. Conroy, D.M. Harrington, P.A. Mayes, R. Czene, K. Hall, P. Teras, L.R. Patel, A.V. Couch, F.J. Olson, J.E. Sawyer, E.J. Roylance, R. Bojesen, S.E. Flyger, H. Lambrechts, D. Baten, A. Matsuo, K. Ito, H. Guénel, P. Truong, T. Keeman, R. Schmidt, M.K. Wu, A.H. Tseng, C.-.C. Cox, A. Cross, S.S. kConFab Investigators, . Andrulis, I.L. Hopper, J.L. Southey, M.C. Wu, P.-.E. Shen, C.-.Y. Fasching, P.A. Ekici, A.B. Muir, K. Lophatananon, A. Brenner, H. Arndt, V. Jones, M.E. Swerdlow, A.J. Hoppe, R. Ko, Y.-.D. Hartman, M. Li, J. Mannermaa, A. Hartikainen, J.M. Benitez, J. González-Neira, A. Haiman, C.A. Dörk, T. Bogdanova, N.V. Teo, S.H. Mohd Taib, N.A. Fletcher, O. Johnson, N. Grip, M. Winqvist, R. Blomqvist, C. Nevanlinna, H. Lindblom, A. Wendt, C. Kristensen, V.N. Nbcs Collaborators, . Tollenaar, R.A.E.M. Heemskerk-Gerritsen, B.A.M. Radice, P. Bonanni, B. Hamann, U. Manoochehri, M. Lacey, J.V. Martinez, M.E. Dunning, A.M. Pharoah, P.D.P. Easton, D.F. Yoo, K.-.Y. Kang, D (2021) Gene-Environment Interactions Relevant to Estrogen and Risk of Breast Cancer: Can Gene-Environment Interactions Be Detected Only among Candidate SNPs from Genome-Wide Association Studies?. Show Abstract full text

In this study we aim to examine gene-environment interactions (GxEs) between genes involved with estrogen metabolism and environmental factors related to estrogen exposure. GxE analyses were conducted with 1970 Korean breast cancer cases and 2052 controls in the case-control study, the Seoul Breast Cancer Study (SEBCS). A total of 11,555 SNPs from the 137 candidate genes were included in the GxE analyses with eight established environmental factors. A replication test was conducted by using an independent population from the Breast Cancer Association Consortium (BCAC), with 62,485 Europeans and 9047 Asians. The GxE tests were performed by using two-step methods in GxEScan software. Two interactions were found in the SEBCS. The first interaction was shown between rs13035764 of NCOA1 and age at menarche in the GE|2df model (<i>p</i>-2df = 1.2 × 10<sup>-3</sup>). The age at menarche before 14 years old was associated with the high risk of breast cancer, and the risk was higher when subjects had homozygous minor allele G. The second GxE was shown between rs851998 near ESR1 and height in the GE|2df model (<i>p</i>-2df = 1.1 × 10<sup>-4</sup>). Height taller than 160 cm was associated with a high risk of breast cancer, and the risk increased when the minor allele was added. The findings were not replicated in the BCAC. These results would suggest specificity in Koreans for breast cancer risk.

Figlioli, G. Bogliolo, M. Catucci, I. Caleca, L. Lasheras, S.V. Pujol, R. Kiiski, J.I. Muranen, T.A. Barnes, D.R. Dennis, J. Michailidou, K. Bolla, M.K. Leslie, G. Aalfs, C.M. ABCTB Investigators, . Adank, M.A. Adlard, J. Agata, S. Cadoo, K. Agnarsson, B.A. Ahearn, T. Aittomäki, K. Ambrosone, C.B. Andrews, L. Anton-Culver, H. Antonenkova, N.N. Arndt, V. Arnold, N. Aronson, K.J. Arun, B.K. Asseryanis, E. Auber, B. Auvinen, P. Azzollini, J. Balmaña, J. Barkardottir, R.B. Barrowdale, D. Barwell, J. Beane Freeman, L.E. Beauparlant, C.J. Beckmann, M.W. Behrens, S. Benitez, J. Berger, R. Bermisheva, M. Blanco, A.M. Blomqvist, C. Bogdanova, N.V. Bojesen, A. Bojesen, S.E. Bonanni, B. Borg, A. Brady, A.F. Brauch, H. Brenner, H. Brüning, T. Burwinkel, B. Buys, S.S. Caldés, T. Caliebe, A. Caligo, M.A. Campa, D. Campbell, I.G. Canzian, F. Castelao, J.E. Chang-Claude, J. Chanock, S.J. Claes, K.B.M. Clarke, C.L. Collavoli, A. Conner, T.A. Cox, D.G. Cybulski, C. Czene, K. Daly, M.B. de la Hoya, M. Devilee, P. Diez, O. Ding, Y.C. Dite, G.S. Ditsch, N. Domchek, S.M. Dorfling, C.M. Dos-Santos-Silva, I. Durda, K. Dwek, M. Eccles, D.M. Ekici, A.B. Eliassen, A.H. Ellberg, C. Eriksson, M. Evans, D.G. Fasching, P.A. Figueroa, J. Flyger, H. Foulkes, W.D. Friebel, T.M. Friedman, E. Gabrielson, M. Gaddam, P. Gago-Dominguez, M. Gao, C. Gapstur, S.M. Garber, J. García-Closas, M. García-Sáenz, J.A. Gaudet, M.M. Gayther, S.A. GEMO Study Collaborators, . Giles, G.G. Glendon, G. Godwin, A.K. Goldberg, M.S. Goldgar, D.E. Guénel, P. Gutierrez-Barrera, A.M. Haeberle, L. Haiman, C.A. Håkansson, N. Hall, P. Hamann, U. Harrington, P.A. Hein, A. Heyworth, J. Hillemanns, P. Hollestelle, A. Hopper, J.L. Hosgood, H.D. Howell, A. Hu, C. Hulick, P.J. Hunter, D.J. Imyanitov, E.N. KConFab, . Isaacs, C. Jakimovska, M. Jakubowska, A. James, P. Janavicius, R. Janni, W. John, E.M. Jones, M.E. Jung, A. Kaaks, R. Karlan, B.Y. Khusnutdinova, E. Kitahara, C.M. Konstantopoulou, I. Koutros, S. Kraft, P. Lambrechts, D. Lazaro, C. Le Marchand, L. Lester, J. Lesueur, F. Lilyquist, J. Loud, J.T. Lu, K.H. Luben, R.N. Lubinski, J. Mannermaa, A. Manoochehri, M. Manoukian, S. Margolin, S. Martens, J.W.M. Maurer, T. Mavroudis, D. Mebirouk, N. Meindl, A. Menon, U. Miller, A. Montagna, M. Nathanson, K.L. Neuhausen, S.L. Newman, W.G. Nguyen-Dumont, T. Nielsen, F.C. Nielsen, S. Nikitina-Zake, L. Offit, K. Olah, E. Olopade, O.I. Olshan, A.F. Olson, J.E. Olsson, H. Osorio, A. Ottini, L. Peissel, B. Peixoto, A. Peto, J. Plaseska-Karanfilska, D. Pocza, T. Presneau, N. Pujana, M.A. Punie, K. Rack, B. Rantala, J. Rashid, M.U. Rau-Murthy, R. Rennert, G. Lejbkowicz, F. Rhenius, V. Romero, A. Rookus, M.A. Ross, E.A. Rossing, M. Rudaitis, V. Ruebner, M. Saloustros, E. Sanden, K. Santamariña, M. Scheuner, M.T. Schmutzler, R.K. Schneider, M. Scott, C. Senter, L. Shah, M. Sharma, P. Shu, X.-.O. Simard, J. Singer, C.F. Sohn, C. Soucy, P. Southey, M.C. Spinelli, J.J. Steele, L. Stoppa-Lyonnet, D. Tapper, W.J. Teixeira, M.R. Terry, M.B. Thomassen, M. Thompson, J. Thull, D.L. Tischkowitz, M. Tollenaar, R.A.E.M. Torres, D. Troester, M.A. Truong, T. Tung, N. Untch, M. Vachon, C.M. van Rensburg, E.J. van Veen, E.M. Vega, A. Viel, A. Wappenschmidt, B. Weitzel, J.N. Wendt, C. Wieme, G. Wolk, A. Yang, X.R. Zheng, W. Ziogas, A. Zorn, K.K. Dunning, A.M. Lush, M. Wang, Q. McGuffog, L. Parsons, M.T. Pharoah, P.D.P. Fostira, F. Toland, A.E. Andrulis, I.L. Ramus, S.J. Swerdlow, A.J. Greene, M.H. Chung, W.K. Milne, R.L. Chenevix-Trench, G. Dörk, T. Schmidt, M.K. Easton, D.F. Radice, P. Hahnen, E. Antoniou, A.C. Couch, F.J. Nevanlinna, H. Surrallés, J. Peterlongo, P (2019) The &lt;i&gt;FANCM&lt;/i&gt;:p.Arg658* truncating variant is associated with risk of triple-negative breast cancer.. Show Abstract full text

Breast cancer is a common disease partially caused by genetic risk factors. Germline pathogenic variants in DNA repair genes <i>BRCA1</i>, <i>BRCA2</i>, <i>PALB2</i>, <i>ATM</i>, and <i>CHEK2</i> are associated with breast cancer risk. <i>FANCM</i>, which encodes for a DNA translocase, has been proposed as a breast cancer predisposition gene, with greater effects for the ER-negative and triple-negative breast cancer (TNBC) subtypes. We tested the three recurrent protein-truncating variants <i>FANCM</i>:p.Arg658*, p.Gln1701*, and p.Arg1931* for association with breast cancer risk in 67,112 cases, 53,766 controls, and 26,662 carriers of pathogenic variants of <i>BRCA1</i> or <i>BRCA2</i>. These three variants were also studied functionally by measuring survival and chromosome fragility in <i>FANCM</i> <sup><i>-/-</i></sup> patient-derived immortalized fibroblasts treated with diepoxybutane or olaparib. We observed that <i>FANCM</i>:p.Arg658* was associated with increased risk of ER-negative disease and TNBC (OR = 2.44, <i>P</i> = 0.034 and OR = 3.79; <i>P</i> = 0.009, respectively). In a country-restricted analysis, we confirmed the associations detected for <i>FANCM</i>:p.Arg658* and found that also <i>FANCM</i>:p.Arg1931* was associated with ER-negative breast cancer risk (OR = 1.96; <i>P</i> = 0.006). The functional results indicated that all three variants were deleterious affecting cell survival and chromosome stability with <i>FANCM</i>:p.Arg658* causing more severe phenotypes. In conclusion, we confirmed that the two rare <i>FANCM</i> deleterious variants p.Arg658* and p.Arg1931* are risk factors for ER-negative and TNBC subtypes. Overall our data suggest that the effect of truncating variants on breast cancer risk may depend on their position in the gene. Cell sensitivity to olaparib exposure, identifies a possible therapeutic option to treat <i>FANCM</i>-associated tumors.

Hurson, A.N. Pal Choudhury, P. Gao, C. Hüsing, A. Eriksson, M. Shi, M. Jones, M.E. Evans, D.G.R. Milne, R.L. Gaudet, M.M. Vachon, C.M. Chasman, D.I. Easton, D.F. Schmidt, M.K. Kraft, P. Garcia-Closas, M. Chatterjee, N. B-CAST Risk Modelling Group, (2022) Prospective evaluation of a breast-cancer risk model integrating classical risk factors and polygenic risk in 15 cohorts from six countries.. Show Abstract full text

<h4>Background</h4>Rigorous evaluation of the calibration and discrimination of breast-cancer risk-prediction models in prospective cohorts is critical for applications under clinical guidelines. We comprehensively evaluated an integrated model incorporating classical risk factors and a 313-variant polygenic risk score (PRS) to predict breast-cancer risk.<h4>Methods</h4>Fifteen prospective cohorts from six countries with 239 340 women (7646 incident breast-cancer cases) of European ancestry aged 19-75 years were included. Calibration of 5-year risk was assessed by comparing expected and observed proportions of cases overall and within risk categories. Risk stratification for women of European ancestry aged 50-70 years in those countries was evaluated by the proportion of women and future cases crossing clinically relevant risk thresholds.<h4>Results</h4>Among women <50 years old, the median (range) expected-to-observed ratio for the integrated model across 15 cohorts was 0.9 (0.7-1.0) overall and 0.9 (0.7-1.4) at the highest-risk decile; among women ≥50 years old, these were 1.0 (0.7-1.3) and 1.2 (0.7-1.6), respectively. The proportion of women identified above a 3% 5-year risk threshold (used for recommending risk-reducing medications in the USA) ranged from 7.0% in Germany (∼841 000 of 12 million) to 17.7% in the USA (∼5.3 of 30 million). At this threshold, 14.7% of US women were reclassified by adding the PRS to classical risk factors, with identification of 12.2% of additional future cases.<h4>Conclusion</h4>Integrating a 313-variant PRS with classical risk factors can improve the identification of European-ancestry women at elevated risk who could benefit from targeted risk-reducing strategies under current clinical guidelines.

Hosseini, M. Ahmadzadeh, H. Toloui, A. Ahmadzadeh, K. Madani Neishaboori, A. Rafiei Alavi, S.N. Gubari, M.I.M. Jones, M.E. Ataei, F. Yousefifard, M. Ataei, N (2022) The value of interleukin levels in the diagnosis of febrile urinary tract infections in children and adolescents; a systematic review and meta-analysis.. Show Abstract full text

<h4>Introduction</h4>In recent years, researchers have been looking for tools and biomarkers to identify urinary tract infections (UTI) in children. Since there exists no systematic reviews and meta-analyses on the matter, the present study intends to determine the diagnostic value of serum and urinary levels of interleukins (IL) in the diagnosis of febrile UTI in children and adolescents.<h4>Methods</h4>Medline, Embase, Scopus, and Web of Science were searched until the end of 2020, using keywords related to UTI and serum and urinary ILs. Two independent researchers included relevant studies and summarized the data. Analyzed data were reported as standardized mean difference (SMD) with 95% confidence interval (CI).<h4>Results</h4>Data from 23 articles were included in the present study. Analyses showed that IL-6, IL-8, IL 1 beta and IL-1 alpha urinary levels are significantly higher in children with UTI than that of other children. Moreover, serum levels of IL-6 and IL-8 in children with UTI were significantly higher than that of healthy children. However, IL-6 and IL-8 serum levels were not significantly different between children with UTI and non-UTI febrile group. Finally, the area under the curve of urinary IL-6 and IL-8 and serum IL-8 levels in the diagnosis of pediatric UTIs were 0.89 (95% CI: 0.86, 0.92), 0.95 (95% CI: 0.92, 0.96) and 0.80 (95% CI: 0.77, 0.84), respectively.<h4>Conclusion</h4>The findings of the present study showed that the diagnostic utility of ILs 8 and 6 urinary levels is most desirable in the detection of febrile UTIs from other febrile conditions in children and adolescents, in comparison with the diagnostic utility of other ILs' urinary and serum levels in the detection of febrile UTI. However, even after nearly 3 decades of research on these biomarkers, their optimal cut-off points in diagnosing pediatric UTIs are still to be determined in further studies.

Dennis, J. Tyrer, J.P. Walker, L.C. Michailidou, K. Dorling, L. Bolla, M.K. Wang, Q. Ahearn, T.U. Andrulis, I.L. Anton-Culver, H. Antonenkova, N.N. Arndt, V. Aronson, K.J. Freeman, L.E.B. Beckmann, M.W. Behrens, S. Benitez, J. Bermisheva, M. Bogdanova, N.V. Bojesen, S.E. Brenner, H. Castelao, J.E. Chang-Claude, J. Chenevix-Trench, G. Clarke, C.L. NBCS Collaborators, . Collée, J.M. CTS Consortium, . Couch, F.J. Cox, A. Cross, S.S. Czene, K. Devilee, P. Dörk, T. Dossus, L. Eliassen, A.H. Eriksson, M. Evans, D.G. Fasching, P.A. Figueroa, J. Fletcher, O. Flyger, H. Fritschi, L. Gabrielson, M. Gago-Dominguez, M. García-Closas, M. Giles, G.G. González-Neira, A. Guénel, P. Hahnen, E. Haiman, C.A. Hall, P. Hollestelle, A. Hoppe, R. Hopper, J.L. Howell, A. ABCTB Investigators, . kConFab/AOCS Investigators, . Jager, A. Jakubowska, A. John, E.M. Johnson, N. Jones, M.E. Jung, A. Kaaks, R. Keeman, R. Khusnutdinova, E. Kitahara, C.M. Ko, Y.-.D. Kosma, V.-.M. Koutros, S. Kraft, P. Kristensen, V.N. Kubelka-Sabit, K. Kurian, A.W. Lacey, J.V. Lambrechts, D. Larson, N.L. Linet, M. Ogrodniczak, A. Mannermaa, A. Manoukian, S. Margolin, S. Mavroudis, D. Milne, R.L. Muranen, T.A. Murphy, R.A. Nevanlinna, H. Olson, J.E. Olsson, H. Park-Simon, T.-.W. Perou, C.M. Peterlongo, P. Plaseska-Karanfilska, D. Pylkäs, K. Rennert, G. Saloustros, E. Sandler, D.P. Sawyer, E.J. Schmidt, M.K. Schmutzler, R.K. Shibli, R. Smeets, A. Soucy, P. Southey, M.C. Swerdlow, A.J. Tamimi, R.M. Taylor, J.A. Teras, L.R. Terry, M.B. Tomlinson, I. Troester, M.A. Truong, T. Vachon, C.M. Wendt, C. Winqvist, R. Wolk, A. Yang, X.R. Zheng, W. Ziogas, A. Simard, J. Dunning, A.M. Pharoah, P.D.P. Easton, D.F (2022) Rare germline copy number variants (CNVs) and breast cancer risk.. Show Abstract full text

Germline copy number variants (CNVs) are pervasive in the human genome but potential disease associations with rare CNVs have not been comprehensively assessed in large datasets. We analysed rare CNVs in genes and non-coding regions for 86,788 breast cancer cases and 76,122 controls of European ancestry with genome-wide array data. Gene burden tests detected the strongest association for deletions in BRCA1 (P = 3.7E-18). Nine other genes were associated with a p-value < 0.01 including known susceptibility genes CHEK2 (P = 0.0008), ATM (P = 0.002) and BRCA2 (P = 0.008). Outside the known genes we detected associations with p-values < 0.001 for either overall or subtype-specific breast cancer at nine deletion regions and four duplication regions. Three of the deletion regions were in established common susceptibility loci. To the best of our knowledge, this is the first genome-wide analysis of rare CNVs in a large breast cancer case-control dataset. We detected associations with exonic deletions in established breast cancer susceptibility genes. We also detected suggestive associations with non-coding CNVs in known and novel loci with large effects sizes. Larger sample sizes will be required to reach robust levels of statistical significance.

Dareng, E.O. Tyrer, J.P. Barnes, D.R. Jones, M.R. Yang, X. Aben, K.K.H. Adank, M.A. Agata, S. Andrulis, I.L. Anton-Culver, H. Antonenkova, N.N. Aravantinos, G. Arun, B.K. Augustinsson, A. Balmaña, J. Bandera, E.V. Barkardottir, R.B. Barrowdale, D. Beckmann, M.W. Beeghly-Fadiel, A. Benitez, J. Bermisheva, M. Bernardini, M.Q. Bjorge, L. Black, A. Bogdanova, N.V. Bonanni, B. Borg, A. Brenton, J.D. Budzilowska, A. Butzow, R. Buys, S.S. Cai, H. Caligo, M.A. Campbell, I. Cannioto, R. Cassingham, H. Chang-Claude, J. Chanock, S.J. Chen, K. Chiew, Y.-.E. Chung, W.K. Claes, K.B.M. Colonna, S. GEMO Study Collaborators, . GC-HBOC Study Collaborators, . EMBRACE Collaborators, . Cook, L.S. Couch, F.J. Daly, M.B. Dao, F. Davies, E. de la Hoya, M. de Putter, R. Dennis, J. DePersia, A. Devilee, P. Diez, O. Ding, Y.C. Doherty, J.A. Domchek, S.M. Dörk, T. du Bois, A. Dürst, M. Eccles, D.M. Eliassen, H.A. Engel, C. Evans, G.D. Fasching, P.A. Flanagan, J.M. Fortner, R.T. Machackova, E. Friedman, E. Ganz, P.A. Garber, J. Gensini, F. Giles, G.G. Glendon, G. Godwin, A.K. Goodman, M.T. Greene, M.H. Gronwald, J. OPAL Study Group, . AOCS Group, . Hahnen, E. Haiman, C.A. Håkansson, N. Hamann, U. Hansen, T.V.O. Harris, H.R. Hartman, M. Heitz, F. Hildebrandt, M.A.T. Høgdall, E. Høgdall, C.K. Hopper, J.L. Huang, R.-.Y. Huff, C. Hulick, P.J. Huntsman, D.G. Imyanitov, E.N. KConFab Investigators, . HEBON Investigators, . Isaacs, C. Jakubowska, A. James, P.A. Janavicius, R. Jensen, A. Johannsson, O.T. John, E.M. Jones, M.E. Kang, D. Karlan, B.Y. Karnezis, A. Kelemen, L.E. Khusnutdinova, E. Kiemeney, L.A. Kim, B.-.G. Kjaer, S.K. Komenaka, I. Kupryjanczyk, J. Kurian, A.W. Kwong, A. Lambrechts, D. Larson, M.C. Lazaro, C. Le, N.D. Leslie, G. Lester, J. Lesueur, F. Levine, D.A. Li, L. Li, J. Loud, J.T. Lu, K.H. Lubiński, J. Mai, P.L. Manoukian, S. Marks, J.R. Matsuno, R.K. Matsuo, K. May, T. McGuffog, L. McLaughlin, J.R. McNeish, I.A. Mebirouk, N. Menon, U. Miller, A. Milne, R.L. Minlikeeva, A. Modugno, F. Montagna, M. Moysich, K.B. Munro, E. Nathanson, K.L. Neuhausen, S.L. Nevanlinna, H. Yie, J.N.Y. Nielsen, H.R. Nielsen, F.C. Nikitina-Zake, L. Odunsi, K. Offit, K. Olah, E. Olbrecht, S. Olopade, O.I. Olson, S.H. Olsson, H. Osorio, A. Papi, L. Park, S.K. Parsons, M.T. Pathak, H. Pedersen, I.S. Peixoto, A. Pejovic, T. Perez-Segura, P. Permuth, J.B. Peshkin, B. Peterlongo, P. Piskorz, A. Prokofyeva, D. Radice, P. Rantala, J. Riggan, M.J. Risch, H.A. Rodriguez-Antona, C. Ross, E. Rossing, M.A. Runnebaum, I. Sandler, D.P. Santamariña, M. Soucy, P. Schmutzler, R.K. Setiawan, V.W. Shan, K. Sieh, W. Simard, J. Singer, C.F. Sokolenko, A.P. Song, H. Southey, M.C. Steed, H. Stoppa-Lyonnet, D. Sutphen, R. Swerdlow, A.J. Tan, Y.Y. Teixeira, M.R. Teo, S.H. Terry, K.L. Terry, M.B. OCAC Consortium, . CIMBA Consortium, . Thomassen, M. Thompson, P.J. Thomsen, L.C.V. Thull, D.L. Tischkowitz, M. Titus, L. Toland, A.E. Torres, D. Trabert, B. Travis, R. Tung, N. Tworoger, S.S. Valen, E. van Altena, A.M. van der Hout, A.H. Van Nieuwenhuysen, E. van Rensburg, E.J. Vega, A. Edwards, D.V. Vierkant, R.A. Wang, F. Wappenschmidt, B. Webb, P.M. Weinberg, C.R. Weitzel, J.N. Wentzensen, N. White, E. Whittemore, A.S. Winham, S.J. Wolk, A. Woo, Y.-.L. Wu, A.H. Yan, L. Yannoukakos, D. Zavaglia, K.M. Zheng, W. Ziogas, A. Zorn, K.K. Kleibl, Z. Easton, D. Lawrenson, K. DeFazio, A. Sellers, T.A. Ramus, S.J. Pearce, C.L. Monteiro, A.N. Cunningham, J. Goode, E.L. Schildkraut, J.M. Berchuck, A. Chenevix-Trench, G. Gayther, S.A. Antoniou, A.C. Pharoah, P.D.P (2022) Polygenic risk modeling for prediction of epithelial ovarian cancer risk.. Show Abstract full text

Polygenic risk scores (PRS) for epithelial ovarian cancer (EOC) have the potential to improve risk stratification. Joint estimation of Single Nucleotide Polymorphism (SNP) effects in models could improve predictive performance over standard approaches of PRS construction. Here, we implemented computationally efficient, penalized, logistic regression models (lasso, elastic net, stepwise) to individual level genotype data and a Bayesian framework with continuous shrinkage, "select and shrink for summary statistics" (S4), to summary level data for epithelial non-mucinous ovarian cancer risk prediction. We developed the models in a dataset consisting of 23,564 non-mucinous EOC cases and 40,138 controls participating in the Ovarian Cancer Association Consortium (OCAC) and validated the best models in three populations of different ancestries: prospective data from 198,101 women of European ancestries; 7,669 women of East Asian ancestries; 1,072 women of African ancestries, and in 18,915 BRCA1 and 12,337 BRCA2 pathogenic variant carriers of European ancestries. In the external validation data, the model with the strongest association for non-mucinous EOC risk derived from the OCAC model development data was the S4 model (27,240 SNPs) with odds ratios (OR) of 1.38 (95% CI: 1.28-1.48, AUC: 0.588) per unit standard deviation, in women of European ancestries; 1.14 (95% CI: 1.08-1.19, AUC: 0.538) in women of East Asian ancestries; 1.38 (95% CI: 1.21-1.58, AUC: 0.593) in women of African ancestries; hazard ratios of 1.36 (95% CI: 1.29-1.43, AUC: 0.592) in BRCA1 pathogenic variant carriers and 1.49 (95% CI: 1.35-1.64, AUC: 0.624) in BRCA2 pathogenic variant carriers. Incorporation of the S4 PRS in risk prediction models for ovarian cancer may have clinical utility in ovarian cancer prevention programs.

Ahearn, T.U. Zhang, H. Michailidou, K. Milne, R.L. Bolla, M.K. Dennis, J. Dunning, A.M. Lush, M. Wang, Q. Andrulis, I.L. Anton-Culver, H. Arndt, V. Aronson, K.J. Auer, P.L. Augustinsson, A. Baten, A. Becher, H. Behrens, S. Benitez, J. Bermisheva, M. Blomqvist, C. Bojesen, S.E. Bonanni, B. Børresen-Dale, A.-.L. Brauch, H. Brenner, H. Brooks-Wilson, A. Brüning, T. Burwinkel, B. Buys, S.S. Canzian, F. Castelao, J.E. Chang-Claude, J. Chanock, S.J. Chenevix-Trench, G. Clarke, C.L. NBCS Collaborators, . Collée, J.M. Cox, A. Cross, S.S. Czene, K. Daly, M.B. Devilee, P. Dörk, T. Dwek, M. Eccles, D.M. Evans, D.G. Fasching, P.A. Figueroa, J. Floris, G. Gago-Dominguez, M. Gapstur, S.M. García-Sáenz, J.A. Gaudet, M.M. Giles, G.G. Goldberg, M.S. González-Neira, A. Alnæs, G.I.G. Grip, M. Guénel, P. Haiman, C.A. Hall, P. Hamann, U. Harkness, E.F. Heemskerk-Gerritsen, B.A.M. Holleczek, B. Hollestelle, A. Hooning, M.J. Hoover, R.N. Hopper, J.L. Howell, A. ABCTB Investigators, . kConFab/AOCS Investigators, . Jakimovska, M. Jakubowska, A. John, E.M. Jones, M.E. Jung, A. Kaaks, R. Kauppila, S. Keeman, R. Khusnutdinova, E. Kitahara, C.M. Ko, Y.-.D. Koutros, S. Kristensen, V.N. Krüger, U. Kubelka-Sabit, K. Kurian, A.W. Kyriacou, K. Lambrechts, D. Lee, D.G. Lindblom, A. Linet, M. Lissowska, J. Llaneza, A. Lo, W.-.Y. MacInnis, R.J. Mannermaa, A. Manoochehri, M. Margolin, S. Martinez, M.E. McLean, C. Meindl, A. Menon, U. Nevanlinna, H. Newman, W.G. Nodora, J. Offit, K. Olsson, H. Orr, N. Park-Simon, T.-.W. Patel, A.V. Peto, J. Pita, G. Plaseska-Karanfilska, D. Prentice, R. Punie, K. Pylkäs, K. Radice, P. Rennert, G. Romero, A. Rüdiger, T. Saloustros, E. Sampson, S. Sandler, D.P. Sawyer, E.J. Schmutzler, R.K. Schoemaker, M.J. Schöttker, B. Sherman, M.E. Shu, X.-.O. Smichkoska, S. Southey, M.C. Spinelli, J.J. Swerdlow, A.J. Tamimi, R.M. Tapper, W.J. Taylor, J.A. Teras, L.R. Terry, M.B. Torres, D. Troester, M.A. Vachon, C.M. van Deurzen, C.H.M. van Veen, E.M. Wagner, P. Weinberg, C.R. Wendt, C. Wesseling, J. Winqvist, R. Wolk, A. Yang, X.R. Zheng, W. Couch, F.J. Simard, J. Kraft, P. Easton, D.F. Pharoah, P.D.P. Schmidt, M.K. García-Closas, M. Chatterjee, N (2022) Common variants in breast cancer risk loci predispose to distinct tumor subtypes.. Show Abstract full text

<h4>Background</h4>Genome-wide association studies (GWAS) have identified multiple common breast cancer susceptibility variants. Many of these variants have differential associations by estrogen receptor (ER) status, but how these variants relate with other tumor features and intrinsic molecular subtypes is unclear.<h4>Methods</h4>Among 106,571 invasive breast cancer cases and 95,762 controls of European ancestry with data on 173 breast cancer variants identified in previous GWAS, we used novel two-stage polytomous logistic regression models to evaluate variants in relation to multiple tumor features (ER, progesterone receptor (PR), human epidermal growth factor receptor 2 (HER2) and grade) adjusting for each other, and to intrinsic-like subtypes.<h4>Results</h4>Eighty-five of 173 variants were associated with at least one tumor feature (false discovery rate < 5%), most commonly ER and grade, followed by PR and HER2. Models for intrinsic-like subtypes found nearly all of these variants (83 of 85) associated at p < 0.05 with risk for at least one luminal-like subtype, and approximately half (41 of 85) of the variants were associated with risk of at least one non-luminal subtype, including 32 variants associated with triple-negative (TN) disease. Ten variants were associated with risk of all subtypes in different magnitude. Five variants were associated with risk of luminal A-like and TN subtypes in opposite directions.<h4>Conclusion</h4>This report demonstrates a high level of complexity in the etiology heterogeneity of breast cancer susceptibility variants and can inform investigations of subtype-specific risk prediction.

Townsend, M.K. Trabert, B. Fortner, R.T. Arslan, A.A. Buring, J.E. Carter, B.D. Giles, G.G. Irvin, S.R. Jones, M.E. Kaaks, R. Kirsh, V.A. Knutsen, S.F. Koh, W.-.P. Lacey, J.V. Langseth, H. Larsson, S.C. Lee, I.-.M. Martínez, M.E. Merritt, M.A. Milne, R.L. O'Brien, K.M. Orlich, M.J. Palmer, J.R. Patel, A.V. Peters, U. Poynter, J.N. Robien, K. Rohan, T.E. Rosenberg, L. Sandin, S. Sandler, D.P. Schouten, L.J. Setiawan, V.W. Swerdlow, A.J. Ursin, G. van den Brandt, P.A. Visvanathan, K. Weiderpass, E. Wolk, A. Yuan, J.-.M. Zeleniuch-Jacquotte, A. Tworoger, S.S. Wentzensen, N (2022) Cohort Profile: The Ovarian Cancer Cohort Consortium (OC3)..
Pal Choudhury, P. Brook, M.N. Hurson, A.N. Lee, A. Mulder, C.V. Coulson, P. Schoemaker, M.J. Jones, M.E. Swerdlow, A.J. Chatterjee, N. Antoniou, A.C. Garcia-Closas, M (2021) Comparative validation of the BOADICEA and Tyrer-Cuzick breast cancer risk models incorporating classical risk factors and polygenic risk in a population-based prospective cohort of women of European ancestry.. Show Abstract full text

<h4>Background</h4>The Breast and Ovarian Analysis of Disease Incidence and Carrier Estimation Algorithm (BOADICEA) and the Tyrer-Cuzick breast cancer risk prediction models are commonly used in clinical practice and have recently been extended to include polygenic risk scores (PRS). In addition, BOADICEA has also been extended to include reproductive and lifestyle factors, which were already part of Tyrer-Cuzick model. We conducted a comparative prospective validation of these models after incorporating the recently developed 313-variant PRS.<h4>Methods</h4>Calibration and discrimination of 5-year absolute risk was assessed in a nested case-control sample of 1337 women of European ancestry (619 incident breast cancer cases) aged 23-75 years from the Generations Study.<h4>Results</h4>The extended BOADICEA model with reproductive/lifestyle factors and PRS was well calibrated across risk deciles; expected-to-observed ratio (E/O) at the highest risk decile :0.97 (95 % CI 0.51 - 1.86) for women younger than 50 years and 1.09 (0.66 - 1.80) for women 50 years or older. Adding reproductive/lifestyle factors and PRS to the BOADICEA model improved discrimination modestly in younger women (area under the curve (AUC) 69.7 % vs. 69.1%) and substantially in older women (AUC 64.6 % vs. 56.8%). The Tyrer-Cuzick model with PRS showed evidence of overestimation at the highest risk decile: E/O = 1.54(0.81 - 2.92) for younger and 1.73 (1.03 - 2.90) for older women.<h4>Conclusion</h4>The extended BOADICEA model identified women in a European-ancestry population at elevated breast cancer risk more accurately than the Tyrer-Cuzick model with PRS. With the increasing availability of PRS, these analyses can inform choice of risk models incorporating PRS for risk stratified breast cancer prevention among women of European ancestry.

Macklin-Doherty, A. Jones, M. Coulson, P. Bruce, C. Chau, I. Alexander, E. Iyengar, S. Taj, M. Cunningham, D. Swerdlow, A (2022) Risk of thyroid disorders in adult and childhood Hodgkin lymphoma survivors 40 years after treatment.. Show Abstract full text

Thyroid abnormalities are well reported following childhood treatment for Hodgkin Lymphoma (HL). Limited information exists for adult patients and after modern treatments. We analyzed risks of thyroid disorders in 237 female participants treated at the Royal Marsden Hospital 1970-2015. Multivariable analyses of risk according to treatment and time-related factors, survival analyses, and Cox regression modeling were undertaken. Overall, 33.8% of patients reported thyroid disorders (hypothyroidism 30.0% and thyroid nodules 6.8%). Cumulative prevalence was 42.9% by 40 years follow-up. Risks were greatest after supradiaphragmatic radiotherapy (RR = 5.0, <i>p</i> < 0.001), and increasing dose (RR = 1.03/Gy, <i>p</i> < 0.001). There was no association with a chemotherapy agent. Risks of thyroid disease were as raised following adult as childhood treatment. There was no trend in risk by decade of supradiaphragmatic radiotherapy treatment. Risks of thyroid disease after supradiaphragmatic radiotherapy are as great after adult as childhood treatment and persist after more recent treatment periods.

Kang, E.Y. Millstein, J. Popovic, G. Meagher, N.S. Bolithon, A. Talhouk, A. Chiu, D.S. Anglesio, M.S. Leung, B. Tang, K. Lambie, N. Pavanello, M. Da-Anoy, A. Lambrechts, D. Loverix, L. Olbrecht, S. Bisinotto, C. Garcia-Donas, J. Ruiz-Llorente, S. Yagüe-Fernandez, M. Edwards, R.P. Elishaev, E. Olawaiye, A. Taylor, S. Ataseven, B. du Bois, A. Harter, P. Lester, J. Høgdall, C.K. Armasu, S.M. Huang, Y. Vierkant, R.A. Wang, C. Winham, S.J. Heublein, S. Kommoss, F.K.F. Cramer, D.W. Sasamoto, N. van-Wagensveld, L. Lycke, M. Mateoiu, C. Joseph, J. Pike, M.C. Odunsi, K. Tseng, C.-.C. Pearce, C.L. Bilic, S. Conrads, T.P. Hartmann, A. Hein, A. Jones, M.E. Leung, Y. Beckmann, M.W. Ruebner, M. Schoemaker, M.J. Terry, K.L. El-Bahrawy, M.A. Coulson, P. Etter, J.L. LaVigne-Mager, K. Andress, J. Grube, M. Fischer, A. Neudeck, N. Robertson, G. Farrell, R. Barlow, E. Quinn, C. Hettiaratchi, A. Casablanca, Y. Erber, R. Stewart, C.J.R. Tan, A. Yu, Y. Boros, J. Brand, A.H. Harnett, P.R. Kennedy, C.J. Nevins, N. Morgan, T. Fasching, P.A. Vergote, I. Swerdlow, A.J. Candido Dos Reis, F.J. Maxwell, G.L. Neuhausen, S.L. Barquin-Garcia, A. Modugno, F. Moysich, K.B. Crowe, P.J. Hirasawa, A. Heitz, F. Karlan, B.Y. Goode, E.L. Sinn, P. Horlings, H.M. Høgdall, E. Sundfeldt, K. Kommoss, S. Staebler, A. Wu, A.H. Cohen, P.A. DeFazio, A. Lee, C.-.H. Steed, H. Le, N.D. Gayther, S.A. Lawrenson, K. Pharoah, P.D.P. Konecny, G. Cook, L.S. Ramus, S.J. Kelemen, L.E. Köbel, M (2022) MCM3 is a novel proliferation marker associated with longer survival for patients with tubo-ovarian high-grade serous carcinoma.. Show Abstract full text

Tubo-ovarian high-grade serous carcinomas (HGSC) are highly proliferative neoplasms that generally respond well to platinum/taxane chemotherapy. We recently identified minichromosome maintenance complex component 3 (MCM3), which is involved in the initiation of DNA replication and proliferation, as a favorable prognostic marker in HGSC. Our objective was to further validate whether MCM3 mRNA expression and possibly MCM3 protein levels are associated with survival in patients with HGSC. MCM3 mRNA expression was measured using NanoString expression profiling on formalin-fixed and paraffin-embedded tissue (N = 2355 HGSC) and MCM3 protein expression was assessed by immunohistochemistry (N = 522 HGSC) and compared with Ki-67. Kaplan-Meier curves and the Cox proportional hazards model were used to estimate associations with survival. Among chemotherapy-naïve HGSC, higher MCM3 mRNA expression (one standard deviation increase in the score) was associated with longer overall survival (HR = 0.87, 95% CI 0.81-0.92, p < 0.0001, N = 1840) in multivariable analysis. MCM3 mRNA expression was highest in the HGSC C5.PRO molecular subtype, although no interaction was observed between MCM3, survival and molecular subtypes. MCM3 and Ki-67 protein levels were significantly lower after exposure to neoadjuvant chemotherapy compared to chemotherapy-naïve tumors: 37.0% versus 46.4% and 22.9% versus 34.2%, respectively. Among chemotherapy-naïve HGSC, high MCM3 protein levels were also associated with significantly longer disease-specific survival (HR = 0.52, 95% CI 0.36-0.74, p = 0.0003, N = 392) compared to cases with low MCM3 protein levels in multivariable analysis. MCM3 immunohistochemistry is a promising surrogate marker of proliferation in HGSC.

Glubb, D.M. Thompson, D.J. Aben, K.K.H. Alsulimani, A. Amant, F. Annibali, D. Attia, J. Barricarte, A. Beckmann, M.W. Berchuck, A. Bermisheva, M. Bernardini, M.Q. Bischof, K. Bjorge, L. Bodelon, C. Brand, A.H. Brenton, J.D. Brinton, L.A. Bruinsma, F. Buchanan, D.D. Burghaus, S. Butzow, R. Cai, H. Carney, M.E. Chanock, S.J. Chen, C. Chen, X.Q. Chen, Z. Cook, L.S. Cunningham, J.M. De Vivo, I. deFazio, A. Doherty, J.A. Dörk, T. du Bois, A. Dunning, A.M. Dürst, M. Edwards, T. Edwards, R.P. Ekici, A.B. Ewing, A. Fasching, P.A. Ferguson, S. Flanagan, J.M. Fostira, F. Fountzilas, G. Friedenreich, C.M. Gao, B. Gaudet, M.M. Gawełko, J. Gentry-Maharaj, A. Giles, G.G. Glasspool, R. Goodman, M.T. Gronwald, J. Harris, H.R. Harter, P. Hein, A. Heitz, F. Hildebrandt, M.A.T. Hillemanns, P. Høgdall, E. Høgdall, C.K. Holliday, E.G. Huntsman, D.G. Huzarski, T. Jakubowska, A. Jensen, A. Jones, M.E. Karlan, B.Y. Karnezis, A. Kelley, J.L. Khusnutdinova, E. Killeen, J.L. Kjaer, S.K. Klapdor, R. Köbel, M. Konopka, B. Konstantopoulou, I. Kopperud, R.K. Koti, M. Kraft, P. Kupryjanczyk, J. Lambrechts, D. Larson, M.C. Le Marchand, L. Lele, S. Lester, J. Li, A.J. Liang, D. Liebrich, C. Lipworth, L. Lissowska, J. Lu, L. Lu, K.H. Macciotta, A. Mattiello, A. May, T. McAlpine, J.N. McGuire, V. McNeish, I.A. Menon, U. Modugno, F. Moysich, K.B. Nevanlinna, H. Odunsi, K. Olsson, H. Orsulic, S. Osorio, A. Palli, D. Park-Simon, T.-.W. Pearce, C.L. Pejovic, T. Permuth, J.B. Podgorska, A. Ramus, S.J. Rebbeck, T.R. Riggan, M.J. Risch, H.A. Rothstein, J.H. Runnebaum, I.B. Scott, R.J. Sellers, T.A. Senz, J. Setiawan, V.W. Siddiqui, N. Sieh, W. Spiewankiewicz, B. Sutphen, R. Swerdlow, A.J. Szafron, L.M. Teo, S.H. Thompson, P.J. Thomsen, L.C.V. Titus, L. Tone, A. Tumino, R. Turman, C. Vanderstichele, A. Edwards, D.V. Vergote, I. Vierkant, R.A. Wang, Z. Wang-Gohrke, S. Webb, P.M. OPAL Study Group, . AOCS Group, . White, E. Whittemore, A.S. Winham, S.J. Wu, X. Wu, A.H. Yannoukakos, D. Spurdle, A.B. O'Mara, T.A (2021) Cross-Cancer Genome-Wide Association Study of Endometrial Cancer and Epithelial Ovarian Cancer Identifies Genetic Risk Regions Associated with Risk of Both Cancers.. Show Abstract full text

<h4>Background</h4>Accumulating evidence suggests a relationship between endometrial cancer and ovarian cancer. Independent genome-wide association studies (GWAS) for endometrial cancer and ovarian cancer have identified 16 and 27 risk regions, respectively, four of which overlap between the two cancers. We aimed to identify joint endometrial and ovarian cancer risk loci by performing a meta-analysis of GWAS summary statistics from these two cancers.<h4>Methods</h4>Using LDScore regression, we explored the genetic correlation between endometrial cancer and ovarian cancer. To identify loci associated with the risk of both cancers, we implemented a pipeline of statistical genetic analyses (i.e., inverse-variance meta-analysis, colocalization, and M-values) and performed analyses stratified by subtype. Candidate target genes were then prioritized using functional genomic data.<h4>Results</h4>Genetic correlation analysis revealed significant genetic correlation between the two cancers (<i>r<sub>G</sub></i> = 0.43, <i>P</i> = 2.66 × 10<sup>-5</sup>). We found seven loci associated with risk for both cancers (<i>P</i> <sub>Bonferroni</sub> < 2.4 × 10<sup>-9</sup>). In addition, four novel subgenome-wide regions at 7p22.2, 7q22.1, 9p12, and 11q13.3 were identified (<i>P</i> < 5 × 10<sup>-7</sup>). Promoter-associated HiChIP chromatin loops from immortalized endometrium and ovarian cell lines and expression quantitative trait loci data highlighted candidate target genes for further investigation.<h4>Conclusions</h4>Using cross-cancer GWAS meta-analysis, we have identified several joint endometrial and ovarian cancer risk loci and candidate target genes for future functional analysis.<h4>Impact</h4>Our research highlights the shared genetic relationship between endometrial cancer and ovarian cancer. Further studies in larger sample sets are required to confirm our findings.

Heinze, K. Nazeran, T.M. Lee, S. Krämer, P. Cairns, E.S. Chiu, D.S. Leung, S.C. Kang, E.Y. Meagher, N.S. Kennedy, C.J. Boros, J. Kommoss, F. Vollert, H.-.W. Heitz, F. du Bois, A. Harter, P. Grube, M. Kraemer, B. Staebler, A. Kommoss, F.K. Heublein, S. Sinn, H.-.P. Singh, N. Laslavic, A. Elishaev, E. Olawaiye, A. Moysich, K. Modugno, F. Sharma, R. Brand, A.H. Harnett, P.R. DeFazio, A. Fortner, R.T. Lubinski, J. Lener, M. Tołoczko-Grabarek, A. Cybulski, C. Gronwald, H. Gronwald, J. Coulson, P. El-Bahrawy, M.A. Jones, M.E. Schoemaker, M.J. Swerdlow, A.J. Gorringe, K.L. Campbell, I. Cook, L. Gayther, S.A. Carney, M.E. Shvetsov, Y.B. Hernandez, B.Y. Wilkens, L.R. Goodman, M.T. Mateoiu, C. Linder, A. Sundfeldt, K. Kelemen, L.E. Gentry-Maharaj, A. Widschwendter, M. Menon, U. Bolton, K.L. Alsop, J. Shah, M. Jimenez-Linan, M. Pharoah, P.D. Brenton, J.D. Cushing-Haugen, K.L. Harris, H.R. Doherty, J.A. Gilks, B. Ghatage, P. Huntsman, D.G. Nelson, G.S. Tinker, A.V. Lee, C.-.H. Goode, E.L. Nelson, B.H. Ramus, S.J. Kommoss, S. Talhouk, A. Köbel, M. Anglesio, M.S (2022) Validated biomarker assays confirm that ARID1A loss is confounded with MMR deficiency, CD8<sup>+</sup> TIL infiltration, and provides no independent prognostic value in endometriosis-associated ovarian carcinomas.. Show Abstract full text

ARID1A (BAF250a) is a component of the SWI/SNF chromatin modifying complex, plays an important tumour suppressor role, and is considered prognostic in several malignancies. However, in ovarian carcinomas there are contradictory reports on its relationship to outcome, immune response, and correlation with clinicopathological features. We assembled a series of 1623 endometriosis-associated ovarian carcinomas, including 1078 endometrioid (ENOC) and 545 clear cell (CCOC) ovarian carcinomas, through combining resources of the Ovarian Tumor Tissue Analysis (OTTA) Consortium, the Canadian Ovarian Unified Experimental Resource (COEUR), local, and collaborative networks. Validated immunohistochemical surrogate assays for ARID1A mutations were applied to all samples. We investigated associations between ARID1A loss/mutation, clinical features, outcome, CD8<sup>+</sup> tumour-infiltrating lymphocytes (CD8<sup>+</sup> TILs), and DNA mismatch repair deficiency (MMRd). ARID1A loss was observed in 42% of CCOCs and 25% of ENOCs. We found no associations between ARID1A loss and outcomes, stage, age, or CD8<sup>+</sup> TIL status in CCOC. Similarly, we found no association with outcome or stage in endometrioid cases. In ENOC, ARID1A loss was more prevalent in younger patients (p = 0.012) and was associated with MMRd (p < 0.001) and the presence of CD8<sup>+</sup> TILs (p = 0.008). Consistent with MMRd being causative of ARID1A mutations, in a subset of ENOCs we also observed an association with ARID1A loss-of-function mutation as a result of small indels (p = 0.035, versus single nucleotide variants). In ENOC, the association with ARID1A loss, CD8<sup>+</sup> TILs, and age appears confounded by MMRd status. Although this observation does not explicitly rule out a role for ARID1A influence on CD8<sup>+</sup> TIL infiltration in ENOC, given current knowledge regarding MMRd, it seems more likely that effects are dominated by the hypermutation phenotype. This large dataset with consistently applied biomarker assessment now provides a benchmark for the prevalence of ARID1A loss-of-function mutations in endometriosis-associated ovarian cancers and brings clarity to the prognostic significance. © 2021 The Pathological Society of Great Britain and Ireland.

Ghelichkhani, P. Shahsavarinia, K. Gharekhani, A. Taghizadieh, A. Baratloo, A. Fattah, F.H.R. Abbasi, N. Gubari, M.I.M. Faridaalee, G. Dinpanah, H. Yekaninejad, M.S. Esmaeili, A. Jones, M.E. Askarian-Amiri, S. Yousefifard, M. Hosseini, M (2021) Value of Canadian C-spine rule versus the NEXUS criteria in ruling out clinically important cervical spine injuries: derivation of modified Canadian C-spine rule.. Show Abstract full text

<h4>Purpose</h4>Although, Canadian C-spine rule and the National Emergency X-Radiography Utilization Study (NEXUS) criteria in ruling out clinically important cervical spine injuries have been validated using large prospective studies, no consensus exist as to which rule should be endorsed. Therefore, the aim of the present study was to compare the accuracy of the Canadian C-spine and NEXUS criteria in ruling out clinically important cervical spine injuries in trauma patients. Finally, we introduced the modified Canadian C-spine rule.<h4>Methods</h4>A prospective diagnostic accuracy study was conducted on trauma patients referred to four emergency departments of Iran in 2018. Emergency physicians evaluated the patients based on the Canadian C-spine rule and NEXUS criteria in two groups of low risk and high risk for clinically important cervical spine injury. Afterward, all patients underwent cervical imaging. In addition, modified Canadian C-spine rule was derived by removing dangerous mechanism and simple rear-end motor vehicle collision from the model.<h4>Results</h4>Data from 673 patients were included. The area under the curve of the NEXUS criteria, Canadian C-spine, and modified Canadian C-spine rule were 0.76 [95% confidence interval (CI) 0.71-0.81)], 0.78 (95% CI 0.74-0.83), and 0.79 (95% CI 0.74-0.83), respectively. The sensitivities of NEXUS criteria, Canadian C-spine, and modified Canadian C-spine rule were 93.4%, 100.0% and 100.0%, respectively.<h4>Conclusions</h4>The modified Canadian C-spine rule has fewer variables than the original Canadian C-spine rule and is entirely based on physical examination, which seems easier to use in emergency departments.

Ghelichkhani, P. Baikpour, M. Mohammad, K. Rahim Fattah, F.H. Rezaei, N. Ahmadi, N. Darvish Noori Kalaki, S. Gubari, M.I.M. Rafei, A. Koohpayehzadeh, J. Gouya, M.M. Yousefifard, M. Jones, M.E. Hosseini, M (2021) Age, Period and Cohort Analysis of Smoking Prevalence in Iranian Population over a 25-Year Period.. Show Abstract full text

<h4>Background</h4>Current and daily smoking prevalence rates have been have investigated in several cross-sectional studies. However, analyses in terms of age-period-cohort (APC) have not been carried out. We assessed daily smoking dynamics over a 25-year period using the APC model.<h4>Methods</h4>In our analyses, we used data from 214,652 people aged 15 to 64 years, collected by national health surveys conducted in 1990-1991, 1999, 2005, 2007, 2011 and 2016. The Intrinsic Estimator model was used to analyze the impact of APC on daily smoking prevalence.<h4>Results</h4>Males were found to exhibit a higher prevalence of smoking compared to females (26.0% against 2.7%). Prevalence of smoking increased by age, peaking at the age groups of 40-44 in men and 45-49 in women, followed by a decreasing trend. The 1990 period had the highest prevalence in both genders, and the 2016 period had the lowest. The coefficients of birth cohort effects showed different patter19s of fluctuations in the two genders with the maximum and minimum coefficients for men calculated in the 1966-1970 and 1991-95 birth cohorts, and for females the 1931-1935 and 1971-1975 birth cohorts, respectively.<h4>Conclusion</h4>We showed the impact of APC on daily tobacco smoking prevalence, and these factors should be considered when dealing with smoking.

Zhang, H. Ahearn, T.U. Lecarpentier, J. Barnes, D. Beesley, J. Qi, G. Jiang, X. O'Mara, T.A. Zhao, N. Bolla, M.K. Dunning, A.M. Dennis, J. Wang, Q. Ful, Z.A. Aittomäki, K. Andrulis, I.L. Anton-Culver, H. Arndt, V. Aronson, K.J. Arun, B.K. Auer, P.L. Azzollini, J. Barrowdale, D. Becher, H. Beckmann, M.W. Behrens, S. Benitez, J. Bermisheva, M. Bialkowska, K. Blanco, A. Blomqvist, C. Bogdanova, N.V. Bojesen, S.E. Bonanni, B. Bondavalli, D. Borg, A. Brauch, H. Brenner, H. Briceno, I. Broeks, A. Brucker, S.Y. Brüning, T. Burwinkel, B. Buys, S.S. Byers, H. Caldés, T. Caligo, M.A. Calvello, M. Campa, D. Castelao, J.E. Chang-Claude, J. Chanock, S.J. Christiaens, M. Christiansen, H. Chung, W.K. Claes, K.B.M. Clarke, C.L. Cornelissen, S. Couch, F.J. Cox, A. Cross, S.S. Czene, K. Daly, M.B. Devilee, P. Diez, O. Domchek, S.M. Dörk, T. Dwek, M. Eccles, D.M. Ekici, A.B. Evans, D.G. Fasching, P.A. Figueroa, J. Foretova, L. Fostira, F. Friedman, E. Frost, D. Gago-Dominguez, M. Gapstur, S.M. Garber, J. García-Sáenz, J.A. Gaudet, M.M. Gayther, S.A. Giles, G.G. Godwin, A.K. Goldberg, M.S. Goldgar, D.E. González-Neira, A. Greene, M.H. Gronwald, J. Guénel, P. Häberle, L. Hahnen, E. Haiman, C.A. Hake, C.R. Hall, P. Hamann, U. Harkness, E.F. Heemskerk-Gerritsen, B.A.M. Hillemanns, P. Hogervorst, F.B.L. Holleczek, B. Hollestelle, A. Hooning, M.J. Hoover, R.N. Hopper, J.L. Howell, A. Huebner, H. Hulick, P.J. Imyanitov, E.N. kConFab Investigators, . ABCTB Investigators, . Isaacs, C. Izatt, L. Jager, A. Jakimovska, M. Jakubowska, A. James, P. Janavicius, R. Janni, W. John, E.M. Jones, M.E. Jung, A. Kaaks, R. Kapoor, P.M. Karlan, B.Y. Keeman, R. Khan, S. Khusnutdinova, E. Kitahara, C.M. Ko, Y.-.D. Konstantopoulou, I. Koppert, L.B. Koutros, S. Kristensen, V.N. Laenkholm, A.-.V. Lambrechts, D. Larsson, S.C. Laurent-Puig, P. Lazaro, C. Lazarova, E. Lejbkowicz, F. Leslie, G. Lesueur, F. Lindblom, A. Lissowska, J. Lo, W.-.Y. Loud, J.T. Lubinski, J. Lukomska, A. MacInnis, R.J. Mannermaa, A. Manoochehri, M. Manoukian, S. Margolin, S. Martinez, M.E. Matricardi, L. McGuffog, L. McLean, C. Mebirouk, N. Meindl, A. Menon, U. Miller, A. Mingazheva, E. Montagna, M. Mulligan, A.M. Mulot, C. Muranen, T.A. Nathanson, K.L. Neuhausen, S.L. Nevanlinna, H. Neven, P. Newman, W.G. Nielsen, F.C. Nikitina-Zake, L. Nodora, J. Offit, K. Olah, E. Olopade, O.I. Olsson, H. Orr, N. Papi, L. Papp, J. Park-Simon, T.-.W. Parsons, M.T. Peissel, B. Peixoto, A. Peshkin, B. Peterlongo, P. Peto, J. Phillips, K.-.A. Piedmonte, M. Plaseska-Karanfilska, D. Prajzendanc, K. Prentice, R. Prokofyeva, D. Rack, B. Radice, P. Ramus, S.J. Rantala, J. Rashid, M.U. Rennert, G. Rennert, H.S. Risch, H.A. Romero, A. Rookus, M.A. Rübner, M. Rüdiger, T. Saloustros, E. Sampson, S. Sandler, D.P. Sawyer, E.J. Scheuner, M.T. Schmutzler, R.K. Schneeweiss, A. Schoemaker, M.J. Schöttker, B. Schürmann, P. Senter, L. Sharma, P. Sherman, M.E. Shu, X.-.O. Singer, C.F. Smichkoska, S. Soucy, P. Southey, M.C. Spinelli, J.J. Stone, J. Stoppa-Lyonnet, D. EMBRACE Study, . GEMO Study Collaborators, . Swerdlow, A.J. Szabo, C.I. Tamimi, R.M. Tapper, W.J. Taylor, J.A. Teixeira, M.R. Terry, M. Thomassen, M. Thull, D.L. Tischkowitz, M. Toland, A.E. Tollenaar, R.A.E.M. Tomlinson, I. Torres, D. Troester, M.A. Truong, T. Tung, N. Untch, M. Vachon, C.M. van den Ouweland, A.M.W. van der Kolk, L.E. van Veen, E.M. vanRensburg, E.J. Vega, A. Wappenschmidt, B. Weinberg, C.R. Weitzel, J.N. Wildiers, H. Winqvist, R. Wolk, A. Yang, X.R. Yannoukakos, D. Zheng, W. Zorn, K.K. Milne, R.L. Kraft, P. Simard, J. Pharoah, P.D.P. Michailidou, K. Antoniou, A.C. Schmidt, M.K. Chenevix-Trench, G. Easton, D.F. Chatterjee, N. García-Closas, M (2020) Genome-wide association study identifies 32 novel breast cancer susceptibility loci from overall and subtype-specific analyses.. Show Abstract full text

Breast cancer susceptibility variants frequently show heterogeneity in associations by tumor subtype<sup>1-3</sup>. To identify novel loci, we performed a genome-wide association study including 133,384 breast cancer cases and 113,789 controls, plus 18,908 BRCA1 mutation carriers (9,414 with breast cancer) of European ancestry, using both standard and novel methodologies that account for underlying tumor heterogeneity by estrogen receptor, progesterone receptor and human epidermal growth factor receptor 2 status and tumor grade. We identified 32 novel susceptibility loci (P < 5.0 × 10<sup>-8</sup>), 15 of which showed evidence for associations with at least one tumor feature (false discovery rate < 0.05). Five loci showed associations (P < 0.05) in opposite directions between luminal and non-luminal subtypes. In silico analyses showed that these five loci contained cell-specific enhancers that differed between normal luminal and basal mammary cells. The genetic correlations between five intrinsic-like subtypes ranged from 0.35 to 0.80. The proportion of genome-wide chip heritability explained by all known susceptibility loci was 54.2% for luminal A-like disease and 37.6% for triple-negative disease. The odds ratios of polygenic risk scores, which included 330 variants, for the highest 1% of quantiles compared with middle quantiles were 5.63 and 3.02 for luminal A-like and triple-negative disease, respectively. These findings provide an improved understanding of genetic predisposition to breast cancer subtypes and will inform the development of subtype-specific polygenic risk scores.

Coignard, J. Lush, M. Beesley, J. O'Mara, T.A. Dennis, J. Tyrer, J.P. Barnes, D.R. McGuffog, L. Leslie, G. Bolla, M.K. Adank, M.A. Agata, S. Ahearn, T. Aittomäki, K. Andrulis, I.L. Anton-Culver, H. Arndt, V. Arnold, N. Aronson, K.J. Arun, B.K. Augustinsson, A. Azzollini, J. Barrowdale, D. Baynes, C. Becher, H. Bermisheva, M. Bernstein, L. Białkowska, K. Blomqvist, C. Bojesen, S.E. Bonanni, B. Borg, A. Brauch, H. Brenner, H. Burwinkel, B. Buys, S.S. Caldés, T. Caligo, M.A. Campa, D. Carter, B.D. Castelao, J.E. Chang-Claude, J. Chanock, S.J. Chung, W.K. Claes, K.B.M. Clarke, C.L. GEMO Study Collaborators, . EMBRACE Collaborators, . Collée, J.M. Conroy, D.M. Czene, K. Daly, M.B. Devilee, P. Diez, O. Ding, Y.C. Domchek, S.M. Dörk, T. Dos-Santos-Silva, I. Dunning, A.M. Dwek, M. Eccles, D.M. Eliassen, A.H. Engel, C. Eriksson, M. Evans, D.G. Fasching, P.A. Flyger, H. Fostira, F. Friedman, E. Fritschi, L. Frost, D. Gago-Dominguez, M. Gapstur, S.M. Garber, J. Garcia-Barberan, V. García-Closas, M. García-Sáenz, J.A. Gaudet, M.M. Gayther, S.A. Gehrig, A. Georgoulias, V. Giles, G.G. Godwin, A.K. Goldberg, M.S. Goldgar, D.E. González-Neira, A. Greene, M.H. Guénel, P. Haeberle, L. Hahnen, E. Haiman, C.A. Håkansson, N. Hall, P. Hamann, U. Harrington, P.A. Hart, S.N. He, W. Hogervorst, F.B.L. Hollestelle, A. Hopper, J.L. Horcasitas, D.J. Hulick, P.J. Hunter, D.J. Imyanitov, E.N. KConFab Investigators, . HEBON Investigators, . ABCTB Investigators, . Jager, A. Jakubowska, A. James, P.A. Jensen, U.B. John, E.M. Jones, M.E. Kaaks, R. Kapoor, P.M. Karlan, B.Y. Keeman, R. Khusnutdinova, E. Kiiski, J.I. Ko, Y.-.D. Kosma, V.-.M. Kraft, P. Kurian, A.W. Laitman, Y. Lambrechts, D. Le Marchand, L. Lester, J. Lesueur, F. Lindstrom, T. Lopez-Fernández, A. Loud, J.T. Luccarini, C. Mannermaa, A. Manoukian, S. Margolin, S. Martens, J.W.M. Mebirouk, N. Meindl, A. Miller, A. Milne, R.L. Montagna, M. Nathanson, K.L. Neuhausen, S.L. Nevanlinna, H. Nielsen, F.C. O'Brien, K.M. Olopade, O.I. Olson, J.E. Olsson, H. Osorio, A. Ottini, L. Park-Simon, T.-.W. Parsons, M.T. Pedersen, I.S. Peshkin, B. Peterlongo, P. Peto, J. Pharoah, P.D.P. Phillips, K.-.A. Polley, E.C. Poppe, B. Presneau, N. Pujana, M.A. Punie, K. Radice, P. Rantala, J. Rashid, M.U. Rennert, G. Rennert, H.S. Robson, M. Romero, A. Rossing, M. Saloustros, E. Sandler, D.P. Santella, R. Scheuner, M.T. Schmidt, M.K. Schmidt, G. Scott, C. Sharma, P. Soucy, P. Southey, M.C. Spinelli, J.J. Steinsnyder, Z. Stone, J. Stoppa-Lyonnet, D. Swerdlow, A. Tamimi, R.M. Tapper, W.J. Taylor, J.A. Terry, M.B. Teulé, A. Thull, D.L. Tischkowitz, M. Toland, A.E. Torres, D. Trainer, A.H. Truong, T. Tung, N. Vachon, C.M. Vega, A. Vijai, J. Wang, Q. Wappenschmidt, B. Weinberg, C.R. Weitzel, J.N. Wendt, C. Wolk, A. Yadav, S. Yang, X.R. Yannoukakos, D. Zheng, W. Ziogas, A. Zorn, K.K. Park, S.K. Thomassen, M. Offit, K. Schmutzler, R.K. Couch, F.J. Simard, J. Chenevix-Trench, G. Easton, D.F. Andrieu, N. Antoniou, A.C (2021) A case-only study to identify genetic modifiers of breast cancer risk for BRCA1/BRCA2 mutation carriers.. Show Abstract full text

Breast cancer (BC) risk for BRCA1 and BRCA2 mutation carriers varies by genetic and familial factors. About 50 common variants have been shown to modify BC risk for mutation carriers. All but three, were identified in general population studies. Other mutation carrier-specific susceptibility variants may exist but studies of mutation carriers have so far been underpowered. We conduct a novel case-only genome-wide association study comparing genotype frequencies between 60,212 general population BC cases and 13,007 cases with BRCA1 or BRCA2 mutations. We identify robust novel associations for 2 variants with BC for BRCA1 and 3 for BRCA2 mutation carriers, P < 10<sup>-8</sup>, at 5 loci, which are not associated with risk in the general population. They include rs60882887 at 11p11.2 where MADD, SP11 and EIF1, genes previously implicated in BC biology, are predicted as potential targets. These findings will contribute towards customising BC polygenic risk scores for BRCA1 and BRCA2 mutation carriers.

Trabert, B. Tworoger, S.S. O'Brien, K.M. Townsend, M.K. Fortner, R.T. Iversen, E.S. Hartge, P. White, E. Amiano, P. Arslan, A.A. Bernstein, L. Brinton, L.A. Buring, J.E. Dossus, L. Fraser, G.E. Gaudet, M.M. Giles, G.G. Gram, I.T. Harris, H.R. Bolton, J.H. Idahl, A. Jones, M.E. Kaaks, R. Kirsh, V.A. Knutsen, S.F. Kvaskoff, M. Lacey, J.V. Lee, I.-.M. Milne, R.L. Onland-Moret, N.C. Overvad, K. Patel, A.V. Peters, U. Poynter, J.N. Riboli, E. Robien, K. Rohan, T.E. Sandler, D.P. Schairer, C. Schouten, L.J. Setiawan, V.W. Swerdlow, A.J. Travis, R.C. Trichopoulou, A. van den Brandt, P.A. Visvanathan, K. Wilkens, L.R. Wolk, A. Zeleniuch-Jacquotte, A. Wentzensen, N. Ovarian Cancer Cohort Consortium (OC3), (2020) The Risk of Ovarian Cancer Increases with an Increase in the Lifetime Number of Ovulatory Cycles: An Analysis from the Ovarian Cancer Cohort Consortium (OC3).. Show Abstract full text

Repeated exposure to the acute proinflammatory environment that follows ovulation at the ovarian surface and distal fallopian tube over a woman's reproductive years may increase ovarian cancer risk. To address this, analyses included individual-level data from 558,709 naturally menopausal women across 20 prospective cohorts, among whom 3,246 developed invasive epithelial ovarian cancer (2,045 serous, 319 endometrioid, 184 mucinous, 121 clear cell, 577 other/unknown). Cox models were used to estimate multivariable-adjusted HRs between lifetime ovulatory cycles (LOC) and its components and ovarian cancer risk overall and by histotype. Women in the 90th percentile of LOC (>514 cycles) were almost twice as likely to be diagnosed with ovarian cancer than women in the 10th percentile (<294) [HR (95% confidence interval): 1.92 (1.60-2.30)]. Risk increased 14% per 5-year increase in LOC (60 cycles) [(1.10-1.17)]; this association remained after adjustment for LOC components: number of pregnancies and oral contraceptive use [1.08 (1.04-1.12)]. The association varied by histotype, with increased risk of serous [1.13 (1.09-1.17)], endometrioid [1.20 (1.10-1.32)], and clear cell [1.37 (1.18-1.58)], but not mucinous [0.99 (0.88-1.10), P-heterogeneity = 0.01] tumors. Heterogeneity across histotypes was reduced [P-heterogeneity = 0.15] with adjustment for LOC components [1.08 serous, 1.11 endometrioid, 1.26 clear cell, 0.94 mucinous]. Although the 10-year absolute risk of ovarian cancer is small, it roughly doubles as the number of LOC rises from approximately 300 to 500. The consistency and linearity of effects strongly support the hypothesis that each ovulation leads to small increases in the risk of most ovarian cancers, a risk that cumulates through life, suggesting this as an important area for identifying intervention strategies. SIGNIFICANCE: Although ovarian cancer is rare, risk of most ovarian cancers doubles as the number of lifetime ovulatory cycles increases from approximately 300 to 500. Thus, identifying an important area for cancer prevention research.

Fachal, L. Aschard, H. Beesley, J. Barnes, D.R. Allen, J. Kar, S. Pooley, K.A. Dennis, J. Michailidou, K. Turman, C. Soucy, P. Lemaçon, A. Lush, M. Tyrer, J.P. Ghoussaini, M. Moradi Marjaneh, M. Jiang, X. Agata, S. Aittomäki, K. Alonso, M.R. Andrulis, I.L. Anton-Culver, H. Antonenkova, N.N. Arason, A. Arndt, V. Aronson, K.J. Arun, B.K. Auber, B. Auer, P.L. Azzollini, J. Balmaña, J. Barkardottir, R.B. Barrowdale, D. Beeghly-Fadiel, A. Benitez, J. Bermisheva, M. Białkowska, K. Blanco, A.M. Blomqvist, C. Blot, W. Bogdanova, N.V. Bojesen, S.E. Bolla, M.K. Bonanni, B. Borg, A. Bosse, K. Brauch, H. Brenner, H. Briceno, I. Brock, I.W. Brooks-Wilson, A. Brüning, T. Burwinkel, B. Buys, S.S. Cai, Q. Caldés, T. Caligo, M.A. Camp, N.J. Campbell, I. Canzian, F. Carroll, J.S. Carter, B.D. Castelao, J.E. Chiquette, J. Christiansen, H. Chung, W.K. Claes, K.B.M. Clarke, C.L. GEMO Study Collaborators, . EMBRACE Collaborators, . Collée, J.M. Cornelissen, S. Couch, F.J. Cox, A. Cross, S.S. Cybulski, C. Czene, K. Daly, M.B. de la Hoya, M. Devilee, P. Diez, O. Ding, Y.C. Dite, G.S. Domchek, S.M. Dörk, T. Dos-Santos-Silva, I. Droit, A. Dubois, S. Dumont, M. Duran, M. Durcan, L. Dwek, M. Eccles, D.M. Engel, C. Eriksson, M. Evans, D.G. Fasching, P.A. Fletcher, O. Floris, G. Flyger, H. Foretova, L. Foulkes, W.D. Friedman, E. Fritschi, L. Frost, D. Gabrielson, M. Gago-Dominguez, M. Gambino, G. Ganz, P.A. Gapstur, S.M. Garber, J. García-Sáenz, J.A. Gaudet, M.M. Georgoulias, V. Giles, G.G. Glendon, G. Godwin, A.K. Goldberg, M.S. Goldgar, D.E. González-Neira, A. Tibiletti, M.G. Greene, M.H. Grip, M. Gronwald, J. Grundy, A. Guénel, P. Hahnen, E. Haiman, C.A. Håkansson, N. Hall, P. Hamann, U. Harrington, P.A. Hartikainen, J.M. Hartman, M. He, W. Healey, C.S. Heemskerk-Gerritsen, B.A.M. Heyworth, J. Hillemanns, P. Hogervorst, F.B.L. Hollestelle, A. Hooning, M.J. Hopper, J.L. Howell, A. Huang, G. Hulick, P.J. Imyanitov, E.N. KConFab Investigators, . HEBON Investigators, . ABCTB Investigators, . Isaacs, C. Iwasaki, M. Jager, A. Jakimovska, M. Jakubowska, A. James, P.A. Janavicius, R. Jankowitz, R.C. John, E.M. Johnson, N. Jones, M.E. Jukkola-Vuorinen, A. Jung, A. Kaaks, R. Kang, D. Kapoor, P.M. Karlan, B.Y. Keeman, R. Kerin, M.J. Khusnutdinova, E. Kiiski, J.I. Kirk, J. Kitahara, C.M. Ko, Y.-.D. Konstantopoulou, I. Kosma, V.-.M. Koutros, S. Kubelka-Sabit, K. Kwong, A. Kyriacou, K. Laitman, Y. Lambrechts, D. Lee, E. Leslie, G. Lester, J. Lesueur, F. Lindblom, A. Lo, W.-.Y. Long, J. Lophatananon, A. Loud, J.T. Lubiński, J. MacInnis, R.J. Maishman, T. Makalic, E. Mannermaa, A. Manoochehri, M. Manoukian, S. Margolin, S. Martinez, M.E. Matsuo, K. Maurer, T. Mavroudis, D. Mayes, R. McGuffog, L. McLean, C. Mebirouk, N. Meindl, A. Miller, A. Miller, N. Montagna, M. Moreno, F. Muir, K. Mulligan, A.M. Muñoz-Garzon, V.M. Muranen, T.A. Narod, S.A. Nassir, R. Nathanson, K.L. Neuhausen, S.L. Nevanlinna, H. Neven, P. Nielsen, F.C. Nikitina-Zake, L. Norman, A. Offit, K. Olah, E. Olopade, O.I. Olsson, H. Orr, N. Osorio, A. Pankratz, V.S. Papp, J. Park, S.K. Park-Simon, T.-.W. Parsons, M.T. Paul, J. Pedersen, I.S. Peissel, B. Peshkin, B. Peterlongo, P. Peto, J. Plaseska-Karanfilska, D. Prajzendanc, K. Prentice, R. Presneau, N. Prokofyeva, D. Pujana, M.A. Pylkäs, K. Radice, P. Ramus, S.J. Rantala, J. Rau-Murthy, R. Rennert, G. Risch, H.A. Robson, M. Romero, A. Rossing, M. Saloustros, E. Sánchez-Herrero, E. Sandler, D.P. Santamariña, M. Santamariña, M. Saunders, C. Sawyer, E.J. Scheuner, M.T. Schmidt, D.F. Schmutzler, R.K. Schneeweiss, A. Schoemaker, M.J. Schöttker, B. Schürmann, P. Scott, C. Scott, R.J. Senter, L. Seynaeve, C.M. Shah, M. Sharma, P. Shen, C.-.Y. Shu, X.-.O. Singer, C.F. Slavin, T.P. Smichkoska, S. Southey, M.C. Spinelli, J.J. Spurdle, A.B. Stone, J. Stoppa-Lyonnet, D. Sutter, C. Swerdlow, A.J. Tamimi, R.M. Tan, Y.Y. Tapper, W.J. Taylor, J.A. Teixeira, M.R. Tengström, M. Teo, S.H. Terry, M.B. Teulé, A. Thomassen, M. Thull, D.L. Tischkowitz, M. Toland, A.E. Tollenaar, R.A.E.M. Tomlinson, I. Torres, D. Torres-Mejía, G. Troester, M.A. Truong, T. Tung, N. Tzardi, M. Ulmer, H.-.U. Vachon, C.M. van Asperen, C.J. van der Kolk, L.E. van Rensburg, E.J. Vega, A. Viel, A. Vijai, J. Vogel, M.J. Wang, Q. Wappenschmidt, B. Weinberg, C.R. Weitzel, J.N. Wendt, C. Wildiers, H. Winqvist, R. Wolk, A. Wu, A.H. Yannoukakos, D. Zhang, Y. Zheng, W. Hunter, D. Pharoah, P.D.P. Chang-Claude, J. García-Closas, M. Schmidt, M.K. Milne, R.L. Kristensen, V.N. French, J.D. Edwards, S.L. Antoniou, A.C. Chenevix-Trench, G. Simard, J. Easton, D.F. Kraft, P. Dunning, A.M (2020) Fine-mapping of 150 breast cancer risk regions identifies 191 likely target genes.. Show Abstract full text

Genome-wide association studies have identified breast cancer risk variants in over 150 genomic regions, but the mechanisms underlying risk remain largely unknown. These regions were explored by combining association analysis with in silico genomic feature annotations. We defined 205 independent risk-associated signals with the set of credible causal variants in each one. In parallel, we used a Bayesian approach (PAINTOR) that combines genetic association, linkage disequilibrium and enriched genomic features to determine variants with high posterior probabilities of being causal. Potentially causal variants were significantly over-represented in active gene regulatory regions and transcription factor binding sites. We applied our INQUSIT pipeline for prioritizing genes as targets of those potentially causal variants, using gene expression (expression quantitative trait loci), chromatin interaction and functional annotations. Known cancer drivers, transcription factors and genes in the developmental, apoptosis, immune system and DNA integrity checkpoint gene ontology pathways were over-represented among the highest-confidence target genes.

dos Santos Silva, I. Jones, M. Malveiro, F. Swerdlow, A (1999) Mortality in the Portuguese thorotrast study.. Show Abstract full text

The Portuguese Thorotrast study cohort consists of a group of patients who received Thorotrast for diagnostic reasons between 1929 and 1956, and a group of similar patients who were given nonradioactive contrast agents. The cohort members were identified from medical records that contained information on reasons for the radiological investigation, type of procedure employed, and name and dose of the contrast medium used. This cohort was assembled in 1961, but its follow-up was interrupted in 1976. We have now reactivated this cohort and extended its follow-up through the end of 1996. Similar methods were used to follow up and ascertain cause of death for the Thorotrast-exposed and unexposed subjects. A total of 1931 patients who received Thorotrast systemically and 2258 unexposed subjects were initially identified from medical records. We were able to successfully follow up 58.6% (1131) of the Thorotrast patients and 45.7% (1032) of the unexposed patients. By the end of 1996, 92.2% of the Thorotrast patients and 75.2% of the unexposed patients were dead. Mortality from all causes was increased in the Thorotrast patients compared to those who were not exposed. This excess in mortality increased with time since exposure, peaking 30 years after administration of Thorotrast. The rise in overall mortality was essentially due to neoplasms [relative risk (RR) adjusted for sex, age and period = 6.04, 95% CI = 4.42-8.26], nonmalignant liver disorders (RR = 5.67, 95% CI 3.13-10.3) and nonmalignant hematological conditions (RR = 14.2, 95% CI = 2.54-79.3). The increase in mortality from neoplasms was explained mainly by increases in the risk of liver cancer (RR = 70.8, 95% CI = 19.9-251.3) and, to a much lesser extent, leukemia (RR = 15.2, 95% CI = 1. 28-181.7).

Swerdlow, A.J. Bruce, C. Cooke, R. Coulson, P. Jones, M.E (2022) Infertility and risk of breast cancer in men: a national case-control study in England and Wales.. Show Abstract full text

<h4>Purpose</h4>Breast cancer is uncommon in men and its aetiology is largely unknown, reflecting the limited size of studies thus far conducted. In general, number of children fathered has been found a risk factor inconsistently, and infertility not. We therefore investigated in a case-control study, the relation of risk of breast cancer in men to infertility and number of children.<h4>Patients and methods</h4>We conducted a national case-control study in England and Wales, interviewing 1998 cases incident 2005-17 and 1597 male controls, which included questions on infertility and offspring.<h4>Results</h4>Risk of breast cancer was statistically significantly associated with male-origin infertility (OR = 2.03 (95% confidence interval (CI) 1.18-3.49)) but not if a couple's infertility had been diagnosed as of origin from the female partner (OR = 0.86 (0.51-1.45)). Risk was statistically significantly raised for men who had not fathered any children (OR = 1.50 (95% CI 1.21-1.86)) compared with men who were fathers. These associations were statistically significantly present for invasive tumours but not statistically significant for in situ tumours.<h4>Conclusion</h4>Our data give strong evidence that risk of breast cancer is increased for men who are infertile. The reason is not clear and needs investigation.

Swerdlow, A.J. Bruce, C. Cooke, R. Coulson, P. Schoemaker, M.J. Jones, M.E (2022) Risk of breast cancer in men in relation to weight change: A national case-control study in England and Wales.. Show Abstract full text

Breast cancer is uncommon in men and knowledge about its causation limited. Obesity is a risk factor but there has been no investigation of whether weight change is an independent risk factor, as it is in women. In a national case-control study, 1998 men with breast cancer incident in England and Wales during 2005 to 2017 and 1597 male controls were interviewed about risk factors for breast cancer including anthropometric factors at several ages. Relative risks of breast cancer in relation to changes in body mass index (BMI) and waist/height ratios at these ages were obtained by logistic regression modelling. There were significant trends of increasing breast cancer risk with increase in BMI from age 20 to 40 (odds ratio [OR] 1.11 [95% confidence interval (CI) 1.05-1.17] per 2 kg/m<sup>2</sup> increase in BMI; P < .001), and from age 40 to 60 (OR 1.12 [1.04-1.20]; P = .003), and with increase in self-reported adiposity compared to peers at age 11 to BMI compared with peers at age 20 (OR 1.19 [1.09-1.30]; P < .001). Increase in waist/height ratio from age 20 to 5 years before diagnosis was also highly significantly associated with risk (OR 1.13 [1.08-1.19]; P < .001). The associations with increases in BMI and waist/height ratio were significant independently of each other and of BMI or waist/height ratio at the start of the period of change analysed, and effects were similar for invasive and in situ tumours separately. Increases in BMI and abdominal obesity are each risk factors for breast cancer in men, independently of obesity per se. These associations might relate to increasing oestrogen levels with weight gain, but this needs investigation.

Swerdlow, A.J. Bruce, C. Cooke, R. Coulson, P. Griffin, J. Butlin, A. Smith, B. Swerdlow, M.J. Jones, M.E (2021) Obesity and Breast Cancer Risk in Men: A National Case-Control Study in England and Wales.. Show Abstract full text

<h4>Background</h4>Breast cancer is rare in men, and information on its causes is very limited from studies that have generally been small. Adult obesity has been shown as a risk factor, but more detailed anthropometric relations have not been investigated.<h4>Methods</h4>We conducted an interview population-based case-control study of breast cancer in men in England and Wales including 1998 cases incident during 2005-2017 at ages younger than 80 years and 1597 male controls, with questions asked about a range of anthropometric variables at several ages. All tests of statistical significance were 2-sided.<h4>Results</h4>Risk of breast cancer statistically significantly increased with increasing body mass index (BMI) at ages 20 (odds ratio [OR] = 1.07, 95% confidence interval [CI] = 1.02 to 1.12 per 2-unit change in BMI), 40 (OR = 1.11, 95% CI = 1.07 to 1.16), and 60 (OR = 1.14, 95% CI = 1.09 to 1.19) years, but there was also an indication of raised risk for the lowest BMIs. Large waist circumference 5 years before interview was more strongly associated than was BMI with risk, and each showed independent associations. Associations were similar for invasive and in situ tumors separately and stronger for HER2-positive than HER2-negative tumors. Of the tumors, 99% were estrogen receptor positive.<h4>Conclusions</h4>Obesity at all adult ages, particularly recent abdominal obesity, is associated with raised risk of breast cancer in men, probably because of the conversion of testosterone to estrogen by aromatase in adipose tissue. The association is particularly strong for HER2-expressing tumors.

Schmidt, M.K. Hogervorst, F. van Hien, R. Cornelissen, S. Broeks, A. Adank, M.A. Meijers, H. Waisfisz, Q. Hollestelle, A. Schutte, M. van den Ouweland, A. Hooning, M. Andrulis, I.L. Anton-Culver, H. Antonenkova, N.N. Antoniou, A.C. Arndt, V. Bermisheva, M. Bogdanova, N.V. Bolla, M.K. Brauch, H. Brenner, H. Brüning, T. Burwinkel, B. Chang-Claude, J. Chenevix-Trench, G. Couch, F.J. Cox, A. Cross, S.S. Czene, K. Dunning, A.M. Fasching, P.A. Figueroa, J. Fletcher, O. Flyger, H. Galle, E. García-Closas, M. Giles, G.G. Haeberle, L. Hall, P. Hillemanns, P. Hopper, J.L. Jakubowska, A. John, E.M. Jones, M. Khusnutdinova, E. Knight, J.A. Kosma, V.-.M. Kristensen, V. Lee, A. Lindblom, A. Lubinski, J. Mannermaa, A. Margolin, S. Meindl, A. Milne, R.L. Muranen, T.A. Newcomb, P.A. Offit, K. Park-Simon, T.-.W. Peto, J. Pharoah, P.D.P. Robson, M. Rudolph, A. Sawyer, E.J. Schmutzler, R.K. Seynaeve, C. Soens, J. Southey, M.C. Spurdle, A.B. Surowy, H. Swerdlow, A. Tollenaar, R.A.E.M. Tomlinson, I. Trentham-Dietz, A. Vachon, C. Wang, Q. Whittemore, A.S. Ziogas, A. van der Kolk, L. Nevanlinna, H. Dörk, T. Bojesen, S. Easton, D.F (2016) Age- and Tumor Subtype-Specific Breast Cancer Risk Estimates for CHEK2*1100delC Carriers.. Show Abstract full text

<h4>Purpose</h4>CHEK2*1100delC is a well-established breast cancer risk variant that is most prevalent in European populations; however, there are limited data on risk of breast cancer by age and tumor subtype, which limits its usefulness in breast cancer risk prediction. We aimed to generate tumor subtype- and age-specific risk estimates by using data from the Breast Cancer Association Consortium, including 44,777 patients with breast cancer and 42,997 controls from 33 studies genotyped for CHEK2*1100delC.<h4>Patients and methods</h4>CHEK2*1100delC genotyping was mostly done by a custom Taqman assay. Breast cancer odds ratios (ORs) for CHEK2*1100delC carriers versus noncarriers were estimated by using logistic regression and adjusted for study (categorical) and age. Main analyses included patients with invasive breast cancer from population- and hospital-based studies.<h4>Results</h4>Proportions of heterozygous CHEK2*1100delC carriers in controls, in patients with breast cancer from population- and hospital-based studies, and in patients with breast cancer from familial- and clinical genetics center-based studies were 0.5%, 1.3%, and 3.0%, respectively. The estimated OR for invasive breast cancer was 2.26 (95%CI, 1.90 to 2.69; P = 2.3 × 10(-20)). The OR was higher for estrogen receptor (ER)-positive disease (2.55 [95%CI, 2.10 to 3.10; P = 4.9 × 10(-21)]) than it was for ER-negative disease (1.32 [95%CI, 0.93 to 1.88; P = .12]; P interaction = 9.9 × 10(-4)). The OR significantly declined with attained age for breast cancer overall (P = .001) and for ER-positive tumors (P = .001). Estimated cumulative risks for development of ER-positive and ER-negative tumors by age 80 in CHEK2*1100delC carriers were 20% and 3%, respectively, compared with 9% and 2%, respectively, in the general population of the United Kingdom.<h4>Conclusion</h4>These CHEK2*1100delC breast cancer risk estimates provide a basis for incorporating CHEK2*1100delC into breast cancer risk prediction models and into guidelines for intensified screening and follow-up.

Wang, X. Kapoor, P.M. Auer, P.L. Dennis, J. Dunning, A.M. Wang, Q. Lush, M. Michailidou, K. Bolla, M.K. Aronson, K.J. Murphy, R.A. Brooks-Wilson, A. Lee, D.G. Cordina-Duverger, E. Guénel, P. Truong, T. Mulot, C. Teras, L.R. Patel, A.V. Dossus, L. Kaaks, R. Hoppe, R. Lo, W.-.Y. Brüning, T. Hamann, U. Czene, K. Gabrielson, M. Hall, P. Eriksson, M. Jung, A. Becher, H. Couch, F.J. Larson, N.L. Olson, J.E. Ruddy, K.J. Giles, G.G. MacInnis, R.J. Southey, M.C. Le Marchand, L. Wilkens, L.R. Haiman, C.A. Olsson, H. Augustinsson, A. Krüger, U. Wagner, P. Scott, C. Winham, S.J. Vachon, C.M. Perou, C.M. Olshan, A.F. Troester, M.A. Hunter, D.J. Eliassen, H.A. Tamimi, R.M. Brantley, K. Andrulis, I.L. Figueroa, J. Chanock, S.J. Ahearn, T.U. García-Closas, M. Evans, G.D. Newman, W.G. van Veen, E.M. Howell, A. Wolk, A. Håkansson, N. Anton-Culver, H. Ziogas, A. Jones, M.E. Orr, N. Schoemaker, M.J. Swerdlow, A.J. Kitahara, C.M. Linet, M. Prentice, R.L. Easton, D.F. Milne, R.L. Kraft, P. Chang-Claude, J. Lindström, S (2022) Genome-wide interaction analysis of menopausal hormone therapy use and breast cancer risk among 62,370 women.. Show Abstract full text

Use of menopausal hormone therapy (MHT) is associated with increased risk for breast cancer. However, the relevant mechanisms and its interaction with genetic variants are not fully understood. We conducted a genome-wide interaction analysis between MHT use and genetic variants for breast cancer risk in 27,585 cases and 34,785 controls from 26 observational studies. All women were post-menopausal and of European ancestry. Multivariable logistic regression models were used to test for multiplicative interactions between genetic variants and current MHT use. We considered interaction p-values < 5 × 10<sup>-8</sup> as genome-wide significant, and p-values < 1 × 10<sup>-5</sup> as suggestive. Linkage disequilibrium (LD)-based clumping was performed to identify independent candidate variants. None of the 9.7 million genetic variants tested for interactions with MHT use reached genome-wide significance. Only 213 variants, representing 18 independent loci, had p-values < 1 × 10<sup>5</sup>. The strongest evidence was found for rs4674019 (p-value = 2.27 × 10<sup>-7</sup>), which showed genome-wide significant interaction (p-value = 3.8 × 10<sup>-8</sup>) with current MHT use when analysis was restricted to population-based studies only. Limiting the analyses to combined estrogen-progesterone MHT use only or to estrogen receptor (ER) positive cases did not identify any genome-wide significant evidence of interactions. In this large genome-wide SNP-MHT interaction study of breast cancer, we found no strong support for common genetic variants modifying the effect of MHT on breast cancer risk. These results suggest that common genetic variation has limited impact on the observed MHT-breast cancer risk association.

Huntley, C. Torr, B. Sud, A. Rowlands, C.F. Way, R. Snape, K. Hanson, H. Swanton, C. Broggio, J. Lucassen, A. McCartney, M. Houlston, R.S. Hingorani, A.D. Jones, M.E. Turnbull, C (2023) Utility of polygenic risk scores in UK cancer screening: a modelling analysis.. Show Abstract full text

<h4>Background</h4>It is proposed that, through restriction to individuals delineated as high risk, polygenic risk scores (PRSs) might enable more efficient targeting of existing cancer screening programmes and enable extension into new age ranges and disease types. To address this proposition, we present an overview of the performance of PRS tools (ie, models and sets of single nucleotide polymorphisms) alongside harms and benefits of PRS-stratified cancer screening for eight example cancers (breast, prostate, colorectal, pancreas, ovary, kidney, lung, and testicular cancer).<h4>Methods</h4>For this modelling analysis, we used age-stratified cancer incidences for the UK population from the National Cancer Registration Dataset (2016-18) and published estimates of the area under the receiver operating characteristic curve for current, future, and optimised PRS for each of the eight cancer types. For each of five PRS-defined high-risk quantiles (ie, the top 50%, 20%, 10%, 5%, and 1%) and according to each of the three PRS tools (ie, current, future, and optimised) for the eight cancers, we calculated the relative proportion of cancers arising, the odds ratios of a cancer arising compared with the UK population average, and the lifetime cancer risk. We examined maximal attainable rates of cancer detection by age stratum from combining PRS-based stratification with cancer screening tools and modelled the maximal impact on cancer-specific survival of hypothetical new UK programmes of PRS-stratified screening.<h4>Findings</h4>The PRS-defined high-risk quintile (20%) of the population was estimated to capture 37% of breast cancer cases, 46% of prostate cancer cases, 34% of colorectal cancer cases, 29% of pancreatic cancer cases, 26% of ovarian cancer cases, 22% of renal cancer cases, 26% of lung cancer cases, and 47% of testicular cancer cases. Extending UK screening programmes to a PRS-defined high-risk quintile including people aged 40-49 years for breast cancer, 50-59 years for colorectal cancer, and 60-69 years for prostate cancer has the potential to avert, respectively, a maximum of 102, 188, and 158 deaths annually. Unstratified screening of the full population aged 48-49 years for breast cancer, 58-59 years for colorectal cancer, and 68-69 years for prostate cancer would use equivalent resources and avert, respectively, an estimated maximum of 80, 155, and 95 deaths annually. These maximal modelled numbers will be substantially attenuated by incomplete population uptake of PRS profiling and cancer screening, interval cancers, non-European ancestry, and other factors.<h4>Interpretation</h4>Under favourable assumptions, our modelling suggests modest potential efficiency gain in cancer case detection and deaths averted for hypothetical new PRS-stratified screening programmes for breast, prostate, and colorectal cancer. Restriction of screening to high-risk quantiles means many or most incident cancers will arise in those assigned as being low-risk. To quantify real-world clinical impact, costs, and harms, UK-specific cluster-randomised trials are required.<h4>Funding</h4>The Wellcome Trust.

Visvanathan, K. Mondul, A.M. Zeleniuch-Jacquotte, A. Wang, M. Gail, M.H. Yaun, S.-.S. Weinstein, S.J. McCullough, M.L. Eliassen, A.H. Cook, N.R. Agnoli, C. Almquist, M. Black, A. Buring, J.E. Chen, C. Chen, Y. Clendenen, T. Dossus, L. Fedirko, V. Gierach, G.L. Giovannucci, E.L. Goodman, G.E. Goodman, M.T. Guénel, P. Hallmans, G. Hankinson, S.E. Horst, R.L. Hou, T. Huang, W.-.Y. Jones, M.E. Joshu, C.E. Kaaks, R. Krogh, V. Kühn, T. Kvaskoff, M. Lee, I.-.M. Mahamat-Saleh, Y. Malm, J. Manjer, J. Maskarinec, G. Millen, A.E. Mukhtar, T.K. Neuhouser, M.L. Robsahm, T.E. Schoemaker, M.J. Sieri, S. Sund, M. Swerdlow, A.J. Thomson, C.A. Ursin, G. Wactawski-Wende, J. Wang, Y. Wilkens, L.R. Wu, Y. Zoltick, E. Willett, W.C. Smith-Warner, S.A. Ziegler, R.G (2023) Circulating vitamin D and breast cancer risk: an international pooling project of 17 cohorts.. Show Abstract full text

Laboratory and animal research support a protective role for vitamin D in breast carcinogenesis, but epidemiologic studies have been inconclusive. To examine comprehensively the relationship of circulating 25-hydroxyvitamin D [25(OH)D] to subsequent breast cancer incidence, we harmonized and pooled participant-level data from 10 U.S. and 7 European prospective cohorts. Included were 10,484 invasive breast cancer cases and 12,953 matched controls. Median age (interdecile range) was 57 (42-68) years at blood collection and 63 (49-75) years at breast cancer diagnosis. Prediagnostic circulating 25(OH)D was either newly measured using a widely accepted immunoassay and laboratory or, if previously measured by the cohort, calibrated to this assay to permit using a common metric. Study-specific relative risks (RRs) for season-standardized 25(OH)D concentrations were estimated by conditional logistic regression and combined by random-effects models. Circulating 25(OH)D increased from a median of 22.6 nmol/L in consortium-wide decile 1 to 93.2 nmol/L in decile 10. Breast cancer risk in each decile was not statistically significantly different from risk in decile 5 in models adjusted for breast cancer risk factors, and no trend was apparent (P-trend = 0.64). Compared to women with sufficient 25(OH)D based on Institute of Medicine guidelines (50- < 62.5 nmol/L), RRs were not statistically significantly different at either low concentrations (< 20 nmol/L, 3% of controls) or high concentrations (100- < 125 nmol/L, 3% of controls; ≥ 125 nmol/L, 0.7% of controls). RR per 25 nmol/L increase in 25(OH)D was 0.99 [95% confidence intervaI (CI) 0.95-1.03]. Associations remained null across subgroups, including those defined by body mass index, physical activity, latitude, and season of blood collection. Although none of the associations by tumor characteristics reached statistical significance, suggestive inverse associations were seen for distant and triple negative tumors. Circulating 25(OH)D, comparably measured in 17 international cohorts and season-standardized, was not related to subsequent incidence of invasive breast cancer over a broad range in vitamin D status.

Swerdlow, A.J. Jones, M.E. Slater, S.D. Burden, A.C.F. Botha, J.L. Waugh, N.R. Morris, A.D. Gatling, W. Gillespie, K.M. Patterson, C.C. Schoemaker, M.J (2023) Cancer incidence and mortality in 23 000 patients with type 1 diabetes in the UK: Long-term follow-up.. Show Abstract full text

Type 2 diabetes is associated with raised risk of several cancers, but for type 1 diabetes risk data are fewer and inconsistent We assembled a cohort of 23 473 UK patients with insulin-treated diabetes diagnosed at ages <30, almost all of whom will have had type 1 diabetes, and for comparison 5058 diagnosed at ages 30 to 49, of whom we estimate two-thirds will have had type 2, and followed them for an average of 30 years for cancer incidence and mortality compared with general population rates. Patients aged <30 at diabetes diagnosis had significantly raised risks only for ovarian (standardised incidence ratio = 1.58; 95% confidence interval 1.16-2.11; P < .01) and vulval (3.55; 1.94-5.96; P < .001) cancers, with greatest risk when diabetes was diagnosed at ages 10-14. Risks of cancer overall (0.89; 0.84-0.95; P < .001) and sites including lung and larynx were significantly diminished. Patients diagnosed with diabetes at ages 30 to 49 had significantly raised risks of liver (1.76;1.08-2.72) and kidney (1.46;1.03-2.00) cancers, and reduced risk of cancer overall (0.89; 0.84-0.95). The raised ovarian and vulval cancer risks in patients with type 1 diabetes, especially with diabetes diagnosed around pubertal ages, suggest possible susceptibility of these organs at puberty to metabolic disruption at diabetes onset. Reduced risk of cancer overall, particularly smoking and alcohol-related sites, might reflect adoption of a healthy lifestyle.

Robbins, H.A. Alcala, K. Moez, E.K. Guida, F. Thomas, S. Zahed, H. Warkentin, M.T. Smith-Byrne, K. Brhane, Y. Muller, D. Feng, X. Albanes, D. Aldrich, M.C. Arslan, A.A. Bassett, J. Berg, C.D. Cai, Q. Chen, C. Davies, M.P.A. Diergaarde, B. Field, J.K. Freedman, N.D. Huang, W.-.Y. Johansson, M. Jones, M. Koh, W.-.P. Lam, S. Lan, Q. Langhammer, A. Liao, L.M. Liu, G. Malekzadeh, R. Milne, R.L. Montuenga, L.M. Rohan, T. Sesso, H.D. Severi, G. Sheikh, M. Sinha, R. Shu, X.-.O. Stevens, V.L. Tammemägi, M.C. Tinker, L.F. Visvanathan, K. Wang, Y. Wang, R. Weinstein, S.J. White, E. Wilson, D. Yuan, J.-.M. Zhang, X. Zheng, W. Amos, C.I. Brennan, P. Johansson, M. Hung, R.J (2023) Design and methodological considerations for biomarker discovery and validation in the Integrative Analysis of Lung Cancer Etiology and Risk (INTEGRAL) Program.. Show Abstract full text

The Integrative Analysis of Lung Cancer Etiology and Risk (INTEGRAL) program is an NCI-funded initiative with an objective to develop tools to optimize low-dose CT (LDCT) lung cancer screening. Here, we describe the rationale and design for the Risk Biomarker and Nodule Malignancy projects within INTEGRAL. The overarching goal of these projects is to systematically investigate circulating protein markers to include on a panel for use (i) pre-LDCT, to identify people likely to benefit from screening, and (ii) post-LDCT, to differentiate benign versus malignant nodules. To identify informative proteins, the Risk Biomarker project measured 1161 proteins in a nested-case control study within 2 prospective cohorts (n = 252 lung cancer cases and 252 controls) and replicated associations for a subset of proteins in 4 cohorts (n = 479 cases and 479 controls). Eligible participants had a current or former history of smoking and cases were diagnosed up to 3 years following blood draw. The Nodule Malignancy project measured 1078 proteins among participants with a heavy smoking history within four LDCT screening studies (n = 425 cases diagnosed up to 5 years following blood draw, 430 benign-nodule controls, and 398 nodule-free controls). The INTEGRAL panel will enable absolute quantification of 21 proteins. We will evaluate its performance in the Risk Biomarker project using a case-cohort study including 14 cohorts (n = 1696 cases and 2926 subcohort representatives), and in the Nodule Malignancy project within five LDCT screening studies (n = 675 cases, 680 benign-nodule controls, and 648 nodule-free controls). Future progress to advance lung cancer early detection biomarkers will require carefully designed validation, translational, and comparative studies.

Nichols, H.B. House, M.G. Yarosh, R. Mitra, S. Goldberg, M. Bertrand, K.A. Eliassen, A.H. Giles, G.G. Jones, M.E. Milne, R.L. O'Brien, K.M. Palmer, J.R. Sandin, S. Willett, W.C. Yin, W. Sandler, D.P. Swerdlow, A.J. Schoemaker, M.J (2023) Hypertensive conditions of pregnancy, preterm birth, and premenopausal breast cancer risk: a premenopausal breast cancer collaborative group analysis.. Show Abstract full text

<h4>Purpose</h4>Women with preeclampsia are more likely to deliver preterm. Reports of inverse associations between preeclampsia and breast cancer risk, and positive associations between preterm birth and breast cancer risk are difficult to reconcile. We investigated the co-occurrence of preeclampsia/gestational hypertension with preterm birth and breast cancer risk using data from the Premenopausal Breast Cancer Collaborative Group.<h4>Methods</h4>Across 6 cohorts, 3096 premenopausal breast cancers were diagnosed among 184,866 parous women. We estimated multivariable hazard ratios (HR) and 95% confidence intervals (CI) for premenopausal breast cancer risk using Cox proportional hazards regression.<h4>Results</h4>Overall, preterm birth was not associated (HR 1.02, 95% CI 0.92, 1.14), and preeclampsia was inversely associated (HR 0.86, 95% CI 0.76, 0.99), with premenopausal breast cancer risk. In stratified analyses using data from 3 cohorts, preterm birth associations with breast cancer risk were modified by hypertensive conditions in first pregnancies (P-interaction = 0.09). Preterm birth was positively associated with premenopausal breast cancer in strata of women with preeclampsia or gestational hypertension (HR 1.52, 95% CI: 1.06, 2.18), but not among women with normotensive pregnancy (HR = 1.09, 95% CI: 0.93, 1.28). When stratified by preterm birth, the inverse association with preeclampsia was more apparent, but not statistically different (P-interaction = 0.2), among women who did not deliver preterm (HR = 0.82, 95% CI 0.68, 1.00) than those who did (HR = 1.07, 95% CI 0.73, 1.56).<h4>Conclusion</h4>Findings support an overall inverse association of preeclampsia history with premenopausal breast cancer risk. Estimates for preterm birth and breast cancer may vary according to other conditions of pregnancy.

Mueller, S.H. Lai, A.G. Valkovskaya, M. Michailidou, K. Bolla, M.K. Wang, Q. Dennis, J. Lush, M. Abu-Ful, Z. Ahearn, T.U. Andrulis, I.L. Anton-Culver, H. Antonenkova, N.N. Arndt, V. Aronson, K.J. Augustinsson, A. Baert, T. Freeman, L.E.B. Beckmann, M.W. Behrens, S. Benitez, J. Bermisheva, M. Blomqvist, C. Bogdanova, N.V. Bojesen, S.E. Bonanni, B. Brenner, H. Brucker, S.Y. Buys, S.S. Castelao, J.E. Chan, T.L. Chang-Claude, J. Chanock, S.J. Choi, J.-.Y. Chung, W.K. NBCS Collaborators, . Colonna, S.V. CTS Consortium, . Cornelissen, S. Couch, F.J. Czene, K. Daly, M.B. Devilee, P. Dörk, T. Dossus, L. Dwek, M. Eccles, D.M. Ekici, A.B. Eliassen, A.H. Engel, C. Evans, D.G. Fasching, P.A. Fletcher, O. Flyger, H. Gago-Dominguez, M. Gao, Y.-.T. García-Closas, M. García-Sáenz, J.A. Genkinger, J. Gentry-Maharaj, A. Grassmann, F. Guénel, P. Gündert, M. Haeberle, L. Hahnen, E. Haiman, C.A. Håkansson, N. Hall, P. Harkness, E.F. Harrington, P.A. Hartikainen, J.M. Hartman, M. Hein, A. Ho, W.-.K. Hooning, M.J. Hoppe, R. Hopper, J.L. Houlston, R.S. Howell, A. Hunter, D.J. Huo, D. ABCTB Investigators, . Ito, H. Iwasaki, M. Jakubowska, A. Janni, W. John, E.M. Jones, M.E. Jung, A. Kaaks, R. Kang, D. Khusnutdinova, E.K. Kim, S.-.W. Kitahara, C.M. Koutros, S. Kraft, P. Kristensen, V.N. Kubelka-Sabit, K. Kurian, A.W. Kwong, A. Lacey, J.V. Lambrechts, D. Le Marchand, L. Li, J. Linet, M. Lo, W.-.Y. Long, J. Lophatananon, A. Mannermaa, A. Manoochehri, M. Margolin, S. Matsuo, K. Mavroudis, D. Menon, U. Muir, K. Murphy, R.A. Nevanlinna, H. Newman, W.G. Niederacher, D. O'Brien, K.M. Obi, N. Offit, K. Olopade, O.I. Olshan, A.F. Olsson, H. Park, S.K. Patel, A.V. Patel, A. Perou, C.M. Peto, J. Pharoah, P.D.P. Plaseska-Karanfilska, D. Presneau, N. Rack, B. Radice, P. Ramachandran, D. Rashid, M.U. Rennert, G. Romero, A. Ruddy, K.J. Ruebner, M. Saloustros, E. Sandler, D.P. Sawyer, E.J. Schmidt, M.K. Schmutzler, R.K. Schneider, M.O. Scott, C. Shah, M. Sharma, P. Shen, C.-.Y. Shu, X.-.O. Simard, J. Surowy, H. Tamimi, R.M. Tapper, W.J. Taylor, J.A. Teo, S.H. Teras, L.R. Toland, A.E. Tollenaar, R.A.E.M. Torres, D. Torres-Mejía, G. Troester, M.A. Truong, T. Vachon, C.M. Vijai, J. Weinberg, C.R. Wendt, C. Winqvist, R. Wolk, A. Wu, A.H. Yamaji, T. Yang, X.R. Yu, J.-.C. Zheng, W. Ziogas, A. Ziv, E. Dunning, A.M. Easton, D.F. Hemingway, H. Hamann, U. Kuchenbaecker, K.B (2023) Aggregation tests identify new gene associations with breast cancer in populations with diverse ancestry.. Show Abstract full text

<h4>Background</h4>Low-frequency variants play an important role in breast cancer (BC) susceptibility. Gene-based methods can increase power by combining multiple variants in the same gene and help identify target genes.<h4>Methods</h4>We evaluated the potential of gene-based aggregation in the Breast Cancer Association Consortium cohorts including 83,471 cases and 59,199 controls. Low-frequency variants were aggregated for individual genes' coding and regulatory regions. Association results in European ancestry samples were compared to single-marker association results in the same cohort. Gene-based associations were also combined in meta-analysis across individuals with European, Asian, African, and Latin American and Hispanic ancestry.<h4>Results</h4>In European ancestry samples, 14 genes were significantly associated (q < 0.05) with BC. Of those, two genes, FMNL3 (P = 6.11 × 10<sup>-6</sup>) and AC058822.1 (P = 1.47 × 10<sup>-4</sup>), represent new associations. High FMNL3 expression has previously been linked to poor prognosis in several other cancers. Meta-analysis of samples with diverse ancestry discovered further associations including established candidate genes ESR1 and CBLB. Furthermore, literature review and database query found further support for a biologically plausible link with cancer for genes CBLB, FMNL3, FGFR2, LSP1, MAP3K1, and SRGAP2C.<h4>Conclusions</h4>Using extended gene-based aggregation tests including coding and regulatory variation, we report identification of plausible target genes for previously identified single-marker associations with BC as well as the discovery of novel genes implicated in BC development. Including multi ancestral cohorts in this study enabled the identification of otherwise missed disease associations as ESR1 (P = 1.31 × 10<sup>-5</sup>), demonstrating the importance of diversifying study cohorts.

Lopes Cardozo, J.M.N. Andrulis, I.L. Bojesen, S.E. Dörk, T. Eccles, D.M. Fasching, P.A. Hooning, M.J. Keeman, R. Nevanlinna, H. Rutgers, E.J.T. Easton, D.F. Hall, P. Pharoah, P.D.P. van 't Veer, L.J. Schmidt, M.K. Breast Cancer Association Consortium and MINDACT Collaborators, (2023) Associations of a Breast Cancer Polygenic Risk Score With Tumor Characteristics and Survival.. Show Abstract full text

<h4>Purpose</h4>A polygenic risk score (PRS) consisting of 313 common genetic variants (PRS<sub>313</sub>) is associated with risk of breast cancer and contralateral breast cancer. This study aimed to evaluate the association of the PRS<sub>313</sub> with clinicopathologic characteristics of, and survival following, breast cancer.<h4>Methods</h4>Women with invasive breast cancer were included, 98,397 of European ancestry and 12,920 of Asian ancestry, from the Breast Cancer Association Consortium (BCAC), and 683 women from the European MINDACT trial. Associations between PRS<sub>313</sub> and clinicopathologic characteristics, including the 70-gene signature for MINDACT, were evaluated using logistic regression analyses. Associations of PRS<sub>313</sub> (continuous, per standard deviation) with overall survival (OS) and breast cancer-specific survival (BCSS) were evaluated with Cox regression, adjusted for clinicopathologic characteristics and treatment.<h4>Results</h4>The PRS<sub>313</sub> was associated with more favorable tumor characteristics. In BCAC, increasing PRS<sub>313</sub> was associated with lower grade, hormone receptor-positive status, and smaller tumor size. In MINDACT, PRS<sub>313</sub> was associated with a low risk 70-gene signature. In European women from BCAC, higher PRS<sub>313</sub> was associated with better OS and BCSS: hazard ratio (HR) 0.96 (95% CI, 0.94 to 0.97) and 0.96 (95% CI, 0.94 to 0.98), but the association disappeared after adjustment for clinicopathologic characteristics (and treatment): OS HR, 1.01 (95% CI, 0.98 to 1.05) and BCSS HR, 1.02 (95% CI, 0.98 to 1.07). The results in MINDACT and Asian women from BCAC were consistent.<h4>Conclusion</h4>An increased PRS<sub>313</sub> is associated with favorable tumor characteristics, but is not independently associated with prognosis. Thus, PRS<sub>313</sub> has no role in the clinical management of primary breast cancer at the time of diagnosis. Nevertheless, breast cancer mortality rates will be higher for women with higher PRS<sub>313</sub> as increasing PRS<sub>313</sub> is associated with an increased risk of disease. This information is crucial for modeling effective stratified screening programs.

Köbel, M. Kang, E.-.Y. Weir, A. Rambau, P.F. Lee, C.-.H. Nelson, G.S. Ghatage, P. Meagher, N.S. Riggan, M.J. Alsop, J. Anglesio, M.S. Beckmann, M.W. Bisinotto, C. Boisen, M. Boros, J. Brand, A.H. Brooks-Wilson, A. Carney, M.E. Coulson, P. Courtney-Brooks, M. Cushing-Haugen, K.L. Cybulski, C. Deen, S. El-Bahrawy, M.A. Elishaev, E. Erber, R. Fereday, S. AOCS Group, . Fischer, A. Gayther, S.A. Barquin-Garcia, A. Gentry-Maharaj, A. Gilks, C.B. Gronwald, H. Grube, M. Harnett, P.R. Harris, H.R. Hartkopf, A.D. Hartmann, A. Hein, A. Hendley, J. Hernandez, B.Y. Huang, Y. Jakubowska, A. Jimenez-Linan, M. Jones, M.E. Kennedy, C.J. Kluz, T. Koziak, J.M. Lesnock, J. Lester, J. Lubiński, J. Longacre, T.A. Lycke, M. Mateoiu, C. McCauley, B.M. McGuire, V. Ney, B. Olawaiye, A. Orsulic, S. Osorio, A. Paz-Ares, L. Ramón Y Cajal, T. Rothstein, J.H. Ruebner, M. Schoemaker, M.J. Shah, M. Sharma, R. Sherman, M.E. Shvetsov, Y.B. Singh, N. Steed, H. Storr, S.J. Talhouk, A. Traficante, N. Wang, C. Whittemore, A.S. Widschwendter, M. Wilkens, L.R. Winham, S.J. Benitez, J. Berchuck, A. Bowtell, D.D. Candido Dos Reis, F.J. Campbell, I. Cook, L.S. DeFazio, A. Doherty, J.A. Fasching, P.A. Fortner, R.T. García, M.J. Goodman, M.T. Goode, E.L. Gronwald, J. Huntsman, D.G. Karlan, B.Y. Kelemen, L.E. Kommoss, S. Le, N.D. Martin, S.G. Menon, U. Modugno, F. Pharoah, P.D. Schildkraut, J.M. Sieh, W. Staebler, A. Sundfeldt, K. Swerdlow, A.J. Ramus, S.J. Brenton, J.D (2023) p53 and ovarian carcinoma survival: an Ovarian Tumor Tissue Analysis consortium study.. Show Abstract full text

Our objective was to test whether p53 expression status is associated with survival for women diagnosed with the most common ovarian carcinoma histotypes (high-grade serous carcinoma [HGSC], endometrioid carcinoma [EC], and clear cell carcinoma [CCC]) using a large multi-institutional cohort from the Ovarian Tumor Tissue Analysis (OTTA) consortium. p53 expression was assessed on 6,678 cases represented on tissue microarrays from 25 participating OTTA study sites using a previously validated immunohistochemical (IHC) assay as a surrogate for the presence and functional effect of TP53 mutations. Three abnormal expression patterns (overexpression, complete absence, and cytoplasmic) and the normal (wild type) pattern were recorded. Survival analyses were performed by histotype. The frequency of abnormal p53 expression was 93.4% (4,630/4,957) in HGSC compared to 11.9% (116/973) in EC and 11.5% (86/748) in CCC. In HGSC, there were no differences in overall survival across the abnormal p53 expression patterns. However, in EC and CCC, abnormal p53 expression was associated with an increased risk of death for women diagnosed with EC in multivariate analysis compared to normal p53 as the reference (hazard ratio [HR] = 2.18, 95% confidence interval [CI] 1.36-3.47, p = 0.0011) and with CCC (HR = 1.57, 95% CI 1.11-2.22, p = 0.012). Abnormal p53 was also associated with shorter overall survival in The International Federation of Gynecology and Obstetrics stage I/II EC and CCC. Our study provides further evidence that functional groups of TP53 mutations assessed by abnormal surrogate p53 IHC patterns are not associated with survival in HGSC. In contrast, we validate that abnormal p53 IHC is a strong independent prognostic marker for EC and demonstrate for the first time an independent prognostic association of abnormal p53 IHC with overall survival in patients with CCC.

Kang, E.-.Y. Weir, A. Meagher, N.S. Farrington, K. Nelson, G.S. Ghatage, P. Lee, C.-.H. Riggan, M.J. Bolithon, A. Popovic, G. Leung, B. Tang, K. Lambie, N. Millstein, J. Alsop, J. Anglesio, M.S. Ataseven, B. Barlow, E. Beckmann, M.W. Berger, J. Bisinotto, C. Bösmüller, H. Boros, J. Brand, A.H. Brooks-Wilson, A. Brucker, S.Y. Carney, M.E. Casablanca, Y. Cazorla-Jiménez, A. Cohen, P.A. Conrads, T.P. Cook, L.S. Coulson, P. Courtney-Brooks, M. Cramer, D.W. Crowe, P. Cunningham, J.M. Cybulski, C. Darcy, K.M. El-Bahrawy, M.A. Elishaev, E. Erber, R. Farrell, R. Fereday, S. Fischer, A. García, M.J. Gayther, S.A. Gentry-Maharaj, A. Gilks, C.B. AOCS Group, . Grube, M. Harnett, P.R. Harrington, S.P. Harter, P. Hartmann, A. Hecht, J.L. Heikaus, S. Hein, A. Heitz, F. Hendley, J. Hernandez, B.Y. Polo, S.H. Heublein, S. Hirasawa, A. Høgdall, E. Høgdall, C.K. Horlings, H.M. Huntsman, D.G. Huzarski, T. Jewell, A. Jimenez-Linan, M. Jones, M.E. Kaufmann, S.H. Kennedy, C.J. Khabele, D. Kommoss, F.K.F. Kruitwagen, R.F.P.M. Lambrechts, D. Le, N.D. Lener, M. Lester, J. Leung, Y. Linder, A. Loverix, L. Lubiński, J. Madan, R. Maxwell, G.L. Modugno, F. Neuhausen, S.L. Olawaiye, A. Olbrecht, S. Orsulic, S. Palacios, J. Pearce, C.L. Pike, M.C. Quinn, C.M. Mohan, G.R. Rodríguez-Antona, C. Ruebner, M. Ryan, A. Salfinger, S.G. Sasamoto, N. Schildkraut, J.M. Schoemaker, M.J. Shah, M. Sharma, R. Shvetsov, Y.B. Singh, N. Sonke, G.S. Steele, L. Stewart, C.J.R. Sundfeldt, K. Swerdlow, A.J. Talhouk, A. Tan, A. Taylor, S.E. Terry, K.L. Tołoczko, A. Traficante, N. Van de Vijver, K.K. van der Aa, M.A. Van Gorp, T. Van Nieuwenhuysen, E. van-Wagensveld, L. Vergote, I. Vierkant, R.A. Wang, C. Wilkens, L.R. Winham, S.J. Wu, A.H. Benitez, J. Berchuck, A. Candido Dos Reis, F.J. DeFazio, A. Fasching, P.A. Goode, E.L. Goodman, M.T. Gronwald, J. Karlan, B.Y. Kommoss, S. Menon, U. Sinn, H.-.P. Staebler, A. Brenton, J.D. Bowtell, D.D. Pharoah, P.D.P. Ramus, S.J. Köbel, M (2023) CCNE1 and survival of patients with tubo-ovarian high-grade serous carcinoma: An Ovarian Tumor Tissue Analysis consortium study.. Show Abstract full text

<h4>Background</h4>Cyclin E1 (CCNE1) is a potential predictive marker and therapeutic target in tubo-ovarian high-grade serous carcinoma (HGSC). Smaller studies have revealed unfavorable associations for CCNE1 amplification and CCNE1 overexpression with survival, but to date no large-scale, histotype-specific validation has been performed. The hypothesis was that high-level amplification of CCNE1 and CCNE1 overexpression, as well as a combination of the two, are linked to shorter overall survival in HGSC.<h4>Methods</h4>Within the Ovarian Tumor Tissue Analysis consortium, amplification status and protein level in 3029 HGSC cases and mRNA expression in 2419 samples were investigated.<h4>Results</h4>High-level amplification (>8 copies by chromogenic in situ hybridization) was found in 8.6% of HGSC and overexpression (>60% with at least 5% demonstrating strong intensity by immunohistochemistry) was found in 22.4%. CCNE1 high-level amplification and overexpression both were linked to shorter overall survival in multivariate survival analysis adjusted for age and stage, with hazard stratification by study (hazard ratio [HR], 1.26; 95% CI, 1.08-1.47, p = .034, and HR, 1.18; 95% CI, 1.05-1.32, p = .015, respectively). This was also true for cases with combined high-level amplification/overexpression (HR, 1.26; 95% CI, 1.09-1.47, p = .033). CCNE1 mRNA expression was not associated with overall survival (HR, 1.00 per 1-SD increase; 95% CI, 0.94-1.06; p = .58). CCNE1 high-level amplification is mutually exclusive with the presence of germline BRCA1/2 pathogenic variants and shows an inverse association to RB1 loss.<h4>Conclusion</h4>This study provides large-scale validation that CCNE1 high-level amplification is associated with shorter survival, supporting its utility as a prognostic biomarker in HGSC.

Chen, H. Fan, S. Stone, J. Thompson, D.J. Douglas, J. Li, S. Scott, C. Bolla, M.K. Wang, Q. Dennis, J. Michailidou, K. Li, C. Peters, U. Hopper, J.L. Southey, M.C. Nguyen-Dumont, T. Nguyen, T.L. Fasching, P.A. Behrens, A. Cadby, G. Murphy, R.A. Aronson, K. Howell, A. Astley, S. Couch, F. Olson, J. Milne, R.L. Giles, G.G. Haiman, C.A. Maskarinec, G. Winham, S. John, E.M. Kurian, A. Eliassen, H. Andrulis, I. Evans, D.G. Newman, W.G. Hall, P. Czene, K. Swerdlow, A. Jones, M. Pollan, M. Fernandez-Navarro, P. McConnell, D.S. Kristensen, V.N. NBCS Investigators, . Rothstein, J.H. Wang, P. Habel, L.A. Sieh, W. Dunning, A.M. Pharoah, P.D.P. Easton, D.F. Gierach, G.L. Tamimi, R.M. Vachon, C.M. Lindström, S (2022) Genome-wide and transcriptome-wide association studies of mammographic density phenotypes reveal novel loci.. Show Abstract full text

<h4>Background</h4>Mammographic density (MD) phenotypes, including percent density (PMD), area of dense tissue (DA), and area of non-dense tissue (NDA), are associated with breast cancer risk. Twin studies suggest that MD phenotypes are highly heritable. However, only a small proportion of their variance is explained by identified genetic variants.<h4>Methods</h4>We conducted a genome-wide association study, as well as a transcriptome-wide association study (TWAS), of age- and BMI-adjusted DA, NDA, and PMD in up to 27,900 European-ancestry women from the MODE/BCAC consortia.<h4>Results</h4>We identified 28 genome-wide significant loci for MD phenotypes, including nine novel signals (5q11.2, 5q14.1, 5q31.1, 5q33.3, 5q35.1, 7p11.2, 8q24.13, 12p11.2, 16q12.2). Further, 45% of all known breast cancer SNPs were associated with at least one MD phenotype at p < 0.05. TWAS further identified two novel genes (SHOX2 and CRISPLD2) whose genetically predicted expression was significantly associated with MD phenotypes.<h4>Conclusions</h4>Our findings provided novel insight into the genetic background of MD phenotypes, and further demonstrated their shared genetic basis with breast cancer.

Milne, R.L. Kuchenbaecker, K.B. Michailidou, K. Beesley, J. Kar, S. Lindström, S. Hui, S. Lemaçon, A. Soucy, P. Dennis, J. Jiang, X. Rostamianfar, A. Finucane, H. Bolla, M.K. McGuffog, L. Wang, Q. Aalfs, C.M. ABCTB Investigators, . Adams, M. Adlard, J. Agata, S. Ahmed, S. Ahsan, H. Aittomäki, K. Al-Ejeh, F. Allen, J. Ambrosone, C.B. Amos, C.I. Andrulis, I.L. Anton-Culver, H. Antonenkova, N.N. Arndt, V. Arnold, N. Aronson, K.J. Auber, B. Auer, P.L. Ausems, M.G.E.M. Azzollini, J. Bacot, F. Balmaña, J. Barile, M. Barjhoux, L. Barkardottir, R.B. Barrdahl, M. Barnes, D. Barrowdale, D. Baynes, C. Beckmann, M.W. Benitez, J. Bermisheva, M. Bernstein, L. Bignon, Y.-.J. Blazer, K.R. Blok, M.J. Blomqvist, C. Blot, W. Bobolis, K. Boeckx, B. Bogdanova, N.V. Bojesen, A. Bojesen, S.E. Bonanni, B. Børresen-Dale, A.-.L. Bozsik, A. Bradbury, A.R. Brand, J.S. Brauch, H. Brenner, H. Bressac-de Paillerets, B. Brewer, C. Brinton, L. Broberg, P. Brooks-Wilson, A. Brunet, J. Brüning, T. Burwinkel, B. Buys, S.S. Byun, J. Cai, Q. Caldés, T. Caligo, M.A. Campbell, I. Canzian, F. Caron, O. Carracedo, A. Carter, B.D. Castelao, J.E. Castera, L. Caux-Moncoutier, V. Chan, S.B. Chang-Claude, J. Chanock, S.J. Chen, X. Cheng, T.-.Y.D. Chiquette, J. Christiansen, H. Claes, K.B.M. Clarke, C.L. Conner, T. Conroy, D.M. Cook, J. Cordina-Duverger, E. Cornelissen, S. Coupier, I. Cox, A. Cox, D.G. Cross, S.S. Cuk, K. Cunningham, J.M. Czene, K. Daly, M.B. Damiola, F. Darabi, H. Davidson, R. De Leeneer, K. Devilee, P. Dicks, E. Diez, O. Ding, Y.C. Ditsch, N. Doheny, K.F. Domchek, S.M. Dorfling, C.M. Dörk, T. Dos-Santos-Silva, I. Dubois, S. Dugué, P.-.A. Dumont, M. Dunning, A.M. Durcan, L. Dwek, M. Dworniczak, B. Eccles, D. Eeles, R. Ehrencrona, H. Eilber, U. Ejlertsen, B. Ekici, A.B. Eliassen, A.H. EMBRACE, . Engel, C. Eriksson, M. Fachal, L. Faivre, L. Fasching, P.A. Faust, U. Figueroa, J. Flesch-Janys, D. Fletcher, O. Flyger, H. Foulkes, W.D. Friedman, E. Fritschi, L. Frost, D. Gabrielson, M. Gaddam, P. Gammon, M.D. Ganz, P.A. Gapstur, S.M. Garber, J. Garcia-Barberan, V. García-Sáenz, J.A. Gaudet, M.M. Gauthier-Villars, M. Gehrig, A. GEMO Study Collaborators, . Georgoulias, V. Gerdes, A.-.M. Giles, G.G. Glendon, G. Godwin, A.K. Goldberg, M.S. Goldgar, D.E. González-Neira, A. Goodfellow, P. Greene, M.H. Alnæs, G.I.G. Grip, M. Gronwald, J. Grundy, A. Gschwantler-Kaulich, D. Guénel, P. Guo, Q. Haeberle, L. Hahnen, E. Haiman, C.A. Håkansson, N. Hallberg, E. Hamann, U. Hamel, N. Hankinson, S. Hansen, T.V.O. Harrington, P. Hart, S.N. Hartikainen, J.M. Healey, C.S. HEBON, . Hein, A. Helbig, S. Henderson, A. Heyworth, J. Hicks, B. Hillemanns, P. Hodgson, S. Hogervorst, F.B. Hollestelle, A. Hooning, M.J. Hoover, B. Hopper, J.L. Hu, C. Huang, G. Hulick, P.J. Humphreys, K. Hunter, D.J. Imyanitov, E.N. Isaacs, C. Iwasaki, M. Izatt, L. Jakubowska, A. James, P. Janavicius, R. Janni, W. Jensen, U.B. John, E.M. Johnson, N. Jones, K. Jones, M. Jukkola-Vuorinen, A. Kaaks, R. Kabisch, M. Kaczmarek, K. Kang, D. Kast, K. kConFab/AOCS Investigators, . Keeman, R. Kerin, M.J. Kets, C.M. Keupers, M. Khan, S. Khusnutdinova, E. Kiiski, J.I. Kim, S.-.W. Knight, J.A. Konstantopoulou, I. Kosma, V.-.M. Kristensen, V.N. Kruse, T.A. Kwong, A. Lænkholm, A.-.V. Laitman, Y. Lalloo, F. Lambrechts, D. Landsman, K. Lasset, C. Lazaro, C. Le Marchand, L. Lecarpentier, J. Lee, A. Lee, E. Lee, J.W. Lee, M.H. Lejbkowicz, F. Lesueur, F. Li, J. Lilyquist, J. Lincoln, A. Lindblom, A. Lissowska, J. Lo, W.-.Y. Loibl, S. Long, J. Loud, J.T. Lubinski, J. Luccarini, C. Lush, M. MacInnis, R.J. Maishman, T. Makalic, E. Kostovska, I.M. Malone, K.E. Manoukian, S. Manson, J.E. Margolin, S. Martens, J.W.M. Martinez, M.E. Matsuo, K. Mavroudis, D. Mazoyer, S. McLean, C. Meijers-Heijboer, H. Menéndez, P. Meyer, J. Miao, H. Miller, A. Miller, N. Mitchell, G. Montagna, M. Muir, K. Mulligan, A.M. Mulot, C. Nadesan, S. Nathanson, K.L. NBSC Collaborators, . Neuhausen, S.L. Nevanlinna, H. Nevelsteen, I. Niederacher, D. Nielsen, S.F. Nordestgaard, B.G. Norman, A. Nussbaum, R.L. Olah, E. Olopade, O.I. Olson, J.E. Olswold, C. Ong, K.-.R. Oosterwijk, J.C. Orr, N. Osorio, A. Pankratz, V.S. Papi, L. Park-Simon, T.-.W. Paulsson-Karlsson, Y. Lloyd, R. Pedersen, I.S. Peissel, B. Peixoto, A. Perez, J.I.A. Peterlongo, P. Peto, J. Pfeiler, G. Phelan, C.M. Pinchev, M. Plaseska-Karanfilska, D. Poppe, B. Porteous, M.E. Prentice, R. Presneau, N. Prokofieva, D. Pugh, E. Pujana, M.A. Pylkäs, K. Rack, B. Radice, P. Rahman, N. Rantala, J. Rappaport-Fuerhauser, C. Rennert, G. Rennert, H.S. Rhenius, V. Rhiem, K. Richardson, A. Rodriguez, G.C. Romero, A. Romm, J. Rookus, M.A. Rudolph, A. Ruediger, T. Saloustros, E. Sanders, J. Sandler, D.P. Sangrajrang, S. Sawyer, E.J. Schmidt, D.F. Schoemaker, M.J. Schumacher, F. Schürmann, P. Schwentner, L. Scott, C. Scott, R.J. Seal, S. Senter, L. Seynaeve, C. Shah, M. Sharma, P. Shen, C.-.Y. Sheng, X. Shimelis, H. Shrubsole, M.J. Shu, X.-.O. Side, L.E. Singer, C.F. Sohn, C. Southey, M.C. Spinelli, J.J. Spurdle, A.B. Stegmaier, C. Stoppa-Lyonnet, D. Sukiennicki, G. Surowy, H. Sutter, C. Swerdlow, A. Szabo, C.I. Tamimi, R.M. Tan, Y.Y. Taylor, J.A. Tejada, M.-.I. Tengström, M. Teo, S.H. Terry, M.B. Tessier, D.C. Teulé, A. Thöne, K. Thull, D.L. Tibiletti, M.G. Tihomirova, L. Tischkowitz, M. Toland, A.E. Tollenaar, R.A.E.M. Tomlinson, I. Tong, L. Torres, D. Tranchant, M. Truong, T. Tucker, K. Tung, N. Tyrer, J. Ulmer, H.-.U. Vachon, C. van Asperen, C.J. Van Den Berg, D. van den Ouweland, A.M.W. van Rensburg, E.J. Varesco, L. Varon-Mateeva, R. Vega, A. Viel, A. Vijai, J. Vincent, D. Vollenweider, J. Walker, L. Wang, Z. Wang-Gohrke, S. Wappenschmidt, B. Weinberg, C.R. Weitzel, J.N. Wendt, C. Wesseling, J. Whittemore, A.S. Wijnen, J.T. Willett, W. Winqvist, R. Wolk, A. Wu, A.H. Xia, L. Yang, X.R. Yannoukakos, D. Zaffaroni, D. Zheng, W. Zhu, B. Ziogas, A. Ziv, E. Zorn, K.K. Gago-Dominguez, M. Mannermaa, A. Olsson, H. Teixeira, M.R. Stone, J. Offit, K. Ottini, L. Park, S.K. Thomassen, M. Hall, P. Meindl, A. Schmutzler, R.K. Droit, A. Bader, G.D. Pharoah, P.D.P. Couch, F.J. Easton, D.F. Kraft, P. Chenevix-Trench, G. García-Closas, M. Schmidt, M.K. Antoniou, A.C. Simard, J (2017) Identification of ten variants associated with risk of estrogen-receptor-negative breast cancer.. Show Abstract full text

Most common breast cancer susceptibility variants have been identified through genome-wide association studies (GWAS) of predominantly estrogen receptor (ER)-positive disease. We conducted a GWAS using 21,468 ER-negative cases and 100,594 controls combined with 18,908 BRCA1 mutation carriers (9,414 with breast cancer), all of European origin. We identified independent associations at P < 5 × 10<sup>-8</sup> with ten variants at nine new loci. At P < 0.05, we replicated associations with 10 of 11 variants previously reported in ER-negative disease or BRCA1 mutation carrier GWAS and observed consistent associations with ER-negative disease for 105 susceptibility variants identified by other studies. These 125 variants explain approximately 16% of the familial risk of this breast cancer subtype. There was high genetic correlation (0.72) between risk of ER-negative breast cancer and breast cancer risk for BRCA1 mutation carriers. These findings may lead to improved risk prediction and inform further fine-mapping and functional work to better understand the biological basis of ER-negative breast cancer.

Morra, A. Schreurs, M.A.C. Andrulis, I.L. Anton-Culver, H. Augustinsson, A. Beckmann, M.W. Behrens, S. Bojesen, S.E. Bojesen, S.E. Bolla, M.K. Brauch, H. Broeks, A. Buys, S.S. Camp, N.J. Castelao, J.E. Cessna, M.H. Chang-Claude, J. Chung, W.K. NBCS Collaborators, . Colonna, S.V. Couch, F.J. Cox, A. Cross, S.S. Czene, K. Daly, M.B. Dennis, J. Devilee, P. Dörk, T. Dunning, A.M. Dwek, M. Easton, D.F. Eccles, D.M. Eriksson, M. Evans, D.G. Fasching, P.A. Fehm, T.N. Figueroa, J.D. Flyger, H. Gabrielson, M. Gago-Dominguez, M. García-Closas, M. García-Sáenz, J.A. Genkinger, J. Grassmann, F. Gündert, M. Hahnen, E. Haiman, C.A. Hamann, U. Harrington, P.A. Hartikainen, J.M. Hoppe, R. Hopper, J.L. Houlston, R.S. Howell, A. Howell, A. ABCTB Investigators, . kConFab Investigators, . Jakubowska, A. Janni, W. Jernström, H. John, E.M. Johnson, N. Jones, M.E. Kristensen, V.N. Kurian, A.W. Lambrechts, D. Le Marchand, L. Lindblom, A. Lubiński, J. Lux, M.P. Mannermaa, A. Mavroudis, D. Mulligan, A.M. Muranen, T.A. Nevanlinna, H. Nevelsteen, I. Neven, P. Newman, W.G. Obi, N. Offit, K. Olshan, A.F. Park-Simon, T.-.W. Patel, A.V. Peterlongo, P. Phillips, K.-.A. Plaseska-Karanfilska, D. Polley, E.C. Presneau, N. Pylkäs, K. Rack, B. Radice, P. Rashid, M.U. Rhenius, V. Robson, M. Romero, A. Saloustros, E. Sawyer, E.J. Schmutzler, R.K. Schuetze, S. Scott, C. Shah, M. Smichkoska, S. Southey, M.C. Tapper, W.J. Teras, L.R. Tollenaar, R.A.E.M. Tomczyk, K. Tomlinson, I. Troester, M.A. Vachon, C.M. van Veen, E.M. Wang, Q. Wendt, C. Wildiers, H. Winqvist, R. Ziogas, A. Hall, P. Pharoah, P.D.P. Adank, M.A. Hollestelle, A. Schmidt, M.K. Hooning, M.J (2023) Association of the CHEK2 c.1100delC variant, radiotherapy, and systemic treatment with contralateral breast cancer risk and breast cancer-specific survival.. Show Abstract full text

<h4>Background</h4>Breast cancer (BC) patients with a germline CHEK2 c.1100delC variant have an increased risk of contralateral BC (CBC) and worse BC-specific survival (BCSS) compared to non-carriers.<h4>Aim</h4>To assessed the associations of CHEK2 c.1100delC, radiotherapy, and systemic treatment with CBC risk and BCSS.<h4>Methods</h4>Analyses were based on 82,701 women diagnosed with a first primary invasive BC including 963 CHEK2 c.1100delC carriers; median follow-up was 9.1 years. Differential associations with treatment by CHEK2 c.1100delC status were tested by including interaction terms in a multivariable Cox regression model. A multi-state model was used for further insight into the relation between CHEK2 c.1100delC status, treatment, CBC risk and death.<h4>Results</h4>There was no evidence for differential associations of therapy with CBC risk by CHEK2 c.1100delC status. The strongest association with reduced CBC risk was observed for the combination of chemotherapy and endocrine therapy [HR (95% CI): 0.66 (0.55-0.78)]. No association was observed with radiotherapy. Results from the multi-state model showed shorter BCSS for CHEK2 c.1100delC carriers versus non-carriers also after accounting for CBC occurrence [HR (95% CI): 1.30 (1.09-1.56)].<h4>Conclusion</h4>Systemic therapy was associated with reduced CBC risk irrespective of CHEK2 c.1100delC status. Moreover, CHEK2 c.1100delC carriers had shorter BCSS, which appears not to be fully explained by their CBC risk.

Levi, H. Carmi, S. Rosset, S. Yerushalmi, R. Zick, A. Yablonski-Peretz, T. BCAC Consortium, . Wang, Q. Bolla, M.K. Dennis, J. Michailidou, K. Lush, M. Ahearn, T. Andrulis, I.L. Anton-Culver, H. Antoniou, A.C. Arndt, V. Augustinsson, A. Auvinen, P. Beane Freeman, L. Beckmann, M. Behrens, S. Bermisheva, M. Bodelon, C. Bogdanova, N.V. Bojesen, S.E. Brenner, H. Byers, H. Camp, N. Castelao, J. Chang-Claude, J. Chirlaque, M.-.D. Chung, W. Clarke, C. NBCS Collaborators, . Collee, M.J. Colonna, S. CTS Consortium, . Couch, F. Cox, A. Cross, S.S. Czene, K. Daly, M. Devilee, P. Dork, T. Dossus, L. Eccles, D.M. Eliassen, A.H. Eriksson, M. Evans, G. Fasching, P. Fletcher, O. Flyger, H. Fritschi, L. Gabrielson, M. Gago-Dominguez, M. García-Closas, M. Garcia-Saenz, J.A. Genkinger, J. Giles, G.G. Goldberg, M. Guénel, P. Hall, P. Hamann, U. He, W. Hillemanns, P. Hollestelle, A. Hoppe, R. Hopper, J. ABCTB Investigators, . Jakovchevska, S. Jakubowska, A. Jernström, H. John, E. Johnson, N. Jones, M. Vijai, J. Kaaks, R. Khusnutdinova, E. Kitahara, C. Koutros, S. Kristensen, V. Kurian, A.W. Lacey, J. Lambrechts, D. Le Marchand, L. Lejbkowicz, F. Lindblom, A. Loibl, S. Lori, A. Lubinski, J. Mannermaa, A. Manoochehri, M. Mavroudis, D. Menon, U. Mulligan, A. Murphy, R. Nevelsteen, I. Newman, W.G. Obi, N. O'Brien, K. Offit, K. Olshan, A. Plaseska-Karanfilska, D. Olson, J. Panico, S. Park-Simon, T.-.W. Patel, A. Peterlongo, P. Rack, B. Radice, P. Rennert, G. Rhenius, V. Romero, A. Saloustros, E. Sandler, D. Schmidt, M.K. Schwentner, L. Shah, M. Sharma, P. Simard, J. Southey, M. Stone, J. Tapper, W.J. Taylor, J. Teras, L. Toland, A.E. Troester, M. Truong, T. van der Kolk, L.E. Weinberg, C. Wendt, C. Yang, X.R. Zheng, W. Ziogas, A. Dunning, A.M. Pharoah, P. Easton, D.F. Ben-Sachar, S. Elefant, N. Shamir, R. Elkon, R (2023) Evaluation of European-based polygenic risk score for breast cancer in Ashkenazi Jewish women in Israel.. Show Abstract full text

<h4>Background</h4>Polygenic risk score (PRS), calculated based on genome-wide association studies (GWASs), can improve breast cancer (BC) risk assessment. To date, most BC GWASs have been performed in individuals of European (EUR) ancestry, and the generalisation of EUR-based PRS to other populations is a major challenge. In this study, we examined the performance of EUR-based BC PRS models in Ashkenazi Jewish (AJ) women.<h4>Methods</h4>We generated PRSs based on data on EUR women from the Breast Cancer Association Consortium (BCAC). We tested the performance of the PRSs in a cohort of 2161 AJ women from Israel (1437 cases and 724 controls) from BCAC (BCAC cohort from Israel (BCAC-IL)). In addition, we tested the performance of these EUR-based BC PRSs, as well as the established 313-SNP EUR BC PRS, in an independent cohort of 181 AJ women from Hadassah Medical Center (HMC) in Israel.<h4>Results</h4>In the BCAC-IL cohort, the highest OR per 1 SD was 1.56 (±0.09). The OR for AJ women at the top 10% of the PRS distribution compared with the middle quintile was 2.10 (±0.24). In the HMC cohort, the OR per 1 SD of the EUR-based PRS that performed best in the BCAC-IL cohort was 1.58±0.27. The OR per 1 SD of the commonly used 313-SNP BC PRS was 1.64 (±0.28).<h4>Conclusions</h4>Extant EUR GWAS data can be used for generating PRSs that identify AJ women with markedly elevated risk of BC and therefore hold promise for improving BC risk assessment in AJ women.

Saner, F.A.M. Takahashi, K. Budden, T. Pandey, A. Ariyaratne, D. Zwimpfer, T.A. Meagher, N.S. Fereday, S. Twomey, L. Pishas, K.I. Hoang, T. Bolithon, A. Traficante, N. Australian Ovarian Cancer Study Group, . Alsop, K. Christie, E.L. Kang, E.-.Y. Nelson, G.S. Ghatage, P. Lee, C.-.H. Riggan, M.J. Alsop, J. Beckmann, M.W. Boros, J. Brand, A.H. Brooks-Wilson, A. Carney, M.E. Coulson, P. Courtney-Brooks, M. Cushing-Haugen, K.L. Cybulski, C. El-Bahrawy, M.A. Elishaev, E. Erber, R. Gayther, S.A. Gentry-Maharaj, A. Gilks, C.B. Harnett, P.R. Harris, H.R. Hartmann, A. Hein, A. Hendley, J. Hernandez, B.Y. Jakubowska, A. Jimenez-Linan, M. Jones, M.E. Kaufmann, S.H. Kennedy, C.J. Kluz, T. Koziak, J.M. Kristjansdottir, B. Le, N.D. Lener, M. Lester, J. Lubiński, J. Mateoiu, C. Orsulic, S. Ruebner, M. Schoemaker, M.J. Shah, M. Sharma, R. Sherman, M.E. Shvetsov, Y.B. Soong, T.R. Steed, H. Sukumvanich, P. Talhouk, A. Taylor, S.E. Vierkant, R.A. Wang, C. Widschwendter, M. Wilkens, L.R. Winham, S.J. Anglesio, M.S. Berchuck, A. Brenton, J.D. Campbell, I. Cook, L.S. Doherty, J.A. Fasching, P.A. Fortner, R.T. Goodman, M.T. Gronwald, J. Huntsman, D.G. Karlan, B.Y. Kelemen, L.E. Menon, U. Modugno, F. Pharoah, P.D.P. Schildkraut, J.M. Sundfeldt, K. Swerdlow, A.J. Goode, E.L. DeFazio, A. Köbel, M. Ramus, S.J. Bowtell, D.D.L. Garsed, D.W (2024) Concurrent RB1 Loss and BRCA Deficiency Predicts Enhanced Immunologic Response and Long-term Survival in Tubo-ovarian High-grade Serous Carcinoma.. Show Abstract full text

<h4>Purpose</h4>The purpose of this study was to evaluate RB1 expression and survival across ovarian carcinoma histotypes and how co-occurrence of BRCA1 or BRCA2 (BRCA) alterations and RB1 loss influences survival in tubo-ovarian high-grade serous carcinoma (HGSC).<h4>Experimental design</h4>RB1 protein expression was classified by immunohistochemistry in ovarian carcinomas of 7,436 patients from the Ovarian Tumor Tissue Analysis consortium. We examined RB1 expression and germline BRCA status in a subset of 1,134 HGSC, and related genotype to overall survival (OS), tumor-infiltrating CD8+ lymphocytes, and transcriptomic subtypes. Using CRISPR-Cas9, we deleted RB1 in HGSC cells with and without BRCA1 alterations to model co-loss with treatment response. We performed whole-genome and transcriptome data analyses on 126 patients with primary HGSC to characterize tumors with concurrent BRCA deficiency and RB1 loss.<h4>Results</h4>RB1 loss was associated with longer OS in HGSC but with poorer prognosis in endometrioid ovarian carcinoma. Patients with HGSC harboring both RB1 loss and pathogenic germline BRCA variants had superior OS compared with patients with either alteration alone, and their median OS was three times longer than those without pathogenic BRCA variants and retained RB1 expression (9.3 vs. 3.1 years). Enhanced sensitivity to cisplatin and paclitaxel was seen in BRCA1-altered cells with RB1 knockout. Combined RB1 loss and BRCA deficiency correlated with transcriptional markers of enhanced IFN response, cell-cycle deregulation, and reduced epithelial-mesenchymal transition. CD8+ lymphocytes were most prevalent in BRCA-deficient HGSC with co-loss of RB1.<h4>Conclusions</h4>Co-occurrence of RB1 loss and BRCA deficiency was associated with exceptionally long survival in patients with HGSC, potentially due to better treatment response and immune stimulation.

Rowlands, C.F. Garrett, A. Allen, S. Durkie, M. Burghel, G.J. Robinson, R. Callaway, A. Field, J. Frugtniet, B. Palmer-Smith, S. Grant, J. Pagan, J. McDevitt, T. McVeigh, T.P. Hanson, H. Whiffin, N. Jones, M. Turnbull, C. CanVIG-UK, (2024) The PS4-likelihood ratio calculator: flexible allocation of evidence weighting for case-control data in variant classification.. Show Abstract full text

<h4>Background</h4>The 2015 American College of Medical Genetics/Association of Molecular Pathology (ACMG/AMP) variant classification framework specifies that case-control observations can be scored as 'strong' evidence (PS4) towards pathogenicity.<h4>Methods</h4>We developed the PS4-likelihood ratio calculator (PS4-LRCalc) for quantitative evidence assignment based on the observed variant frequencies in cases and controls. Binomial likelihoods are computed for two models, each defined by prespecified OR thresholds. Model 1 represents the hypothesis of association between variant and phenotype (eg, OR≥5) and model 2 represents the hypothesis of non-association (eg, OR≤1).<h4>Results</h4>PS4-LRCalc enables continuous quantitation of evidence for variant classification expressed as a likelihood ratio (LR), which can be log-converted into log LR (evidence points). Using PS4-LRCalc, observed data can be used to quantify evidence towards either pathogenicity or benignity. Variants can also be evaluated against models of different penetrance. The approach is applicable to balanced data sets generated for more common phenotypes and smaller data sets more typical in very rare disease variant evaluation.<h4>Conclusion</h4>PS4-LRCalc enables flexible evidence quantitation on a continuous scale for observed case-control data. The converted LR is amenable to incorporation into the now widely used 2018 updated Bayesian ACMG/AMP framework.

Von Holle, A. Adami, H.-.O. Baglietto, L. Berrington de Gonzalez, A. Bertrand, K.A. Blot, W. Chen, Y. DeHart, J.C. Dossus, L. Eliassen, A.H. Fournier, A. Garcia-Closas, M. Giles, G. Guevara, M. Hankinson, S.E. Heath, A. Jones, M.E. Joshu, C.E. Kaaks, R. Kirsh, V.A. Kitahara, C.M. Koh, W.-.P. Linet, M.S. Park, H.L. Masala, G. Mellemkjaer, L. Milne, R.L. O'Brien, K.M. Palmer, J.R. Riboli, E. Rohan, T.E. Shrubsole, M.J. Sund, M. Tamimi, R. Tin Tin, S. Visvanathan, K. Vermeulen, R.C. Weiderpass, E. Willett, W.C. Yuan, J.-.M. Zeleniuch-Jacquotte, A. Nichols, H.B. Sandler, D.P. Swerdlow, A.J. Schoemaker, M.J. Weinberg, C.R (2024) BMI and breast cancer risk around age at menopause.. Show Abstract full text

<h4>Background</h4>A high body mass index (BMI, kg/m<sup>2</sup>) is associated with decreased risk of breast cancer before menopause, but increased risk after menopause. Exactly when this reversal occurs in relation to menopause is unclear. Locating that change point could provide insight into the role of adiposity in breast cancer etiology.<h4>Methods</h4>We examined the association between BMI and breast cancer risk in the Premenopausal Breast Cancer Collaborative Group, from age 45 up to breast cancer diagnosis, loss to follow-up, death, or age 55, whichever came first. Analyses included 609,880 women in 16 prospective studies, including 9956 who developed breast cancer before age 55. We fitted three BMI hazard ratio (HR) models over age-time: constant, linear, or nonlinear (via splines), applying piecewise exponential additive mixed models, with age as the primary time scale. We divided person-time into four strata: premenopause; postmenopause due to natural menopause; postmenopause because of interventional loss of ovarian function (bilateral oophorectomy (BO) or chemotherapy); postmenopause due to hysterectomy without BO. Sensitivity analyses included stratifying by BMI in young adulthood, or excluding women using menopausal hormone therapy.<h4>Results</h4>The constant BMI HR model provided the best fit for all four menopausal status groups. Under this model, the estimated association between a five-unit increment in BMI and breast cancer risk was HR=0.87 (95% CI: 0.85, 0.89) before menopause, HR=1.00 (95% CI: 0.96, 1.04) after natural menopause, HR=0.99 (95% CI: 0.93, 1.05) after interventional loss of ovarian function, and HR=0.88 (95% CI: 0.76, 1.02) after hysterectomy without BO.<h4>Conclusion</h4>The BMI breast cancer HRs remained less than or near one during the 45-55 year age range indicating that the transition to a positive association between BMI and risk occurs after age 55.

Haas, C.B. Chen, H. Harrison, T. Fan, S. Gago-Dominguez, M. Castelao, J.E. Bolla, M.K. Wang, Q. Dennis, J. Michailidou, K. Dunning, A.M. Easton, D.F. Antoniou, A.C. Hall, P. Czene, K. Andrulis, I.L. Mulligan, A.M. Milne, R.L. Fasching, P.A. Haeberle, L. Garcia-Closas, M. Ahearn, T. Gierach, G.L. Haiman, C. Maskarinec, G. Couch, F.J. Olson, J.E. John, E.M. Chenevix-Trench, G. Berrington de Gonzalez, A. Jones, M. Stone, J. Murphy, R. Aronson, K.J. Wernli, K.J. Hsu, L. Vachon, C. Tamimi, R.M. Lindström, S (2024) Disentangling the relationships of body mass index and circulating sex hormone concentrations in mammographic density using Mendelian randomization.. Show Abstract full text

<h4>Purpose</h4>Mammographic density phenotypes, adjusted for age and body mass index (BMI), are strong predictors of breast cancer risk. BMI is associated with mammographic density measures, but the role of circulating sex hormone concentrations is less clear. We investigated the relationship between BMI, circulating sex hormone concentrations, and mammographic density phenotypes using Mendelian randomization (MR).<h4>Methods</h4>We applied two-sample MR approaches to assess the association between genetically predicted circulating concentrations of sex hormones [estradiol, testosterone, sex hormone-binding globulin (SHBG)], BMI, and mammographic density phenotypes (dense and non-dense area). We created instrumental variables from large European ancestry-based genome-wide association studies and applied estimates to mammographic density phenotypes in up to 14,000 women of European ancestry. We performed analyses overall and by menopausal status.<h4>Results</h4>Genetically predicted BMI was positively associated with non-dense area (IVW: β = 1.79; 95% CI = 1.58, 2.00; p = 9.57 × 10<sup>-63</sup>) and inversely associated with dense area (IVW: β = - 0.37; 95% CI = - 0.51,- 0.23; p = 4.7 × 10<sup>-7</sup>). We observed weak evidence for an association of circulating sex hormone concentrations with mammographic density phenotypes, specifically inverse associations between genetically predicted testosterone concentration and dense area (β = - 0.22; 95% CI = - 0.38, - 0.053; p = 0.009) and between genetically predicted estradiol concentration and non-dense area (β =  - 3.32; 95% CI = - 5.83, - 0.82; p = 0.009), although results were not consistent across a range of MR approaches.<h4>Conclusion</h4>Our findings support a positive causal association between BMI and mammographic non-dense area and an inverse association between BMI and dense area. Evidence was weaker and inconsistent for a causal effect of circulating sex hormone concentrations on mammographic density phenotypes. Based on our findings, associations between circulating sex hormone concentrations and mammographic density phenotypes are weak at best.

Dareng, E.O. Coetzee, S.G. Tyrer, J.P. Peng, P.-.C. Rosenow, W. Chen, S. Davis, B.D. Dezem, F.S. Seo, J.-.H. Nameki, R. Reyes, A.L. Aben, K.K.H. Anton-Culver, H. Antonenkova, N.N. Aravantinos, G. Bandera, E.V. Beane Freeman, L.E. Beckmann, M.W. Beeghly-Fadiel, A. Benitez, J. Bernardini, M.Q. Bjorge, L. Black, A. Bogdanova, N.V. Bolton, K.L. Brenton, J.D. Budzilowska, A. Butzow, R. Cai, H. Campbell, I. Cannioto, R. Chang-Claude, J. Chanock, S.J. Chen, K. Chenevix-Trench, G. AOCS Group, . Chiew, Y.-.E. Cook, L.S. DeFazio, A. Dennis, J. Doherty, J.A. Dörk, T. du Bois, A. Dürst, M. Eccles, D.M. Ene, G. Fasching, P.A. Flanagan, J.M. Fortner, R.T. Fostira, F. Gentry-Maharaj, A. Giles, G.G. Goodman, M.T. Gronwald, J. Haiman, C.A. Håkansson, N. Heitz, F. Hildebrandt, M.A.T. Høgdall, E. Høgdall, C.K. Huang, R.-.Y. Jensen, A. Jones, M.E. Kang, D. Karlan, B.Y. Karnezis, A.N. Kelemen, L.E. Kennedy, C.J. Khusnutdinova, E.K. Kiemeney, L.A. Kjaer, S.K. Kupryjanczyk, J. Labrie, M. Lambrechts, D. Larson, M.C. Le, N.D. Lester, J. Li, L. Lubiński, J. Lush, M. Marks, J.R. Matsuo, K. May, T. McLaughlin, J.R. McNeish, I.A. Menon, U. Missmer, S. Modugno, F. Moffitt, M. Monteiro, A.N. Moysich, K.B. Narod, S.A. Nguyen-Dumont, T. Odunsi, K. Olsson, H. Onland-Moret, N.C. Park, S.K. Pejovic, T. Permuth, J.B. Piskorz, A. Prokofyeva, D. Riggan, M.J. Risch, H.A. Rodríguez-Antona, C. Rossing, M.A. Sandler, D.P. Setiawan, V.W. Shan, K. Song, H. Southey, M.C. Steed, H. Sutphen, R. Swerdlow, A.J. Teo, S.H. Terry, K.L. Thompson, P.J. Vestrheim Thomsen, L.C. Titus, L. Trabert, B. Travis, R. Tworoger, S.S. Valen, E. Van Nieuwenhuysen, E. Edwards, D.V. Vierkant, R.A. Webb, P.M. OPAL Study Group, . Weinberg, C.R. Weise, R.M. Wentzensen, N. White, E. Winham, S.J. Wolk, A. Woo, Y.-.L. Wu, A.H. Yan, L. Yannoukakos, D. Zeinomar, N. Zheng, W. Ziogas, A. Berchuck, A. Goode, E.L. Huntsman, D.G. Pearce, C.L. Ramus, S.J. Sellers, T.A. Ovarian Cancer Association Consortium (OCAC), . Freedman, M.L. Lawrenson, K. Schildkraut, J.M. Hazelett, D. Plummer, J.T. Kar, S. Jones, M.R. Pharoah, P.D.P. Gayther, S.A (2024) Integrative multi-omics analyses to identify the genetic and functional mechanisms underlying ovarian cancer risk regions.. Show Abstract full text

To identify credible causal risk variants (CCVs) associated with different histotypes of epithelial ovarian cancer (EOC), we performed genome-wide association analysis for 470,825 genotyped and 10,163,797 imputed SNPs in 25,981 EOC cases and 105,724 controls of European origin. We identified five histotype-specific EOC risk regions (p value <5 × 10<sup>-8</sup>) and confirmed previously reported associations for 27 risk regions. Conditional analyses identified an additional 11 signals independent of the primary signal at six risk regions (p value <10<sup>-5</sup>). Fine mapping identified 4,008 CCVs in these regions, of which 1,452 CCVs were located in ovarian cancer-related chromatin marks with significant enrichment in active enhancers, active promoters, and active regions for CCVs from each EOC histotype. Transcriptome-wide association and colocalization analyses across histotypes using tissue-specific and cross-tissue datasets identified 86 candidate susceptibility genes in known EOC risk regions and 32 genes in 23 additional genomic regions that may represent novel EOC risk loci (false discovery rate <0.05). Finally, by integrating genome-wide HiChIP interactome analysis with transcriptome-wide association study (TWAS), variant effect predictor, transcription factor ChIP-seq, and motifbreakR data, we identified candidate gene-CCV interactions at each locus. This included risk loci where TWAS identified one or more candidate susceptibility genes (e.g., HOXD-AS2, HOXD8, and HOXD3 at 2q31) and other loci where no candidate gene was identified (e.g., MYC and PVT1 at 8q24) by TWAS. In summary, this study describes a functional framework and provides a greater understanding of the biological significance of risk alleles and candidate gene targets at EOC susceptibility loci identified by a genome-wide association study.

Timmins, I.R. Jones, M.E. O'Brien, K.M. Adami, H.-.O. Aune, D. Baglietto, L. Bertrand, K.A. Brantley, K.D. Chen, Y. Clague DeHart, J. Clendenen, T.V. Dossus, L. Eliassen, A.H. Fletcher, O. Fournier, A. Håkansson, N. Hankinson, S.E. Houlston, R.S. Joshu, C.E. Kirsh, V.A. Kitahara, C.M. Koh, W.-.P. Linet, M.S. Park, H.L. Lynch, B.M. May, A.M. Mellemkjær, L. Milne, R.L. Palmer, J.R. Ricceri, F. Rohan, T.E. Ruddy, K.J. Sánchez, M.-.J. Shu, X.-.O. Smith-Byrne, K. Steindorf, K. Sund, M. Vachon, C.M. Vatten, L.J. Visvanathan, K. Weiderpass, E. Willett, W.C. Wolk, A. Yuan, J.-.M. Zheng, W. Nichols, H.B. Sandler, D.P. Swerdlow, A.J. Schoemaker, M.J (2024) International Pooled Analysis of Leisure-Time Physical Activity and Premenopausal Breast Cancer in Women From 19 Cohorts.. Show Abstract full text

PURPOSE: There is strong evidence that leisure-time physical activity is protective against postmenopausal breast cancer risk but the association with premenopausal breast cancer is less clear. The purpose of this study was to examine the association of physical activity with the risk of developing premenopausal breast cancer. METHODS: We pooled individual-level data on self-reported leisure-time physical activity across 19 cohort studies comprising 547,601 premenopausal women, with 10,231 incident cases of breast cancer. Multivariable Cox regression was used to estimate hazard ratios (HRs) and 95% CIs for associations of leisure-time physical activity with breast cancer incidence. HRs for high versus low levels of activity were based on a comparison of risk at the 90th versus 10th percentiles of activity. We assessed the linearity of the relationship and examined subtype-specific associations and effect modification across strata of breast cancer risk factors, including adiposity. RESULTS: Over a median 11.5 years of follow-up (IQR, 8.0-16.1 years), high versus low levels of leisure-time physical activity were associated with a 6% (HR, 0.94 [95% CI, 0.89 to 0.99]) and a 10% (HR, 0.90 [95% CI, 0.85 to 0.95]) reduction in breast cancer risk, before and after adjustment for BMI, respectively. Tests of nonlinearity suggested an approximately linear relationship (Pnonlinearity = .94). The inverse association was particularly strong for human epidermal growth factor receptor 2-enriched breast cancer (HR, 0.57 [95% CI, 0.39 to 0.84]; Phet = .07). Associations did not vary significantly across strata of breast cancer risk factors, including subgroups of adiposity. CONCLUSION: This large, pooled analysis of cohort studies adds to evidence that engagement in higher levels of leisure-time physical activity may lead to reduced premenopausal breast cancer risk.

Middha, P. Wang, X. Behrens, S. Bolla, M.K. Wang, Q. Dennis, J. Michailidou, K. Ahearn, T.U. Andrulis, I.L. Anton-Culver, H. Arndt, V. Aronson, K.J. Auer, P.L. Augustinsson, A. Baert, T. Freeman, L.E.B. Becher, H. Beckmann, M.W. Benitez, J. Bojesen, S.E. Brauch, H. Brenner, H. Brooks-Wilson, A. Campa, D. Canzian, F. Carracedo, A. Castelao, J.E. Chanock, S.J. Chenevix-Trench, G. CTS Consortium, . Cordina-Duverger, E. Couch, F.J. Cox, A. Cross, S.S. Czene, K. Dossus, L. Dugué, P.-.A. Eliassen, A.H. Eriksson, M. Evans, D.G. Fasching, P.A. Figueroa, J.D. Fletcher, O. Flyger, H. Gabrielson, M. Gago-Dominguez, M. Giles, G.G. González-Neira, A. Grassmann, F. Grundy, A. Guénel, P. Haiman, C.A. Håkansson, N. Hall, P. Hamann, U. Hankinson, S.E. Harkness, E.F. Holleczek, B. Hoppe, R. Hopper, J.L. Houlston, R.S. Howell, A. Howell, A. Hunter, D.J. Ingvar, C. ABCTB Investigators, . kConFab Investigators, . Isaksson, K. Jernström, H. John, E.M. Jones, M.E. Kaaks, R. Keeman, R. Kitahara, C.M. Ko, Y.-.D. Koutros, S. Kurian, A.W. Lacey, J.V. Lambrechts, D. Larson, N.L. Larsson, S. Le Marchand, L. Lejbkowicz, F. Li, S. Linet, M. Lissowska, J. Martinez, M.E. Maurer, T. Mulligan, A.M. Mulot, C. Murphy, R.A. Newman, W.G. Nielsen, S.F. Nordestgaard, B.G. Norman, A. O'Brien, K.M. Olson, J.E. Patel, A.V. Prentice, R. Rees-Punia, E. Rennert, G. Rhenius, V. Ruddy, K.J. Sandler, D.P. Scott, C.G. Shah, M. Shu, X.-.O. Smeets, A. Southey, M.C. Stone, J. Tamimi, R.M. Taylor, J.A. Teras, L.R. Tomczyk, K. Troester, M.A. Truong, T. Vachon, C.M. Wang, S.S. Weinberg, C.R. Wildiers, H. Willett, W. Winham, S.J. Wolk, A. Yang, X.R. Zamora, M.P. Zheng, W. Ziogas, A. Dunning, A.M. Pharoah, P.D.P. García-Closas, M. Schmidt, M.K. Kraft, P. Milne, R.L. Lindström, S. Easton, D.F. Chang-Claude, J (2023) A genome-wide gene-environment interaction study of breast cancer risk for women of European ancestry.. Show Abstract full text

<h4>Background</h4>Genome-wide studies of gene-environment interactions (G×E) may identify variants associated with disease risk in conjunction with lifestyle/environmental exposures. We conducted a genome-wide G×E analysis of ~ 7.6 million common variants and seven lifestyle/environmental risk factors for breast cancer risk overall and for estrogen receptor positive (ER +) breast cancer.<h4>Methods</h4>Analyses were conducted using 72,285 breast cancer cases and 80,354 controls of European ancestry from the Breast Cancer Association Consortium. Gene-environment interactions were evaluated using standard unconditional logistic regression models and likelihood ratio tests for breast cancer risk overall and for ER + breast cancer. Bayesian False Discovery Probability was employed to assess the noteworthiness of each SNP-risk factor pairs.<h4>Results</h4>Assuming a 1 × 10<sup>-5</sup> prior probability of a true association for each SNP-risk factor pairs and a Bayesian False Discovery Probability < 15%, we identified two independent SNP-risk factor pairs: rs80018847(9p13)-LINGO2 and adult height in association with overall breast cancer risk (OR<sub>int</sub> = 0.94, 95% CI 0.92-0.96), and rs4770552(13q12)-SPATA13 and age at menarche for ER + breast cancer risk (OR<sub>int</sub> = 0.91, 95% CI 0.88-0.94).<h4>Conclusions</h4>Overall, the contribution of G×E interactions to the heritability of breast cancer is very small. At the population level, multiplicative G×E interactions do not make an important contribution to risk prediction in breast cancer.

Bagherpour-Kalo, M. Jones, M.E. Darabi, P. Hosseini, M (2024) Water pipe smoking and stroke: A systematic review and meta-analysis.. Show Abstract full text

<h4>Objective</h4>Despite the damaging effects of water pipe on physical health, there is little information about the potential harmful effects of this tobacco on stroke. This study aims to investigate the relationship between water pipe smoking and stroke.<h4>Method</h4>A systematic review was conducted including Ovid SP, Embase, Pubmed, Web of Science, Scopus, and Google Scholar databases with focus on cohort, case-control, and cross-sectional studies. We reviewed all studies reporting on water pipe smoking and stroke. The funnel plot and the Egger regression test were used to assess publication bias.<h4>Results</h4>In the four eligible studies, there were a total of 2759 participants that 555 patients had at least once experienced stroke. Meta-analysis revealed positive association between water pipe smoking and stroke with pooled adjusted OR 2.79 (95% CI: 1.74-3.84; I2=0,p=.741${I^2}\; = \;\;0,{\mathrm{\;}}p\;\; = {\mathrm{\;\;}}.741$ ) and the funnel plot shows asymmetry publication bias.<h4>Conclusions</h4>We found a higher effect of water pipe smoking on stroke compared to cigarette smoking and concluded that water pipe increases the risk of stroke by 2.79. Hence, because most of the water pipe consumer society is young, especially women, policies and decisions need to be taken to control the supply of this tobacco to the market and more provide education on the health problem of water pipe smoking.<h4>Implications</h4>This study provides a higher effect of water pipe smoking on stroke. Physicians and researchers who intend to study in the field of stroke should better examine the effects of water pipe (including time of use, dose-response, long-term effects, and risk factors) on stroke.

Yousefifard, M. Toloui, A. Forouzannia, S.A. Ataei, N. Hossein, H. Khaneh, A.Z.S. Ghahfarokhi, M.K. Jones, M.E. Hosseini, M (2022) Prevalence and Mortality of Post-traumatic Acute Kidney Injury in Children; a Systematic Review and Meta-analysis.
Yousefifard, M. Ahmadzadeh, K. Neishaboori, A.M. Alavi, S.N.R. Alavi, S.R.R. Ahmadzadeh, H. Toloui, A. Gubari, M.I.M. Jones, M.E. Ataei, N. Fazel, M. Hosseini, M (2024) Review Paper A Systematic Review on the Risk Factors of Steroid-sensitive Nephrotic Syndrome Relapse in the Pediatric Popul.