Hassan, H.
Allen, I.
Sofianopoulou, E.
Walburga, Y.
Turnbull, C.
Eccles, D.M.
Tischkowitz, M.
Pharoah, P.
Antoniou, A.C.
(2024). Long-term outcomes of hysterectomy with bilateral salpingo-oophorectomy: a systematic review and meta-analysis. Am j obstet gynecol,
Vol.230
(1),
pp. 44-57.
show abstract
OBJECTIVE: This study aimed to provide an up-to-date systematic review of "the long-term outcomes of bilateral salpingo-oophorectomy at the time of hysterectomy" and perform a meta-analysis for the reported associations. DATA SOURCES: Our study updated a previous systematic review by searching the literature using PubMed, Web of Science, and Embase for publications between January 2015 and August 2022. STUDY ELIGIBILITY CRITERIA: Our study included studies of women who had a hysterectomy with bilateral salpingo-oophorectomy vs women who had a hysterectomy with ovarian conservation or no surgery. METHODS: The quality of the evidence was assessed using the Grading of Recommendations, Assessment, Development and Evaluations. Adjusted hazard ratios were extracted and combined to obtain fixed effect estimates. RESULTS: Compared with hysterectomy or no surgery, hysterectomy with bilateral salpingo-oophorectomy in young women was associated with decreased risk of breast cancer (hazard ratio, 0.78; 95% confidence interval, 0.73-0.84) but with an increased risk of colorectal cancer (hazard ratio, 1.27; 95% confidence interval, 1.10-1.47). In addition, it was associated with an increased risk of total cardiovascular diseases, coronary heart disease, and stroke with hazard ratios of 1.18 (95% confidence interval, 1.11-1.25), 1.17 (95% confidence interval, 1.10-1.25), and 1.20 (95% confidence interval, 1.10-1.31), respectively. Compared with no surgery, hysterectomy with bilateral salpingo-oophorectomy before the age of 50 years was associated with an increased risk of hyperlipidemia (hazard ratio, 1.44; 95% confidence interval, 1.25-1.65), diabetes mellitus (hazard ratio, 1.16; 95% confidence interval, 1.09-1.24), hypertension (hazard ratio, 1.13; 95% confidence interval, 1.06-1.20), dementia (hazard ratio, 1.70; 95% confidence interval, 1.07-2.69), and depression (hazard ratio, 1.39; 95% confidence interval, 1.22-1.60). The evidence on the association with all-cause mortality in young women showed substantial heterogeneity between the studies (I2=85%; P<.01). CONCLUSION: Hysterectomy with bilateral salpingo-oophorectomy was associated with multiple long-term outcomes. The benefits of the addition of bilateral salpingo-oophorectomy to hysterectomy should be balanced against the risks..
Sosinsky, A.
Ambrose, J.
Cross, W.
Turnbull, C.
Henderson, S.
Jones, L.
Hamblin, A.
Arumugam, P.
Chan, G.
Chubb, D.
Noyvert, B.
Mitchell, J.
Walker, S.
Bowman, K.
Pasko, D.
Buongermino Pereira, M.
Volkova, N.
Rueda-Martin, A.
Perez-Gil, D.
Lopez, J.
Pullinger, J.
Siddiq, A.
Zainy, T.
Choudhury, T.
Yavorska, O.
Fowler, T.
Bentley, D.
Kingsley, C.
Hing, S.
Deans, Z.
Rendon, A.
Hill, S.
Caulfield, M.
Murugaesu, N.
(2024). Insights for precision oncology from the integration of genomic and clinical data of 13,880 tumors from the 100,000 Genomes Cancer Programme. Nat med,
Vol.30
(1),
pp. 279-289.
show abstract
full text
The Cancer Programme of the 100,000 Genomes Project was an initiative to provide whole-genome sequencing (WGS) for patients with cancer, evaluating opportunities for precision cancer care within the UK National Healthcare System (NHS). Genomics England, alongside NHS England, analyzed WGS data from 13,880 solid tumors spanning 33 cancer types, integrating genomic data with real-world treatment and outcome data, within a secure Research Environment. Incidence of somatic mutations in genes recommended for standard-of-care testing varied across cancer types. For instance, in glioblastoma multiforme, small variants were present in 94% of cases and copy number aberrations in at least one gene in 58% of cases, while sarcoma demonstrated the highest occurrence of actionable structural variants (13%). Homologous recombination deficiency was identified in 40% of high-grade serous ovarian cancer cases with 30% linked to pathogenic germline variants, highlighting the value of combined somatic and germline analysis. The linkage of WGS and longitudinal life course clinical data allowed the assessment of treatment outcomes for patients stratified according to pangenomic markers. Our findings demonstrate the utility of linking genomic and real-world clinical data to enable survival analysis to identify cancer genes that affect prognosis and advance our understanding of how cancer genomics impacts patient outcomes..
Wade, I.
Witkowski, L.
Ahmed, A.
Rowlands, C.F.
McVeigh, T.P.
Tischkowitz, M.D.
Foulkes, W.D.
Turnbull, C.
(2024). Using cancer phenotype sex-specificity to enable unbiased penetrance estimation of SMARCA4 pathogenic variants for small cell carcinoma of the ovary, hypercalcemic type (SCCOHT). Genet med,
,
p. 101287.
show abstract
PURPOSE: Small cell carcinoma of the ovary, hypercalcemic type (SCCOHT) is an extremely rare, highly aggressive cancer (mean age of onset, 24 years). Nearly all cases are associated with somatic or germline pathogenic variants (GPVs) in SMARCA4. Early bilateral oophorectomy is recommended for unaffected females with a SMARCA4 GPV. However, the penetrance of SMARCA4 GPVs for SCCOHT is highly uncertain and subject to ascertainment bias. METHODS: Leveraging the early-onset, sex-specific, highly morbid nature of SCCOHT, we hypothesized that the penetrance for SCCOHT could be quantified from the deficit in SMARCA4 GPVs in females compared to males in UK Biobank, a population cohort for which recruitment was restricted to those age 40-69. We also analyzed pedigrees ascertained internationally by the Montreal-based SCCOHT-SMARCA4 Registry. RESULTS: We observed SMARCA4 GPVs in 8/210,182 (0.0038%) female and 18/179,210 (0.0100%) male participants in UK Biobank (p = 0.028), representing a male:female odds ratio of 2.64 (95%CI 1.09-7.02), implying a penetrance of 62% for SCCOHT (given absence of other SMARCA4-related female-specific early morbid diseases). A deficit of GPVs in females in UK Biobank was also demonstrated for BRCA1 and TP53. CONCLUSION: Our findings support bilateral oophorectomy in early adulthood as a rational choice for at-risk females with SMARCA4 GPVs..
Huntley, C.
Loong, L.
Mallinson, C.
Bethell, R.
Rahman, T.
Alhaddad, N.
Tulloch, O.
Zhou, X.
Lee, J.
Eves, P.
GMSA Lynch Consortium,
McRonald, F.
Torr, B.
Burn, J.
Shaw, A.
Morris, E.J.
Monahan, K.
Hardy, S.
Turnbull, C.
(2024). The comprehensive English National Lynch Syndrome Registry: development and description of a new genomics data resource. Eclinicalmedicine,
Vol.69,
p. 102465.
show abstract
full text
BACKGROUND: Lynch Syndrome (LS) is a cancer predisposition syndrome caused by constitutional pathogenic variants in the mismatch repair (MMR) genes. To date, fragmentation of clinical and genomic data has restricted understanding of national LS ascertainment and outcomes, and precluded evaluation of NICE guidance on testing and management. To address this, via collaboration between researchers, the National Disease Registration Service (NDRS), NHS Genomic Medicine Service Alliances (GMSAs), and NHS Regional Clinical Genetics Services, a comprehensive registry of LS carriers in England has been established. METHODS: For comprehensive ascertainment of retrospectively identified MMR pathogenic variant (PV) carriers (diagnosed prior to January 1, 2023), information was retrieved from all clinical genetics services across England, then restructured, amalgamated, and validated via a team of trained experts in NDRS. An online submission portal was established for prospective ascertainment from January 1, 2023. The resulting data, stored in a secure database in NDRS, were used to investigate the demographic and genetic characteristics of the cohort, censored at July 25, 2023. Cancer outcomes were investigated via linkage to the National Cancer Registration Dataset (NCRD). FINDINGS: A total of 11,722 retrospective and 570 prospective data submissions were received, resulting in a comprehensive English National Lynch Syndrome Registry (ENLSR) comprising 9030 unique individuals. The most frequently identified pathogenic MMR genes were MSH2 and MLH1 at 37.2% (n = 3362) and 29.1% (n = 2624), respectively. 35.9% (n = 3239) of the ENLSR cohort received their LS diagnosis before their first cancer diagnosis (presumptive predictive germline test). Of these, 6.3% (n = 204) developed colorectal cancer, at a median age of initial diagnosis of 51 (IQR 40-62), compared to 73 years (IQR 64-80) in the general population (p < 0.0001). INTERPRETATION: The ENLSR represents the first comprehensive national registry of PV carriers in England and one of the largest cohorts of MMR PV carriers worldwide. The establishment of a secure, centralised infrastructure and mechanism for routine registration of newly identified carriers ensures sustainability of the data resource. FUNDING: This work was funded by the Wellcome Trust, Cancer Research UK and Bowel Cancer UK. The funder of this study had no role in study design, data collection, data analysis, data interpretation, or writing of the report..
Olvera-León, R.
Zhang, F.
Offord, V.
Zhao, Y.
Tan, H.K.
Gupta, P.
Pal, T.
Robles-Espinoza, C.D.
Arriaga-González, F.G.
Matsuyama, L.S.
Delage, E.
Dicks, E.
Ezquina, S.
Rowlands, C.F.
Turnbull, C.
Pharoah, P.
Perry, J.R.
Jasin, M.
Waters, A.J.
Adams, D.J.
(2024). High-resolution functional mapping of RAD51C by saturation genome editing. Cell,
Vol.187
(20),
pp. 5719-5734.e19.
show abstract
Pathogenic variants in RAD51C confer an elevated risk of breast and ovarian cancer, while individuals homozygous for specific RAD51C alleles may develop Fanconi anemia. Using saturation genome editing (SGE), we functionally assess 9,188 unique variants, including >99.5% of all possible coding sequence single-nucleotide alterations. By computing changes in variant abundance and Gaussian mixture modeling (GMM), we functionally classify 3,094 variants to be disruptive and use clinical truth sets to reveal an accuracy/concordance of variant classification >99.9%. Cell fitness was the primary assay readout allowing us to observe a phenomenon where specific missense variants exhibit distinct depletion kinetics potentially suggesting that they represent hypomorphic alleles. We further explored our exhaustive functional map, revealing critical residues on the RAD51C structure and resolving variants found in cancer-segregating kindred. Furthermore, through interrogation of UK Biobank and a large multi-center ovarian cancer cohort, we find significant associations between SGE-depleted variants and cancer diagnoses..
McRonald, F.E.
Pethick, J.
Santaniello, F.
Shand, B.
Tyson, A.
Tulloch, O.
Goel, S.
Lüchtenborg, M.
Borthwick, G.M.
Turnbull, C.
Shaw, A.C.
Monahan, K.J.
Frayling, I.M.
Hardy, S.
Burn, J.
(2024). Identification of people with Lynch syndrome from those presenting with colorectal cancer in England: baseline analysis of the diagnostic pathway. Eur j hum genet,
Vol.32
(5),
pp. 529-538.
show abstract
It is believed that >95% of people with Lynch syndrome (LS) remain undiagnosed. Within the National Health Service (NHS) in England, formal guidelines issued in 2017 state that all colorectal cancers (CRC) should be tested for DNA Mismatch Repair deficiency (dMMR). We used a comprehensive population-level national dataset to analyse implementation of the agreed diagnostic pathway at a baseline point 2 years post-publication of official guidelines. Using real-world data collected and curated by the National Cancer Registration and Analysis Service (NCRAS), we retrospectively followed up all people diagnosed with CRC in England in 2019. Nationwide laboratory diagnostic data incorporated somatic (tumour) testing for dMMR (via immunohistochemistry or microsatellite instability), somatic testing for MLH1 promoter methylation and BRAF status, and constitutional (germline) testing of MMR genes. Only 44% of CRCs were screened for dMMR; these figures varied over four-fold with respect to geography. Of those CRCs identified as dMMR, only 51% underwent subsequent diagnostic testing. Overall, only 1.3% of patients with colorectal cancer had a germline MMR genetic test performed; up to 37% of these tests occurred outside of NICE guidelines. The low rates of molecular diagnostic testing in CRC support the premise that Lynch syndrome is underdiagnosed, with significant attrition at all stages of the testing pathway. Applying our methodology to subsequent years' data will allow ongoing monitoring and analysis of the impact of recent investment. If the diagnostic guidelines were fully implemented, we estimate that up to 700 additional people with LS could be identified each year..
Allen, S.
Garrett, A.
Muffley, L.
Fayer, S.
Foreman, J.
Adams, D.J.
Hurles, M.
Rubin, A.F.
Roth, F.P.
Starita, L.M.
Biesecker, L.G.
Turnbull, C.
(2024). Workshop report: the clinical application of data from multiplex assays of variant effect (MAVEs), 12 July 2023. Eur j hum genet,
Vol.32
(5),
pp. 593-600.
show abstract
full text
Clinical classification of genomic variants identified on sequencing is often challenging, with many variants classified as Variants of Uncertain Significance (VUS) on account of insufficient evidence. Advances in sequencing and gene synthesis has made feasible multiplexed assays of variant effect (MAVEs), which quantify the functional impact of many thousands of genomic variants in a single experiment. These assays and the functional evidence they generate have the potential to empower more accurate clinical variant classification. However, there are many outstanding challenges and opportunities that require joint resolution and specification, thus necessitating communication between the research scientists who have designed and performed MAVEs and the clinicians and diagnostic scientists who will apply their data to clinical variant classification. In the ‘Clinical Application of MAVE Data’ workshop, held on 12th July 2023 at the Wellcome Connecting Science Conference Centre in between two relevant research meetings, ‘Curating the Clinical Genome 2023’ and the ‘Mutational Scanning Symposium 2023’, 44 key scientific and/or clinical stakeholders were brought together to consider important questions relating to clinical application of MAVE data, such as quantitative validation, variant truth-sets, platforms and standards for dissemination of MAVE data. The outcomes and possible next steps that were discussed encompassed development of focused workshops to develop consensus recommendations, creating a MAVE evaluation working group, and collaboration of ClinVar and MaveDB to enact software changes that support enhanced functional data submission..
Allen, I.
Hassan, H.
Joko-Fru, W.Y.
Huntley, C.
Loong, L.
Rahman, T.
Torr, B.
Bacon, A.
Knott, C.
Jose, S.
Vernon, S.
Lüchtenborg, M.
Pethick, J.
Lavelle, K.
McRonald, F.
Eccles, D.
Morris, E.J.
Hardy, S.
Turnbull, C.
Tischkowitz, M.
Pharoah, P.
Antoniou, A.C.
(2024). Risks of second primary cancers among 584,965 female and male breast cancer survivors in England: a 25-year retrospective cohort study. Lancet reg health eur,
Vol.40,
p. 100903.
show abstract
full text
BACKGROUND: Second primary cancers (SPCs) after breast cancer (BC) present an increasing public health burden, with little existing research on socio-demographic, tumour, and treatment effects. We addressed this in the largest BC survivor cohort to date, using a novel linkage of National Disease Registration Service datasets. METHODS: The cohort included 581,403 female and 3562 male BC survivors diagnosed between 1995 and 2019. We estimated standardized incidence ratios (SIRs) for combined and site-specific SPCs using incidences for England, overall and by age at BC and socioeconomic status. We estimated incidences and Kaplan-Meier cumulative risks stratified by age at BC, and assessed risk variation by socio-demographic, tumour, and treatment characteristics using Cox regression. FINDINGS: Both genders were at elevated contralateral breast (SIR: 2.02 (95% CI: 1.99-2.06) females; 55.4 (35.5-82.4) males) and non-breast (1.10 (1.09-1.11) females, 1.10 (1.00-1.20) males) SPC risks. Non-breast SPC risks were higher for females younger at BC diagnosis (SIR: 1.34 (1.31-1.38) <50 y, 1.07 (1.06-1.09) ≥50 y) and more socioeconomically deprived (SIR: 1.00 (0.98-1.02) least deprived quintile, 1.34 (1.30-1.37) most). INTERPRETATION: Enhanced SPC surveillance may benefit BC survivors, although specific recommendations require more detailed multifactorial risk and cost-benefit analyses. The associations between deprivation and SPC risks could provide clinical management insights. FUNDING: CRUK Catalyst Award CanGene-CanVar (C61296/A27223). Cancer Research UK grant: PPRPGM-Nov 20∖100,002. This work was supported by core funding from the NIHR Cambridge Biomedical Research Centre (NIHR203312)]. The views expressed are those of the author(s) and not necessarily those of the NIHR or the Department of Health and Social Care..
Horton, R.
Wright, C.F.
Firth, H.V.
Turnbull, C.
Lachmann, R.
Houlston, R.S.
Lucassen, A.
(2024). Challenges of using whole genome sequencing in population newborn screening. Bmj,
Vol.384,
p. e077060.
McDevitt, T.
Durkie, M.
Arnold, N.
Burghel, G.J.
Butler, S.
Claes, K.B.
Logan, P.
Robinson, R.
Sheils, K.
Wolstenholme, N.
Hanson, H.
Turnbull, C.
Hume, S.
(2024). EMQN best practice guidelines for genetic testing in hereditary breast and ovarian cancer. Eur j hum genet,
Vol.32
(5),
pp. 479-488.
show abstract
Hereditary Breast and Ovarian Cancer (HBOC) is a genetic condition associated with increased risk of cancers. The past decade has brought about significant changes to hereditary breast and ovarian cancer (HBOC) diagnostic testing with new treatments, testing methods and strategies, and evolving information on genetic associations. These best practice guidelines have been produced to assist clinical laboratories in effectively addressing the complexities of HBOC testing, while taking into account advancements since the last guidelines were published in 2007. These guidelines summarise cancer risk data from recent studies for the most commonly tested high and moderate risk HBOC genes for laboratories to refer to as a guide. Furthermore, recommendations are provided for somatic and germline testing services with regards to clinical referral, laboratory analyses, variant interpretation, and reporting. The guidelines present recommendations where 'must' is assigned to advocate that the recommendation is essential; and 'should' is assigned to advocate that the recommendation is highly advised but may not be universally applicable. Recommendations are presented in the form of shaded italicised statements throughout the document, and in the form of a table in supplementary materials (Table S4). Finally, for the purposes of encouraging standardisation and aiding implementation of recommendations, example report wording covering the essential points to be included is provided for the most common HBOC referral and reporting scenarios. These guidelines are aimed primarily at genomic scientists working in diagnostic testing laboratories..
Parsons, M.T.
de la Hoya, M.
Richardson, M.E.
Tudini, E.
Anderson, M.
Berkofsky-Fessler, W.
Caputo, S.M.
Chan, R.C.
Cline, M.S.
Feng, B.-.
Fortuno, C.
Gomez-Garcia, E.
Hadler, J.
Hiraki, S.
Holdren, M.
Houdayer, C.
Hruska, K.
James, P.
Karam, R.
Leong, H.S.
Martins, A.
Mensenkamp, A.R.
Monteiro, A.N.
Nathan, V.
O'Connor, R.
Pedersen, I.S.
Pesaran, T.
Radice, P.
Schmidt, G.
Southey, M.
Tavtigian, S.
Thompson, B.A.
Toland, A.E.
Turnbull, C.
Vogel, M.J.
Weyandt, J.
Wiggins, G.A.
Zec, L.
Couch, F.J.
Walker, L.C.
Vreeswijk, M.P.
Goldgar, D.E.
Spurdle, A.B.
(2024). Evidence-based recommendations for gene-specific ACMG/AMP variant classification from the ClinGen ENIGMA BRCA1 and BRCA2 Variant Curation Expert Panel. Am j hum genet,
Vol.111
(9),
pp. 2044-2058.
show abstract
The ENIGMA research consortium develops and applies methods to determine clinical significance of variants in hereditary breast and ovarian cancer genes. An ENIGMA BRCA1/2 classification sub-group, formed in 2015 as a ClinGen external expert panel, evolved into a ClinGen internal Variant Curation Expert Panel (VCEP) to align with Food and Drug Administration recognized processes for ClinVar contributions. The VCEP reviewed American College of Medical Genetics and Genomics/Association of Molecular Pathology (ACMG/AMP) classification criteria for relevance to interpreting BRCA1 and BRCA2 variants. Statistical methods were used to calibrate evidence strength for different data types. Pilot specifications were tested on 40 variants and documentation revised for clarity and ease of use. The original criterion descriptions for 13 evidence codes were considered non-applicable or overlapping with other criteria. Scenario of use was extended or re-purposed for eight codes. Extensive analysis and/or data review informed specification descriptions and weights for all codes. Specifications were applied to pilot variants with pre-existing ClinVar classification as follows: 13 uncertain significance or conflicting, 14 pathogenic and/or likely pathogenic, and 13 benign and/or likely benign. Review resolved classification for 11/13 uncertain significance or conflicting variants and retained or improved confidence in classification for the remaining variants. Alignment of pre-existing ENIGMA research classification processes with ACMG/AMP classification guidelines highlighted several gaps in the research processes and the baseline ACMG/AMP criteria. Calibration of evidence strength was key to justify utility and strength of different data types for gene-specific application. The gene-specific criteria demonstrated value for improving ACMG/AMP-aligned classification of BRCA1 and BRCA2 variants..
Tsoulaki, O.
Tischkowitz, M.
Antoniou, A.C.
Musgrave, H.
Rea, G.
Gandhi, A.
Cox, K.
Irvine, T.
Holcombe, S.
Eccles, D.
Turnbull, C.
Cutress, R.
Meeting Attendees,
Archer, S.
Hanson, H.
(2024). Joint ABS-UKCGG-CanGene-CanVar consensus regarding the use of CanRisk in clinical practice. Br j cancer,
Vol.130
(12),
pp. 2027-2036.
show abstract
full text
BACKGROUND: The CanRisk tool, which operationalises the Breast and Ovarian Analysis of Disease Incidence and Carrier Estimation Algorithm (BOADICEA) is used by Clinical Geneticists, Genetic Counsellors, Breast Oncologists, Surgeons and Family History Nurses for breast cancer risk assessments both nationally and internationally. There are currently no guidelines with respect to the day-to-day clinical application of CanRisk and differing inputs to the model can result in different recommendations for practice. METHODS: To address this gap, the UK Cancer Genetics Group in collaboration with the Association of Breast Surgery and the CanGene-CanVar programme held a workshop on 16th of May 2023, with the aim of establishing best practice guidelines. RESULTS: Using a pre-workshop survey followed by structured discussion and in-meeting polling, we achieved consensus for UK best practice in use of CanRisk in making recommendations for breast cancer surveillance, eligibility for genetic testing and the input of available information to undertake an individualised risk assessment. CONCLUSIONS: Whilst consensus recommendations were achieved, the meeting highlighted some of the barriers limiting the use of CanRisk in clinical practice and identified areas that require further work and collaboration with relevant national bodies and policy makers to incorporate wider use of CanRisk into routine breast cancer risk assessments..
Huntley, C.
Torr, B.
Kavanaugh, G.
George, A.
Hanson, H.
Snape, K.
Broggio, J.
Glasgow, L.
Tischkowitz, M.
Evans, D.G.
Antoniou, A.C.
Turnbull, C.
(2024). Breast cancer risk assessment for prescription of menopausal hormone therapy in women with a family history of breast cancer: an epidemiological modelling study. Br j gen pract,
Vol.74
(746),
pp. e610-e618.
show abstract
full text
BACKGROUND: Menopausal hormone therapy (MHT) can alleviate menopausal symptoms but has been associated with an increased risk of breast cancer. MHT prescription should be preceded by individualised risk/benefit evaluation; however, data outlining the impact of family history alongside different MHT therapeutic approaches are lacking. AIM: To quantify the risks associated with MHT use in women with varying breast cancer family histories of developing and dying from breast cancer. DESIGN AND SETTING: An epidemiological modelling study for women in England using the BOADICEA breast cancer prediction model and data relating to MHT use and breast cancer risk taken from research by the Collaborative Group on Hormonal Factors in Breast Cancer. METHOD: The risk of developing and dying from breast cancer between the ages of 50 and 80 years was modelled in women with four different breast cancer family history profiles: 'average', 'modest', 'intermediate', and 'strong' by using 1) background risks of breast cancer by age and family history, 2) relative risks for breast cancer associated with MHT use, and 3) 10-year breast cancer-specific net mortality rates. This study modelled use of combined oestrogen-progestogen MHT (cyclical or continuous) and oestrogen-only MHT. RESULTS: For a woman of 'average' family history taking no MHT, the cumulative breast cancer risk (age 50-80 years) is 9.8%, and the risk of dying from the breast cancer is 1.7%. In this model, 5 years' exposure to combined-cyclical MHT (age 50-55 years) was calculated to increase these risks to 11.0% and 1.8%, respectively. For a woman with a 'strong' family history taking no MHT, the cumulative breast cancer risk is 19.6% (age 50-80 years), and the risk of dying from the breast cancer is 3.2%. With 5 years' exposure to MHT (age 50-55 years), this model showed that these risks increase to 22.4% and 3.5%, respectively. CONCLUSION: In this model, both family history and MHT are associated with increased risk of breast cancer. Estimates of the risks of breast cancer associated with MHT for women with different family histories can be used to support decision making around MHT prescription for women experiencing menopausal symptoms..
Rowlands, C.F.
Allen, S.
Balmaña, J.
Domchek, S.M.
Evans, D.G.
Hanson, H.
Hoogerbrugge, N.
James, P.A.
Nathanson, K.L.
Robson, M.
Tischkowitz, M.
Foulkes, W.D.
Turnbull, C.
(2024). Population-based germline breast cancer gene association studies and meta-analysis to inform wider mainstream testing. Ann oncol,
Vol.35
(10),
pp. 892-901.
show abstract
full text
BACKGROUND: Germline genetic testing, previously restricted to familial and young-onset breast cancer, is now offered increasingly broadly to patients with 'population-type' breast cancer in mainstream oncology clinics, with wide variation in the genes included. PATIENTS AND METHODS: Weighted meta-analysis was carried out for three population-based case-control studies (BRIDGES, CARRIERS and UK Biobank) comprising in total 101 397 women with breast cancer and 312 944 women without breast cancer, to quantify 37 putative breast cancer susceptibility genes (BCSGs) for the frequency of pathogenic variants (PVs) in unselected, 'population-type' breast cancer cases and their association with breast cancer and its subtypes. RESULTS: Meta-analysed odds ratios (ORs) and frequencies of PVs in 'population-type' breast cancer cases were generated for BRCA1 (OR 8.73, 95% confidence interval (CI) 7.47-10.20; 1 in 101), BRCA2 (OR 5.68, 95% CI 5.13-6.30; 1 in 68) and PALB2 (OR 4.30, 95% CI 3.68-5.03; 1 in 187). For both CHEK2 (OR 2.40, 95% CI 2.21-2.62; 1 in 73) and ATM (OR 2.16, 95% CI 1.93-2.41; 1 in 132) subgroup analysis showed a stronger association with oestrogen receptor-positive disease. The magnitude of association and frequency of PVs were low for RAD51C (OR 1.53, 95% CI 1.29-2.04; 1 in 913), RAD51D (OR 1.76, 95% CI 1.29-2.41; 1 in 1079) and BARD1 (OR 2.34, 95% CI 1.85-2.97; 1 in 672); frequencies and associations were higher when the analysis was restricted to triple-negative breast cancers. The PV frequency in 'population-type' breast cancer cases was very low for 'syndromic' BCSGs TP53 (1 in 1844), STK11 (1 in 11 525), CDH1 (1 in 2668), PTEN (1 in 3755) and NF1 (1 in 1470), with metrics of association also modest ranging from OR 3.62 (95% CI 1.98-6.61) for TP53 down to OR 1.60 (95% CI 0.48-5.30) for STK11. CONCLUSIONS: These metrics reflecting 'population-type' breast cancer will be informative in defining the appropriate gene set as we continue to expand to germline testing to an increasingly unselected group of breast cancer cases..
Torr, B.
Jones, C.
Kavanaugh, G.
Hamill, M.
Allen, S.
Choi, S.
Garrett, A.
Valganon-Petrizan, M.
MacMahon, S.
Yuan, L.
Way, R.
Harder, H.
Gold, R.
Taylor, A.
Gabe, R.
Lucassen, A.
Manchanda, R.
Fallowfield, L.
Jenkins, V.
Gandhi, A.
Evans, D.G.
George, A.
Hubank, M.
Kemp, Z.
Bremner, S.
Turnbull, C.
(2024). BRCA-DIRECT digital pathway for diagnostic germline genetic testing within a UK breast oncology setting: a randomised, non-inferiority trial. Br j cancer,
Vol.131
(9),
pp. 1506-1515.
show abstract
full text
BACKGROUND: Genetic testing to identify germline high-risk pathogenic variants in breast cancer susceptibility genes is increasingly part of the breast cancer diagnostic pathway. Novel patient-centred pathways may offer opportunity to expand capacity and reduce turnaround time. METHODS: We recruited 1140 women with unselected breast cancer to undergo germline genetic testing through the BRCA-DIRECT pathway (which includes a digital platform, postal saliva sampling and a genetic counsellor telephone helpline). Ahead of consenting to the test, participants were randomised to receive information about genetic testing digitally (569/1140, 49.9%) or via a pre-test genetic counselling consultation (571/1140, 50.1%). RESULTS: 1001 (87.8%) participants progressed to receive their pre-test information and consented to testing. The primary outcome, uptake of genetic testing, was higher amongst participants randomised to receive digital information compared with those randomised to a pre-test genetic counselling consultation (90.8% (95% CI: 88.5% to 93.1%) vs 84.7% (95% CI: 81.8% to 87.6%), p = 0.002, adjusted for participant age and site). Non-inferiority was observed in relation to patient knowledge, anxiety, and satisfaction. CONCLUSIONS: Findings demonstrate that standardised, digital information offers a non-inferior alternative to conventional genetic counselling, and an end-to-end patient-centred, digital pathway (supported by genetic counselling hotline) could feasibly be implemented into breast oncology settings. CLINICAL TRIAL REGISTRATION: The study is registered with, and protocol available on, ClinicalTrials.gov (NCT04842799)..
van de Haar, J.
Roepman, P.
Andre, F.
Balmaña, J.
Castro, E.
Chakravarty, D.
Curigliano, G.
Czarnecka, A.M.
Dienstmann, R.
Horak, P.
Italiano, A.
Marchiò, C.
Monkhorst, K.
Pritchard, C.C.
Reardon, B.
Russnes, H.E.
Sirohi, B.
Sosinsky, A.
Spanic, T.
Turnbull, C.
Van Allen, E.
Westphalen, C.B.
Tamborero, D.
Mateo, J.
(2024). ESMO Recommendations on clinical reporting of genomic test results for solid cancers. Ann oncol,
Vol.35
(11),
pp. 954-967.
show abstract
BACKGROUND: Genomic tumour profiling has a crucial role in the management of patients with solid cancers, as it helps selecting and prioritising therapeutic interventions based on prognostic and predictive biomarkers, as well as identifying markers of hereditary cancers. Harmonised approaches to interpret the results of genomic testing are needed to support physicians in their decision making, prevent inequalities in precision medicine and maximise patient benefit from available cancer management options. METHODS: The European Society for Medical Oncology (ESMO) Translational Research and Precision Medicine Working Group assembled a group of international experts to propose recommendations for preparing clinical genomic reports for solid cancers. These recommendations aim to foster best practices in integrating genomic testing within clinical settings. After review of available evidence, several rounds of surveys and focused discussions were conducted to reach consensus on the recommendation statements. Only consensus recommendations were reported. Recommendation statements were graded in two tiers based on their clinical importance: level A (required to maintain common standards in reporting) and level B (optional but necessary to achieve ideal practice). RESULTS: Genomics reports should present key information in a front page(s) followed by supplementary information in one or more appendices. Reports should be structured into sections: (i) patient and sample details; (ii) assay and data analysis characteristics; (iii) sample-specific assay performance and quality control; (iv) genomic alterations and their functional annotation; (v) clinical actionability assessment and matching to potential therapy indications; and (vi) summary of the main findings. Specific recommendations to prepare each of these sections are made. CONCLUSIONS: We present a set of recommendations aimed at structuring genomics reports to enhance physician comprehension of genomic profiling results for solid cancers. Communication between ordering physicians and professionals reporting genomic data is key to minimise uncertainties and to optimise the impact of genomic tests in patient care..
McLaughlin, S.
Amir, H.
Garrido, N.
Turnbull, C.
Rouncefield-Swales, A.
Swadźba-Kwaśny, M.
Morgan, K.
(2024). Evaluating the Impact of Project-Based Learning in Supporting Students with the A-Level Chemistry Curriculum in Northern Ireland. Journal of chemical education,
Vol.101
(2),
pp. 537-546.
Allen, S.
Loong, L.
Garrett, A.
Torr, B.
Durkie, M.
Drummond, J.
Callaway, A.
Robinson, R.
Burghel, G.J.
Hanson, H.
Field, J.
McDevitt, T.
McVeigh, T.P.
Bedenham, T.
Bowles, C.
Bradshaw, K.
Brooks, C.
Butler, S.
Del Rey Jimenez, J.C.
Hawkes, L.
Stinton, V.
MacMahon, S.
Owens, M.
Palmer-Smith, S.
Smith, K.
Tellez, J.
Valganon-Petrizan, M.
Waskiewicz, E.
Yau, M.
Eccles, D.M.
Tischkowitz, M.
Goel, S.
McRonald, F.
Antoniou, A.C.
Morris, E.
Hardy, S.
Turnbull, C.
(2024). Recommendations for laboratory workflow that better support centralised amalgamation of genomic variant data: findings from CanVIG-UK national molecular laboratory survey. J med genet,
Vol.61
(4),
pp. 305-312.
show abstract
full text
BACKGROUND: National and international amalgamation of genomic data offers opportunity for research and audit, including analyses enabling improved classification of variants of uncertain significance. Review of individual-level data from National Health Service (NHS) testing of cancer susceptibility genes (2002-2023) submitted to the National Disease Registration Service revealed heterogeneity across participating laboratories regarding (1) the structure, quality and completeness of submitted data, and (2) the ease with which that data could be assembled locally for submission. METHODS: In May 2023, we undertook a closed online survey of 51 clinical scientists who provided consensus responses representing all 17 of 17 NHS molecular genetic laboratories in England and Wales which undertake NHS diagnostic analyses of cancer susceptibility genes. The survey included 18 questions relating to 'next-generation sequencing workflow' (11), 'variant classification' (3) and 'phenotypical context' (4). RESULTS: Widely differing processes were reported for transfer of variant data into their local LIMS (Laboratory Information Management System), for the formatting in which the variants are stored in the LIMS and which classes of variants are retained in the local LIMS. Differing local provisions and workflow for variant classifications were also reported, including the resources provided and the mechanisms by which classifications are stored. CONCLUSION: The survey responses illustrate heterogeneous laboratory workflow for preparation of genomic variant data from local LIMS for centralised submission. Workflow is often labour-intensive and inefficient, involving multiple manual steps which introduce opportunities for error. These survey findings and adoption of the concomitant recommendations may support improvement in laboratory dataflows, better facilitating submission of data for central amalgamation..
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. J med genet,
Vol.61
(10),
pp. 983-991.
show abstract
full text
BACKGROUND: 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. METHODS: 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). RESULTS: 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. CONCLUSION: 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..
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.
Rowlands, C.F.
McVeigh, T.
Hanson, H.
Turnbull, C.
CanVIG-UK,
(2024). Classification of Variants of Reduced Penetrance in High Penetrance Cancer Susceptibility Genes: Framework for Genetics Clinicians and Clinical Scientists by CanVIG-UK (Cancer Variant Interpretation Group-UK). Genet med,
,
p. 101305.
show abstract
PURPOSE: Current practice is to report and manage likely pathogenic/pathogenic variants in a given cancer susceptibility gene (CSG) as though having equivalent penetrance, despite increasing evidence of inter-variant variability in risk associations. Using existing variant interpretation approaches, largely based on full-penetrance models, variants where reduced penetrance is suspected may be classified inconsistently and/or as variants of uncertain significance (VUS). We aimed to develop a national consensus approach for such variants within the Cancer Variant Interpretation Group UK (CanVIG-UK) multidisciplinary network. METHODS: A series of surveys and live polls were conducted during and between CanVIG-UK monthly meetings on various scenarios potentially indicating reduced penetrance. These informed the iterative development of a framework for the classification of variants of reduced penetrance by the CanVIG-UK Steering and Advisory Group (CStAG) working group. RESULTS: CanVIG-UK recommendations for amendment of the 2015 ACMG/AMP variant interpretation framework were developed for variants where (A) Active evidence suggests a reduced penetrance effect size (e.g. from case-control or segregation data) (B) Reduced penetrance effect is inferred from weaker/potentially-inconsistent observed data. CONCLUSIONS: CanVIG-UK propose a framework for the classification of variants of reduced penetrance in high-penetrance genes. These principles, whilst developed for CSGs, are potentially applicable to other clinical contexts..
Allen, I.
Hassan, H.
Walburga, Y.
Huntley, C.
Loong, L.
Rahman, T.
Allen, S.
Garrett, A.
Torr, B.
Bacon, A.
Knott, C.
Jose, S.
Vernon, S.
Lüchtenborg, M.
Pethick, J.
Santaniello, F.
Goel, S.
Wang, Y.-.
Lavelle, K.
McRonald, F.
Eccles, D.
Morris, E.
Hardy, S.
Turnbull, C.
Tischkowitz, M.
Pharoah, P.
Antoniou, A.C.
(2024). Second Primary Cancer Risks After Breast Cancer in BRCA1 and BRCA2 Pathogenic Variant Carriers. J clin oncol,
,
p. JCO2401146.
show abstract
PURPOSE: Second primary cancer (SPC) risks after breast cancer (BC) in BRCA1/BRCA2 pathogenic variant (PV) carriers are uncertain. We estimated relative and absolute risks using a novel linkage of genetic testing data to population-scale National Disease Registration Service and Hospital Episode Statistics electronic health records. METHODS: We followed 25,811 females and 480 males diagnosed with BC and tested for germline BRCA1/BRCA2 PVs in NHS Clinical Genetics centers in England between 1995 and 2019 until SPC diagnosis, death, migration, contralateral breast/ovarian surgery plus 1 year, or the 31st of December 2020. We estimated standardized incidence ratios (SIRs) using English population incidences, hazard ratios (HRs) comparing carriers to noncarriers using Cox regression, and Kaplan-Meier 10-year cumulative risks. RESULTS: There were 1,840 BRCA1 and 1,750 BRCA2 female PV carriers. Compared with population incidences, BRCA1 carriers had elevated contralateral BC (CBC; SIR, 15.6 [95% CI, 11.8 to 20.2]), ovarian (SIR, 44.0 [95% CI, 31.4 to 59.9]), combined nonbreast/ovarian (SIR, 2.18 [95% CI, 1.59 to 2.92]), colorectal (SIR, 4.80 [95% CI, 2.62 to 8.05]), and endometrial (SIR, 2.92 [95% CI, 1.07 to 6.35]) SPC risks. BRCA2 carriers had elevated CBC (SIR, 7.70 [95% CI, 5.45 to 10.6]), ovarian (SIR, 16.8 [95% CI, 10.3 to 26.0]), pancreatic (SIR, 5.42 [95% CI, 2.09 to 12.5]), and combined nonbreast/ovarian (SIR, 1.68 [95% CI, 1.24 to 2.23]) SPC risks. Compared with females without BRCA1/BRCA2 PVs on testing, BRCA1 carriers had elevated CBC (HR, 3.60 [95% CI, 2.65 to 4.90]), ovarian (HR, 33.0 [95% CI, 19.1 to 57.1]), combined nonbreast/ovarian (HR, 1.45 [95% CI, 1.05 to 2.01]), and colorectal (HR, 2.93 [95% CI, 1.53 to 5.62]) SPC risks. BRCA2 carriers had elevated CBC (HR, 2.40 [95% CI, 1.70 to 3.40]), ovarian (HR, 12.0 [95% CI, 6.70 to 21.5]), and pancreatic (HR, 3.56 [95% CI, 1.34 to 9.48]) SPC risks. Ten-year cumulative CBC, ovarian, and combined nonbreast/ovarian cancer risks were 16%/6.3%/7.8% (BRCA1 carriers), 12%/3.0%/6.2% (BRCA2 carriers), and 3.6%/0.4%/4.9% (noncarriers). Male BRCA2 carriers had higher CBC (HR, 13.1 [95% CI, 1.19 to 146]) and prostate (HR, 5.61 [95% CI, 1.96 to 16.0]) SPC risks than noncarriers. CONCLUSION: Survivors of BC carrying BRCA1 and BRCA2 PVs are at high SPC risk. They may benefit from enhanced surveillance and risk-reduction measures..
Casolino, R.
Beer, P.A.
Chakravarty, D.
Davis, M.B.
Malapelle, U.
Mazzarella, L.
Normanno, N.
Pauli, C.
Subbiah, V.
Turnbull, C.
Westphalen, C.B.
Biankin, A.V.
(2024). Interpreting and integrating genomic tests results in clinical cancer care: Overview and practical guidance. Ca cancer j clin,
Vol.74
(3),
pp. 264-285.
show abstract
The last decade has seen rapid progress in the use of genomic tests, including gene panels, whole-exome sequencing, and whole-genome sequencing, in research and clinical cancer care. These advances have created expansive opportunities to characterize the molecular attributes of cancer, revealing a subset of cancer-associated aberrations called driver mutations. The identification of these driver mutations can unearth vulnerabilities of cancer cells to targeted therapeutics, which has led to the development and approval of novel diagnostics and personalized interventions in various malignancies. The applications of this modern approach, often referred to as precision oncology or precision cancer medicine, are already becoming a staple in cancer care and will expand exponentially over the coming years. Although genomic tests can lead to better outcomes by informing cancer risk, prognosis, and therapeutic selection, they remain underutilized in routine cancer care. A contributing factor is a lack of understanding of their clinical utility and the difficulty of results interpretation by the broad oncology community. Practical guidelines on how to interpret and integrate genomic information in the clinical setting, addressed to clinicians without expertise in cancer genomics, are currently limited. Building upon the genomic foundations of cancer and the concept of precision oncology, the authors have developed practical guidance to aid the interpretation of genomic test results that help inform clinical decision making for patients with cancer. They also discuss the challenges that prevent the wider implementation of precision oncology..
Hanson, H.
Durkie, M.
Lalloo, F.
Izatt, L.
McVeigh, T.P.
Cook, J.A.
Brewer, C.
Drummond, J.
Butler, S.
Cranston, T.
Casey, R.
Tan, T.
Morganstein, D.
Eccles, D.M.
Tischkowitz, M.
Turnbull, C.
Woodward, E.R.
Maher, E.R.
UK Cancer Genetics Centres,
(2023). UK recommendations for SDHA germline genetic testing and surveillance in clinical practice. J med genet,
Vol.60
(2),
pp. 107-111.
show abstract
full text
SDHA pathogenic germline variants (PGVs) are identified in up to 10% of patients with paraganglioma and phaeochromocytoma and up to 30% with wild-type gastrointestinal stromal tumours. Most SDHA PGV carriers present with an apparently sporadic tumour, but often the pathogenic variant has been inherited from parent who has the variant, but has not developed any clinical features. Studies of SDHA PGV carriers suggest that lifetime penetrance for SDHA-associated tumours is low, particularly when identified outside the context of a family history. Current recommended surveillance for SDHA PGV carriers follows an intensive protocol. With increasing implementation of tumour and germline large panel and whole-genome sequencing, it is likely more SDHA PGV carriers will be identified in patients with tumours not strongly associated with SDHA, or outside the context of a strong family history. This creates a complex situation about what to recommend in clinical practice considering low penetrance for tumour development, surveillance burden and patient anxiety. An expert SDHA working group was formed to discuss and consider this situation. This paper outlines the recommendations from this working group for testing and management of SDHA PGV carriers in clinical practice..
Kuzbari, Z.
Bandlamudi, C.
Loveday, C.
Garrett, A.
Mehine, M.
George, A.
Hanson, H.
Snape, K.
Kulkarni, A.
Allen, S.
Jezdic, S.
Ferrandino, R.
Westphalen, C.B.
Castro, E.
Rodon, J.
Mateo, J.
Burghel, G.J.
Berger, M.F.
Mandelker, D.
Turnbull, C.
(2023). Germline-focused analysis of tumour-detected variants in 49,264 cancer patients: ESMO Precision Medicine Working Group recommendations. Ann oncol,
Vol.34
(3),
pp. 215-227.
show abstract
full text
BACKGROUND: The European Society for Medical Oncology Precision Medicine Working Group (ESMO PMWG) was reconvened to update its 2018/19 recommendations on follow-up of putative germline variants detected on tumour-only sequencing, which were based on an analysis of 17 152 cancers. METHODS: We analysed an expanded dataset including 49 264 paired tumour-normal samples. We applied filters to tumour-detected variants based on variant allele frequency, predicted pathogenicity and population variant frequency. For 58 cancer-susceptibility genes, we then examined the proportion of filtered tumour-detected variants of true germline origin [germline conversion rate (GCR)]. We conducted subanalyses based on the age of cancer diagnosis, specific tumour types and 'on-tumour' status (established tumour-gene association). RESULTS: Analysis of 45 472 nonhypermutated solid malignancy tumour samples yielded 21 351 filtered tumour-detected variants of which 3515 were of true germline origin. 3.1% of true germline pathogenic variants were absent from the filtered tumour-detected variants. For genes such as BRCA1, BRCA2 and PALB2, the GCR in filtered tumour-detected variants was >80%; conversely for TP53, APC and STK11 this GCR was <2%. CONCLUSION: Strategic germline-focused analysis can prioritise a subset of tumour-detected variants for which germline follow-up will produce the highest yield of most actionable true germline variants. We present updated recommendations around germline follow-up of tumour-only sequencing including (i) revision to 5% for the minimum per-gene GCR, (ii) inclusion of actionable intermediate penetrance genes ATM and CHEK2, (iii) definition of a set of seven 'most actionable' cancer-susceptibility genes (BRCA1, BRCA2, PALB2, MLH1, MSH2, MSH6 and RET) in which germline follow-up is recommended regardless of tumour type..
Fowler, D.M.
Adams, D.J.
Gloyn, A.L.
Hahn, W.C.
Marks, D.S.
Muffley, L.A.
Neal, J.T.
Roth, F.P.
Rubin, A.F.
Starita, L.M.
Hurles, M.E.
(2023). An Atlas of Variant Effects to understand the genome at nucleotide resolution. Genome biol,
Vol.24
(1),
p. 147.
show abstract
Sequencing has revealed hundreds of millions of human genetic variants, and continued efforts will only add to this variant avalanche. Insufficient information exists to interpret the effects of most variants, limiting opportunities for precision medicine and comprehension of genome function. A solution lies in experimental assessment of the functional effect of variants, which can reveal their biological and clinical impact. However, variant effect assays have generally been undertaken reactively for individual variants only after and, in most cases long after, their first observation. Now, multiplexed assays of variant effect can characterise massive numbers of variants simultaneously, yielding variant effect maps that reveal the function of every possible single nucleotide change in a gene or regulatory element. Generating maps for every protein encoding gene and regulatory element in the human genome would create an 'Atlas' of variant effect maps and transform our understanding of genetics and usher in a new era of nucleotide-resolution functional knowledge of the genome. An Atlas would reveal the fundamental biology of the human genome, inform human evolution, empower the development and use of therapeutics and maximize the utility of genomics for diagnosing and treating disease. The Atlas of Variant Effects Alliance is an international collaborative group comprising hundreds of researchers, technologists and clinicians dedicated to realising an Atlas of Variant Effects to help deliver on the promise of genomics..
Speight, B.
Hanson, H.
Turnbull, C.
Hardy, S.
Drummond, J.
Khorashad, J.
Wragg, C.
Page, P.
Parkin, N.W.
Rio-Machin, A.
Fitzgibbon, J.
Kulasekararaj, A.G.
Hamblin, A.
Talley, P.
McVeigh, T.P.
Snape, K.
Consensus Meeting Attendees,
(2023). Germline predisposition to haematological malignancies: Best practice consensus guidelines from the UK Cancer Genetics Group (UKCGG), CanGene-CanVar and the NHS England Haematological Oncology Working Group. Br j haematol,
Vol.201
(1),
pp. 25-34.
show abstract
full text
The implementation of whole genome sequencing and large somatic gene panels in haematological malignancies is identifying an increasing number of individuals with either potential or confirmed germline predisposition to haematological malignancy. There are currently no national or international best practice guidelines with respect to management of carriers of such variants or of their at-risk relatives. To address this gap, the UK Cancer Genetics Group (UKCGG), CanGene-CanVar and the NHS England Haematological Oncology Working Group held a workshop over two days on 28-29th April 2022, with the aim of establishing consensus guidelines on relevant clinical and laboratory pathways. The workshop focussed on the management of disease-causing germline variation in the following genes: DDX41, CEBPA, RUNX1, ANKRD26, ETV6, GATA2. Using a pre-workshop survey followed by structured discussion and in-meeting polling, we achieved consensus for UK best practice in several areas. In particular, high consensus was achieved on issues regarding standardised reporting, variant classification, multidisciplinary team working and patient support. The best practice recommendations from this meeting may be applicable to an expanding number of other genes in this setting..
Hill, W.
Lim, E.L.
Weeden, C.E.
Lee, C.
Augustine, M.
Chen, K.
Kuan, F.-.
Marongiu, F.
Evans, E.J.
Moore, D.A.
Rodrigues, F.S.
Pich, O.
Bakker, B.
Cha, H.
Myers, R.
van Maldegem, F.
Boumelha, J.
Veeriah, S.
Rowan, A.
Naceur-Lombardelli, C.
Karasaki, T.
Sivakumar, M.
De, S.
Caswell, D.R.
Nagano, A.
Black, J.R.
Martínez-Ruiz, C.
Ryu, M.H.
Huff, R.D.
Li, S.
Favé, M.-.
Magness, A.
Suárez-Bonnet, A.
Priestnall, S.L.
Lüchtenborg, M.
Lavelle, K.
Pethick, J.
Hardy, S.
McRonald, F.E.
Lin, M.-.
Troccoli, C.I.
Ghosh, M.
Miller, Y.E.
Merrick, D.T.
Keith, R.L.
Al Bakir, M.
Bailey, C.
Hill, M.S.
Saal, L.H.
Chen, Y.
George, A.M.
Abbosh, C.
Kanu, N.
Lee, S.-.
McGranahan, N.
Berg, C.D.
Sasieni, P.
Houlston, R.
Turnbull, C.
Lam, S.
Awadalla, P.
Grönroos, E.
Downward, J.
Jacks, T.
Carlsten, C.
Malanchi, I.
Hackshaw, A.
Litchfield, K.
TRACERx Consortium,
DeGregori, J.
Jamal-Hanjani, M.
Swanton, C.
(2023). Lung adenocarcinoma promotion by air pollutants. Nature,
Vol.616
(7955),
pp. 159-167.
show abstract
full text
A complete understanding of how exposure to environmental substances promotes cancer formation is lacking. More than 70 years ago, tumorigenesis was proposed to occur in a two-step process: an initiating step that induces mutations in healthy cells, followed by a promoter step that triggers cancer development1. Here we propose that environmental particulate matter measuring ≤2.5 μm (PM2.5), known to be associated with lung cancer risk, promotes lung cancer by acting on cells that harbour pre-existing oncogenic mutations in healthy lung tissue. Focusing on EGFR-driven lung cancer, which is more common in never-smokers or light smokers, we found a significant association between PM2.5 levels and the incidence of lung cancer for 32,957 EGFR-driven lung cancer cases in four within-country cohorts. Functional mouse models revealed that air pollutants cause an influx of macrophages into the lung and release of interleukin-1β. This process results in a progenitor-like cell state within EGFR mutant lung alveolar type II epithelial cells that fuels tumorigenesis. Ultradeep mutational profiling of histologically normal lung tissue from 295 individuals across 3 clinical cohorts revealed oncogenic EGFR and KRAS driver mutations in 18% and 53% of healthy tissue samples, respectively. These findings collectively support a tumour-promoting role for PM2.5 air pollutants and provide impetus for public health policy initiatives to address air pollution to reduce disease burden..
Hanson, H.
Kulkarni, A.
Loong, L.
Kavanaugh, G.
Torr, B.
Allen, S.
Ahmed, M.
Antoniou, A.C.
Cleaver, R.
Dabir, T.
Evans, D.G.
Golightly, E.
Jewell, R.
Kohut, K.
Manchanda, R.
Murray, A.
Murray, J.
Ong, K.-.
Rosenthal, A.N.
Woodward, E.R.
Eccles, D.M.
Turnbull, C.
Tischkowitz, M.
Consensus meeting attendees,
Lalloo, F.
(2023). UK consensus recommendations for clinical management of cancer risk for women with germline pathogenic variants in cancer predisposition genes: RAD51C, RAD51D, BRIP1 and PALB2. J med genet,
Vol.60
(5),
pp. 417-429.
show abstract
full text
Germline pathogenic variants (GPVs) in the cancer predisposition genes BRCA1, BRCA2, MLH1, MSH2, MSH6, BRIP1, PALB2, RAD51D and RAD51C are identified in approximately 15% of patients with ovarian cancer (OC). While there are clear guidelines around clinical management of cancer risk in patients with GPV in BRCA1, BRCA2, MLH1, MSH2 and MSH6, there are few guidelines on how to manage the more moderate OC risk in patients with GPV in BRIP1, PALB2, RAD51D and RAD51C, with clinical questions about appropriateness and timing of risk-reducing gynaecological surgery. Furthermore, while recognition of RAD51C and RAD51D as OC predisposition genes has been established for several years, an association with breast cancer (BC) has only more recently been described and clinical management of this risk has been unclear. With expansion of genetic testing of these genes to all patients with non-mucinous OC, new data on BC risk and improved estimates of OC risk, the UK Cancer Genetics Group and CanGene-CanVar project convened a 2-day meeting to reach a national consensus on clinical management of BRIP1, PALB2, RAD51D and RAD51C carriers in clinical practice. In this paper, we present a summary of the processes used to reach and agree on a consensus, as well as the key recommendations from the meeting..
Turnbull, C.
Boomsma, A.
Milte, R.
Stanton, T.R.
Hordacre, B.
(2023). Safety and Adverse Events following Non-invasive Electrical Brain Stimulation in Stroke: A Systematic Review. Top stroke rehabil,
Vol.30
(4),
pp. 355-367.
show abstract
BACKGROUND: Noninvasive electrical stimulation (ES) could have therapeutic potential in stroke recovery. However, there is no comprehensive evaluation of adverse events. This study systematically searched the literature to document frequency and prevalence of adverse events. A secondary aim was to explore associations between adverse events and ES parameters or participant characteristics.Methods: Databases were searched for studies evaluating ES in adults with stroke. All included studies were required to report on adverse events. Extracted data were: (1) study design; (2) adverse events; (3) participant characteristics; (4) ES parameters. RESULTS: Seventy-five studies were included. Adverse events were minor in nature. The most frequently reported adverse events were tingling (37.3% of papers), burning (18.7%), headaches (14.7%) and fatigue (14.7%). Cathodal stimulation was associated with greater frequency of itching (p = .02), intensities of 1-2 mA with increased tingling (p = .04) and discomfort (p = .03), and current density <0.4mA/cm2 with greater discomfort (p = .03). Tingling was the most prevalent adverse event (18.1% of participants), with prevalence data not differing between active and control conditions (all p ≥ 0.37). Individual participants were more likely to report adverse events with increasing current density (r = 0.99, p = .001). Two severe adverse events were noted (a seizure and percutaneous endoscopic gastrostomy placement). CONCLUSION: ES appears safe in people with stroke as reported adverse events were predominantly minor in nature. An adverse events questionnaire is proposed to enable a more comprehensive and nuanced analysis of the frequency and prevalence of adverse events..
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. Lancet oncol,
Vol.24
(6),
pp. 658-668.
show abstract
full text
BACKGROUND: 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). METHODS: 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. FINDINGS: 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. INTERPRETATION: 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. FUNDING: The Wellcome Trust..
Loong, L.
Huntley, C.
McRonald, F.
Santaniello, F.
Pethick, J.
Torr, B.
Allen, S.
Tulloch, O.
Goel, S.
Shand, B.
Rahman, T.
Luchtenborg, M.
Garrett, A.
Barber, R.
Bedenham, T.
Bourn, D.
Bradshaw, K.
Brooks, C.
Bruty, J.
Burghel, G.J.
Butler, S.
Buxton, C.
Callaway, A.
Callaway, J.
Drummond, J.
Durkie, M.
Field, J.
Jenkins, L.
McVeigh, T.P.
Mountford, R.
Nyanhete, R.
Petrides, E.
Robinson, R.
Scott, T.
Stinton, V.
Tellez, J.
Wallace, A.J.
Yarram-Smith, L.
Sahan, K.
Hallowell, N.
Eccles, D.M.
Pharoah, P.
Tischkowitz, M.
Antoniou, A.C.
Evans, D.G.
Lalloo, F.
Norbury, G.
Morris, E.
Burn, J.
Hardy, S.
Turnbull, C.
(2023). Germline mismatch repair (MMR) gene analyses from English NHS regional molecular genomics laboratories 1996-2020: development of a national resource of patient-level genomics laboratory records. J med genet,
Vol.60
(7),
pp. 669-678.
show abstract
full text
OBJECTIVE: To describe national patterns of National Health Service (NHS) analysis of mismatch repair (MMR) genes in England using individual-level data submitted to the National Disease Registration Service (NDRS) by the NHS regional molecular genetics laboratories. DESIGN: Laboratories submitted individual-level patient data to NDRS against a prescribed data model, including (1) patient identifiers, (2) test episode data, (3) per-gene results and (4) detected sequence variants. Individualised per-laboratory algorithms were designed and applied in NDRS to extract and map the data to the common data model. Laboratory-level MMR activity audit data from the Clinical Molecular Genetics Society/Association of Clinical Genomic Science were used to assess early years' missing data. RESULTS: Individual-level data from patients undergoing NHS MMR germline genetic testing were submitted from all 13 English laboratories performing MMR analyses, comprising in total 16 722 patients (9649 full-gene, 7073 targeted), with the earliest submission from 2000. The NDRS dataset is estimated to comprise >60% of NHS MMR analyses performed since inception of NHS MMR analysis, with complete national data for full-gene analyses for 2016 onwards. Out of 9649 full-gene tests, 2724 had an abnormal result, approximately 70% of which were (likely) pathogenic. Data linkage to the National Cancer Registry demonstrated colorectal cancer was the most frequent cancer type in which full-gene analysis was performed. CONCLUSION: The NDRS MMR dataset is a unique national pan-laboratory amalgamation of individual-level clinical and genomic patient data with pseudonymised identifiers enabling linkage to other national datasets. This growing resource will enable longitudinal research and can form the basis of a live national genomic disease registry..
Allen, I.
Hassan, H.
Sofianopoulou, E.
Eccles, D.
Turnbull, C.
Tischkowitz, M.
Pharoah, P.
Antoniou, A.C.
(2023). Risks of second non-breast primaries following breast cancer in women: a systematic review and meta-analysis. Breast cancer res,
Vol.25
(1),
p. 18.
show abstract
BACKGROUND: Second primary cancer incidence is rising among breast cancer survivors. We examined the risks of non-breast second primaries, in combination and at specific cancer sites, through a systematic review and meta-analysis. METHODS: We conducted a systematic search of PubMed, Embase, and Web of Science, seeking studies published by March 2022. We included studies that reported standardized incidence ratios (SIRs), with associated standard errors, assessing the combined risk of second non-breast primaries following breast cancer. We performed meta-analyses of combined second primary risks, stratifying by age, follow-up duration, and geographic region. We also assessed second primary risks at several specific sites, stratifying by age. The inverse variance method with DerSimonian-Laird estimators was used in all meta-analyses, assuming a random-effects model. Associated biases and study quality were evaluated using the Newcastle-Ottawa scale. RESULTS: One prospective and twenty-seven retrospective cohort studies were identified. SIRs for second non-breast primaries combined ranged from 0.84 to 1.84. The summary SIR estimate was 1.24 (95% CI 1.14-1.36, I2: 99%). This varied by age: the estimate was 1.59 (95% CI 1.36-1.85) when breast cancer was diagnosed before age 50, which was significantly higher than in women first diagnosed at 50 or over (SIR: 1.13, 95% CI 1.01-1.36, p for difference: < 0.001). SPC risks were also significantly higher when based on Asian, rather than European, registries (Asia-SIR: 1.47, 95% CI 1.29-1.67. Europe-SIR: 1.16, 95% CI 1.04-1.28). There were significantly increased risks of second thyroid (SIR: 1.89, 95% CI 1.49-2.38), corpus uteri (SIR: 1.84, 95% CI 1.53-2.23), ovary (SIR: 1.53, 95% CI 1.35-1.73), kidney (SIR: 1.43, 95% CI 1.17-1.73), oesophagus (SIR: 1.39, 95% CI 1.26-1.55), skin (melanoma) (SIR: 1.34, 95% CI 1.18-1.52), blood (leukaemia) (SIR: 1.30, 95% CI 1.17-1.45), lung (SIR: 1.25, 95% CI 1.03-1.51), stomach (SIR: 1.23, 95% CI 1.12-1.36) and bladder (SIR: 1.15, 95% CI 1.05-1.26) primaries. CONCLUSIONS: Breast cancer survivors are at significantly increased risk of second primaries at many sites. Risks are higher for those diagnosed with breast cancer before age 50 and in Asian breast cancer survivors compared to European breast cancer survivors. This study is limited by a lack of data on potentially confounding variables. The conclusions may inform clinical management decisions following breast cancer, although specific clinical recommendations lie outside the scope of this review..
Cook, S.
Pethick, J.
Kibbi, N.
Hollestein, L.
Lavelle, K.
de Vere Hunt, I.
Turnbull, C.
Rous, B.
Husain, A.
Burn, J.
Lüchtenborg, M.
Santaniello, F.
McRonald, F.
Hardy, S.
Linos, E.
Venables, Z.
Rajan, N.
(2023). Sebaceous carcinoma epidemiology, associated malignancies and Lynch/Muir-Torre syndrome screening in England from 2008 to 2018. J am acad dermatol,
Vol.89
(6),
pp. 1129-1135.
show abstract
BACKGROUND: Sebaceous carcinomas (SC) may be associated with the cancer predisposition syndrome Muir-Torre/Lynch syndrome (MTS/LS), identifiable by SC mismatch repair (MMR) screening; however, there is limited data on MMR status of SC. OBJECTIVE: To describe the epidemiology of SC, copresentation of other cancers, and population level frequency of MMR screening in SC. METHODS: A population-based retrospective cohort study of SC patients in the National Cancer Registration and Analysis Service in England. RESULTS: This study included 1077 SC cases (739 extraocular, 338 periocular). Age-standardized incidence rates (ASIR) were higher in men compared with women, 2.74 (95% CI, 2.52-9.69) per 1,000,000 person-years for men versus 1.47 person-years (95% CI, 1.4-1.62) for women. Of the patients, 19% (210/1077) developed at least one MTS/LS-associated malignancy. MMR immunohistochemical screening was performed in only 20% (220/1077) of SC tumors; of these, 32% (70/219) of tumors were MMR deficient. LIMITATIONS: Retrospective design. CONCLUSIONS: Incorporation of MMR screening into clinical practice guidelines for the management of SC will increase the opportunity for MTS/LS diagnoses, with implications for cancer surveillance, chemoprevention with aspirin, and immunotherapy treatment targeted to MTS/LS cancers..
Sud, A.
Horton, R.H.
Hingorani, A.D.
Tzoulaki, I.
Turnbull, C.
Houlston, R.S.
Lucassen, A.
(2023). Realistic expectations are key to realising the benefits of polygenic scores. Bmj,
Vol.380,
p. e073149.
show abstract
full text
We must not let enthusiasm around polygenic scores allow us to forget other factors that are bigger, more modifiable, and relevant for everyone, argue Amit Sud, Rachel Horton, and colleagues.
Garrett, A.
Loveday, C.
King, L.
Butler, S.
Robinson, R.
Horton, C.
Yussuf, A.
Choi, S.
Torr, B.
Durkie, M.
Burghel, G.J.
Drummond, J.
Berry, I.
Wallace, A.
Callaway, A.
Eccles, D.
Tischkowitz, M.
Tatton-Brown, K.
Snape, K.
McVeigh, T.
Izatt, L.
Woodward, E.R.
Burnichon, N.
Gimenez-Roqueplo, A.-.
Mazzarotto, F.
Whiffin, N.
Ware, J.
Hanson, H.
Pesaran, T.
LaDuca, H.
Buffet, A.
Maher, E.R.
Turnbull, C.
Cancer Variant Interpretation Group UK (CanVIG-UK),
(2022). Quantifying evidence toward pathogenicity for rare phenotypes: The case of succinate dehydrogenase genes, SDHB and SDHD. Genet med,
Vol.24
(1),
pp. 41-50.
show abstract
full text
PURPOSE: The weight of the evidence to attach to observation of a novel rare missense variant in SDHB or SDHD in individuals with the rare neuroendocrine tumors, pheochromocytomas and paragangliomas (PCC/PGL), is uncertain. METHODS: We compared the frequency of SDHB and SDHD very rare missense variants (VRMVs) in 6328 and 5847 cases of PCC/PGL, respectively, with that of population controls to generate a pan-gene VRMV likelihood ratio (LR). Via windowing analysis, we measured regional enrichments of VRMVs to calculate the domain-specific VRMV-LR (DS-VRMV-LR). We also calculated subphenotypic LRs for variant pathogenicity for various clinical, histologic, and molecular features. RESULTS: We estimated the pan-gene VRMV-LR to be 76.2 (54.8-105.9) for SDHB and 14.8 (8.7-25.0) for SDHD. Clustering analysis revealed an SDHB enriched region (ɑɑ 177-260, P = .001) for which the DS-VRMV-LR was 127.2 (64.9-249.4) and an SDHD enriched region (ɑɑ 70-114, P = .000003) for which the DS-VRMV-LR was 33.9 (14.8-77.8). Subphenotypic LRs exceeded 6 for invasive disease (SDHB), head-and-neck disease (SDHD), multiple tumors (SDHD), family history of PCC/PGL, loss of SDHB staining on immunohistochemistry, and succinate-to-fumarate ratio >97 (SDHB, SDHD). CONCLUSION: Using methodology generalizable to other gene-phenotype dyads, the LRs relating to rarity and phenotypic specificity for a single observation in PCC/PGL of a SDHB/SDHD VRMV can afford substantial evidence toward pathogenicity..
Westphalen, C.B.
Fine, A.D.
André, F.
Ganesan, S.
Heinemann, V.
Rouleau, E.
Turnbull, C.
Garcia Palacios, L.
Lopez, J.-.
Sokol, E.S.
Mateo, J.
(2022). Pan-cancer Analysis of Homologous Recombination Repair-associated Gene Alterations and Genome-wide Loss-of-Heterozygosity Score. Clin cancer res,
Vol.28
(7),
pp. 1412-1421.
show abstract
PURPOSE: To study associations across tumor types between genome-wide loss of heterozygosity (gLOH) and alterations in homologous recombination repair (HRR)-associated genes beyond BRCA1 and BRCA2. EXPERIMENTAL DESIGN: Genomic profiling using a targeted next-generation sequencing assay examining 324-465 genes (FoundationOne, FoundationOne Heme, and FoundationOne CDx; Foundation Medicine, Inc.) was performed in a cohort of 160,790 samples across different tumor types. Zygosity predictions and gLOH status were calculated and linked with alterations in 18 HRR-associated genes (BRCA1, BRCA2, PALB2, BARD1, ATR, ATRX, ATM, BAP1, RAD51B, RAD51C, RAD51D, BRIP1, NBN, CHEK1, CHEK2, FANCA, FANCC, MRE11) and other genomic features, using Fisher's exact test and Mann-Whitney U tests. RESULTS: We identified a strong correlation between elevated gLOH and biallelic alterations in a core set of HRR-associated genes beyond BRCA1 and BRCA2, such as BARD1, PALB2, FANCC, RAD51C, and RAD51D (particularly in breast, ovarian, pancreatic, and prostate cancer). Monoallelic/heterozygous alterations in HRR-associated genes were not associated with elevated gLOH. gLOH was also independently associated with TP53 loss. Co-occurrence of TP53 loss and alterations in HRR-associated genes, and combined loss of TP53-PTEN or TP53-RB1, was associated with a higher gLOH than each of the events separately. CONCLUSIONS: Biallelic alterations in core HRR-associated genes are frequent, strongly associated with elevated gLOH, and enriched in breast, ovarian, pancreatic, and prostate cancer. This analysis could inform the design of the next generation of clinical trials examining DNA repair-targeting agents, including PARP inhibitors..
Grasso, C.
Popovic, M.
Isaevska, E.
Lazzarato, F.
Fiano, V.
Zugna, D.
Pluta, J.
Weathers, B.
D'Andrea, K.
Almstrup, K.
Anson-Cartwright, L.
Bishop, D.T.
Chanock, S.J.
Chen, C.
Cortessis, V.K.
Dalgaard, M.D.
Daneshmand, S.
Ferlin, A.
Foresta, C.
Frone, M.N.
Gamulin, M.
Gietema, J.A.
Greene, M.H.
Grotmol, T.
Hamilton, R.J.
Haugen, T.B.
Hauser, R.
Karlsson, R.
Kiemeney, L.A.
Lessel, D.
Lista, P.
Lothe, R.A.
Loveday, C.
Meijer, C.
Nead, K.T.
Nsengimana, J.
Skotheim, R.I.
Turnbull, C.
Vaughn, D.J.
Wiklund, F.
Zheng, T.
Zitella, A.
Schwartz, S.M.
McGlynn, K.A.
Kanetsky, P.A.
Nathanson, K.L.
Richiardi, L.
(2022). Association Study between Polymorphisms in DNA Methylation-Related Genes and Testicular Germ Cell Tumor Risk. Cancer epidemiol biomarkers prev,
Vol.31
(9),
pp. 1769-1779.
show abstract
full text
BACKGROUND: Testicular germ cell tumors (TGCT), histologically classified as seminomas and nonseminomas, are believed to arise from primordial gonocytes, with the maturation process blocked when they are subjected to DNA methylation reprogramming. SNPs in DNA methylation machinery and folate-dependent one-carbon metabolism genes have been postulated to influence the proper establishment of DNA methylation. METHODS: In this pathway-focused investigation, we evaluated the association between 273 selected tag SNPs from 28 DNA methylation-related genes and TGCT risk. We carried out association analysis at individual SNP and gene-based level using summary statistics from the Genome Wide Association Study meta-analysis recently conducted by the international Testicular Cancer Consortium on 10,156 TGCT cases and 179,683 controls. RESULTS: In individual SNP analyses, seven SNPs, four mapping within MTHFR, were associated with TGCT risk after correction for multiple testing (q ≤ 0.05). Queries of public databases showed that three of these SNPs were associated with MTHFR changes in enzymatic activity (rs1801133) or expression level in testis tissue (rs12121543, rs1476413). Gene-based analyses revealed MTHFR (q = 8.4 × 10-4), methyl-CpG-binding protein 2 (MECP2; q = 2 × 10-3), and ZBTB4 (q = 0.03) as the top TGCT-associated genes. Stratifying by tumor histology, four MTHFR SNPs were associated with seminoma. In gene-based analysis MTHFR was associated with risk of seminoma (q = 2.8 × 10-4), but not with nonseminomatous tumors (q = 0.22). CONCLUSIONS: Genetic variants within MTHFR, potentially having an impact on the DNA methylation pattern, are associated with TGCT risk. IMPACT: This finding suggests that TGCT pathogenesis could be associated with the folate cycle status, and this relation could be partly due to hereditary factors..
Loong, L.
Cubuk, C.
Choi, S.
Allen, S.
Torr, B.
Garrett, A.
Loveday, C.
Durkie, M.
Callaway, A.
Burghel, G.J.
Drummond, J.
Robinson, R.
Berry, I.R.
Wallace, A.
Eccles, D.M.
Tischkowitz, M.
Ellard, S.
Ware, J.S.
Hanson, H.
Turnbull, C.
CanVIG-UK,
(2022). Quantifying prediction of pathogenicity for within-codon concordance (PM5) using 7541 functional classifications of BRCA1 and MSH2 missense variants. Genet med,
Vol.24
(3),
pp. 552-563.
show abstract
full text
PURPOSE: Conditions and thresholds applied for evidence weighting of within-codon concordance (PM5) for pathogenicity vary widely between laboratories and expert groups. Because of the sparseness of available clinical classifications, there is little evidence for variation in practice. METHODS: We used as a truthset 7541 dichotomous functional classifications of BRCA1 and MSH2, spanning 311 codons of BRCA1 and 918 codons of MSH2, generated from large-scale functional assays that have been shown to correlate excellently with clinical classifications. We assessed PM5 at 5 stringencies with incorporation of 8 in silico tools. For each analysis, we quantified a positive likelihood ratio (pLR, true positive rate/false positive rate), the predictive value of PM5-lookup in ClinVar compared with the functional truthset. RESULTS: pLR was 16.3 (10.6-24.9) for variants for which there was exactly 1 additional colocated deleterious variant on ClinVar, and the variant under examination was equally or more damaging when analyzed using BLOSUM62. pLR was 71.5 (37.8-135.3) for variants for which there were 2 or more colocated deleterious ClinVar variants, and the variant under examination was equally or more damaging than at least 1 colocated variant when analyzed using BLOSUM62. CONCLUSION: These analyses support the graded use of PM5, with potential to use it at higher evidence weighting where more stringent criteria are met..
Pujol, P.
Yauy, K.
Coffy, A.
Duforet-Frebourg, N.
Gabteni, S.
Daurès, J.-.
Penault Llorca, F.
Thomas, F.
Hughes, K.
Turnbull, C.
Galibert, V.
Rideau, C.
Corsini, C.
Collet, L.
You, B.
Geneviève, D.
Philippe, N.
(2022). Predominance of BRCA2 Mutation and Estrogen Receptor Positivity in Unselected Breast Cancer with BRCA1 or BRCA2 Mutation. Cancers (basel),
Vol.14
(13).
show abstract
full text
BACKGROUND: Poly(ADP-ribose) polymerase 1 inhibitor (PARPi) agents can improve progression-free survival of patients with breast cancer who carry a germline BRCA1 or BRCA2 pathogenic or likely pathogenic variant (gBRCA) in both the metastatic and adjuvant setting. Therefore, we need to reassess the frequency of gBRCA1 and gBRCA2 in order to redefine the criteria for women and tumor phenotype that should be tested. OBJECTIVE: We studied the relative distribution of gBRCA1 and gBRCA2 in unselected populations of women with breast cancer and in unaffected individuals. We also analyzed the proportion of estrogen receptor (ER)-positive (ER+) tumors in unselected breast cancer patients with gBRCA. DESIGN: We performed a meta-analysis of studies of unselected breast cancer that analyzed the relative contribution of gBRCA1 versus gBRCA2 among unselected breast cancer cases in gBRCA carriers. We then performed a meta-analysis of gBRCA carriage in unaffected individuals from genome-wide population studies, the gnomAD databank, and case-control studies. RESULTS: The BRCA2 gene was involved in 54% of breast cancer cases in unselected patients with gBRCA (n = 108,699) and 60% of unaffected individuals (n = 238,973) as compared with 38% of the largest gBRCA family cohort (n = 29,700). The meta-analysis showed that 1.66% (95% CI 1.08-2.54) and 1.71% (95% CI 1.33-2.2) of unselected breast cancer patients carried gBRCA1 and gBRCA2, respectively. In a population of unaffected individuals, the frequency of heterozygosity for gBRCA1 and gBRCA2 was estimated at 1/434 and 1/288, respectively. Nearly 0.5% of unaffected individuals in the studied populations carried a gBRCA. Carriage of a gBRCA was 2.5% for patients with ER+ tumors (95% CI 1.5-4.1) and 5.7% (95% CI 5.1-6.2) for those with ER- tumors. Overall, 58% of breast tumors occurring in women carrying a gBRCA were ER+ (n = 86,870). CONCLUSIONS: This meta-analysis showed that gBRCA2 carriage is predominant in unselected breast cancer patients and unaffected individuals. ER+ tumors among women with gBRCA-related breast cancer are predominant and have been underestimated. Because PARPi agents improve progression-free survival with ER+ gBRCA breast cancer in most clinical trials, breast cancer should be considered, regardless of ER status, for BRCA1/2 screening for therapeutic purposes..
Palles, C.
West, H.D.
Chew, E.
Galavotti, S.
Flensburg, C.
Grolleman, J.E.
Jansen, E.A.
Curley, H.
Chegwidden, L.
Arbe-Barnes, E.H.
Lander, N.
Truscott, R.
Pagan, J.
Bajel, A.
Sherwood, K.
Martin, L.
Thomas, H.
Georgiou, D.
Fostira, F.
Goldberg, Y.
Adams, D.J.
van der Biezen, S.A.
Christie, M.
Clendenning, M.
Thomas, L.E.
Deltas, C.
Dimovski, A.J.
Dymerska, D.
Lubinski, J.
Mahmood, K.
van der Post, R.S.
Sanders, M.
Weitz, J.
Taylor, J.C.
Turnbull, C.
Vreede, L.
van Wezel, T.
Whalley, C.
Arnedo-Pac, C.
Caravagna, G.
Cross, W.
Chubb, D.
Frangou, A.
Gruber, A.J.
Kinnersley, B.
Noyvert, B.
Church, D.
Graham, T.
Houlston, R.
Lopez-Bigas, N.
Sottoriva, A.
Wedge, D.
Genomics England Research Consortium,
CORGI Consortium,
WGS500 Consortium,
Jenkins, M.A.
Kuiper, R.P.
Roberts, A.W.
Cheadle, J.P.
Ligtenberg, M.J.
Hoogerbrugge, N.
Koelzer, V.H.
Rivas, A.D.
Winship, I.M.
Ponte, C.R.
Buchanan, D.D.
Power, D.G.
Green, A.
Tomlinson, I.P.
Sampson, J.R.
Majewski, I.J.
de Voer, R.M.
(2022). Germline MBD4 deficiency causes a multi-tumor predisposition syndrome. Am j hum genet,
Vol.109
(5),
pp. 953-960.
show abstract
full text
We report an autosomal recessive, multi-organ tumor predisposition syndrome, caused by bi-allelic loss-of-function germline variants in the base excision repair (BER) gene MBD4. We identified five individuals with bi-allelic MBD4 variants within four families and these individuals had a personal and/or family history of adenomatous colorectal polyposis, acute myeloid leukemia, and uveal melanoma. MBD4 encodes a glycosylase involved in repair of G:T mismatches resulting from deamination of 5'-methylcytosine. The colorectal adenomas from MBD4-deficient individuals showed a mutator phenotype attributable to mutational signature SBS1, consistent with the function of MBD4. MBD4-deficient polyps harbored somatic mutations in similar driver genes to sporadic colorectal tumors, although AMER1 mutations were more common and KRAS mutations less frequent. Our findings expand the role of BER deficiencies in tumor predisposition. Inclusion of MBD4 in genetic testing for polyposis and multi-tumor phenotypes is warranted to improve disease management..
Loong, L.
Garrett, A.
Allen, S.
Choi, S.
Durkie, M.
Callaway, A.
Drummond, J.
Burghel, G.J.
Robinson, R.
Torr, B.
Berry, I.R.
Wallace, A.J.
Eccles, D.M.
Ellard, S.
Baple, E.
Evans, D.G.
Woodward, E.R.
Kulkarni, A.
Lalloo, F.
Tischkowitz, M.
Lucassen, A.
Hanson, H.
Turnbull, C.
CanVIG-UK,
(2022). Reclassification of clinically-detected sequence variants: Framework for genetic clinicians and clinical scientists by CanVIG-UK (Cancer Variant Interpretation Group UK). Genet med,
Vol.24
(9),
pp. 1867-1877.
show abstract
full text
PURPOSE: Variant classifications may change over time, driven by emergence of fresh or contradictory evidence or evolution in weighing or combination of evidence items. For variant classifications above the actionability threshold, which is classification of likely pathogenic or pathogenic, clinical actions may be irreversible, such as risk-reducing surgery or prenatal interventions. Variant reclassification up or down across the actionability threshold can therefore have significant clinical consequences. Laboratory approaches to variant reinterpretation and reclassification vary widely. METHODS: Cancer Variant Interpretation Group UK is a multidisciplinary network of clinical scientists and genetic clinicians from across the 24 Molecular Diagnostic Laboratories and Clinical Genetics Services of the United Kingdom (NHS) and Republic of Ireland. We undertook surveys, polls, and national meetings of Cancer Variant Interpretation Group UK to evaluate opinions about clinical and laboratory management regarding variant reclassification. RESULTS: We generated a consensus framework on variant reclassification applicable to cancer susceptibility genes and other clinical areas, which provides explicit recommendations for clinical and laboratory management of variant reclassification scenarios on the basis of the nature of the new evidence, the magnitude of evidence shift, and the final classification score. CONCLUSION: In this framework, clinical and laboratory resources are targeted for maximal clinical effect and minimal patient harm, as appropriate to all resource-constrained health care settings..
Fallowfield, L.
Solis-Trapala, I.
Starkings, R.
May, S.
Matthews, L.
Eccles, D.
Evans, D.G.
Turnbull, C.
Crawford, G.
Jenkins, V.
(2022). Talking about Risk, UncertaintieS of Testing IN Genetics (TRUSTING): development and evaluation of an educational programme for healthcare professionals about BRCA1 & BRCA2 testing. Br j cancer,
Vol.127
(6),
pp. 1116-1122.
show abstract
BACKGROUND: Mainstreaming of germline testing demands that all healthcare professionals have good communication skills, but few have genetic testing and counselling experience. We developed and evaluated educational workshops-Talking about Risk & UncertaintieS of Testing IN Genetics (TRUSTING). Contents included: presentations and exercises, an interview with a geneticist about BRCA testing, screening and prevention implications, filmed interactions between surgeons, a genetic counsellor and geneticists with a fictitious family (proband had a BRCA2 pathogenic variant with triple-negative breast cancer, her older sister-BRCA2 heterozygous, and cousin-negative for BRCA2 variant). METHODS: Twenty-one surgeons, 5 oncologists, 18 nurses and 9 genetic counsellors participated. Knowledge (18 item MCQ), communication skills (responses to 6 questions from proband and relatives) and self-confidence (discussing 9 genetic testing issues) were assessed pre- and post workshop. RESULTS: Knowledge scores improved significantly post workshop (mean change = 7.06; 95% confidence interval (CI) 6.37-7.74; P < 0.001), as did communication (mean change = 5.38; 95% CI 4.37-6.38; P < 0.001) and self-confidence (P < 0.001). DISCUSSION: Healthcare professionals' knowledge and self-confidence when discussing the risks and uncertainties in genetics are often poor. TRUSTING workshops significantly enhanced attendees' navigation of communication difficulties encountered and will be rolled out more widely..
Allen, I.
Hassan, H.
Sofianopoulou, E.
Eccles, D.
Turnbull, C.
Tischkowitz, M.
Pharoah, P.
Antoniou, A.C.
(2022). Risk of developing a second primary cancer in male breast cancer survivors: a systematic review and meta-analysis. Br j cancer,
Vol.127
(9),
pp. 1660-1669.
show abstract
BACKGROUND: With increasing survival after cancer diagnoses, second primary cancers (SPCs) are becoming more prevalent. We investigated the incidence and site of non-breast SPC risks following male breast cancer (BC). METHODS: PubMed, Embase and Web of Science were systematically searched for studies reporting standardised incidence ratios (SIRs) for SPCs published by March 2022. Meta-analyses used the generic inverse-variance method, assuming a random-effects model. We evaluated SIRs for overall SPCs, site-specific risks, by age at BC onset, time since BC onset and geographic region. We assessed study quality using routine techniques. RESULTS: Eight population-based retrospective cohort studies were identified. SIRs ranged from 1.05 to 2.17. The summary SIR estimate was 1.27 (95% CI: 1.03-1.56, I2: 86%), and there were increased colorectal (SIR: 1.29, 95% CI: 1.03-1.61), pancreatic (SIR: 1.64, 95% CI: 1.05-2.55) and thyroid (SIR: 5.58, 95% CI: 1.04-30.05) SPC risks. When an outlying study was excluded, the summary SIR for men diagnosed with BC before age 50 was 1.50 (95% CI: 1.21-1.85), significantly higher than men diagnosed at older ages (SIR: 1.14, 95% CI: 0.98-1.33). CONCLUSIONS: Male BC survivors are at elevated risks of developing second primary colorectal, pancreatic and thyroid cancers. The estimates may assist their clinical management and guide decisions on genetic testing..
Torr, B.
Jones, C.
Choi, S.
Allen, S.
Kavanaugh, G.
Hamill, M.
Garrett, A.
MacMahon, S.
Loong, L.
Reay, A.
Yuan, L.
Valganon Petrizan, M.
Monson, K.
Perry, N.
Fallowfield, L.
Jenkins, V.
Gold, R.
Taylor, A.
Gabe, R.
Wiggins, J.
Lucassen, A.
Manchanda, R.
Gandhi, A.
George, A.
Hubank, M.
Kemp, Z.
Evans, D.G.
Bremner, S.
Turnbull, C.
(2022). A digital pathway for genetic testing in UK NHS patients with cancer: BRCA-DIRECT randomised study internal pilot. J med genet,
Vol.59
(12),
pp. 1179-1188.
show abstract
full text
BACKGROUND: Germline genetic testing affords multiple opportunities for women with breast cancer, however, current UK NHS models for delivery of germline genetic testing are clinician-intensive and only a minority of breast cancer cases access testing. METHODS: We designed a rapid, digital pathway, supported by a genetics specialist hotline, for delivery of germline testing of BRCA1/BRCA2/PALB2 (BRCA-testing), integrated into routine UK NHS breast cancer care. We piloted the pathway, as part of the larger BRCA-DIRECT study, in 130 unselected patients with breast cancer and gathered preliminary data from a randomised comparison of delivery of pretest information digitally (fully digital pathway) or via telephone consultation with a genetics professional (partially digital pathway). RESULTS: Uptake of genetic testing was 98.4%, with good satisfaction reported for both the fully and partially digital pathways. Similar outcomes were observed in both arms regarding patient knowledge score and anxiety, with <5% of patients contacting the genetics specialist hotline. All progression criteria established for continuation of the study were met. CONCLUSION: Pilot data indicate preliminary demonstration of feasibility and acceptability of a fully digital pathway for BRCA-testing and support proceeding to a full powered study for evaluation of non-inferiority of the fully digital pathway, detailed quantitative assessment of outcomes and operational economic analyses. TRIAL REGISTRATION NUMBER: ISRCTN87845055..
Loveday, C.
Garrett, A.
Law, P.
Hanks, S.
Poyastro-Pearson, E.
Adlard, J.W.
Barwell, J.
Berg, J.
Brady, A.F.
Brewer, C.
Chapman, C.
Cook, J.
Davidson, R.
Donaldson, A.
Douglas, F.
Greenhalgh, L.
Henderson, A.
Izatt, L.
Kumar, A.
Lalloo, F.
Miedzybrodzka, Z.
Morrison, P.J.
Paterson, J.
Porteous, M.
Rogers, M.T.
Walker, L.
Breast and Ovarian Cancer Susceptibility Collaboration,
Eccles, D.
Evans, D.G.
Snape, K.
Hanson, H.
Houlston, R.S.
Turnbull, C.
(2022). Analysis of rare disruptive germline mutations in 2135 enriched BRCA-negative breast cancers excludes additional high-impact susceptibility genes. Ann oncol,
Vol.33
(12),
pp. 1318-1327.
show abstract
full text
BACKGROUND: Breast cancer has a significant heritable basis, of which ∼60% remains unexplained. Testing for BRCA1/BRCA2 offers useful discrimination of breast cancer risk within families, and identification of additional breast cancer susceptibility genes could offer clinical utility. PATIENTS AND METHODS: We included 2135 invasive breast cancer cases recruited via the Breast and Ovarian Cancer Susceptibility study, a retrospective UK study of familial breast cancer. ELIGIBILITY CRITERIA: female, BRCA-negative, white European ethnicity, and one of: (i) breast cancer family history, (ii) bilateral disease, (iii) young age of onset (<30 years), and (iv) concomitant ovarian cancer. We undertook exome sequencing of cases and carried out gene-level burden testing of rare damaging variants against those from 51 377 ethnicity-matched population controls from gnomAD. RESULTS: 159/2135 (7.4%) cases had a qualifying variant in an established breast cancer susceptibility gene, with minimal evidence of signal in other cancer susceptibility genes. Known breast cancer susceptibility genes PALB2, CHEK2, and ATM were the only genes to retain statistical significance after correcting for multiple testing. Due to the enrichment of hereditary cases in the series, we had good power (>80%) to detect a gene of BRCA1-like risk [odds ratio (OR) = 10.6] down to a population minor allele frequency of 4.6 × 10-5 (1 in 10 799, less than one-tenth that of BRCA1)and of PALB2-like risk (OR = 5.0) down to a population minor allele frequency of 2.8 × 10-4 (1 in 1779, less than half that of PALB2). Power was lower for identification of novel moderate penetrance genes (OR = 2-3) like CHEK2 and ATM. CONCLUSIONS: This is the largest case-control whole-exome analysis of enriched breast cancer published to date. Whilst additional breast cancer susceptibility genes likely exist, those of high penetrance are likely to be of very low mutational frequency. Contention exists regarding the clinical utility of such genes..
Zhang, P.
Kitchen-Smith, I.
Xiong, L.
Stracquadanio, G.
Brown, K.
Richter, P.H.
Wallace, M.D.
Bond, E.
Sahgal, N.
Moore, S.
Nornes, S.
De Val, S.
Surakhy, M.
Sims, D.
Wang, X.
Bell, D.A.
Zeron-Medina, J.
Jiang, Y.
Ryan, A.J.
Selfe, J.L.
Shipley, J.
Kar, S.
Pharoah, P.D.
Loveday, C.
Jansen, R.
Grochola, L.F.
Palles, C.
Protheroe, A.
Millar, V.
Ebner, D.V.
Pagadala, M.
Blagden, S.P.
Maughan, T.S.
Domingo, E.
Tomlinson, I.
Turnbull, C.
Carter, H.
Bond, G.L.
(2021). Germline and Somatic Genetic Variants in the p53 Pathway Interact to Affect Cancer Risk, Progression, and Drug Response. Cancer res,
Vol.81
(7),
pp. 1667-1680.
show abstract
full text
Insights into oncogenesis derived from cancer susceptibility loci (SNP) hold the potential to facilitate better cancer management and treatment through precision oncology. However, therapeutic insights have thus far been limited by our current lack of understanding regarding both interactions of these loci with somatic cancer driver mutations and their influence on tumorigenesis. For example, although both germline and somatic genetic variation to the p53 tumor suppressor pathway are known to promote tumorigenesis, little is known about the extent to which such variants cooperate to alter pathway activity. Here we hypothesize that cancer risk-associated germline variants interact with somatic TP53 mutational status to modify cancer risk, progression, and response to therapy. Focusing on a cancer risk SNP (rs78378222) with a well-documented ability to directly influence p53 activity as well as integration of germline datasets relating to cancer susceptibility with tumor data capturing somatically-acquired genetic variation provided supportive evidence for this hypothesis. Integration of germline and somatic genetic data enabled identification of a novel entry point for therapeutic manipulation of p53 activities. A cluster of cancer risk SNPs resulted in increased expression of prosurvival p53 target gene KITLG and attenuation of p53-mediated responses to genotoxic therapies, which were reversed by pharmacologic inhibition of the prosurvival c-KIT signal. Together, our results offer evidence of how cancer susceptibility SNPs can interact with cancer driver genes to affect cancer progression and identify novel combinatorial therapies. SIGNIFICANCE: These results offer evidence of how cancer susceptibility SNPs can interact with cancer driver genes to affect cancer progression and present novel therapeutic targets..
Turnbull, C.
(2021). Effect of COVID-19 on colorectal cancer care in England. Lancet gastroenterol hepatol,
Vol.6
(3),
pp. 152-154.
full text
Pujol, P.
Barberis, M.
Beer, P.
Friedman, E.
Piulats, J.M.
Capoluongo, E.D.
Garcia Foncillas, J.
Ray-Coquard, I.
Penault-Llorca, F.
Foulkes, W.D.
Turnbull, C.
Hanson, H.
Narod, S.
Arun, B.K.
Aapro, M.S.
Mandel, J.-.
Normanno, N.
Lambrechts, D.
Vergote, I.
Anahory, M.
Baertschi, B.
Baudry, K.
Bignon, Y.-.
Bollet, M.
Corsini, C.
Cussenot, O.
De la Motte Rouge, T.
Duboys de Labarre, M.
Duchamp, F.
Duriez, C.
Fizazi, K.
Galibert, V.
Gladieff, L.
Gligorov, J.
Hammel, P.
Imbert-Bouteille, M.
Jacot, W.
Kogut-Kubiak, T.
Lamy, P.-.
Nambot, S.
Neuzillet, Y.
Olschwang, S.
Rebillard, X.
Rey, J.-.
Rideau, C.
Spano, J.-.
Thomas, F.
Treilleux, I.
Vandromme, M.
Vendrell, J.
Vintraud, M.
Zarca, D.
Hughes, K.S.
Alés Martínez, J.E.
(2021). Clinical practice guidelines for BRCA1 and BRCA2 genetic testing. Eur j cancer,
Vol.146,
pp. 30-47.
show abstract
full text
BRCA1 and BRCA2 gene pathogenic variants account for most hereditary breast cancer and are increasingly used to determine eligibility for PARP inhibitor (PARPi) therapy of BRCA-related cancer. Because issues of BRCA testing in clinical practice now overlap with both preventive and therapeutic management, updated and comprehensive practice guidelines for BRCA genotyping are needed. The integrative recommendations for BRCA testing presented here aim to (1) identify individuals who may benefit from genetic counselling and risk-reducing strategies; (2) update germline and tumour-testing indications for PARPi-approved therapies; (3) provide testing recommendations for personalised management of early and metastatic breast cancer; and (4) address the issues of rapid process and tumour analysis. An international group of experts, including geneticists, medical and surgical oncologists, pathologists, ethicists and patient representatives, was commissioned by the French Society of Predictive and Personalised Medicine (SFMPP). The group followed a methodology based on specific formal guidelines development, including (1) evaluating the likelihood of BRCAm from a combined systematic review of the literature, risk assessment models and expert quotations, and (2) therapeutic values of BRCAm status for PARPi therapy in BRCA-related cancer and for management of early and advanced breast cancer. These international guidelines may help clinicians comprehensively update and standardise BRCA testing practices..
Odor, P.M.
Bampoe, S.
Moonesinghe, S.R.
Andrade, J.
Pandit, J.J.
Lucas, D.N.
Pan-London Perioperative Audit and Research Network (PLAN), for the DREAMY Investigators Group,
(2021). General anaesthetic and airway management practice for obstetric surgery in England: a prospective, multicentre observational study. Anaesthesia,
Vol.76
(4),
pp. 460-471.
show abstract
full text
There are no current descriptions of general anaesthesia characteristics for obstetric surgery, despite recent changes to patient baseline characteristics and airway management guidelines. This analysis of data from the direct reporting of awareness in maternity patients' (DREAMY) study of accidental awareness during obstetric anaesthesia aimed to describe practice for obstetric general anaesthesia in England and compare with earlier surveys and best-practice recommendations. Consenting patients who received general anaesthesia for obstetric surgery in 72 hospitals from May 2017 to August 2018 were included. Baseline characteristics, airway management, anaesthetic techniques and major complications were collected. Descriptive analysis, binary logistic regression modelling and comparisons with earlier data were conducted. Data were collected from 3117 procedures, including 2554 (81.9%) caesarean deliveries. Thiopental was the induction drug in 1649 (52.9%) patients, compared with propofol in 1419 (45.5%). Suxamethonium was the neuromuscular blocking drug for tracheal intubation in 2631 (86.1%), compared with rocuronium in 367 (11.8%). Difficult tracheal intubation was reported in 1 in 19 (95%CI 1 in 16-22) and failed intubation in 1 in 312 (95%CI 1 in 169-667). Obese patients were over-represented compared with national baselines and associated with difficult, but not failed intubation. There was more evidence of change in practice for induction drugs (increased use of propofol) than neuromuscular blocking drugs (suxamethonium remains the most popular). There was evidence of improvement in practice, with increased monitoring and reversal of neuromuscular blockade (although this remains suboptimal). Despite a high risk of difficult intubation in this population, videolaryngoscopy was rarely used (1.9%)..
Garrett, A.
Durkie, M.
Callaway, A.
Burghel, G.J.
Robinson, R.
Drummond, J.
Torr, B.
Cubuk, C.
Berry, I.R.
Wallace, A.J.
Ellard, S.
Eccles, D.M.
Tischkowitz, M.
Hanson, H.
Turnbull, C.
CanVIG-UK,
(2021). Combining evidence for and against pathogenicity for variants in cancer susceptibility genes: CanVIG-UK consensus recommendations. J med genet,
Vol.58
(5),
pp. 297-304.
show abstract
full text
Accurate classification of variants in cancer susceptibility genes (CSGs) is key for correct estimation of cancer risk and management of patients. Consistency in the weighting assigned to individual elements of evidence has been much improved by the American College of Medical Genetics (ACMG) 2015 framework for variant classification, UK Association for Clinical Genomic Science (UK-ACGS) Best Practice Guidelines and subsequent Cancer Variant Interpretation Group UK (CanVIG-UK) consensus specification for CSGs. However, considerable inconsistency persists regarding practice in the combination of evidence elements. CanVIG-UK is a national subspecialist multidisciplinary network for cancer susceptibility genomic variant interpretation, comprising clinical scientist and clinical geneticist representation from each of the 25 diagnostic laboratories/clinical genetic units across the UK and Republic of Ireland. Here, we summarise the aggregated evidence elements and combinations possible within different variant classification schemata currently employed for CSGs (ACMG, UK-ACGS, CanVIG-UK and ClinGen gene-specific guidance for PTEN, TP53 and CDH1). We present consensus recommendations from CanVIG-UK regarding (1) consistent scoring for combinations of evidence elements using a validated numerical 'exponent score' (2) new combinations of evidence elements constituting likely pathogenic' and 'pathogenic' classification categories, (3) which evidence elements can and cannot be used in combination for specific variant types and (4) classification of variants for which there are evidence elements for both pathogenicity and benignity..
George, A.
Turnbull, C.
(2021). Tumor-only sequencing for oncology management: Germline-focused analysis and implications. Genes chromosomes cancer,
Vol.60
(5),
pp. 352-357.
show abstract
full text
As patients are now routinely having large somatic genomic testing panels undertaken as part of routine management, there is the rising likelihood of uncovering the presence of a germline pathogenic variant. This may be found on testing undertaken on plasma (ctDNA) or tissue. This has led to the need for clear guidelines for oncologists about how to manage such results, including which variants require validation, how this should be undertaken, and what potential problems may arise. This requires an understanding of the limits of testing, and the pitfalls that may be encountered. In this review, we assess the frequency of detecting germline variants through tumor-only sequencing, the necessary considerations for such information to be analyzed and the role of the molecular tumor board in considering results. We assess the additional considerations for interpretation of the underlying tumor, use of ctDNA or tissue for testing, clonal hematopoiesis, and hypermutation..
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. Gut,
Vol.70
(6),
pp. 1053-1060.
show abstract
full text
OBJECTIVE: 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. DESIGN: 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. RESULTS: 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%. CONCLUSIONS: 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..
Odor, P.M.
Bampoe, S.
Lucas, D.N.
Moonesinghe, S.R.
Andrade, J.
Pandit, J.J.
Pan-London Peri-operative Audit and Research Network (PLAN), for the DREAMY Investigators Group,
(2021). Incidence of accidental awareness during general anaesthesia in obstetrics: a multicentre, prospective cohort study. Anaesthesia,
Vol.76
(6),
pp. 759-776.
show abstract
full text
General anaesthesia for obstetric surgery has distinct characteristics that may contribute towards a higher risk of accidental awareness during general anaesthesia. The primary aim of this study was to investigate the incidence, experience and psychological implications of unintended conscious awareness during general anaesthesia in obstetric patients. From May 2017 to August 2018, 3115 consenting patients receiving general anaesthesia for obstetric surgery in 72 hospitals in England were recruited to the study. Patients received three repetitions of standardised questioning over 30 days, with responses indicating memories during general anaesthesia that were verified using interviews and record interrogation. A total of 12 patients had certain/probable or possible awareness, an incidence of 1 in 256 (95%CI 149-500) for all obstetric surgery. The incidence was 1 in 212 (95%CI 122-417) for caesarean section surgery. Distressing experiences were reported by seven (58.3%) patients, paralysis by five (41.7%) and paralysis with pain by two (16.7%). Accidental awareness occurred during induction and emergence in nine (75%) of the patients who reported awareness. Factors associated with accidental awareness during general anaesthesia were: high BMI (25-30 kg.m-2 ); low BMI (<18.5 kg.m-2 ); out-of-hours surgery; and use of ketamine or thiopental for induction. Standardised psychological impact scores at 30 days were significantly higher in awareness patients (median (IQR [range]) 15 (2.7-52.0 [2-56]) than in patients without awareness 3 (1-9 [0-64]), p = 0.010. Four patients had a provisional diagnosis of post-traumatic stress disorder. We conclude that direct postoperative questioning reveals high rates of accidental awareness during general anaesthesia for obstetric surgery, which has implications for anaesthetic practice, consent and follow-up..
Brown, D.W.
Lan, Q.
Rothman, N.
Pluta, J.
Almstrup, K.
Dalgaard, M.D.
Greene, M.H.
Grotmol, T.
Loveday, C.
Schwartz, S.M.
Turnbull, C.
Wiklund, F.
Kanetsky, P.A.
Nathanson, K.L.
McGlynn, K.A.
Machiela, M.J.
Testicular Cancer Consortium,
(2021). Genetically Inferred Telomere Length and Testicular Germ Cell Tumor Risk. Cancer epidemiol biomarkers prev,
Vol.30
(6),
pp. 1275-1278.
show abstract
BACKGROUND: Studies evaluating the association between peripheral blood leukocyte telomere length (LTL) and testicular germ cell tumor (TGCT) risk have produced conflicting results. METHODS: Using available genotype data from the Testicular Cancer Consortium (TECAC), polygenic risk score and Mendelian randomization analyses of genetic variants previously associated with LTL were used to assess potential etiologic associations between telomere length and TGCT risk. RESULTS: Genetically inferred telomere length was not associated with TGCT risk among 2,049 cases and 6,921 controls with individual-level genotype data (OR, 1.02; 95% confidence interval, 0.97-1.07). Mendelian randomization analyses using summary statistic data further indicated no evidence for an association between telomere length and TGCT risk among all available TECAC participants (3,558 cases and 13,971 controls). CONCLUSIONS: Our analyses in the largest molecular genetic testicular cancer study to date provide no evidence for an association between genetically inferred peripheral blood LTL and TGCT risk. IMPACT: The lack of evidence for an overall association indicates that peripheral blood LTL is likely not a strong biomarker for TGCT risk..
Hartley, T.
Lane, N.D.
Steer, J.
Elliott, M.W.
Sovani, M.P.
Curtis, H.J.
Fuller, E.R.
Murphy, P.B.
Shrikrishna, D.
Lewis, K.E.
Ward, N.R.
Turnbull, C.D.
Hart, N.
Bourke, S.C.
(2021). The Noninvasive Ventilation Outcomes (NIVO) score: prediction of in-hospital mortality in exacerbations of COPD requiring assisted ventilation. Eur respir j,
Vol.58
(2).
show abstract
full text
INTRODUCTION: Acute exacerbations of COPD (AECOPD) complicated by acute (acidaemic) hypercapnic respiratory failure (AHRF) requiring ventilation are common. When applied appropriately, ventilation substantially reduces mortality. Despite this, there is evidence of poor practice and prognostic pessimism. A clinical prediction tool could improve decision making regarding ventilation, but none is routinely used. METHODS: Consecutive patients admitted with AECOPD and AHRF treated with assisted ventilation (principally noninvasive ventilation) were identified in two hospitals serving differing populations. Known and potential prognostic indices were identified a priori. A prediction tool for in-hospital death was derived using multivariable regression analysis. Prospective, external validation was performed in a temporally separate, geographically diverse 10-centre study. The trial methodology adhered to TRIPOD (Transparent Reporting of a Multivariable Prediction Model for Individual Prognosis or Diagnosis) recommendations. RESULTS: Derivation cohort: n=489, in-hospital mortality 25.4%; validation cohort: n=733, in-hospital mortality 20.1%. Using six simple categorised variables (extended Medical Research Council Dyspnoea score 1-4/5a/5b, time from admission to acidaemia >12 h, pH <7.25, presence of atrial fibrillation, Glasgow coma scale ≤14 and chest radiograph consolidation), a simple scoring system with strong prediction of in-hospital mortality is achieved. The resultant Noninvasive Ventilation Outcomes (NIVO) score had area under the receiver operating curve of 0.79 and offers good calibration and discrimination across stratified risk groups in its validation cohort. DISCUSSION: The NIVO score outperformed pre-specified comparator scores. It is validated in a generalisable cohort and works despite the heterogeneity inherent to both this patient group and this intervention. Potential applications include informing discussions with patients and their families, aiding treatment escalation decisions, challenging pessimism and comparing risk-adjusted outcomes across centres..
Cubuk, C.
Garrett, A.
Choi, S.
King, L.
Loveday, C.
Torr, B.
Burghel, G.J.
Durkie, M.
Callaway, A.
Robinson, R.
Drummond, J.
Berry, I.
Wallace, A.
Eccles, D.
Tischkowitz, M.
Whiffin, N.
Ware, J.S.
Hanson, H.
Turnbull, C.
CanVIG-Uk,
(2021). Clinical likelihood ratios and balanced accuracy for 44 in silico tools against multiple large-scale functional assays of cancer susceptibility genes. Genet med,
Vol.23
(11),
pp. 2096-2104.
show abstract
full text
PURPOSE: Where multiple in silico tools are concordant, the American College of Medical Genetics and Genomics/Association for Molecular Pathology (ACMG/AMP) framework affords supporting evidence toward pathogenicity or benignity, equivalent to a likelihood ratio of ~2. However, limited availability of "clinical truth sets" and prior use in tool training limits their utility for evaluation of tool performance. METHODS: We created a truth set of 9,436 missense variants classified as deleterious or tolerated in clinically validated high-throughput functional assays for BRCA1, BRCA2, MSH2, PTEN, and TP53 to evaluate predictive performance for 44 recommended/commonly used in silico tools. RESULTS: Over two-thirds of the tool-threshold combinations examined had specificity of <50%, thus substantially overcalling deleteriousness. REVEL scores of 0.8-1.0 had a Positive Likelihood Ratio (PLR) of 6.74 (5.24-8.82) compared to scores <0.7 and scores of 0-0.4 had a Negative Likelihood Ratio (NLR) of 34.3 (31.5-37.3) compared to scores of >0.7. For Meta-SNP, the equivalent PLR = 42.9 (14.4-406) and NLR = 19.4 (15.6-24.9). CONCLUSION: Against these clinically validated "functional truth sets," there was wide variation in the predictive performance of commonly used in silico tools. Overall, REVEL and Meta-SNP had best balanced accuracy and might potentially be used at stronger evidence weighting than current ACMG/AMP prescription, in particular for predictions of benignity..
100,000 Genomes Project Pilot Investigators,
Smedley, D.
Smith, K.R.
Martin, A.
Thomas, E.A.
McDonagh, E.M.
Cipriani, V.
Ellingford, J.M.
Arno, G.
Tucci, A.
Vandrovcova, J.
Chan, G.
Williams, H.J.
Ratnaike, T.
Wei, W.
Stirrups, K.
Ibanez, K.
Moutsianas, L.
Wielscher, M.
Need, A.
Barnes, M.R.
Vestito, L.
Buchanan, J.
Wordsworth, S.
Ashford, S.
Rehmström, K.
Li, E.
Fuller, G.
Twiss, P.
Spasic-Boskovic, O.
Halsall, S.
Floto, R.A.
Poole, K.
Wagner, A.
Mehta, S.G.
Gurnell, M.
Burrows, N.
James, R.
Penkett, C.
Dewhurst, E.
Gräf, S.
Mapeta, R.
Kasanicki, M.
Haworth, A.
Savage, H.
Babcock, M.
Reese, M.G.
Bale, M.
Baple, E.
Boustred, C.
Brittain, H.
de Burca, A.
Bleda, M.
Devereau, A.
Halai, D.
Haraldsdottir, E.
Hyder, Z.
Kasperaviciute, D.
Patch, C.
Polychronopoulos, D.
Matchan, A.
Sultana, R.
Ryten, M.
Tavares, A.L.
Tregidgo, C.
Turnbull, C.
Welland, M.
Wood, S.
Snow, C.
Williams, E.
Leigh, S.
Foulger, R.E.
Daugherty, L.C.
Niblock, O.
Leong, I.U.
Wright, C.F.
Davies, J.
Crichton, C.
Welch, J.
Woods, K.
Abulhoul, L.
Aurora, P.
Bockenhauer, D.
Broomfield, A.
Cleary, M.A.
Lam, T.
Dattani, M.
Footitt, E.
Ganesan, V.
Grunewald, S.
Compeyrot-Lacassagne, S.
Muntoni, F.
Pilkington, C.
Quinlivan, R.
Thapar, N.
Wallis, C.
Wedderburn, L.R.
Worth, A.
Bueser, T.
Compton, C.
Deshpande, C.
Fassihi, H.
Haque, E.
Izatt, L.
Josifova, D.
Mohammed, S.
Robert, L.
Rose, S.
Ruddy, D.
Sarkany, R.
Say, G.
Shaw, A.C.
Wolejko, A.
Habib, B.
Burns, G.
Hunter, S.
Grocock, R.J.
Humphray, S.J.
Robinson, P.N.
Haendel, M.
Simpson, M.A.
Banka, S.
Clayton-Smith, J.
Douzgou, S.
Hall, G.
Thomas, H.B.
O'Keefe, R.T.
Michaelides, M.
Moore, A.T.
Malka, S.
Pontikos, N.
Browning, A.C.
Straub, V.
Gorman, G.S.
Horvath, R.
Quinton, R.
Schaefer, A.M.
Yu-Wai-Man, P.
Turnbull, D.M.
McFarland, R.
Taylor, R.W.
O'Connor, E.
Yip, J.
Newland, K.
Morris, H.R.
Polke, J.
Wood, N.W.
Campbell, C.
Camps, C.
Gibson, K.
Koelling, N.
Lester, T.
Németh, A.H.
Palles, C.
Patel, S.
Roy, N.B.
Sen, A.
Taylor, J.
Cacheiro, P.
Jacobsen, J.O.
Seaby, E.G.
Davison, V.
Chitty, L.
Douglas, A.
Naresh, K.
McMullan, D.
Ellard, S.
Temple, I.K.
Mumford, A.D.
Wilson, G.
Beales, P.
Bitner-Glindzicz, M.
Black, G.
Bradley, J.R.
Brennan, P.
Burn, J.
Chinnery, P.F.
Elliott, P.
Flinter, F.
Houlden, H.
Irving, M.
Newman, W.
Rahman, S.
Sayer, J.A.
Taylor, J.C.
Webster, A.R.
Wilkie, A.O.
Ouwehand, W.H.
Raymond, F.L.
Chisholm, J.
Hill, S.
Bentley, D.
Scott, R.H.
Fowler, T.
Rendon, A.
Caulfield, M.
(2021). 100,000 Genomes Pilot on Rare-Disease Diagnosis in Health Care - Preliminary Report. N engl j med,
Vol.385
(20),
pp. 1868-1880.
show abstract
full text
BACKGROUND: The U.K. 100,000 Genomes Project is in the process of investigating the role of genome sequencing in patients with undiagnosed rare diseases after usual care and the alignment of this research with health care implementation in the U.K. National Health Service. Other parts of this project focus on patients with cancer and infection. METHODS: We conducted a pilot study involving 4660 participants from 2183 families, among whom 161 disorders covering a broad spectrum of rare diseases were present. We collected data on clinical features with the use of Human Phenotype Ontology terms, undertook genome sequencing, applied automated variant prioritization on the basis of applied virtual gene panels and phenotypes, and identified novel pathogenic variants through research analysis. RESULTS: Diagnostic yields varied among family structures and were highest in family trios (both parents and a proband) and families with larger pedigrees. Diagnostic yields were much higher for disorders likely to have a monogenic cause (35%) than for disorders likely to have a complex cause (11%). Diagnostic yields for intellectual disability, hearing disorders, and vision disorders ranged from 40 to 55%. We made genetic diagnoses in 25% of the probands. A total of 14% of the diagnoses were made by means of the combination of research and automated approaches, which was critical for cases in which we found etiologic noncoding, structural, and mitochondrial genome variants and coding variants poorly covered by exome sequencing. Cohortwide burden testing across 57,000 genomes enabled the discovery of three new disease genes and 19 new associations. Of the genetic diagnoses that we made, 25% had immediate ramifications for clinical decision making for the patients or their relatives. CONCLUSIONS: Our pilot study of genome sequencing in a national health care system showed an increase in diagnostic yield across a range of rare diseases. (Funded by the National Institute for Health Research and others.)..
Sud, A.
Turnbull, C.
Houlston, R.
(2021). Will polygenic risk scores for cancer ever be clinically useful?. Npj precis oncol,
Vol.5
(1),
p. 40.
full text
Pluta, J.
Pyle, L.C.
Nead, K.T.
Wilf, R.
Li, M.
Mitra, N.
Weathers, B.
D'Andrea, K.
Almstrup, K.
Anson-Cartwright, L.
Benitez, J.
Brown, C.D.
Chanock, S.
Chen, C.
Cortessis, V.K.
Ferlin, A.
Foresta, C.
Gamulin, M.
Gietema, J.A.
Grasso, C.
Greene, M.H.
Grotmol, T.
Hamilton, R.J.
Haugen, T.B.
Hauser, R.
Hildebrandt, M.A.
Johnson, M.E.
Karlsson, R.
Kiemeney, L.A.
Lessel, D.
Lothe, R.A.
Loud, J.T.
Loveday, C.
Martin-Gimeno, P.
Meijer, C.
Nsengimana, J.
Quinn, D.I.
Rafnar, T.
Ramdas, S.
Richiardi, L.
Skotheim, R.I.
Stefansson, K.
Turnbull, C.
Vaughn, D.J.
Wiklund, F.
Wu, X.
Yang, D.
Zheng, T.
Wells, A.D.
Grant, S.F.
Rajpert-De Meyts, E.
Schwartz, S.M.
Bishop, D.T.
McGlynn, K.A.
Kanetsky, P.A.
Nathanson, K.L.
Testicular Cancer Consortium,
(2021). Identification of 22 susceptibility loci associated with testicular germ cell tumors. Nat commun,
Vol.12
(1),
p. 4487.
show abstract
full text
Testicular germ cell tumors (TGCT) are the most common tumor in young white men and have a high heritability. In this study, the international Testicular Cancer Consortium assemble 10,156 and 179,683 men with and without TGCT, respectively, for a genome-wide association study. This meta-analysis identifies 22 TGCT susceptibility loci, bringing the total to 78, which account for 44% of disease heritability. Men with a polygenic risk score (PRS) in the 95th percentile have a 6.8-fold increased risk of TGCT compared to men with median scores. Among men with independent TGCT risk factors such as cryptorchidism, the PRS may guide screening decisions with the goal of reducing treatment-related complications causing long-term morbidity in survivors. These findings emphasize the interconnected nature of two known pathways that promote TGCT susceptibility: male germ cell development within its somatic niche and regulation of chromosomal division and structure, and implicate an additional biological pathway, mRNA translation..
Zhang, Y.D.
Hurson, A.N.
Zhang, H.
Choudhury, P.P.
Easton, D.F.
Milne, R.L.
Simard, J.
Hall, P.
Michailidou, K.
Dennis, J.
Schmidt, M.K.
Chang-Claude, J.
Gharahkhani, P.
Whiteman, D.
Campbell, P.T.
Hoffmeister, M.
Jenkins, M.
Peters, U.
Hsu, L.
Gruber, S.B.
Casey, G.
Schmit, S.L.
O'Mara, T.A.
Spurdle, A.B.
Thompson, D.J.
Tomlinson, I.
De Vivo, I.
Landi, M.T.
Law, M.H.
Iles, M.M.
Demenais, F.
Kumar, R.
MacGregor, S.
Bishop, D.T.
Ward, S.V.
Bondy, M.L.
Houlston, R.
Wiencke, J.K.
Melin, B.
Barnholtz-Sloan, J.
Kinnersley, B.
Wrensch, M.R.
Amos, C.I.
Hung, R.J.
Brennan, P.
McKay, J.
Caporaso, N.E.
Berndt, S.I.
Birmann, B.M.
Camp, N.J.
Kraft, P.
Rothman, N.
Slager, S.L.
Berchuck, A.
Pharoah, P.D.
Sellers, T.A.
Gayther, S.A.
Pearce, C.L.
Goode, E.L.
Schildkraut, J.M.
Moysich, K.B.
Amundadottir, L.T.
Jacobs, E.J.
Klein, A.P.
Petersen, G.M.
Risch, H.A.
Stolzenberg-Solomon, R.Z.
Wolpin, B.M.
Li, D.
Eeles, R.A.
Haiman, C.A.
Kote-Jarai, Z.
Schumacher, F.R.
Al Olama, A.A.
Purdue, M.P.
Scelo, G.
Dalgaard, M.D.
Greene, M.H.
Grotmol, T.
Kanetsky, P.A.
McGlynn, K.A.
Nathanson, K.L.
Turnbull, C.
Wiklund, F.
Breast Cancer Association Consortium (BCAC),
Barrett’s and Esophageal Adenocarcinoma Consortium (BEACON),
Colon Cancer Family Registry (CCFR),
Transdisciplinary Studies of Genetic Variation in Colorectal Cancer (CORECT),
Endometrial Cancer Association Consortium (ECAC),
Genetics and Epidemiology of Colorectal Cancer Consortium (GECCO),
Melanoma Genetics Consortium (GenoMEL),
Glioma International Case-Control Study (GICC),
International Lung Cancer Consortium (ILCCO),
Integrative Analysis of Lung Cancer Etiology and Risk (INTEGRAL) Consortium,
International Consortium of Investigators Working on Non-Hodgkin’s Lymphoma Epidemiologic Studies (InterLymph),
Ovarian Cancer Association Consortium (OCAC),
Oral Cancer GWAS,
Pancreatic Cancer Case-Control Consortium (PanC4),
Pancreatic Cancer Cohort Consortium (PanScan),
Prostate Cancer Association Group to Investigate Cancer Associated Alterations in the Genome (PRACTICAL),
Renal Cancer GWAS,
Testicular Cancer Consortium (TECAC),
Chanock, S.J.
Chatterjee, N.
Garcia-Closas, M.
(2020). Assessment of polygenic architecture and risk prediction based on common variants across fourteen cancers. Nat commun,
Vol.11
(1),
p. 3353.
show abstract
full text
Genome-wide association studies (GWAS) have led to the identification of hundreds of susceptibility loci across cancers, but the impact of further studies remains uncertain. Here we analyse summary-level data from GWAS of European ancestry across fourteen cancer sites to estimate the number of common susceptibility variants (polygenicity) and underlying effect-size distribution. All cancers show a high degree of polygenicity, involving at a minimum of thousands of loci. We project that sample sizes required to explain 80% of GWAS heritability vary from 60,000 cases for testicular to over 1,000,000 cases for lung cancer. The maximum relative risk achievable for subjects at the 99th risk percentile of underlying polygenic risk scores (PRS), compared to average risk, ranges from 12 for testicular to 2.5 for ovarian cancer. We show that PRS have potential for risk stratification for cancers of breast, colon and prostate, but less so for others because of modest heritability and lower incidence..
Loveday, C.
Litchfield, K.
Proszek, P.Z.
Cornish, A.J.
Santo, F.
Levy, M.
Macintyre, G.
Holryod, A.
Broderick, P.
Dudakia, D.
Benton, B.
Bakir, M.A.
Hiley, C.
Grist, E.
Swanton, C.
Huddart, R.
Powles, T.
Chowdhury, S.
Shipley, J.
O'Connor, S.
Brenton, J.D.
Reid, A.
de Castro, D.G.
Houlston, R.S.
Turnbull, C.
(2020). Genomic landscape of platinum resistant and sensitive testicular cancers. Nat commun,
Vol.11
(1),
p. 2189.
show abstract
full text
While most testicular germ cell tumours (TGCTs) exhibit exquisite sensitivity to platinum chemotherapy, ~10% are platinum resistant. To gain insight into the underlying mechanisms, we undertake whole exome sequencing and copy number analysis in 40 tumours from 26 cases with platinum-resistant TGCT, and combine this with published genomic data on an additional 624 TGCTs. We integrate analyses for driver mutations, mutational burden, global, arm-level and focal copy number (CN) events, and SNV and CN signatures. Albeit preliminary and observational in nature, these analyses provide support for a possible mechanistic link between early driver mutations in RAS and KIT and the widespread copy number events by which TGCT is characterised..
Melis, D.
Carvalho, D.
Barbaro-Dieber, T.
Espay, A.J.
Gambello, M.J.
Gener, B.
Gerkes, E.
Hitzert, M.M.
Hove, H.B.
Jansen, S.
Jira, P.E.
Lachlan, K.
Menke, L.A.
Narayanan, V.
Ortiz, D.
Overwater, E.
Posmyk, R.
Ramsey, K.
Rossi, A.
Sandoval, R.L.
Stumpel, C.
Stuurman, K.E.
Cordeddu, V.
Turnpenny, P.
Strisciuglio, P.
Tartaglia, M.
Unger, S.
Waters, T.
Turnbull, C.
Hennekam, R.C.
(2020). Primrose syndrome: Characterization of the phenotype in 42 patients. Clin genet,
Vol.97
(6),
pp. 890-901.
show abstract
Primrose syndrome (PS; MIM# 259050) is characterized by intellectual disability (ID), macrocephaly, unusual facial features (frontal bossing, deeply set eyes, down-slanting palpebral fissures), calcified external ears, sparse body hair and distal muscle wasting. The syndrome is caused by de novo heterozygous missense variants in ZBTB20. Most of the 29 published patients are adults as characteristics appear more recognizable with age. We present 13 hitherto unpublished individuals and summarize the clinical and molecular findings in all 42 patients. Several signs and symptoms of PS develop during childhood, but the cardinal features, such as calcification of the external ears, cystic bone lesions, muscle wasting, and contractures typically develop between 10 and 16 years of age. Biochemically, anemia and increased alpha-fetoprotein levels are often present. Two adult males with PS developed a testicular tumor. Although PS should be regarded as a progressive entity, there are no indications that cognition becomes more impaired with age. No obvious genotype-phenotype correlation is present. A subgroup of patients with ZBTB20 variants may be associated with mild, nonspecific ID. Metabolic investigations suggest a disturbed mitochondrial fatty acid oxidation. We suggest a regular surveillance in all adult males with PS until it is clear whether or not there is a truly elevated risk of testicular cancer..
Peto, J.
Carpenter, J.
Smith, G.D.
Duffy, S.
Houlston, R.
Hunter, D.J.
McPherson, K.
Pearce, N.
Romer, P.
Sasieni, P.
Turnbull, C.
(2020). Weekly COVID-19 testing with household quarantine and contact tracing is feasible and would probably end the epidemic. R soc open sci,
Vol.7
(6),
p. 200915.
show abstract
full text
The COVID-19 epidemic can probably be ended and normal life restored, perhaps quite quickly, by weekly SARS-CoV-2 RNA testing together with household quarantine and systematic contact tracing. Isolated outbreaks could then be contained by contact tracing, supplemented if necessary by temporary local reintroduction of population testing or lockdown. Leading public health experts have recommended that this should be tried in a demonstration project in which a medium-sized city introduces weekly testing and lifts lockdown completely. The idea was not considered by the groups whose predictions have guided UK policy, so we have examined the statistical case for such a study. The combination of regular testing with strict household quarantine, which was not analysed in their models, has remarkable power to reduce transmission to the community from other household members as well as providing earlier diagnosis and facilitating rapid contact tracing..
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.-.
Scott, S.
Pharoah, P.D.
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. Ann oncol,
Vol.31
(8),
pp. 1065-1074.
show abstract
full text
BACKGROUND: 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. PATIENTS AND METHODS: 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. RESULTS: 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. CONCLUSIONS: 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..
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. Lancet oncol,
Vol.21
(8),
pp. 1035-1044.
show abstract
full text
BACKGROUND: 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. METHODS: 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. FINDINGS: 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. INTERPRETATION: 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. FUNDING: None..
Hallifax, R.J.
Porter, B.M.
Elder, P.J.
Evans, S.B.
Turnbull, C.D.
Hynes, G.
Lardner, R.
Archer, K.
Bettinson, H.V.
Nickol, A.H.
Flight, W.G.
Chapman, S.J.
Hardinge, M.
Hoyles, R.K.
Saunders, P.
Sykes, A.
Wrightson, J.M.
Moore, A.
Ho, L.-.
Fraser, E.
Pavord, I.D.
Talbot, N.P.
Bafadhel, M.
Petousi, N.
Rahman, N.M.
Oxford Respiratory Group,
(2020). Successful awake proning is associated with improved clinical outcomes in patients with COVID-19: single-centre high-dependency unit experience. Bmj open respir res,
Vol.7
(1).
show abstract
The SARS-CoV-2 can lead to severe illness with COVID-19. Outcomes of patients requiring mechanical ventilation are poor. Awake proning in COVID-19 improves oxygenation, but on data clinical outcomes is limited. This single-centre retrospective study aimed to assess whether successful awake proning of patients with COVID-19, requiring respiratory support (continuous positive airways pressure (CPAP) or high-flow nasal oxygen (HFNO)) on a respiratory high-dependency unit (HDU), is associated with improved outcomes. HDU care included awake proning by respiratory physiotherapists. Of 565 patients admitted with COVID-19, 71 (12.6%) were managed on the respiratory HDU, with 48 of these (67.6%) requiring respiratory support. Patients managed with CPAP alone 22/48 (45.8%) were significantly less likely to die than patients who required transfer onto HFNO 26/48 (54.2%): CPAP mortality 36.4%; HFNO mortality 69.2%, (p=0.023); however, multivariate analysis demonstrated that increasing age and the inability to awake prone were the only independent predictors of COVID-19 mortality. The mortality of patients with COVID-19 requiring respiratory support is considerable. Data from our cohort managed on HDU show that CPAP and awake proning are possible in a selected population of COVID-19, and may be useful. Further prospective studies are required..
Tang, I.
Turnbull, C.D.
Sen, D.
Craig, S.
Kohler, M.
Stradling, J.R.
(2020). Effect of CPAP on cardiovascular events in minimally symptomatic OSA: long-term follow-up of the MOSAIC randomised controlled trial. Bmj open respir res,
Vol.7
(1).
show abstract
The effect of continuous positive airway pressure (CPAP) on cardiovascular events is uncertain in minimally symptomatic obstructive sleep apnoea. Previous 2-year follow-up data from the Multicentre Obstructive Sleep Apnoea Intervention Cardiovascular (MOSAIC) trial showed a marginal reduction in cardiovascular events with CPAP therapy. We now present long-term MOSAIC study follow-up data. Median (first quartile, third quartile) follow-up was 5.0 (2.2, 5.0) and 3.7 (1.5, 5.0) years for CPAP and standard care, respectively. Compared to standard care, CPAP had no statistically significant effect on the risk of cardiovascular events (HR=0.83, p=0.54, 95% CI 0.46-1.51)..
Alamar, M.C.
Anastasiadi, M.
Lopez-Cobollo, R.
Bennett, M.H.
Thompson, A.J.
Turnbull, C.G.
Mohareb, F.
Terry, L.A.
(2020). Transcriptome and phytohormone changes associated with ethylene-induced onion bulb dormancy. Postharvest biol technol,
Vol.168,
p. 111267.
show abstract
full text
Control of dormancy and sprouting in onion bulbs is commercially important for postharvest management. Although ethylene application is sometimes used to extend dormancy, the underlying mechanisms regulating dormancy transition remain unclear. Since the sprout leaves emerge from the bulb baseplate, we used this tissue to assess the impact of ethylene treatment and storage time on the hormone profile and the transcriptome. Reads from 30 libraries were assembled and annotated, with 94,840 unigenes retained after filtering. The de novo transcriptome assembly was of high quality and continuity (N50: 1809 bp, GC content: 36.21 %), and was used to analyse differential expression and Gene Onotologies. Across two years, applied ethylene resulted in delayed dormancy break and reduced post-dormancy sprout vigour. Ethylene supplementation enhanced endogenous ethylene production and caused a transient climacteric-like increase in respiration. Significant changes in hormone and associated transcript profiles occurred through storage and in response to ethylene. In particular, abscisic acid (ABA) and its metabolite phaseic acid (PA) increased under ethylene during the longer dormancy period; however, cytokinin increases observed during storage appeared largely independent of ethylene treatment. Several hormone-related transcripts showed differential expression over time and/or in response to ethylene. Expression of ethylene biosynthesis (ACO), receptor (EIN4) and transcription factor (EIL3) genes were modified by ethylene, as were ABA biosynthesis genes such NCED, and cytokinin biosynthesis genes such as LOG and CKX. We conclude that ethylene substantially modifies expression of genes in several phytohormone pathways, and some of these changes may underlie the dormancy-extending effects of exogenous ethylene..
Lee, L.Y.
Cazier, J.-.
Starkey, T.
Briggs, S.E.
Arnold, R.
Bisht, V.
Booth, S.
Campton, N.A.
Cheng, V.W.
Collins, G.
Curley, H.M.
Earwaker, P.
Fittall, M.W.
Gennatas, S.
Goel, A.
Hartley, S.
Hughes, D.J.
Kerr, D.
Lee, A.J.
Lee, R.J.
Lee, S.M.
Mckenzie, H.
Middleton, C.P.
Murugaesu, N.
Newsom-Davis, T.
Olsson-Brown, A.C.
Palles, C.
Powles, T.
Protheroe, E.A.
Purshouse, K.
Sharma-Oates, A.
Sivakumar, S.
Smith, A.J.
Topping, O.
Turnbull, C.D.
Várnai, C.
Briggs, A.D.
Middleton, G.
Kerr, R.
UK Coronavirus Cancer Monitoring Project Team,
(2020). COVID-19 prevalence and mortality in patients with cancer and the effect of primary tumour subtype and patient demographics: a prospective cohort study. Lancet oncol,
Vol.21
(10),
pp. 1309-1316.
show abstract
full text
BACKGROUND: Patients with cancer are purported to have poor COVID-19 outcomes. However, cancer is a heterogeneous group of diseases, encompassing a spectrum of tumour subtypes. The aim of this study was to investigate COVID-19 risk according to tumour subtype and patient demographics in patients with cancer in the UK. METHODS: We compared adult patients with cancer enrolled in the UK Coronavirus Cancer Monitoring Project (UKCCMP) cohort between March 18 and May 8, 2020, with a parallel non-COVID-19 UK cancer control population from the UK Office for National Statistics (2017 data). The primary outcome of the study was the effect of primary tumour subtype, age, and sex and on severe acute respiratory syndrome coronavirus 2 (SARS-CoV-2) prevalence and the case-fatality rate during hospital admission. We analysed the effect of tumour subtype and patient demographics (age and sex) on prevalence and mortality from COVID-19 using univariable and multivariable models. FINDINGS: 319 (30·6%) of 1044 patients in the UKCCMP cohort died, 295 (92·5%) of whom had a cause of death recorded as due to COVID-19. The all-cause case-fatality rate in patients with cancer after SARS-CoV-2 infection was significantly associated with increasing age, rising from 0·10 in patients aged 40-49 years to 0·48 in those aged 80 years and older. Patients with haematological malignancies (leukaemia, lymphoma, and myeloma) had a more severe COVID-19 trajectory compared with patients with solid organ tumours (odds ratio [OR] 1·57, 95% CI 1·15-2·15; p<0·0043). Compared with the rest of the UKCCMP cohort, patients with leukaemia showed a significantly increased case-fatality rate (2·25, 1·13-4·57; p=0·023). After correction for age and sex, patients with haematological malignancies who had recent chemotherapy had an increased risk of death during COVID-19-associated hospital admission (OR 2·09, 95% CI 1·09-4·08; p=0·028). INTERPRETATION: Patients with cancer with different tumour types have differing susceptibility to SARS-CoV-2 infection and COVID-19 phenotypes. We generated individualised risk tables for patients with cancer, considering age, sex, and tumour subtype. Our results could be useful to assist physicians in informed risk-benefit discussions to explain COVID-19 risk and enable an evidenced-based approach to national social isolation policies. FUNDING: University of Birmingham and University of Oxford..
Pashayan, N.
Antoniou, A.C.
Ivanus, U.
Esserman, L.J.
Easton, D.F.
French, D.
Sroczynski, G.
Hall, P.
Cuzick, J.
Evans, D.G.
Simard, J.
Garcia-Closas, M.
Schmutzler, R.
Wegwarth, O.
Pharoah, P.
Moorthie, S.
De Montgolfier, S.
Baron, C.
Herceg, Z.
Turnbull, C.
Balleyguier, C.
Rossi, P.G.
Wesseling, J.
Ritchie, D.
Tischkowitz, M.
Broeders, M.
Reisel, D.
Metspalu, A.
Callender, T.
de Koning, H.
Devilee, P.
Delaloge, S.
Schmidt, M.K.
Widschwendter, M.
(2020). Personalized early detection and prevention of breast cancer: ENVISION consensus statement. Nat rev clin oncol,
Vol.17
(11),
pp. 687-705.
show abstract
The European Collaborative on Personalized Early Detection and Prevention of Breast Cancer (ENVISION) brings together several international research consortia working on different aspects of the personalized early detection and prevention of breast cancer. In a consensus conference held in 2019, the members of this network identified research areas requiring development to enable evidence-based personalized interventions that might improve the benefits and reduce the harms of existing breast cancer screening and prevention programmes. The priority areas identified were: 1) breast cancer subtype-specific risk assessment tools applicable to women of all ancestries; 2) intermediate surrogate markers of response to preventive measures; 3) novel non-surgical preventive measures to reduce the incidence of breast cancer of poor prognosis; and 4) hybrid effectiveness-implementation research combined with modelling studies to evaluate the long-term population outcomes of risk-based early detection strategies. The implementation of such programmes would require health-care systems to be open to learning and adapting, the engagement of a diverse range of stakeholders and tailoring to societal norms and values, while also addressing the ethical and legal issues. In this Consensus Statement, we discuss the current state of breast cancer risk prediction, risk-stratified prevention and early detection strategies, and their implementation. Throughout, we highlight priorities for advancing each of these areas..
Garrett, A.
Callaway, A.
Durkie, M.
Cubuk, C.
Alikian, M.
Burghel, G.J.
Robinson, R.
Izatt, L.
Talukdar, S.
Side, L.
Cranston, T.
Palmer-Smith, S.
Baralle, D.
Berry, I.R.
Drummond, J.
Wallace, A.J.
Norbury, G.
Eccles, D.M.
Ellard, S.
Lalloo, F.
Evans, D.G.
Woodward, E.
Tischkowitz, M.
Hanson, H.
Turnbull, C.
CanVIG-UK,
(2020). Cancer Variant Interpretation Group UK (CanVIG-UK): an exemplar national subspecialty multidisciplinary network. J med genet,
Vol.57
(12),
pp. 829-834.
show abstract
full text
Advances in technology have led to a massive expansion in the capacity for genomic analysis, with a commensurate fall in costs. The clinical indications for genomic testing have evolved markedly; the volume of clinical sequencing has increased dramatically; and the range of clinical professionals involved in the process has broadened. There is general acceptance that our early dichotomous paradigms of variants being pathogenic-high risk and benign-no risk are overly simplistic. There is increasing recognition that the clinical interpretation of genomic data requires significant expertise in disease-gene-variant associations specific to each disease area. Inaccurate interpretation can lead to clinical mismanagement, inconsistent information within families and misdirection of resources. It is for this reason that 'national subspecialist multidisciplinary meetings' (MDMs) for genomic interpretation have been articulated as key for the new NHS Genomic Medicine Service, of which Cancer Variant Interpretation Group UK (CanVIG-UK) is an early exemplar. CanVIG-UK was established in 2017 and now has >100 UK members, including at least one clinical diagnostic scientist and one clinical cancer geneticist from each of the 25 regional molecular genetics laboratories of the UK and Ireland. Through CanVIG-UK, we have established national consensus around variant interpretation for cancer susceptibility genes via monthly national teleconferenced MDMs and collaborative data sharing using a secure online portal. We describe here the activities of CanVIG-UK, including exemplar outputs and feedback from the membership..
Lawn, S.
Waddell, E.
Cowain, T.
Turnbull, C.
McMahon, J.
(2020). Implementing national mental health carer partnership standards in South Australia. Aust health rev,
Vol.44
(6),
pp. 880-890.
show abstract
Objective The aim of this study was to describe the current state of carer engagement and partnership in two mental health (MH) services in South Australia and the implementation of the six partnership standards in A Practical Guide to Working with Carers of People with a Mental Illness. Methods Anonymous surveys of carer experiences and clinician self-ratings of their own practice against the six partnership standards were completed by 94 staff and 58 carers within public and private MH in-patient units before and after exposure of clinicians to education about the partnership standards. Descriptive statistical analysis was performed and, where applicable, a comparative analysis used the two-sample Z-test of proportions. Qualitative data was analysed thematically. Results Considerable gaps were evident between carer experiences and clinician self-ratings of their own practice. Overall, the surveys point to the lack of a consistent approach by both public and private services, and suggest potential barriers to fostering carer participation and engagement. Confidentiality was a particularly noted barrier to partnership with carers. Conclusion Significant improvement is needed to meet the partnership standards. Brief exposure to the Guide is not, in itself, sufficient to effect change in the overall attitudes, skills and knowledge of clinical staff about engaging carers. Significantly more focus on staff education, clinical discussions and supervision is needed to meet the MH carer partnership standards. What is known about the topic? Partnership with MH consumers and carers is an established key principle within national MH policies and accreditation standards. Family carers play an important role in supporting consumers' recovery, yet many carers continue to report being excluded, particularly by in-patient clinical staff. What does this paper add? This is the first study to investigate the partnership standards in practice by comparing the perspectives of carers and in-patient MH clinical staff. What are the implications for practitioners? Improving partnership with carers of people with mental illness will require significant MH service leadership support shifts in current practice and culture. In addition, a more nuanced understanding of confidentiality is required to overcome the barriers to involving family carers more meaningfully in care..
Yang, X.
Song, H.
Leslie, G.
Engel, C.
Hahnen, E.
Auber, B.
Horváth, J.
Kast, K.
Niederacher, D.
Turnbull, C.
Houlston, R.
Hanson, H.
Loveday, C.
Dolinsky, J.S.
LaDuca, H.
Ramus, S.J.
Menon, U.
Rosenthal, A.N.
Jacobs, I.
Gayther, S.A.
Dicks, E.
Nevanlinna, H.
Aittomäki, K.
Pelttari, L.M.
Ehrencrona, H.
Borg, Å.
Kvist, A.
Rivera, B.
Hansen, T.V.
Djursby, M.
Lee, A.
Dennis, J.
Bowtell, D.D.
Traficante, N.
Diez, O.
Balmaña, J.
Gruber, S.B.
Chenevix-Trench, G.
Investigators, K.
Jensen, A.
Kjær, S.K.
Høgdall, E.
Castéra, L.
Garber, J.
Janavicius, R.
Osorio, A.
Golmard, L.
Vega, A.
Couch, F.J.
Robson, M.
Gronwald, J.
Domchek, S.M.
Culver, J.O.
de la Hoya, M.
Easton, D.F.
Foulkes, W.D.
Tischkowitz, M.
Meindl, A.
Schmutzler, R.K.
Pharoah, P.D.
Antoniou, A.C.
(2020). Ovarian and Breast Cancer Risks Associated With Pathogenic Variants in RAD51C and RAD51D. J natl cancer inst,
Vol.112
(12),
pp. 1242-1250.
show abstract
full text
BACKGROUND: The purpose of this study was to estimate precise age-specific tubo-ovarian carcinoma (TOC) and breast cancer (BC) risks for carriers of pathogenic variants in RAD51C and RAD51D. METHODS: We analyzed data from 6178 families, 125 with pathogenic variants in RAD51C, and 6690 families, 60 with pathogenic variants in RAD51D. TOC and BC relative and cumulative risks were estimated using complex segregation analysis to model the cancer inheritance patterns in families while adjusting for the mode of ascertainment of each family. All statistical tests were two-sided. RESULTS: Pathogenic variants in both RAD51C and RAD51D were associated with TOC (RAD51C: relative risk [RR] = 7.55, 95% confidence interval [CI] = 5.60 to 10.19; P = 5 × 10-40; RAD51D: RR = 7.60, 95% CI = 5.61 to 10.30; P = 5 × 10-39) and BC (RAD51C: RR = 1.99, 95% CI = 1.39 to 2.85; P = 1.55 × 10-4; RAD51D: RR = 1.83, 95% CI = 1.24 to 2.72; P = .002). For both RAD51C and RAD51D, there was a suggestion that the TOC relative risks increased with age until around age 60 years and decreased thereafter. The estimated cumulative risks of developing TOC to age 80 years were 11% (95% CI = 6% to 21%) for RAD51C and 13% (95% CI = 7% to 23%) for RAD51D pathogenic variant carriers. The estimated cumulative risks of developing BC to 80 years were 21% (95% CI = 15% to 29%) for RAD51C and 20% (95% CI = 14% to 28%) for RAD51D pathogenic variant carriers. Both TOC and BC risks for RAD51C and RAD51D pathogenic variant carriers varied by cancer family history and could be as high as 32-36% for TOC, for carriers with two first-degree relatives diagnosed with TOC, or 44-46% for BC, for carriers with two first-degree relatives diagnosed with BC. CONCLUSIONS: These estimates will facilitate the genetic counseling of RAD51C and RAD51D pathogenic variant carriers and justify the incorporation of RAD51C and RAD51D into cancer risk prediction models..
Escala-Garcia, M.
Abraham, J.
Andrulis, I.L.
Anton-Culver, H.
Arndt, V.
Ashworth, A.
Auer, P.L.
Auvinen, P.
Beckmann, M.W.
Beesley, J.
Behrens, S.
Benitez, J.
Bermisheva, M.
Blomqvist, C.
Blot, W.
Bogdanova, N.V.
Bojesen, S.E.
Bolla, M.K.
Børresen-Dale, A.-.
Brauch, H.
Brenner, H.
Brucker, S.Y.
Burwinkel, B.
Caldas, C.
Canzian, F.
Chang-Claude, J.
Chanock, S.J.
Chin, S.-.
Clarke, C.L.
Couch, F.J.
Cox, A.
Cross, S.S.
Czene, K.
Daly, M.B.
Dennis, J.
Devilee, P.
Dunn, J.A.
Dunning, A.M.
Dwek, M.
Earl, H.M.
Eccles, D.M.
Eliassen, A.H.
Ellberg, C.
Evans, D.G.
Fasching, P.A.
Figueroa, J.
Flyger, H.
Gago-Dominguez, M.
Gapstur, S.M.
García-Closas, M.
García-Sáenz, J.A.
Gaudet, M.M.
George, A.
Giles, G.G.
Goldgar, D.E.
González-Neira, A.
Grip, M.
Guénel, P.
Guo, Q.
Haiman, C.A.
Håkansson, N.
Hamann, U.
Harrington, P.A.
Hiller, L.
Hooning, M.J.
Hopper, J.L.
Howell, A.
Huang, C.-.
Huang, G.
Hunter, D.J.
Jakubowska, A.
John, E.M.
Kaaks, R.
Kapoor, P.M.
Keeman, R.
Kitahara, C.M.
Koppert, L.B.
Kraft, P.
Kristensen, V.N.
Lambrechts, D.
Le Marchand, L.
Lejbkowicz, F.
Lindblom, A.
Lubiński, J.
Mannermaa, A.
Manoochehri, M.
Manoukian, S.
Margolin, S.
Martinez, M.E.
Maurer, T.
Mavroudis, D.
Meindl, A.
Milne, R.L.
Mulligan, A.M.
Neuhausen, S.L.
Nevanlinna, H.
Newman, W.G.
Olshan, A.F.
Olson, J.E.
Olsson, H.
Orr, N.
Peterlongo, P.
Petridis, C.
Prentice, R.L.
Presneau, N.
Punie, K.
Ramachandran, D.
Rennert, G.
Romero, A.
Sachchithananthan, M.
Saloustros, E.
Sawyer, E.J.
Schmutzler, R.K.
Schwentner, L.
Scott, C.
Simard, J.
Sohn, C.
Southey, M.C.
Swerdlow, A.J.
Tamimi, R.M.
Tapper, W.J.
Teixeira, M.R.
Terry, M.B.
Thorne, H.
Tollenaar, R.A.
Tomlinson, I.
Troester, M.A.
Truong, T.
Turnbull, C.
Vachon, C.M.
van der Kolk, L.E.
Wang, Q.
Winqvist, R.
Wolk, A.
Yang, X.R.
Ziogas, A.
Pharoah, P.D.
Hall, P.
Wessels, L.F.
Chenevix-Trench, G.
Bader, G.D.
Dörk, T.
Easton, D.F.
Canisius, S.
Schmidt, M.K.
(2020). A network analysis to identify mediators of germline-driven differences in breast cancer prognosis. Nat commun,
Vol.11
(1),
p. 312.
show abstract
full text
Identifying the underlying genetic drivers of the heritability of breast cancer prognosis remains elusive. We adapt a network-based approach to handle underpowered complex datasets to provide new insights into the potential function of germline variants in breast cancer prognosis. This network-based analysis studies ~7.3 million variants in 84,457 breast cancer patients in relation to breast cancer survival and confirms the results on 12,381 independent patients. Aggregating the prognostic effects of genetic variants across multiple genes, we identify four gene modules associated with survival in estrogen receptor (ER)-negative and one in ER-positive disease. The modules show biological enrichment for cancer-related processes such as G-alpha signaling, circadian clock, angiogenesis, and Rho-GTPases in apoptosis..
Lai, A.G.
Pasea, L.
Banerjee, A.
Hall, G.
Denaxas, S.
Chang, W.H.
Katsoulis, M.
Williams, B.
Pillay, D.
Noursadeghi, M.
Linch, D.
Hughes, D.
Forster, M.D.
Turnbull, C.
Fitzpatrick, N.K.
Boyd, K.
Foster, G.R.
Enver, T.
Nafilyan, V.
Humberstone, B.
Neal, R.D.
Cooper, M.
Jones, M.
Pritchard-Jones, K.
Sullivan, R.
Davie, C.
Lawler, M.
Hemingway, H.
(2020). Estimated impact of the COVID-19 pandemic on cancer services and excess 1-year mortality in people with cancer and multimorbidity: near real-time data on cancer care, cancer deaths and a population-based cohort study. Bmj open,
Vol.10
(11),
p. e043828.
show abstract
full text
OBJECTIVES: To estimate the impact of the COVID-19 pandemic on cancer care services and overall (direct and indirect) excess deaths in people with cancer. METHODS: We employed near real-time weekly data on cancer care to determine the adverse effect of the pandemic on cancer services. We also used these data, together with national death registrations until June 2020 to model deaths, in excess of background (pre-COVID-19) mortality, in people with cancer. Background mortality risks for 24 cancers with and without COVID-19-relevant comorbidities were obtained from population-based primary care cohort (Clinical Practice Research Datalink) on 3 862 012 adults in England. RESULTS: Declines in urgent referrals (median=-70.4%) and chemotherapy attendances (median=-41.5%) to a nadir (lowest point) in the pandemic were observed. By 31 May, these declines have only partially recovered; urgent referrals (median=-44.5%) and chemotherapy attendances (median=-31.2%). There were short-term excess death registrations for cancer (without COVID-19), with peak relative risk (RR) of 1.17 at week ending on 3 April. The peak RR for all-cause deaths was 2.1 from week ending on 17 April. Based on these findings and recent literature, we modelled 40% and 80% of cancer patients being affected by the pandemic in the long-term. At 40% affected, we estimated 1-year total (direct and indirect) excess deaths in people with cancer as between 7165 and 17 910, using RRs of 1.2 and 1.5, respectively, where 78% of excess deaths occured in patients with ≥1 comorbidity. CONCLUSIONS: Dramatic reductions were detected in the demand for, and supply of, cancer services which have not fully recovered with lockdown easing. These may contribute, over a 1-year time horizon, to substantial excess mortality among people with cancer and multimorbidity. It is urgent to understand how the recovery of general practitioner, oncology and other hospital services might best mitigate these long-term excess mortality risks..
Antoniadi, I.
Novák, O.
Gelová, Z.
Johnson, A.
Plíhal, O.
Simerský, R.
Mik, V.
Vain, T.
Mateo-Bonmatí, E.
Karady, M.
Pernisová, M.
Plačková, L.
Opassathian, K.
Hejátko, J.
Robert, S.
Friml, J.
Doležal, K.
Ljung, K.
Turnbull, C.
(2020). Cell-surface receptors enable perception of extracellular cytokinins. Nat commun,
Vol.11
(1),
p. 4284.
show abstract
full text
Cytokinins are mobile multifunctional plant hormones with roles in development and stress resilience. Although their Histidine Kinase receptors are substantially localised to the endoplasmic reticulum, cellular sites of cytokinin perception and importance of spatially heterogeneous cytokinin distribution continue to be debated. Here we show that cytokinin perception by plasma membrane receptors is an effective additional path for cytokinin response. Readout from a Two Component Signalling cytokinin-specific reporter (TCSn::GFP) closely matches intracellular cytokinin content in roots, yet we also find cytokinins in extracellular fluid, potentially enabling action at the cell surface. Cytokinins covalently linked to beads that could not pass the plasma membrane increased expression of both TCSn::GFP and Cytokinin Response Factors. Super-resolution microscopy of GFP-labelled receptors and diminished TCSn::GFP response to immobilised cytokinins in cytokinin receptor mutants, further indicate that receptors can function at the cell surface. We argue that dual intracellular and surface locations may augment flexibility of cytokinin responses..
Parry, G.
Benitez-Alfonso, Y.
Gibbs, D.J.
Grant, M.
Harper, A.
Harrison, C.J.
Kaiserli, E.
Leonelli, S.
May, S.
McKim, S.
Spoel, S.
Turnbull, C.
van der Hoorn, R.A.
Murray, J.
(2020). How to build an effective research network: lessons from two decades of the GARNet plant science community. J exp bot,
Vol.71
(22),
pp. 6881-6889.
show abstract
full text
Successful collaborative research is dependent on excellent ideas and innovative experimental approaches, as well as the provision of appropriate support networks. Collaboration requires venues, infrastructures, training facilities, and, perhaps most importantly, a sustained commitment to work together as a community. These activities do not occur without significant effort, yet can be facilitated and overseen by the leadership of a research network that has a clearly defined role to help build resources for their community. Over the past 20 years, this is a role that the UKRI-BBSRC-funded GARNet network has played in the support of the UK curiosity-driven, discovery-led plant science research community. This article reviews the lessons learnt by GARNet in the hope that they can inform the practical implementation of current and future research networks..
Mandelker, D.
Donoghue, M.
Talukdar, S.
Bandlamudi, C.
Srinivasan, P.
Vivek, M.
Jezdic, S.
Hanson, H.
Snape, K.
Kulkarni, A.
Hawkes, L.
Douillard, J.-.
Wallace, S.E.
Rial-Sebbag, E.
Meric-Bersntam, F.
George, A.
Chubb, D.
Loveday, C.
Ladanyi, M.
Berger, M.F.
Taylor, B.S.
Turnbull, C.
(2019). Germline-focussed analysis of tumour-only sequencing: recommendations from the ESMO Precision Medicine Working Group. Ann oncol,
Vol.30
(8),
pp. 1221-1231.
show abstract
full text
It is increasingly common in oncology practice to perform tumour sequencing using large cancer panels. For pathogenic sequence variants in cancer susceptibility genes identified on tumour-only sequencing, it is often unclear whether they are of somatic or constitutional (germline) origin. There is wide-spread disparity regarding both the extent to which systematic 'germline-focussed analysis' is carried out upon tumour sequencing data and for which variants follow-up analysis of a germline sample is carried out. Here we present analyses of paired sequencing data from 17 152 cancer samples, in which 1494 pathogenic sequence variants were identified across 65 cancer susceptibility genes. From these analyses, the European Society of Medical Oncology Precision Medicine Working Group Germline Subgroup has generated (i) recommendations regarding germline-focussed analyses of tumour-only sequencing data, (ii) indications for germline follow-up testing and (iii) guidance on patient information-giving and consent..
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.-.
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.-.
Chin, S.-.
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.
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.-.
Koppert, L.B.
Kosma, V.-.
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.-.
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.-.
Perez, J.I.
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.
Tomlinson, I.
Torres, D.
Truong, T.
Turman, C.
Turnbull, C.
Ulmer, H.-.
Untch, M.
Vachon, C.
van Asperen, C.J.
van den Ouweland, A.M.
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.
Schmidt, M.K.
(2019). Genome-wide association study of germline variants and breast cancer-specific mortality. Br j cancer,
Vol.120
(6),
pp. 647-657.
show abstract
full text
BACKGROUND: We examined the associations between germline variants and breast cancer mortality using a large meta-analysis of women of European ancestry. METHODS: 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). RESULTS: We did not find any variant associated with breast cancer-specific mortality at P < 5 × 10-8. For ER-positive disease, the most significantly associated variant was chr7:rs4717568 (BFDP = 7%, P = 1.28 × 10-7, 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 × 10-7, 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. CONCLUSIONS: 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..
Spurdle, A.B.
Greville-Heygate, S.
Antoniou, A.C.
Brown, M.
Burke, L.
de la Hoya, M.
Domchek, S.
Dörk, T.
Firth, H.V.
Monteiro, A.N.
Mensenkamp, A.
Parsons, M.T.
Radice, P.
Robson, M.
Tischkowitz, M.
Tudini, E.
Turnbull, C.
Vreeswijk, M.P.
Walker, L.C.
Tavtigian, S.
Eccles, D.M.
(2019). Towards controlled terminology for reporting germline cancer susceptibility variants: an ENIGMA report. J med genet,
Vol.56
(6),
pp. 347-357.
show abstract
The vocabulary currently used to describe genetic variants and their consequences reflects many years of studying and discovering monogenic disease with high penetrance. With the recent rapid expansion of genetic testing brought about by wide availability of high-throughput massively parallel sequencing platforms, accurate variant interpretation has become a major issue. The vocabulary used to describe single genetic variants in silico, in vitro, in vivo and as a contributor to human disease uses terms in common, but the meaning is not necessarily shared across all these contexts. In the setting of cancer genetic tests, the added dimension of using data from genetic sequencing of tumour DNA to direct treatment is an additional source of confusion to those who are not experienced in cancer genetics. The language used to describe variants identified in cancer susceptibility genetic testing typically still reflects an outdated paradigm of Mendelian inheritance with dichotomous outcomes. Cancer is a common disease with complex genetic architecture; an improved lexicon is required to better communicate among scientists, clinicians and patients, the risks and implications of genetic variants detected. This review arises from a recognition of, and discussion about, inconsistencies in vocabulary usage by members of the ENIGMA international multidisciplinary consortium focused on variant classification in breast-ovarian cancer susceptibility genes. It sets out the vocabulary commonly used in genetic variant interpretation and reporting, and suggests a framework for a common vocabulary that may facilitate understanding and clarity in clinical reporting of germline genetic tests for cancer susceptibility..
Loveday, C.
Sud, A.
Litchfield, K.
Levy, M.
Holroyd, A.
Broderick, P.
Kote-Jarai, Z.
Dunning, A.M.
Muir, K.
Peto, J.
Eeles, R.
Easton, D.F.
Dudakia, D.
Orr, N.
Pashayan, N.
UK Testicular Cancer Collaboration,
PRACTICAL Consortium,
Reid, A.
Huddart, R.A.
Houlston, R.S.
Turnbull, C.
(2019). Runs of homozygosity and testicular cancer risk. Andrology,
Vol.7
(4),
pp. 555-564.
show abstract
full text
BACKGROUND: Testicular germ cell tumour (TGCT) is highly heritable but > 50% of the genetic risk remains unexplained. Epidemiological observation of greater relative risk to brothers of men with TGCT compared to sons has long alluded to recessively acting TGCT genetic susceptibility factors, but to date none have been reported. Runs of homozygosity (RoH) are a signature indicating underlying recessively acting alleles and have been associated with increased risk of other cancer types. OBJECTIVE: To examine whether RoH are associated with TGCT risk. METHODS: We performed a genome-wide RoH analysis using GWAS data from 3206 TGCT cases and 7422 controls uniformly genotyped using the OncoArray platform. RESULTS: Global measures of homozygosity were not significantly different between cases and controls, and the frequency of individual consensus RoH was not significantly different between cases and controls, after correction for multiple testing. RoH at three regions, 11p13-11p14.3, 5q14.1-5q22.3 and 13q14.11-13q.14.13, were, however, nominally statistically significant at p < 0.01. Intriguingly, RoH200 at 11p13-11p14.3 encompasses Wilms tumour 1 (WT1), a recognized cancer susceptibility gene with roles in sex determination and developmental transcriptional regulation, processes repeatedly implicated in TGCT aetiology. DISCUSSION AND CONCLUSION: Overall, our data do not support a major role in the risk of TGCT for recessively acting alleles acting through homozygosity, as measured by RoH in outbred populations of cases and controls..
Talukdar, S.
Hawkes, L.
Hanson, H.
Kulkarni, A.
Brady, A.F.
McMullan, D.J.
Ahn, J.W.
Woodward, E.
Turnbull, C.
UK Association for Clinical Genomic Science and UK Cancer Genetics Group,
(2019). Structural Aberrations with Secondary Implications (SASIs): consensus recommendations for reporting of cancer susceptibility genes identified during analysis of Copy Number Variants (CNVs). J med genet,
Vol.56
(11),
pp. 718-726.
show abstract
full text
Clinical testing with chromosomal microarray (CMA) is most commonly undertaken for clinical indications such as intellectual disability, dysmorphic features and/or congenital abnormalities. Identification of a structural aberration (SA) involving a cancer susceptibility gene (CSG) constitutes a type of incidental or secondary finding. Laboratory reporting, risk communication and clinical management of these structural aberrations with secondary implications (SASIs) is currently inconsistent. We undertake meta-analysis of 18 622 instances of CMA performed for unrelated indications in which 106 SASIs are identified involving in total 40 different CSGs. Here we present the recommendations of a joint UK working group representing the British Society of Genomic Medicine, UK Cancer Genetics Group and UK Association for Clinical Genomic Science. SASIs are categorised into four groups, defined by the type of SA and the cancer risk. For each group, recommendations are provided regarding reflex parental testing and cancer risk management..
Thompson, D.J.
Genovese, G.
Halvardson, J.
Ulirsch, J.C.
Wright, D.J.
Terao, C.
Davidsson, O.B.
Day, F.R.
Sulem, P.
Jiang, Y.
Danielsson, M.
Davies, H.
Dennis, J.
Dunlop, M.G.
Easton, D.F.
Fisher, V.A.
Zink, F.
Houlston, R.S.
Ingelsson, M.
Kar, S.
Kerrison, N.D.
Kinnersley, B.
Kristjansson, R.P.
Law, P.J.
Li, R.
Loveday, C.
Mattisson, J.
McCarroll, S.A.
Murakami, Y.
Murray, A.
Olszewski, P.
Rychlicka-Buniowska, E.
Scott, R.A.
Thorsteinsdottir, U.
Tomlinson, I.
Moghadam, B.T.
Turnbull, C.
Wareham, N.J.
Gudbjartsson, D.F.
International Lung Cancer Consortium (INTEGRAL-ILCCO),
Breast Cancer Association Consortium,
Consortium of Investigators of Modifiers of BRCA1/2,
Endometrial Cancer Association Consortium,
Ovarian Cancer Association Consortium,
Prostate Cancer Association Group to Investigate Cancer Associated Alterations in the Genome (PRACTICAL) Consortium,
Kidney Cancer GWAS Meta-Analysis Project,
eQTLGen Consortium,
Biobank-based Integrative Omics Study (BIOS) Consortium,
23andMe Research Team,
Kamatani, Y.
Hoffmann, E.R.
Jackson, S.P.
Stefansson, K.
Auton, A.
Ong, K.K.
Machiela, M.J.
Loh, P.-.
Dumanski, J.P.
Chanock, S.J.
Forsberg, L.A.
Perry, J.R.
(2019). Genetic predisposition to mosaic Y chromosome loss in blood. Nature,
Vol.575
(7784),
pp. 652-657.
show abstract
full text
Mosaic loss of chromosome Y (LOY) in circulating white blood cells is the most common form of clonal mosaicism1-5, yet our knowledge of the causes and consequences of this is limited. Here, using a computational approach, we estimate that 20% of the male population represented in the UK Biobank study (n = 205,011) has detectable LOY. We identify 156 autosomal genetic determinants of LOY, which we replicate in 757,114 men of European and Japanese ancestry. These loci highlight genes that are involved in cell-cycle regulation and cancer susceptibility, as well as somatic drivers of tumour growth and targets of cancer therapy. We demonstrate that genetic susceptibility to LOY is associated with non-haematological effects on health in both men and women, which supports the hypothesis that clonal haematopoiesis is a biomarker of genomic instability in other tissues. Single-cell RNA sequencing identifies dysregulated expression of autosomal genes in leukocytes with LOY and provides insights into why clonal expansion of these cells may occur. Collectively, these data highlight the value of studying clonal mosaicism to uncover fundamental mechanisms that underlie cancer and other ageing-related diseases..
Romagnoni, A.
Jégou, S.
Van Steen, K.
Wainrib, G.
Hugot, J.-.
International Inflammatory Bowel Disease Genetics Consortium (IIBDGC),
(2019). Comparative performances of machine learning methods for classifying Crohn Disease patients using genome-wide genotyping data. Sci rep,
Vol.9
(1),
p. 10351.
show abstract
full text
Crohn Disease (CD) is a complex genetic disorder for which more than 140 genes have been identified using genome wide association studies (GWAS). However, the genetic architecture of the trait remains largely unknown. The recent development of machine learning (ML) approaches incited us to apply them to classify healthy and diseased people according to their genomic information. The Immunochip dataset containing 18,227 CD patients and 34,050 healthy controls enrolled and genotyped by the international Inflammatory Bowel Disease genetic consortium (IIBDGC) has been re-analyzed using a set of ML methods: penalized logistic regression (LR), gradient boosted trees (GBT) and artificial neural networks (NN). The main score used to compare the methods was the Area Under the ROC Curve (AUC) statistics. The impact of quality control (QC), imputing and coding methods on LR results showed that QC methods and imputation of missing genotypes may artificially increase the scores. At the opposite, neither the patient/control ratio nor marker preselection or coding strategies significantly affected the results. LR methods, including Lasso, Ridge and ElasticNet provided similar results with a maximum AUC of 0.80. GBT methods like XGBoost, LightGBM and CatBoost, together with dense NN with one or more hidden layers, provided similar AUC values, suggesting limited epistatic effects in the genetic architecture of the trait. ML methods detected near all the genetic variants previously identified by GWAS among the best predictors plus additional predictors with lower effects. The robustness and complementarity of the different methods are also studied. Compared to LR, non-linear models such as GBT or NN may provide robust complementary approaches to identify and classify genetic markers..
Takahashi, H.
Cornish, A.J.
Sud, A.
Law, P.J.
Disney-Hogg, L.
Calvocoressi, L.
Lu, L.
Hansen, H.M.
Smirnov, I.
Walsh, K.M.
Schramm, J.
Hoffmann, P.
Nöthen, M.M.
Jöckel, K.-.
Schildkraut, J.M.
Simon, M.
Bondy, M.
Wrensch, M.
Wiemels, J.L.
Claus, E.B.
Turnbull, C.
Houlston, R.S.
(2019). Mendelian randomization provides support for obesity as a risk factor for meningioma. Sci rep,
Vol.9
(1),
p. 309.
show abstract
full text
Little is known about the causes of meningioma. Obesity and obesity-related traits have been reported in several epidemiological observational studies to be risk factors for meningioma. We performed an analysis of genetic variants associated with obesity-related traits to assess the relationship with meningioma risk using Mendelian randomization (MR), an approach unaffected by biases from temporal variability and reverse causation that might have affected earlier investigations. We considered 11 obesity-related traits, identified genetic instruments for these factors, and assessed their association with meningioma risk using data from a genome-wide association study comprising 1,606 meningioma patients and 9,823 controls. To evaluate the causal relationship between the obesity-related traits and meningioma risk, we consider the estimated odds ratio (OR) of meningioma for each genetic instrument. We identified positive associations between body mass index (odds ratio [ORSD] = 1.27, 95% confidence interval [CI] = 1.03-1.56, P = 0.028) and body fat percentage (ORSD = 1.28, 95% CI = 1.01-1.63, P = 0.042) with meningioma risk, albeit non-significant after correction for multiple testing. Associations for basal metabolic rate, diastolic blood pressure, fasting glucose, high-density lipoprotein cholesterol, low-density lipoprotein cholesterol, systolic blood pressure, total cholesterol, triglycerides and waist circumference with risk of meningioma were non-significant. Our analysis provides additional support for obesity being associated with an increased risk of meningioma..
Turnbull, C.
Loveday, C.
Izatt, L.
Ellard, S.
(2018). Response to Letter to the Editor: "p Val804Met, the Most Frequent Pathogenic Mutation in RET, Confers a Very Low Lifetime Risk of Medullary Thyroid Cancer". J clin endocrinol metab,
Vol.103
(9),
pp. 3518-3519.
full text
Loveday, C.
Josephs, K.
Chubb, D.
Gunning, A.
Izatt, L.
Tischkowitz, M.
Ellard, S.
Turnbull, C.
(2018). p Val804Met, the Most Frequent Pathogenic Mutation in RET, Confers a Very Low Lifetime Risk of Medullary Thyroid Cancer. J clin endocrinol metab,
Vol.103
(11),
pp. 4275-4282.
show abstract
full text
CONTEXT: To date, penetrance figures for medullary thyroid cancer (MTC) for variants in rearranged during transfection (RET) have been estimated from families ascertained because of the presence of MTC. OBJECTIVE: To gain estimates of penetrance, unbiased by ascertainment, we analyzed 61 RET mutations assigned as disease causing by the American Thyroid Association (ATA) in population whole-exome sequencing data. DESIGN: For the 61 RET mutations, we used analyses of the observed allele frequencies in ∼51,000 individuals from the Exome Aggregation Consortium (ExAC) database that were not contributed via The Cancer Genome Atlas (TCGA; non-TCGA ExAC), assuming lifetime penetrance for MTC of 90%, 50%, and unbounded. SETTING: Population-based. RESULTS: Ten of 61 ATA disease-causing RET mutations were present in the non-TCGA ExAC population with observed frequency consistent with penetrance for MTC of >90%. For p.Val804Met, the lifetime penetrance for MTC, estimated from the allele frequency observed, was 4% [95% confidence interval (CI), 0.9% to 8%]. CONCLUSIONS: Based on penetrance analysis in carrier relatives of p.Val804Met-positive cases of MTC, p.Val804Met is currently understood to have high-lifetime penetrance for MTC (87% by age 70), albeit of later onset of MTC than other RET mutations. Given our unbiased estimate of penetrance for RET p.Val804Met of 4% (95% CI, 0.9% to 8%), the current recommendation by the ATA of prophylactic thyroidectomy as standard for all RET mutation carriers is likely inappropriate..
Wang, Y.
Gray, D.R.
Robbins, A.K.
Crowgey, E.L.
Chanock, S.J.
Greene, M.H.
McGlynn, K.A.
Nathanson, K.
Turnbull, C.
Wang, Z.
Devoto, M.
Barthold, J.S.
Testicular Cancer Consortium,
(2018). Subphenotype meta-analysis of testicular cancer genome-wide association study data suggests a role for RBFOX family genes in cryptorchidism susceptibility. Hum reprod,
Vol.33
(5),
pp. 967-977.
show abstract
STUDY QUESTION: Can subphenotype analysis of genome-wide association study (GWAS) data from subjects with testicular germ cell tumor (TGCT) provide insight into cryptorchidism (undescended testis, UDT) susceptibility? SUMMARY ANSWER: Suggestive intragenic GWAS signals common to UDT, TGCT case-case and TGCT case-control analyses occur in genes encoding RBFOX RNA-binding proteins (RBPs) and their neurodevelopmental targets. WHAT IS KNOWN ALREADY: UDT is a strong risk factor for TGCT, but while genetic risk factors for TGCT are well-known, genetic susceptibility to UDT is poorly understood and appears to be more complex. STUDY DESIGN, SIZE, DURATION: We performed a secondary subphenotype analysis of existing GWAS data from the Testicular Cancer Consortium (TECAC) and compared these results with our previously published UDT GWAS data, and with data previously acquired from studies of the fetal rat gubernaculum. PARTICIPANTS/MATERIALS, SETTING, METHODS: Studies from the National Cancer Institute (NCI), United Kingdom (UK) and University of Pennsylvania (Penn) that enrolled white subjects were the source of the TGCT GWAS data. We completed UDT subphenotype case-case (TGCT/UDT vs TGCT/non-UDT) and case-control (TGCT/UDT vs control), collectively referred to as 'TECAC' analyses, followed by a meta-analysis comprising 129 TGCT/UDT cases, 1771 TGCT/non-UDT cases, and 3967 unaffected controls. We reanalyzed our UDT GWAS results comprising 844 cases and 2718 controls by mapping suggestive UDT and TECAC signals (defined as P < 0.001) to genes using Ingenuity Pathway Analysis (IPA®). We compared associated pathways and enriched gene categories common to all analyses after Benjamini-Hochberg multiple testing correction, and analyzed transcript levels and protein expression using qRT-PCR and rat fetal gubernaculum confocal imaging, respectively. MAIN RESULTS AND THE ROLE OF CHANCE: We found suggestive signals within 19 genes common to all three analyses, including RBFOX1 and RBFOX3, neurodevelopmental paralogs that encode RBPs targeting (U)GCATG-containing transcripts. Ten of the 19 genes participate in neurodevelopment and/or contribute to risk of neurodevelopmental disorders. Experimentally predicted RBFOX gene targets were strongly overrepresented among suggestive intragenic signals for the UDT (117 of 628 (19%), P = 3.5 × 10-24), TECAC case-case (129 of 711 (18%), P = 2.5 × 10-27) and TECAC case-control (117 of 679 (17%), P = 2 × 10-21) analyses, and a majority of the genes common to all three analyses (12 of 19 (63%), P = 3 × 10-9) are predicted RBFOX targets. Rbfox1, Rbfox2 and their encoded proteins are expressed in the rat fetal gubernaculum. Predicted RBFOX targets are also enriched among transcripts differentially regulated in the fetal gubernaculum during normal development (P = 3 × 10-31), in response to in vitro hormonal stimulation (P = 5 × 10-45) and in the cryptorchid LE/orl rat (P = 2 × 10-42). LARGE SCALE DATA: GWAS data included in this study are available in the database of Genotypes and Phenotypes (dbGaP accession numbers phs000986.v1.p1 and phs001349.v1p1). LIMITATIONS, REASONS FOR CAUTION: These GWAS data did not reach genome-wide significance for any individual analysis. UDT appears to have a complex etiology that also includes environmental factors, and such complexity may require much larger sample sizes than are currently available. The current methodology may also introduce bias that favors false discovery of larger genes. WIDER IMPLICATIONS OF THE FINDINGS: Common suggestive intragenic GWAS signals suggest that RBFOX paralogs and other neurodevelopmental genes are potential UDT risk candidates, and potential TGCT susceptibility modifiers. Enrichment of predicted RBFOX targets among differentially expressed transcripts in the fetal gubernaculum additionally suggests a role for this RBP family in regulation of testicular descent. As RBFOX proteins regulate alternative splicing of Calca to generate calcitonin gene-related peptide, a protein linked to development and function of the gubernaculum, additional studies that address the role of these proteins in UDT are warranted. STUDY FUNDING/COMPETING INTEREST(S): The Eunice Kennedy Shriver National Institute for Child Health and Human Development (R01HD060769); National Center for Research Resources (P20RR20173), National Institute of General Medical Sciences (P20GM103464), Nemours Biomedical Research, the Testicular Cancer Consortium (U01CA164947), the Intramural Research Program of the NCI, a support services contract HHSN26120130003C with IMS, Inc., the Abramson Cancer Center at Penn, National Cancer Institute (CA114478), the Institute of Cancer Research, UK and the Wellcome Trust Case-Control Consortium (WTCCC) 2. None of the authors reports a conflict of interest..
Manchanda, R.
Patel, S.
Gordeev, V.S.
Antoniou, A.C.
Smith, S.
Lee, A.
Hopper, J.L.
MacInnis, R.J.
Turnbull, C.
Ramus, S.J.
Gayther, S.A.
Pharoah, P.D.
Menon, U.
Jacobs, I.
Legood, R.
(2018). Cost-effectiveness of Population-Based BRCA1, BRCA2, RAD51C, RAD51D, BRIP1, PALB2 Mutation Testing in Unselected General Population Women. J natl cancer inst,
Vol.110
(7),
pp. 714-725.
show abstract
BACKGROUND: The cost-effectiveness of population-based panel testing for high- and moderate-penetrance ovarian cancer (OC)/breast cancer (BC) gene mutations is unknown. We evaluate the cost-effectiveness of population-based BRCA1/BRCA2/RAD51C/RAD51D/BRIP1/PALB2 mutation testing compared with clinical criteria/family history (FH) testing in unselected general population women. METHODS: A decision-analytic model comparing lifetime costs and effects of criteria/FH-based BRCA1/BRCA2 testing is compared with BRCA1/BRCA2/RAD51C/RAD51D/BRIP1/PALB2 testing in those fulfilling clinical criteria/strong FH of cancer (≥10% BRCA1/BRCA2 probability) and all women age 30 years or older. Analyses are presented for UK and US populations. Identified carriers undergo risk-reducing salpingo-oophorectomy. BRCA1/BRCA2/PALB2 carriers can opt for magnetic resonance imaging/mammography, chemoprevention, or risk-reducing mastectomy. One-way and probabilistic sensitivity analysis (PSA) enabled model uncertainty evaluation. Outcomes include OC, BC, and additional heart disease deaths. Quality-adjusted life-years (QALYs), OC incidence, BC incidence, and incremental cost-effectiveness ratio (ICER) were calculated. The time horizon is lifetime and perspective is payer. RESULTS: Compared with clinical criteria/FH-based BRCA1/BRCA2 testing, clinical criteria/FH-based BRCA1/BRCA2/RAD51C/RAD51D/BRIP1/PALB2 testing is cost-effective (ICER = £7629.65/QALY or $49 282.19/QALY; 0.04 days' life-expectancy gained). Population-based testing for BRCA1/BRCA2/RAD51C/RAD51D/BRIP1/PALB2 mutations is the most cost-effective strategy compared with current policy: ICER = £21 599.96/QALY or $54 769.78/QALY (9.34 or 7.57 days' life-expectancy gained). At £30 000/QALY and $100 000/QALY willingness-to-pay thresholds, population-based BRCA1/BRCA2/RAD51C/RAD51D/BRIP1/PALB2 panel testing is the preferred strategy in 83.7% and 92.7% of PSA simulations; criteria/FH-based panel testing is preferred in 16.2% and 5.8% of simulations, respectively. Population-based BRCA1/BRCA2/RAD51C/RAD51D/BRIP1/PALB2 testing can prevent 1.86%/1.91% of BC and 3.2%/4.88% of OC in UK/US women: 657/655 OC cases and 2420/2386 BC cases prevented per million. CONCLUSIONS: Population-based BRCA1/BRCA2/RAD51C/RAD51D/BRIP1/PALB2 testing is more cost-effective than any clinical criteria/FH-based strategy. Clinical criteria/FH-based BRCA1/BRCA2/RAD51C/RAD51D/BRIP1/PALB2 testing is more cost-effective than BRCA1/BRCA2 testing alone..
Turnbull, C.
(2018). Introducing whole-genome sequencing into routine cancer care: the Genomics England 100 000 Genomes Project. Ann oncol,
Vol.29
(4),
pp. 784-787.
Murray, M.J.
Turnbull, C.
(2018). Testicular cancer in 2017: Sequencing advances understanding. Nature reviews. urology,
Vol.15
(2),
pp. 79-80.
Patel, S.
Legood, R.
Evans, D.G.
Turnbull, C.
Antoniou, A.C.
Menon, U.
Jacobs, I.
Manchanda, R.
(2018). Cost effectiveness of population based BRCA1 founder mutation testing in Sephardi Jewish women. Am j obstet gynecol,
Vol.218
(4),
pp. 431.e1-431.e12.
show abstract
BACKGROUND: Population-based BRCA1/BRCA2 founder-mutation testing has been demonstrated as cost effective compared with family history based testing in Ashkenazi Jewish women. However, only 1 of the 3 Ashkenazi Jewish BRCA1/BRCA2 founder mutations (185delAG[c.68_69delAG]), 5382insC[c.5266dupC]), and 6174delT[c.5946delT]) is found in the Sephardi Jewish population (185delAG[c.68_69delAG]), and the overall prevalence of BRCA mutations in the Sephardi Jewish population is accordingly lower (0.7% compared with 2.5% in the Ashkenazi Jewish population). Cost-effectiveness analyses of BRCA testing have not previously been performed at these lower BRCA prevalence levels seen in the Sephardi Jewish population. Here we present a cost-effectiveness analysis for UK and US populations comparing population testing with clinical criteria/family history-based testing in Sephardi Jewish women. STUDY DESIGN: A Markov model was built comparing the lifetime costs and effects of population-based BRCA1 testing, with testing using family history-based clinical criteria in Sephardi Jewish women aged ≥30 years. BRCA1 carriers identified were offered magnetic resonance imaging/mammograms and risk-reducing surgery. Costs are reported at 2015 prices. Outcomes include breast cancer, ovarian cancer, and excess deaths from heart disease. All costs and outcomes are discounted at 3.5%. The time horizon is lifetime, and perspective is payer. The incremental cost-effectiveness ratio per quality-adjusted life-year was calculated. Parameter uncertainty was evaluated through 1-way and probabilistic sensitivity analysis. RESULTS: Population testing resulted in gain in life expectancy of 12 months (quality-adjusted life-year = 1.00). The baseline discounted incremental cost-effectiveness ratio for UK population-based testing was £67.04/quality-adjusted life-year and for US population was $308.42/quality-adjusted life-year. Results were robust in the 1-way sensitivity analysis. The probabilistic sensitivity analysis showed 100% of simulations were cost effective at £20,000/quality-adjusted life-year UK and the $100,000/quality-adjusted life-year US willingness-to-pay thresholds. Scenario analysis showed that population testing remains cost effective in UK and US populations, even if premenopausal oophorectomy does not reduce breast cancer risk or if hormone replacement therapy compliance is nil. CONCLUSION: Population-based BRCA1 testing is highly cost effective compared with clinical criteria-driven approach in Sephardi Jewish women. This supports changing the paradigm to population-based BRCA testing in the Jewish population, regardless of Ashkenazi/Sephardi ancestry..
Takahashi, H.
Cornish, A.J.
Sud, A.
Law, P.J.
Kinnersley, B.
Ostrom, Q.T.
Labreche, K.
Eckel-Passow, J.E.
Armstrong, G.N.
Claus, E.B.
Il'yasova, D.
Schildkraut, J.
Barnholtz-Sloan, J.S.
Olson, S.H.
Bernstein, J.L.
Lai, R.K.
Schoemaker, M.J.
Simon, M.
Hoffmann, P.
Nöthen, M.M.
Jöckel, K.-.
Chanock, S.
Rajaraman, P.
Johansen, C.
Jenkins, R.B.
Melin, B.S.
Wrensch, M.R.
Sanson, M.
Bondy, M.L.
Turnbull, C.
Houlston, R.S.
(2018). Mendelian randomisation study of the relationship between vitamin D and risk of glioma. Sci rep,
Vol.8
(1),
p. 2339.
show abstract
full text
To examine for a causal relationship between vitamin D and glioma risk we performed an analysis of genetic variants associated with serum 25-hydroxyvitamin D (25(OH)D) levels using Mendelian randomisation (MR), an approach unaffected by biases from confounding. Two-sample MR was undertaken using genome-wide association study data. Single nucleotide polymorphisms (SNPs) associated with 25(OH)D levels were used as instrumental variables (IVs). We calculated MR estimates for the odds ratio (OR) for 25(OH)D levels with glioma using SNP-glioma estimates from 12,488 cases and 18,169 controls, using inverse-variance weighted (IVW) and maximum likelihood estimation (MLE) methods. A non-significant association between 25(OH)D levels and glioma risk was shown using both the IVW (OR = 1.21, 95% confidence interval [CI] = 0.90-1.62, P = 0.201) and MLE (OR = 1.20, 95% CI = 0.98-1.48, P = 0.083) methods. In an exploratory analysis of tumour subtype, an inverse relationship between 25(OH)D levels and glioblastoma (GBM) risk was identified using the MLE method (OR = 0.62, 95% CI = 0.43-0.89, P = 0.010), but not the IVW method (OR = 0.62, 95% CI = 0.37-1.04, P = 0.070). No statistically significant association was shown between 25(OH)D levels and non-GBM glioma. Our results do not provide evidence for a causal relationship between 25(OH)D levels and all forms of glioma risk. More evidence is required to explore the relationship between 25(OH)D levels and risk of GBM..
Taylor, A.
Brady, A.F.
Frayling, I.M.
Hanson, H.
Tischkowitz, M.
Turnbull, C.
Side, L.
UK Cancer Genetics Group (UK-CGG),
(2018). Consensus for genes to be included on cancer panel tests offered by UK genetics services: guidelines of the UK Cancer Genetics Group. J med genet,
Vol.55
(6),
pp. 372-377.
show abstract
Genetic testing for hereditary cancer predisposition has evolved rapidly in recent years with the discovery of new genes, but there is much debate over the clinical utility of testing genes for which there are currently limited data regarding the degree of associated cancer risk. To address the discrepancies that have arisen in the provision of these tests across the UK, the UK Cancer Genetics Group facilitated a 1-day workshop with representation from the majority of National Health Service (NHS) clinical genetics services. Using a preworkshop survey followed by focused discussion of genes without prior majority agreement for inclusion, we achieved consensus for panels of cancer genes with sufficient evidence for clinical utility, to be adopted by all NHS genetics services. To support consistency in the delivery of these tests and advice given to families across the country, we also developed management proposals for individuals who are found to have pathogenic mutations in these genes. However, we fully acknowledge that the decision regarding what test is most appropriate for an individual family rests with the clinician, and will depend on factors including specific phenotypic features and the family structure..
Manchanda, R.
Blyuss, O.
Gaba, F.
Gordeev, V.S.
Jacobs, C.
Burnell, M.
Gan, C.
Taylor, R.
Turnbull, C.
Legood, R.
Zaikin, A.
Antoniou, A.C.
Menon, U.
Jacobs, I.
(2018). Current detection rates and time-to-detection of all identifiable BRCA carriers in the Greater London population. J med genet,
Vol.55
(8),
pp. 538-545.
show abstract
BACKGROUND: BRCA carrier identification offers opportunities for early diagnoses, targeted treatment and cancer prevention. We evaluate BRCA- carrier detection rates in general and Ashkenazi Jewish (AJ) populations across Greater London and estimate time-to-detection of all identifiable BRCA carriers. METHODS: BRCA carrier data from 1993 to 2014 were obtained from National Health Service genetic laboratories and compared with modelled predictions of BRCA prevalence from published literature and geographical data from UK Office for National Statistics. Proportion of BRCA carriers identified was estimated. Prediction models were developed to fit BRCA detection rate data. BRCA carrier identification rates were evaluated for an 'Angelina Jolie effect'. Maps for four Greater London regions were constructed, and their relative BRCA detection rates were compared. Models developed were used to predict future time-to-identify all detectable BRCA carriers in AJ and general populations. RESULTS: Until 2014, only 2.6% (3072/111 742 estimated) general population and 10.9% (548/4985 estimated) AJ population BRCA carriers have been identified in 16 696 608 (AJ=190 997) Greater London population. 57% general population and 54% AJ mutations were identified through cascade testing. Current detection rates mirror linear fit rather than parabolic model and will not identify all BRCA carriers. Addition of unselected ovarian/triple-negative breast cancer testing would take >250 years to identify all BRCA carriers. Doubling current detection rates can identify all 'detectable' BRCA carriers in the general population by year 2181, while parabolic and triple linear rates can identify 'detectable' BRCA carriers by 2084 and 2093, respectively. The linear fit model can identify 'detectable' AJ carriers by 2044. We did not find an Angelina Jolie effect on BRCA carrier detection rates. There was a significant difference in BRCA detection rates between geographical regions over time (P<0.001). CONCLUSIONS: The majority of BRCA carriers have not been identified, missing key opportunities for prevention/earlier diagnosis. Enhanced and new strategies/approaches are needed..
Loveday, C.
Law, P.
Litchfield, K.
Levy, M.
Holroyd, A.
Broderick, P.
Kote-Jarai, Z.
Dunning, A.M.
Muir, K.
Peto, J.
Eeles, R.
Easton, D.F.
Dudakia, D.
Orr, N.
Pashayan, N.
UK Testicular Cancer Collaboration, The PRACTICAL Consortium,
Reid, A.
Huddart, R.A.
Houlston, R.S.
Turnbull, C.
(2018). Large-scale Analysis Demonstrates Familial Testicular Cancer to have Polygenic Aetiology. Eur urol,
Vol.74
(3),
pp. 248-252.
show abstract
full text
UNLABELLED: Testicular germ cell tumour (TGCT) is the most common cancer in young men. Multiplex TGCT families have been well reported and analyses of population cancer registries have demonstrated a four- to eightfold risk to male relatives of TGCT patients. Early linkage analysis and recent large-scale germline exome analysis in TGCT cases demonstrate absence of major high-penetrance TGCT susceptibility gene(s). Serial genome-wide association study analyses in sporadic TGCT have in total reported 49 independent risk loci. To date, it has not been demonstrated whether familial TGCT arises due to enrichment of the same common variants underpinning susceptibility to sporadic TGCT or is due to shared environmental/lifestyle factors or disparate rare genetic TGCT susceptibility factors. Here we present polygenic risk score analysis of 37 TGCT susceptibility single-nucleotide polymorphisms in 236 familial and 3931 sporadic TGCT cases, and 12 368 controls, which demonstrates clear enrichment for TGCT susceptibility alleles in familial compared to sporadic cases (p=0.0001), with the majority of familial cases (84-100%) being attributable to polygenic enrichment. These analyses reveal TGCT as the first rare malignancy of early adulthood in which familial clustering is driven by the aggregate effects of polygenic variation in the absence of a major high-penetrance susceptibility gene. PATIENT SUMMARY: To date, it has been unclear whether familial clusters of testicular germ cell tumour (TGCT) arise due to genetics or shared environmental or lifestyle factors. We present large-scale genetic analyses comparing 236 familial TGCT cases, 3931 isolated TGCT cases, and 12 368 controls. We show that familial TGCT is caused, at least in part, by presence of a higher dose of the same common genetic variants that cause susceptibility to TGCT in general..
Turnbull, C.
Sud, A.
Houlston, R.S.
(2018). Cancer genetics, precision prevention and a call to action. Nat genet,
Vol.50
(9),
pp. 1212-1218.
show abstract
full text
More than 15 years have passed since the identification, through linkage, of 'first-wave' susceptibility genes for common cancers (BRCA1, BRCA2, MLH1 and MSH2). These genes have strong frequency-penetrance profiles, such that the associated clinical utility probably remains relevant regardless of the context of ascertainment. 'Second-wave' genes, not tractable by linkage, were subsequently identified by mutation screening of candidate genes (PALB2, ATM, CHEK2, BRIP1, RAD51C and RAD51D). Their innately weaker frequency-penetrance profiles have rendered delineation of cancer associations, risks and variant pathogenicity challenging, thereby compromising their clinical application. Early germline exome-sequencing endeavors for common cancers did not yield the long-anticipated slew of 'next-wave' genes but instead implied a highly polygenic genomic architecture requiring much larger experiments to make any substantive inroads into gene discovery. As such, the 'genetic economics' of frequency penetrance clearly indicates that focused identification of carriers of first-wave-gene mutations is most impactful for cancer control. With screening, prevention and early detection at the forefront of the cancer management agenda, we propose that the time is nigh for the initiation of national population-testing programs to identify carriers of first-wave gene mutation carriers. To fully deliver a precision prevention program, long-term, large-scale mutation studies that capture longitudinal clinical data and serial biosamples are required..
Robbe, P.
Popitsch, N.
Knight, S.J.
Antoniou, P.
Becq, J.
He, M.
Kanapin, A.
Samsonova, A.
Vavoulis, D.V.
Ross, M.T.
Kingsbury, Z.
Cabes, M.
Ramos, S.D.
Page, S.
Dreau, H.
Ridout, K.
Jones, L.J.
Tuff-Lacey, A.
Henderson, S.
Mason, J.
Buffa, F.M.
Verrill, C.
Maldonado-Perez, D.
Roxanis, I.
Collantes, E.
Browning, L.
Dhar, S.
Damato, S.
Davies, S.
Caulfield, M.
Bentley, D.R.
Taylor, J.C.
Turnbull, C.
Schuh, A.
100,000 Genomes Project,
(2018). Clinical whole-genome sequencing from routine formalin-fixed, paraffin-embedded specimens: pilot study for the 100,000 Genomes Project. Genet med,
Vol.20
(10),
pp. 1196-1205.
show abstract
PURPOSE: Fresh-frozen (FF) tissue is the optimal source of DNA for whole-genome sequencing (WGS) of cancer patients. However, it is not always available, limiting the widespread application of WGS in clinical practice. We explored the viability of using formalin-fixed, paraffin-embedded (FFPE) tissues, available routinely for cancer patients, as a source of DNA for clinical WGS. METHODS: We conducted a prospective study using DNAs from matched FF, FFPE, and peripheral blood germ-line specimens collected from 52 cancer patients (156 samples) following routine diagnostic protocols. We compared somatic variants detected in FFPE and matching FF samples. RESULTS: We found the single-nucleotide variant agreement reached 71% across the genome and somatic copy-number alterations (CNAs) detection from FFPE samples was suboptimal (0.44 median correlation with FF) due to nonuniform coverage. CNA detection was improved significantly with lower reverse crosslinking temperature in FFPE DNA extraction (80 °C or 65 °C depending on the methods). Our final data showed somatic variant detection from FFPE for clinical decision making is possible. We detected 98% of clinically actionable variants (including 30/31 CNAs). CONCLUSION: We present the first prospective WGS study of cancer patients using FFPE specimens collected in a routine clinical environment proving WGS can be applied in the clinic..
Turnbull, C.
(2018). Genomics in medicine. Medicine,
Vol.46
(12),
pp. 774-779.
Turnbull, C.
Scott, R.H.
Thomas, E.
Jones, L.
Murugaesu, N.
Pretty, F.B.
Halai, D.
Baple, E.
Craig, C.
Hamblin, A.
Henderson, S.
Patch, C.
O'Neill, A.
Devereau, A.
Smith, K.
Martin, A.R.
Sosinsky, A.
McDonagh, E.M.
Sultana, R.
Mueller, M.
Smedley, D.
Toms, A.
Dinh, L.
Fowler, T.
Bale, M.
Hubbard, T.
Rendon, A.
Hill, S.
Caulfield, M.J.
100 000 Genomes Project,
(2018). The 100 000 Genomes Project: bringing whole genome sequencing to the NHS. Bmj,
Vol.361,
p. k1687.
Loveday, C.
Litchfield, K.
Levy, M.
Holroyd, A.
Broderick, P.
Kote-Jarai, Z.
Dunning, A.M.
Muir, K.
Peto, J.
Eeles, R.
Easton, D.F.
Dudakia, D.
Orr, N.
Pashayan, N.
Reid, A.
Huddart, R.A.
Houlston, R.S.
Turnbull, C.
(2018). Validation of loci at 2q14 2 and 15q21 3 as risk factors for testicular cancer. Oncotarget,
Vol.9
(16),
pp. 12630-12638.
show abstract
full text
Testicular germ cell tumor (TGCT), the most common cancer in men aged 18 to 45 years, has a strong heritable basis. Genome-wide association studies (GWAS) have proposed single nucleotide polymorphisms (SNPs) at a number of loci influencing TGCT risk. To further evaluate the association of recently proposed risk SNPs with TGCT at 2q14.2, 3q26.2, 7q36.3, 10q26.13 and 15q21.3, we analyzed genotype data on 3,206 cases and 7,422 controls. Our analysis provides independent replication of the associations for risk SNPs at 2q14.2 (rs2713206 at P = 3.03 × 10-2; P-meta = 3.92 × 10-8; nearest gene, TFCP2L1) and rs12912292 at 15q21.3 (P = 7.96 × 10-11; P-meta = 1.55 × 10-19; nearest gene PRTG). Case-only analyses did not reveal specific associations with TGCT histology. TFCP2L1 joins the growing list of genes located within TGCT risk loci with biologically plausible roles in developmental transcriptional regulation, further highlighting the importance of this phenomenon in TGCT oncogenesis..
Telomeres Mendelian Randomization Collaboration,
Haycock, P.C.
Burgess, S.
Nounu, A.
Zheng, J.
Okoli, G.N.
Bowden, J.
Wade, K.H.
Timpson, N.J.
Evans, D.M.
Willeit, P.
Aviv, A.
Gaunt, T.R.
Hemani, G.
Mangino, M.
Ellis, H.P.
Kurian, K.M.
Pooley, K.A.
Eeles, R.A.
Lee, J.E.
Fang, S.
Chen, W.V.
Law, M.H.
Bowdler, L.M.
Iles, M.M.
Yang, Q.
Worrall, B.B.
Markus, H.S.
Hung, R.J.
Amos, C.I.
Spurdle, A.B.
Thompson, D.J.
O'Mara, T.A.
Wolpin, B.
Amundadottir, L.
Stolzenberg-Solomon, R.
Trichopoulou, A.
Onland-Moret, N.C.
Lund, E.
Duell, E.J.
Canzian, F.
Severi, G.
Overvad, K.
Gunter, M.J.
Tumino, R.
Svenson, U.
van Rij, A.
Baas, A.F.
Bown, M.J.
Samani, N.J.
van t'Hof, F.N.
Tromp, G.
Jones, G.T.
Kuivaniemi, H.
Elmore, J.R.
Johansson, M.
Mckay, J.
Scelo, G.
Carreras-Torres, R.
Gaborieau, V.
Brennan, P.
Bracci, P.M.
Neale, R.E.
Olson, S.H.
Gallinger, S.
Li, D.
Petersen, G.M.
Risch, H.A.
Klein, A.P.
Han, J.
Abnet, C.C.
Freedman, N.D.
Taylor, P.R.
Maris, J.M.
Aben, K.K.
Kiemeney, L.A.
Vermeulen, S.H.
Wiencke, J.K.
Walsh, K.M.
Wrensch, M.
Rice, T.
Turnbull, C.
Litchfield, K.
Paternoster, L.
Standl, M.
Abecasis, G.R.
SanGiovanni, J.P.
Li, Y.
Mijatovic, V.
Sapkota, Y.
Low, S.-.
Zondervan, K.T.
Montgomery, G.W.
Nyholt, D.R.
van Heel, D.A.
Hunt, K.
Arking, D.E.
Ashar, F.N.
Sotoodehnia, N.
Woo, D.
Rosand, J.
Comeau, M.E.
Brown, W.M.
Silverman, E.K.
Hokanson, J.E.
Cho, M.H.
Hui, J.
Ferreira, M.A.
Thompson, P.J.
Morrison, A.C.
Felix, J.F.
Smith, N.L.
Christiano, A.M.
Petukhova, L.
Betz, R.C.
Fan, X.
Zhang, X.
Zhu, C.
Langefeld, C.D.
Thompson, S.D.
Wang, F.
Lin, X.
Schwartz, D.A.
Fingerlin, T.
Rotter, J.I.
Cotch, M.F.
Jensen, R.A.
Munz, M.
Dommisch, H.
Schaefer, A.S.
Han, F.
Ollila, H.M.
Hillary, R.P.
Albagha, O.
Ralston, S.H.
Zeng, C.
Zheng, W.
Shu, X.-.
Reis, A.
Uebe, S.
Hüffmeier, U.
Kawamura, Y.
Otowa, T.
Sasaki, T.
Hibberd, M.L.
Davila, S.
Xie, G.
Siminovitch, K.
Bei, J.-.
Zeng, Y.-.
Försti, A.
Chen, B.
Landi, S.
Franke, A.
Fischer, A.
Ellinghaus, D.
Flores, C.
Noth, I.
Ma, S.-.
Foo, J.N.
Liu, J.
Kim, J.-.
Cox, D.G.
Delattre, O.
Mirabeau, O.
Skibola, C.F.
Tang, C.S.
Garcia-Barcelo, M.
Chang, K.-.
Su, W.-.
Chang, Y.-.
Martin, N.G.
Gordon, S.
Wade, T.D.
Lee, C.
Kubo, M.
Cha, P.-.
Nakamura, Y.
Levy, D.
Kimura, M.
Hwang, S.-.
Hunt, S.
Spector, T.
Soranzo, N.
Manichaikul, A.W.
Barr, R.G.
Kahali, B.
Speliotes, E.
Yerges-Armstrong, L.M.
Cheng, C.-.
Jonas, J.B.
Wong, T.Y.
Fogh, I.
Lin, K.
Powell, J.F.
Rice, K.
Relton, C.L.
Martin, R.M.
Davey Smith, G.
(2017). Association Between Telomere Length and Risk of Cancer and Non-Neoplastic Diseases: A Mendelian Randomization Study. Jama oncol,
Vol.3
(5),
pp. 636-651.
show abstract
full text
IMPORTANCE: The causal direction and magnitude of the association between telomere length and incidence of cancer and non-neoplastic diseases is uncertain owing to the susceptibility of observational studies to confounding and reverse causation. OBJECTIVE: To conduct a Mendelian randomization study, using germline genetic variants as instrumental variables, to appraise the causal relevance of telomere length for risk of cancer and non-neoplastic diseases. DATA SOURCES: Genomewide association studies (GWAS) published up to January 15, 2015. STUDY SELECTION: GWAS of noncommunicable diseases that assayed germline genetic variation and did not select cohort or control participants on the basis of preexisting diseases. Of 163 GWAS of noncommunicable diseases identified, summary data from 103 were available. DATA EXTRACTION AND SYNTHESIS: Summary association statistics for single nucleotide polymorphisms (SNPs) that are strongly associated with telomere length in the general population. MAIN OUTCOMES AND MEASURES: Odds ratios (ORs) and 95% confidence intervals (CIs) for disease per standard deviation (SD) higher telomere length due to germline genetic variation. RESULTS: Summary data were available for 35 cancers and 48 non-neoplastic diseases, corresponding to 420 081 cases (median cases, 2526 per disease) and 1 093 105 controls (median, 6789 per disease). Increased telomere length due to germline genetic variation was generally associated with increased risk for site-specific cancers. The strongest associations (ORs [95% CIs] per 1-SD change in genetically increased telomere length) were observed for glioma, 5.27 (3.15-8.81); serous low-malignant-potential ovarian cancer, 4.35 (2.39-7.94); lung adenocarcinoma, 3.19 (2.40-4.22); neuroblastoma, 2.98 (1.92-4.62); bladder cancer, 2.19 (1.32-3.66); melanoma, 1.87 (1.55-2.26); testicular cancer, 1.76 (1.02-3.04); kidney cancer, 1.55 (1.08-2.23); and endometrial cancer, 1.31 (1.07-1.61). Associations were stronger for rarer cancers and at tissue sites with lower rates of stem cell division. There was generally little evidence of association between genetically increased telomere length and risk of psychiatric, autoimmune, inflammatory, diabetic, and other non-neoplastic diseases, except for coronary heart disease (OR, 0.78 [95% CI, 0.67-0.90]), abdominal aortic aneurysm (OR, 0.63 [95% CI, 0.49-0.81]), celiac disease (OR, 0.42 [95% CI, 0.28-0.61]) and interstitial lung disease (OR, 0.09 [95% CI, 0.05-0.15]). CONCLUSIONS AND RELEVANCE: It is likely that longer telomeres increase risk for several cancers but reduce risk for some non-neoplastic diseases, including cardiovascular diseases..
Litchfield, K.
Levy, M.
Orlando, G.
Loveday, C.
Law, P.J.
Migliorini, G.
Holroyd, A.
Broderick, P.
Karlsson, R.
Haugen, T.B.
Kristiansen, W.
Nsengimana, J.
Fenwick, K.
Assiotis, I.
Kote-Jarai, Z.
Dunning, A.M.
Muir, K.
Peto, J.
Eeles, R.
Easton, D.F.
Dudakia, D.
Orr, N.
Pashayan, N.
UK Testicular Cancer Collaboration,
PRACTICAL Consortium,
Bishop, D.T.
Reid, A.
Huddart, R.A.
Shipley, J.
Grotmol, T.
Wiklund, F.
Houlston, R.S.
Turnbull, C.
(2017). Identification of 19 new risk loci and potential regulatory mechanisms influencing susceptibility to testicular germ cell tumor. Nat genet,
Vol.49
(7),
pp. 1133-1140.
show abstract
full text
Genome-wide association studies (GWAS) have transformed understanding of susceptibility to testicular germ cell tumors (TGCTs), but much of the heritability remains unexplained. Here we report a new GWAS, a meta-analysis with previous GWAS and a replication series, totaling 7,319 TGCT cases and 23,082 controls. We identify 19 new TGCT risk loci, roughly doubling the number of known TGCT risk loci to 44. By performing in situ Hi-C in TGCT cells, we provide evidence for a network of physical interactions among all 44 TGCT risk SNPs and candidate causal genes. Our findings implicate widespread disruption of developmental transcriptional regulators as a basis of TGCT susceptibility, consistent with failed primordial germ cell differentiation as an initiating step in oncogenesis. Defective microtubule assembly and dysregulation of KIT-MAPK signaling also feature as recurrently disrupted pathways. Our findings support a polygenic model of risk and provide insight into the biological basis of TGCT..
Wang, Z.
McGlynn, K.A.
Rajpert-De Meyts, E.
Bishop, D.T.
Chung, C.C.
Dalgaard, M.D.
Greene, M.H.
Gupta, R.
Grotmol, T.
Haugen, T.B.
Karlsson, R.
Litchfield, K.
Mitra, N.
Nielsen, K.
Pyle, L.C.
Schwartz, S.M.
Thorsson, V.
Vardhanabhuti, S.
Wiklund, F.
Turnbull, C.
Chanock, S.J.
Kanetsky, P.A.
Nathanson, K.L.
Testicular Cancer Consortium,
(2017). Meta-analysis of five genome-wide association studies identifies multiple new loci associated with testicular germ cell tumor. Nat genet,
Vol.49
(7),
pp. 1141-1147.
show abstract
full text
The international Testicular Cancer Consortium (TECAC) combined five published genome-wide association studies of testicular germ cell tumor (TGCT; 3,558 cases and 13,970 controls) to identify new susceptibility loci. We conducted a fixed-effects meta-analysis, including, to our knowledge, the first analysis of the X chromosome. Eight new loci mapping to 2q14.2, 3q26.2, 4q35.2, 7q36.3, 10q26.13, 15q21.3, 15q22.31, and Xq28 achieved genome-wide significance (P < 5 × 10-8). Most loci harbor biologically plausible candidate genes. We refined previously reported associations at 9p24.3 and 19p12 by identifying one and three additional independent SNPs, respectively. In aggregate, the 39 independent markers identified to date explain 37% of father-to-son familial risk, 8% of which can be attributed to the 12 new signals reported here. Our findings substantially increase the number of known TGCT susceptibility alleles, move the field closer to a comprehensive understanding of the underlying genetic architecture of TGCT, and provide further clues to the etiology of TGCT..
Levy, M.
Hall, D.
Sud, A.
Law, P.
Litchfield, K.
Dudakia, D.
Haugen, T.B.
Karlsson, R.
Reid, A.
Huddart, R.A.
Grotmol, T.
Wiklund, F.
Houlston, R.S.
Turnbull, C.
(2017). Mendelian randomisation analysis provides no evidence for a relationship between adult height and testicular cancer risk. Andrology,
Vol.5
(5),
pp. 914-922.
show abstract
full text
Observational studies have suggested anthropometric traits, particularly increased height are associated with an elevated risk of testicular cancer (testicular germ cell tumour). However, there is an inconsistency between study findings, suggesting the possibility of the influence of confounding factors. To examine the association between anthropometric traits and testicular germ cell tumour using an unbiased approach, we performed a Mendelian randomisation study. We used genotype data from genome wide association studies of testicular germ cell tumour totalling 5518 cases and 19,055 controls. Externally weighted polygenic risk scores were created and used to evaluate associations with testicular germ cell tumour risk per one standard deviation (s.d) increase in genetically-defined adult height, adult BMI, adult waist hip ratio adjusted for BMI (WHRadjBMI), adult hip circumference adjusted for BMI (HIPadjBMI), adult waist circumference adjusted for BMI (WCadjBMI), birth weight (BW) and childhood obesity. Mendelian randomisation analysis did not demonstrate an association between any anthropometric trait and testicular germ cell tumour risk. In particular, despite good power, there was no global evidence for association between height and testicular germ cell tumour. However, three SNPs for adult height individually showed association with testicular germ cell tumour (rs4624820: OR = 1.47, 95% CI: 1.41-1.55, p = 2.7 × 10-57 ; rs12228415: OR = 1.17, 95% CI: 1.11-1.22, p = 3.1 × 10-10 ; rs7568069: OR = 1.13, 95% CI: 1.07-1.18, p = 1.1 × 10-6 ). This Mendelian randomisation analysis, based on the largest testicular germ cell tumour genome wide association dataset to date, does not support a causal etiological association between anthropometric traits and testicular germ cell tumour aetiology. Our findings are more compatible with confounding by shared environmental factors, possibly related to prenatal growth with exposure to these risk factors occurring in utero..
Feng, Y.-.
Cho, K.
Lindstrom, S.
Kraft, P.
Cormack, J.
IGAP Consortium, Colorectal Transdisciplinary Study (CORECT),
Discovery, Biology, and Risk of Inherited Variants in Breast Cancer (DRIVE),
Elucidating Loci Involved in Prostate Cancer Susceptibility (ELLIPSE),
Transdisciplinary Research in Cancer of the Lung (TRICL),
Liang, L.
Driver, J.A.
(2017). Investigating the genetic relationship between Alzheimer's disease and cancer using GWAS summary statistics. Hum genet,
Vol.136
(10),
pp. 1341-1351.
show abstract
full text
Growing evidence from both epidemiology and basic science suggest an inverse association between Alzheimer's disease (AD) and cancer. We examined the genetic relationship between AD and various cancer types using GWAS summary statistics from the IGAP and GAME-ON consortia. Sample size ranged from 9931 to 54,162; SNPs were imputed to the 1000 Genomes European panel. Our results based on cross-trait LD Score regression showed a significant positive genetic correlation between AD and five cancers combined (colon, breast, prostate, ovarian, lung; r g = 0.17, P = 0.04), and specifically with breast cancer (ER-negative and overall; r g = 0.21 and 0.18, P = 0.035 and 0.034) and lung cancer (adenocarcinoma, squamous cell carcinoma and overall; r g = 0.31, 0.38 and 0.30, P = 0.029, 0.016, and 0.006). Estimating the genetic correlation in specific functional categories revealed mixed positive and negative signals, notably stronger at annotations associated with increased enhancer activity. This suggests a role of gene expression regulators in the shared genetic etiology between AD and cancer, and that some shared variants modulate disease risk concordantly while others have effects in opposite directions. Due to power issues, we did not detect cross-phenotype associations at individual SNPs. This genetic overlap is not likely driven by a handful of major loci. Our study is the first to examine the co-heritability of AD and cancer leveraging large-scale GWAS results. The functional categories highlighted in this study need further investigation to illustrate the details of the genetic sharing and to bridge between different levels of associations..
Manchanda, R.
Patel, S.
Antoniou, A.C.
Levy-Lahad, E.
Turnbull, C.
Evans, D.G.
Hopper, J.L.
Macinnis, R.J.
Menon, U.
Jacobs, I.
Legood, R.
(2017). Cost-effectiveness of population based BRCA testing with varying Ashkenazi Jewish ancestry. Am j obstet gynecol,
Vol.217
(5),
pp. 578.e1-578.e12.
show abstract
BACKGROUND: Population-based BRCA1/BRCA2 testing has been found to be cost-effective compared with family history-based testing in Ashkenazi-Jewish women were >30 years old with 4 Ashkenazi-Jewish grandparents. However, individuals may have 1, 2, or 3 Ashkenazi-Jewish grandparents, and cost-effectiveness data are lacking at these lower BRCA prevalence estimates. We present an updated cost-effectiveness analysis of population BRCA1/BRCA2 testing for women with 1, 2, and 3 Ashkenazi-Jewish grandparents. STUDY DESIGN: Decision analysis model. METHODS: Lifetime costs and effects of population and family history-based testing were compared with the use of a decision analysis model. 56% BRCA carriers are missed by family history criteria alone. Analyses were conducted for United Kingdom and United States populations. Model parameters were obtained from the Genetic Cancer Prediction through Population Screening trial and published literature. Model parameters and BRCA population prevalence for individuals with 3, 2, or 1 Ashkenazi-Jewish grandparent were adjusted for the relative frequency of BRCA mutations in the Ashkenazi-Jewish and general populations. Incremental cost-effectiveness ratios were calculated for all Ashkenazi-Jewish grandparent scenarios. Costs, along with outcomes, were discounted at 3.5%. The time horizon of the analysis is "life-time," and perspective is "payer." Probabilistic sensitivity analysis evaluated model uncertainty. RESULTS: Population testing for BRCA mutations is cost-saving in Ashkenazi-Jewish women with 2, 3, or 4 grandparents (22-33 days life-gained) in the United Kingdom and 1, 2, 3, or 4 grandparents (12-26 days life-gained) in the United States populations, respectively. It is also extremely cost-effective in women in the United Kingdom with just 1 Ashkenazi-Jewish grandparent with an incremental cost-effectiveness ratio of £863 per quality-adjusted life-years and 15 days life gained. Results show that population-testing remains cost-effective at the £20,000-30000 per quality-adjusted life-years and $100,000 per quality-adjusted life-years willingness-to-pay thresholds for all 4 Ashkenazi-Jewish grandparent scenarios, with ≥95% simulations found to be cost-effective on probabilistic sensitivity analysis. Population-testing remains cost-effective in the absence of reduction in breast cancer risk from oophorectomy and at lower risk-reducing mastectomy (13%) or risk-reducing salpingo-oophorectomy (20%) rates. CONCLUSION: Population testing for BRCA mutations with varying levels of Ashkenazi-Jewish ancestry is cost-effective in the United Kingdom and the United States. These results support population testing in Ashkenazi-Jewish women with 1-4 Ashkenazi-Jewish grandparent ancestry..
Foulkes, W.D.
Knoppers, B.M.
Turnbull, C.
(2016). Population genetic testing for cancer susceptibility: founder mutations to genomes. Nat rev clin oncol,
Vol.13
(1),
pp. 41-54.
show abstract
The current standard model for identifying carriers of high-risk mutations in cancer-susceptibility genes (CSGs) generally involves a process that is not amenable to population-based testing: access to genetic tests is typically regulated by health-care providers on the basis of a labour-intensive assessment of an individual's personal and family history of cancer, with face-to-face genetic counselling performed before mutation testing. Several studies have shown that application of these selection criteria results in a substantial proportion of mutation carriers being missed. Population-based genetic testing has been proposed as an alternative approach to determining cancer susceptibility, and aims for a more-comprehensive detection of mutation carriers. Herein, we review the existing data on population-based genetic testing, and consider some of the barriers, pitfalls, and challenges related to the possible expansion of this approach. We consider mechanisms by which population-based genetic testing for cancer susceptibility could be delivered, and suggest how such genetic testing might be integrated into existing and emerging health-care structures. The existing models of genetic testing (including issues relating to informed consent) will very likely require considerable alteration if the potential benefits of population-based genetic testing are to be fully realized..
Fehringer, G.
Kraft, P.
Pharoah, P.D.
Eeles, R.A.
Chatterjee, N.
Schumacher, F.R.
Schildkraut, J.M.
Lindström, S.
Brennan, P.
Bickeböller, H.
Houlston, R.S.
Landi, M.T.
Caporaso, N.
Risch, A.
Amin Al Olama, A.
Berndt, S.I.
Giovannucci, E.L.
Grönberg, H.
Kote-Jarai, Z.
Ma, J.
Muir, K.
Stampfer, M.J.
Stevens, V.L.
Wiklund, F.
Willett, W.C.
Goode, E.L.
Permuth, J.B.
Risch, H.A.
Reid, B.M.
Bezieau, S.
Brenner, H.
Chan, A.T.
Chang-Claude, J.
Hudson, T.J.
Kocarnik, J.K.
Newcomb, P.A.
Schoen, R.E.
Slattery, M.L.
White, E.
Adank, M.A.
Ahsan, H.
Aittomäki, K.
Baglietto, L.
Blomquist, C.
Canzian, F.
Czene, K.
Dos-Santos-Silva, I.
Eliassen, A.H.
Figueroa, J.D.
Flesch-Janys, D.
Fletcher, O.
Garcia-Closas, M.
Gaudet, M.M.
Johnson, N.
Hall, P.
Hazra, A.
Hein, R.
Hofman, A.
Hopper, J.L.
Irwanto, A.
Johansson, M.
Kaaks, R.
Kibriya, M.G.
Lichtner, P.
Liu, J.
Lund, E.
Makalic, E.
Meindl, A.
Müller-Myhsok, B.
Muranen, T.A.
Nevanlinna, H.
Peeters, P.H.
Peto, J.
Prentice, R.L.
Rahman, N.
Sanchez, M.J.
Schmidt, D.F.
Schmutzler, R.K.
Southey, M.C.
Tamimi, R.
Travis, R.C.
Turnbull, C.
Uitterlinden, A.G.
Wang, Z.
Whittemore, A.S.
Yang, X.R.
Zheng, W.
Buchanan, D.D.
Casey, G.
Conti, D.V.
Edlund, C.K.
Gallinger, S.
Haile, R.W.
Jenkins, M.
Le Marchand, L.
Li, L.
Lindor, N.M.
Schmit, S.L.
Thibodeau, S.N.
Woods, M.O.
Rafnar, T.
Gudmundsson, J.
Stacey, S.N.
Stefansson, K.
Sulem, P.
Chen, Y.A.
Tyrer, J.P.
Christiani, D.C.
Wei, Y.
Shen, H.
Hu, Z.
Shu, X.-.
Shiraishi, K.
Takahashi, A.
Bossé, Y.
Obeidat, M.
Nickle, D.
Timens, W.
Freedman, M.L.
Li, Q.
Seminara, D.
Chanock, S.J.
Gong, J.
Peters, U.
Gruber, S.B.
Amos, C.I.
Sellers, T.A.
Easton, D.F.
Hunter, D.J.
Haiman, C.A.
Henderson, B.E.
Hung, R.J.
Ovarian Cancer Association Consortium (OCAC),
PRACTICAL Consortium,
Hereditary Breast and Ovarian Cancer Research Group Netherlands (HEBON),
Colorectal Transdisciplinary (CORECT) Study,
African American Breast Cancer Consortium (AABC) and African Ancestry Prostate Cancer Consortium (AAPC),
(2016). Cross-Cancer Genome-Wide Analysis of Lung, Ovary, Breast, Prostate, and Colorectal Cancer Reveals Novel Pleiotropic Associations. Cancer res,
Vol.76
(17),
pp. 5103-5114.
show abstract
full text
Identifying genetic variants with pleiotropic associations can uncover common pathways influencing multiple cancers. We took a two-stage approach to conduct genome-wide association studies for lung, ovary, breast, prostate, and colorectal cancer from the GAME-ON/GECCO Network (61,851 cases, 61,820 controls) to identify pleiotropic loci. Findings were replicated in independent association studies (55,789 cases, 330,490 controls). We identified a novel pleiotropic association at 1q22 involving breast and lung squamous cell carcinoma, with eQTL analysis showing an association with ADAM15/THBS3 gene expression in lung. We also identified a known breast cancer locus CASP8/ALS2CR12 associated with prostate cancer, a known cancer locus at CDKN2B-AS1 with different variants associated with lung adenocarcinoma and prostate cancer, and confirmed the associations of a breast BRCA2 locus with lung and serous ovarian cancer. This is the largest study to date examining pleiotropy across multiple cancer-associated loci, identifying common mechanisms of cancer development and progression. Cancer Res; 76(17); 5103-14. ©2016 AACR..
Foulkes, W.D.
Knoppers, B.M.
Turnbull, C.
(2016). Population genetic testing for cancer susceptibility: founder mutations to genomes (vol 13, pg 41, 2016). Nature reviews clinical oncology,
Vol.13
(1),
p. 1.
Frampton, M.J.
Law, P.
Litchfield, K.
Morris, E.J.
Kerr, D.
Turnbull, C.
Tomlinson, I.P.
Houlston, R.S.
(2016). Implications of polygenic risk for personalised colorectal cancer screening. Ann oncol,
Vol.27
(3),
pp. 429-434.
show abstract
BACKGROUND: We modelled the utility of applying a personalised screening approach for colorectal cancer (CRC) when compared with standard age-based screening. In this personalised screening approach, eligibility is determined by absolute risk which is calculated from age and polygenic risk score (PRS), where the PRS is relative risk attributable to common genetic variation. In contrast, eligibility in age-based screening is determined only by age. DESIGN: We calculated absolute risks of CRC from UK population age structure, incidence and mortality rate data, and a PRS distribution which we derived for the 37 known CRC susceptibility variants. We compared the number of CRC cases potentially detectable by personalised and age-based screening. Using Genome-Wide Complex Trait Analysis to calculate the heritability attributable to common variation, we repeated the analysis assuming all common CRC risk variants were known. RESULTS: Based on the known CRC variants, individuals with a PRS in the top 1% have a 2.9-fold increased CRC risk over the population median. Compared with age-based screening (aged 60: 10-year absolute risk 1.96% in men, 1.19% in women, as per the UK NHS National Bowel Screening Programme), personalised screening of individuals aged 55-69 at the same risk would lead to 16% fewer men and 17% fewer women being eligible for screening with 10% and 8%, respectively, fewer screen-detected cases. If all susceptibility variants were known, individuals with a PRS in the top 1% would have an estimated 7.7-fold increased risk. Personalised screening would then result in 26% fewer men and women being eligible for screening with 7% and 5% fewer screen-detected cases. CONCLUSION: Personalised screening using PRS has the potential to optimise population screening for CRC and to define those likely to maximally benefit from chemoprevention. There are however significant technical and operational details to be addressed before any such programme is introduced..
Bodea, C.A.
Neale, B.M.
Ripke, S.
International IBD Genetics Consortium,
Daly, M.J.
Devlin, B.
Roeder, K.
(2016). A Method to Exploit the Structure of Genetic Ancestry Space to Enhance Case-Control Studies. Am j hum genet,
Vol.98
(5),
pp. 857-868.
show abstract
One goal of human genetics is to understand the genetic basis of disease, a challenge for diseases of complex inheritance because risk alleles are few relative to the vast set of benign variants. Risk variants are often sought by association studies in which allele frequencies in case subjects are contrasted with those from population-based samples used as control subjects. In an ideal world we would know population-level allele frequencies, releasing researchers to focus on case subjects. We argue this ideal is possible, at least theoretically, and we outline a path to achieving it in reality. If such a resource were to exist, it would yield ample savings and would facilitate the effective use of data repositories by removing administrative and technical barriers. We call this concept the Universal Control Repository Network (UNICORN), a means to perform association analyses without necessitating direct access to individual-level control data. Our approach to UNICORN uses existing genetic resources and various statistical tools to analyze these data, including hierarchical clustering with spectral analysis of ancestry; and empirical Bayesian analysis along with Gaussian spatial processes to estimate ancestry-specific allele frequencies. We demonstrate our approach using tens of thousands of control subjects from studies of Crohn disease, showing how it controls false positives, provides power similar to that achieved when all control data are directly accessible, and enhances power when control data are limiting or even imperfectly matched ancestrally. These results highlight how UNICORN can enable reliable, powerful, and convenient genetic association analyses without access to the individual-level data..
Gao, C.
Patel, C.J.
Michailidou, K.
Peters, U.
Gong, J.
Schildkraut, J.
Schumacher, F.R.
Zheng, W.
Boffetta, P.
Stucker, I.
Willett, W.
Gruber, S.
Easton, D.F.
Hunter, D.J.
Sellers, T.A.
Haiman, C.
Henderson, B.E.
Hung, R.J.
Amos, C.
Pierce, B.L.
Lindström, S.
Kraft, P.
the Colorectal Transdisciplinary Study (CORECT); Discovery, Biology and Risk of Inherited Variants in Breast Cancer (DRIVE); Elucidating Loci Involved in Prostate Cancer Susceptibility (ELLIPSE); Follow-up of Ovarian Cancer Genetic Association and Interaction Studies (FOCI); and Transdisciplinary Research in Cancer of the Lung (TRICL),
(2016). Mendelian randomization study of adiposity-related traits and risk of breast, ovarian, prostate, lung and colorectal cancer. Int j epidemiol,
Vol.45
(3),
pp. 896-908.
show abstract
BACKGROUND: Adiposity traits have been associated with risk of many cancers in observational studies, but whether these associations are causal is unclear. Mendelian randomization (MR) uses genetic predictors of risk factors as instrumental variables to eliminate reverse causation and reduce confounding bias. We performed MR analyses to assess the possible causal relationship of birthweight, childhood and adult body mass index (BMI), and waist-hip ratio (WHR) on the risks of breast, ovarian, prostate, colorectal and lung cancers. METHODS: We tested the association between genetic risk scores and each trait using summary statistics from published genome-wide association studies (GWAS) and from 51 537 cancer cases and 61 600 controls in the Genetic Associations and Mechanisms in Oncology (GAME-ON) Consortium. RESULTS: We found an inverse association between the genetic score for childhood BMI and risk of breast cancer [odds ratio (OR) = 0.71 per standard deviation (s.d.) increase in childhood BMI; 95% confidence interval (CI): 0.60, 0.80; P = 6.5 × 10(-5)). We also found the genetic score for adult BMI to be inversely associated with breast cancer risk (OR = 0.66 per s.d. increase in BMI; 95% CI: 0.57, 0.77; P = 2.5 × 10(-7)), and positively associated with ovarian cancer (OR = 1.35; 95% CI: 1.05, 1.72; P = 0.017), lung cancer (OR = 1.27; 95% CI: 1.09, 1.49; P = 2.9 × 10(-3)) and colorectal cancer (OR = 1.39; 95% CI: 1.06, 1.82, P = 0.016). The inverse association between genetically predicted adult BMI and breast cancer risk remained even after adjusting for directional pleiotropy via MR-Egger regression. CONCLUSIONS: Findings from this study provide additional understandings of the complex relationship between adiposity and cancer risks. Our results for breast and lung cancer are particularly interesting, given previous reports of effect heterogeneity by menopausal status and smoking status..
Litchfield, K.
Levy, M.
Huddart, R.A.
Shipley, J.
Turnbull, C.
(2016). The genomic landscape of testicular germ cell tumours: from susceptibility to treatment. Nat rev urol,
Vol.13
(7),
pp. 409-419.
show abstract
full text
The genomic landscape of testicular germ cell tumour (TGCT) can be summarized using four overarching hypotheses. Firstly, TGCT risk is dominated by inherited genetic factors, which determine nearly half of all disease risk and are highly polygenic in nature. Secondly KIT-KITLG signalling is currently the major pathway that is implicated in TGCT formation, both as a predisposition risk factor and a somatic driver event. Results from genome-wide association studies have also consistently suggested that other closely related pathways involved in male germ cell development and sex determination are associated with TGCT risk. Thirdly, the method of disease formation is unique, with tumours universally stemming from a noninvasive precursor lesion, probably of fetal origin, which lies dormant through childhood into adolescence and then eventually begins malignant growth in early adulthood. Formation of a 12p isochromosome, a hallmark of TGCT observed in nearly all tumours, is likely to be a key triggering event for malignant transformation. Finally, TGCT have been shown to have a distinctive somatic mutational profile, with a low rate of point mutations contrasted with frequent large-scale chromosomal gains. These four hypotheses by no means constitute a complete model that explains TGCT tumorigenesis, but advances in genomic technologies have enabled considerable progress in describing and understanding the disease. Further advancing our understanding of the genomic basis of TGCT offers a clear opportunity for clinical benefit in terms of preventing invasive cancer arising in young men, decreasing the burden of chemotherapy-related survivorship issues and reducing mortality in the minority of patients who have treatment-refractory disease..
Khankari, N.K.
Shu, X.-.
Wen, W.
Kraft, P.
Lindström, S.
Peters, U.
Schildkraut, J.
Schumacher, F.
Bofetta, P.
Risch, A.
Bickeböller, H.
Amos, C.I.
Easton, D.
Eeles, R.A.
Gruber, S.B.
Haiman, C.A.
Hunter, D.J.
Chanock, S.J.
Pierce, B.L.
Zheng, W.
Colorectal Transdisciplinary Study (CORECT),
Discovery, Biology, and Risk of Inherited Variants in Breast Cancer (DRIVE),
Elucidating Loci Involved in Prostate Cancer Susceptibility (ELLIPSE),
Transdisciplinary Research in Cancer of the Lung (TRICL),
(2016). Association between Adult Height and Risk of Colorectal, Lung, and Prostate Cancer: Results from Meta-analyses of Prospective Studies and Mendelian Randomization Analyses. Plos med,
Vol.13
(9),
p. e1002118.
show abstract
full text
BACKGROUND: Observational studies examining associations between adult height and risk of colorectal, prostate, and lung cancers have generated mixed results. We conducted meta-analyses using data from prospective cohort studies and further carried out Mendelian randomization analyses, using height-associated genetic variants identified in a genome-wide association study (GWAS), to evaluate the association of adult height with these cancers. METHODS AND FINDINGS: A systematic review of prospective studies was conducted using the PubMed, Embase, and Web of Science databases. Using meta-analyses, results obtained from 62 studies were summarized for the association of a 10-cm increase in height with cancer risk. Mendelian randomization analyses were conducted using summary statistics obtained for 423 genetic variants identified from a recent GWAS of adult height and from a cancer genetics consortium study of multiple cancers that included 47,800 cases and 81,353 controls. For a 10-cm increase in height, the summary relative risks derived from the meta-analyses of prospective studies were 1.12 (95% CI 1.10, 1.15), 1.07 (95% CI 1.05, 1.10), and 1.06 (95% CI 1.02, 1.11) for colorectal, prostate, and lung cancers, respectively. Mendelian randomization analyses showed increased risks of colorectal (odds ratio [OR] = 1.58, 95% CI 1.14, 2.18) and lung cancer (OR = 1.10, 95% CI 1.00, 1.22) associated with each 10-cm increase in genetically predicted height. No association was observed for prostate cancer (OR = 1.03, 95% CI 0.92, 1.15). Our meta-analysis was limited to published studies. The sample size for the Mendelian randomization analysis of colorectal cancer was relatively small, thus affecting the precision of the point estimate. CONCLUSIONS: Our study provides evidence for a potential causal association of adult height with the risk of colorectal and lung cancers and suggests that certain genetic factors and biological pathways affecting adult height may also affect the risk of these cancers..
Litchfield, K.
Levy, M.
Dudakia, D.
Proszek, P.
Shipley, C.
Basten, S.
Rapley, E.
Bishop, D.T.
Reid, A.
Huddart, R.
Broderick, P.
Castro, D.G.
O'Connor, S.
Giles, R.H.
Houlston, R.S.
Turnbull, C.
(2016). Rare disruptive mutations in ciliary function genes contribute to testicular cancer susceptibility. Nat commun,
Vol.7,
p. 13840.
show abstract
full text
Testicular germ cell tumour (TGCT) is the most common cancer in young men. Here we sought to identify risk factors for TGCT by performing whole-exome sequencing on 328 TGCT cases from 153 families, 634 sporadic TGCT cases and 1,644 controls. We search for genes that are recurrently affected by rare variants (minor allele frequency <0.01) with potentially damaging effects and evidence of segregation in families. A total of 8.7% of TGCT families carry rare disruptive mutations in the cilia-microtubule genes (CMG) as compared with 0.5% of controls (P=2.1 × 10-8). The most significantly mutated CMG is DNAAF1 with biallelic inactivation and loss of DNAAF1 expression shown in tumours from carriers. DNAAF1 mutation as a cause of TGCT is supported by a dnaaf1hu255h(+/-) zebrafish model, which has a 94% risk of TGCT. Our data implicate cilia-microtubule inactivation as a cause of TGCT and provide evidence for CMGs as cancer susceptibility genes..
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.-.
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.-.
Claes, K.B.
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.-.
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.
Hart, S.
Hartikainen, J.M.
Hartman, M.
Hassan, N.
Healey, S.
Hogervorst, F.B.
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.-.
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.-.
Shrubsole, M.J.
Shu, X.-.
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.
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.-.
Pharoah, P.D.
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. Breast cancer res,
Vol.18
(1),
p. 64.
show abstract
full text
BACKGROUND: Multiple recent genome-wide association studies (GWAS) have identified a single nucleotide polymorphism (SNP), rs10771399, at 12p11 that is associated with breast cancer risk. METHOD: 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. RESULTS: 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. CONCLUSION: 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..
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.
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.-.
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.
Hart, S.
Healey, S.
Heikkinen, T.
Henderson, B.E.
Herzog, J.
Hogervorst, F.B.
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.-.
Kibriya, M.G.
Knight, J.A.
Ko, Y.-.
Konstantopoulou, I.
Kosma, V.-.
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.
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.-.
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.-.
Terry, M.B.
Thomassen, M.
Thompson, D.
Tihomirova, L.
Toland, A.E.
Tollenaar, R.A.
Tomlinson, I.
Truong, T.
Tsimiklis, H.
Teulé, A.
Tumino, R.
Tung, N.
Turnbull, C.
Ursin, G.
van Deurzen, C.H.
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.
Monteiro, A.A.
García-Closas, M.
Easton, D.F.
Antoniou, A.C.
(2016). Identification of four novel susceptibility loci for oestrogen receptor negative breast cancer. Nat commun,
Vol.7,
p. 11375.
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..
Taylor-Weiner, A.
Zack, T.
O'Donnell, E.
Guerriero, J.L.
Bernard, B.
Reddy, A.
Han, G.C.
AlDubayan, S.
Amin-Mansour, A.
Schumacher, S.E.
Litchfield, K.
Turnbull, C.
Gabriel, S.
Beroukhim, R.
Getz, G.
Carter, S.L.
Hirsch, M.S.
Letai, A.
Sweeney, C.
Van Allen, E.M.
(2016). Genomic evolution and chemoresistance in germ-cell tumours. Nature,
Vol.540
(7631),
pp. 114-118.
show abstract
full text
Germ-cell tumours (GCTs) are derived from germ cells and occur most frequently in the testes. GCTs are histologically heterogeneous and distinctly curable with chemotherapy. Gains of chromosome arm 12p and aneuploidy are nearly universal in GCTs, but specific somatic genomic features driving tumour initiation, chemosensitivity and progression are incompletely characterized. Here, using clinical whole-exome and transcriptome sequencing of precursor, primary (testicular and mediastinal) and chemoresistant metastatic human GCTs, we show that the primary somatic feature of GCTs is highly recurrent chromosome arm level amplifications and reciprocal deletions (reciprocal loss of heterozygosity), variations that are significantly enriched in GCTs compared to 19 other cancer types. These tumours also acquire KRAS mutations during the development from precursor to primary disease, and primary testicular GCTs (TGCTs) are uniformly wild type for TP53. In addition, by functional measurement of apoptotic signalling (BH3 profiling) of fresh tumour and adjacent tissue, we find that primary TGCTs have high mitochondrial priming that facilitates chemotherapy-induced apoptosis. Finally, by phylogenetic analysis of serial TGCTs that emerge with chemotherapy resistance, we show how TGCTs gain additional reciprocal loss of heterozygosity and that this is associated with loss of pluripotency markers (NANOG and POU5F1) in chemoresistant teratomas or transformed carcinomas. Our results demonstrate the distinct genomic features underlying the origins of this disease and associated with the chemosensitivity phenotype, as well as the rare progression to chemoresistance. These results identify the convergence of cancer genomics, mitochondrial priming and GCT evolution, and may provide insights into chemosensitivity and resistance in other cancers..
Litchfield, K.
Shipley, J.
Turnbull, C.
(2015). Common variants identified in genome-wide association studies of testicular germ cell tumour: an update, biological insights and clinical application. Andrology,
Vol.3
(1),
pp. 34-46.
show abstract
Testicular germ cell tumour (TGCT) is the most common cause of cancer in young men (aged 15-45 years) in many populations. Multiple genome-wide association studies (GWAS) of TGCT have now been conducted, yielding over 25 disease-associated single-nucleotide polymorphism (SNP)s at 19 independent loci. The genes at these loci have provided rich biological and genetic insight into possible mechanisms underlying testicular germ cell oncogenesis. In this review, we summarize these mechanisms which can be grouped into five distinct categories: KIT/KITLG signalling, other pathways of male germ cell development/differentiation, telomerase function, microtubule assembly and DNA damage repair. The TGCT risk markers identified through GWAS include individual SNPs carrying per allele odds ratios (OR) in excess of 2.5. These ORs are among the highest reported in GWAS of any cancer type, hence suggesting a potential clinical utility in risk determination. Here, we present analysis of such an approach, using polygenic risk scores to calculate the combined effect of all risk loci on overall TGCT risk and discuss how a potential screening strategy may fit within a broader clinical context..
Lin, W.-.
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.-.
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.-.
Hartikainen, J.M.
kConFab Investigators,
Australian Ovarian Cancer Study Group,
Wu, A.H.
Tseng, 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.
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.
Jager, A.
Li, J.
Liu, J.
Humphreys, K.
Shu, X.-.
Lu, W.
Gao, Y.-.
Cai, H.
Cross, S.S.
Reed, M.W.
Blot, W.
Signorello, L.B.
Cai, Q.
Pharoah, P.D.
Perkins, B.
Shah, M.
Blows, F.M.
Kang, D.
Yoo, K.-.
Noh, D.-.
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.-.
Hsiung, C.-.
Wu, P.-.
Ding, S.-.
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.
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. Hum mol genet,
Vol.24
(1),
pp. 285-298.
show abstract
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..
Vilhjálmsson, B.J.
Yang, J.
Finucane, H.K.
Gusev, A.
Lindström, S.
Ripke, S.
Genovese, G.
Loh, P.-.
Bhatia, G.
Do, R.
Hayeck, T.
Won, H.-.
Schizophrenia Working Group of the Psychiatric Genomics Consortium, Discovery, Biology, and Risk of Inherited Variants in Breast Cancer (DRIVE) study,
Kathiresan, S.
Pato, M.
Pato, C.
Tamimi, R.
Stahl, E.
Zaitlen, N.
Pasaniuc, B.
Belbin, G.
Kenny, E.E.
Schierup, M.H.
De Jager, P.
Patsopoulos, N.A.
McCarroll, S.
Daly, M.
Purcell, S.
Chasman, D.
Neale, B.
Goddard, M.
Visscher, P.M.
Kraft, P.
Patterson, N.
Price, A.L.
(2015). Modeling Linkage Disequilibrium Increases Accuracy of Polygenic Risk Scores. Am j hum genet,
Vol.97
(4),
pp. 576-592.
show abstract
Polygenic risk scores have shown great promise in predicting complex disease risk and will become more accurate as training sample sizes increase. The standard approach for calculating risk scores involves linkage disequilibrium (LD)-based marker pruning and applying a p value threshold to association statistics, but this discards information and can reduce predictive accuracy. We introduce LDpred, a method that infers the posterior mean effect size of each marker by using a prior on effect sizes and LD information from an external reference panel. Theory and simulations show that LDpred outperforms the approach of pruning followed by thresholding, particularly at large sample sizes. Accordingly, predicted R(2) increased from 20.1% to 25.3% in a large schizophrenia dataset and from 9.8% to 12.0% in a large multiple sclerosis dataset. A similar relative improvement in accuracy was observed for three additional large disease datasets and for non-European schizophrenia samples. The advantage of LDpred over existing methods will grow as sample sizes increase..
Michailidou, K.
Beesley, J.
Lindstrom, S.
Canisius, S.
Dennis, J.
Lush, M.J.
Maranian, M.J.
Bolla, M.K.
Wang, Q.
Shah, M.
Perkins, B.J.
Czene, K.
Eriksson, M.
Darabi, H.
Brand, J.S.
Bojesen, S.E.
Nordestgaard, B.G.
Flyger, H.
Nielsen, S.F.
Rahman, N.
Turnbull, C.
BOCS,
Fletcher, O.
Peto, J.
Gibson, L.
dos-Santos-Silva, I.
Chang-Claude, J.
Flesch-Janys, D.
Rudolph, A.
Eilber, U.
Behrens, S.
Nevanlinna, H.
Muranen, T.A.
Aittomäki, K.
Blomqvist, C.
Khan, S.
Aaltonen, K.
Ahsan, H.
Kibriya, M.G.
Whittemore, A.S.
John, E.M.
Malone, K.E.
Gammon, M.D.
Santella, R.M.
Ursin, G.
Makalic, E.
Schmidt, D.F.
Casey, G.
Hunter, D.J.
Gapstur, S.M.
Gaudet, M.M.
Diver, W.R.
Haiman, C.A.
Schumacher, F.
Henderson, B.E.
Le Marchand, L.
Berg, C.D.
Chanock, S.J.
Figueroa, J.
Hoover, R.N.
Lambrechts, D.
Neven, P.
Wildiers, H.
van Limbergen, E.
Schmidt, M.K.
Broeks, A.
Verhoef, S.
Cornelissen, S.
Couch, F.J.
Olson, J.E.
Hallberg, E.
Vachon, C.
Waisfisz, Q.
Meijers-Heijboer, H.
Adank, M.A.
van der Luijt, R.B.
Li, J.
Liu, J.
Humphreys, K.
Kang, D.
Choi, J.-.
Park, S.K.
Yoo, K.-.
Matsuo, K.
Ito, H.
Iwata, H.
Tajima, K.
Guénel, P.
Truong, T.
Mulot, C.
Sanchez, M.
Burwinkel, B.
Marme, F.
Surowy, H.
Sohn, C.
Wu, A.H.
Tseng, C.-.
Van Den Berg, D.
Stram, D.O.
González-Neira, A.
Benitez, J.
Zamora, M.P.
Perez, J.I.
Shu, X.-.
Lu, W.
Gao, Y.-.
Cai, H.
Cox, A.
Cross, S.S.
Reed, M.W.
Andrulis, I.L.
Knight, J.A.
Glendon, G.
Mulligan, A.M.
Sawyer, E.J.
Tomlinson, I.
Kerin, M.J.
Miller, N.
kConFab Investigators,
AOCS Group,
Lindblom, A.
Margolin, S.
Teo, S.H.
Yip, C.H.
Taib, N.A.
Tan, G.-.
Hooning, M.J.
Hollestelle, A.
Martens, J.W.
Collée, J.M.
Blot, W.
Signorello, L.B.
Cai, Q.
Hopper, J.L.
Southey, M.C.
Tsimiklis, H.
Apicella, C.
Shen, C.-.
Hsiung, C.-.
Wu, P.-.
Hou, M.-.
Kristensen, V.N.
Nord, S.
Alnaes, G.I.
NBCS,
Giles, G.G.
Milne, R.L.
McLean, C.
Canzian, F.
Trichopoulos, D.
Peeters, P.
Lund, E.
Sund, M.
Khaw, K.-.
Gunter, M.J.
Palli, D.
Mortensen, L.M.
Dossus, L.
Huerta, J.-.
Meindl, A.
Schmutzler, R.K.
Sutter, C.
Yang, R.
Muir, K.
Lophatananon, A.
Stewart-Brown, S.
Siriwanarangsan, P.
Hartman, M.
Miao, H.
Chia, K.S.
Chan, C.W.
Fasching, P.A.
Hein, A.
Beckmann, M.W.
Haeberle, L.
Brenner, H.
Dieffenbach, A.K.
Arndt, V.
Stegmaier, C.
Ashworth, A.
Orr, N.
Schoemaker, M.J.
Swerdlow, A.J.
Brinton, L.
Garcia-Closas, M.
Zheng, W.
Halverson, S.L.
Shrubsole, M.
Long, J.
Goldberg, M.S.
Labrèche, F.
Dumont, M.
Winqvist, R.
Pylkäs, K.
Jukkola-Vuorinen, A.
Grip, M.
Brauch, H.
Hamann, U.
Brüning, T.
GENICA Network,
Radice, P.
Peterlongo, P.
Manoukian, S.
Bernard, L.
Bogdanova, N.V.
Dörk, T.
Mannermaa, A.
Kataja, V.
Kosma, V.-.
Hartikainen, J.M.
Devilee, P.
Tollenaar, R.A.
Seynaeve, C.
Van Asperen, C.J.
Jakubowska, A.
Lubinski, J.
Jaworska, K.
Huzarski, T.
Sangrajrang, S.
Gaborieau, V.
Brennan, P.
McKay, J.
Slager, S.
Toland, A.E.
Ambrosone, C.B.
Yannoukakos, D.
Kabisch, M.
Torres, D.
Neuhausen, S.L.
Anton-Culver, H.
Luccarini, C.
Baynes, C.
Ahmed, S.
Healey, C.S.
Tessier, D.C.
Vincent, D.
Bacot, F.
Pita, G.
Alonso, M.R.
Álvarez, N.
Herrero, D.
Simard, J.
Pharoah, P.P.
Kraft, P.
Dunning, A.M.
Chenevix-Trench, G.
Hall, P.
Easton, D.F.
(2015). Genome-wide association analysis of more than 120,000 individuals identifies 15 new susceptibility loci for breast cancer. Nat genet,
Vol.47
(4),
pp. 373-380.
show abstract
full text
Genome-wide association studies (GWAS) and large-scale replication studies have identified common variants in 79 loci associated with breast cancer, explaining ∼14% of the familial risk of the disease. To identify new susceptibility loci, we performed a meta-analysis of 11 GWAS, comprising 15,748 breast cancer cases and 18,084 controls together with 46,785 cases and 42,892 controls from 41 studies genotyped on a 211,155-marker custom array (iCOGS). Analyses were restricted to women of European ancestry. We generated genotypes for more than 11 million SNPs by imputation using the 1000 Genomes Project reference panel, and we identified 15 new loci associated with breast cancer at P < 5 × 10(-8). Combining association analysis with ChIP-seq chromatin binding data in mammary cell lines and ChIA-PET chromatin interaction data from ENCODE, we identified likely target genes in two regions: SETBP1 at 18q12.3 and RNF115 and PDZK1 at 1q21.1. One association appears to be driven by an amino acid substitution encoded in EXO1..
Mavaddat, N.
Pharoah, P.D.
Michailidou, K.
Tyrer, J.
Brook, M.N.
Bolla, M.K.
Wang, Q.
Dennis, J.
Dunning, A.M.
Shah, M.
Luben, R.
Brown, J.
Bojesen, S.E.
Nordestgaard, B.G.
Nielsen, S.F.
Flyger, H.
Czene, K.
Darabi, H.
Eriksson, M.
Peto, J.
Dos-Santos-Silva, I.
Dudbridge, F.
Johnson, N.
Schmidt, M.K.
Broeks, A.
Verhoef, S.
Rutgers, E.J.
Swerdlow, A.
Ashworth, A.
Orr, N.
Schoemaker, M.J.
Figueroa, J.
Chanock, S.J.
Brinton, L.
Lissowska, J.
Couch, F.J.
Olson, J.E.
Vachon, C.
Pankratz, V.S.
Lambrechts, D.
Wildiers, H.
Van Ongeval, C.
van Limbergen, E.
Kristensen, V.
Grenaker Alnæs, G.
Nord, S.
Borresen-Dale, A.-.
Nevanlinna, H.
Muranen, T.A.
Aittomäki, K.
Blomqvist, C.
Chang-Claude, J.
Rudolph, A.
Seibold, P.
Flesch-Janys, D.
Fasching, P.A.
Haeberle, L.
Ekici, A.B.
Beckmann, M.W.
Burwinkel, B.
Marme, F.
Schneeweiss, A.
Sohn, C.
Trentham-Dietz, A.
Newcomb, P.
Titus, L.
Egan, K.M.
Hunter, D.J.
Lindstrom, S.
Tamimi, R.M.
Kraft, P.
Rahman, N.
Turnbull, C.
Renwick, A.
Seal, S.
Li, J.
Liu, J.
Humphreys, K.
Benitez, J.
Pilar Zamora, M.
Arias Perez, J.I.
Menéndez, P.
Jakubowska, A.
Lubinski, J.
Jaworska-Bieniek, K.
Durda, K.
Bogdanova, N.V.
Antonenkova, N.N.
Dörk, T.
Anton-Culver, H.
Neuhausen, S.L.
Ziogas, A.
Bernstein, L.
Devilee, P.
Tollenaar, R.A.
Seynaeve, C.
van Asperen, C.J.
Cox, A.
Cross, S.S.
Reed, M.W.
Khusnutdinova, E.
Bermisheva, M.
Prokofyeva, D.
Takhirova, Z.
Meindl, A.
Schmutzler, R.K.
Sutter, C.
Yang, R.
Schürmann, P.
Bremer, M.
Christiansen, H.
Park-Simon, T.-.
Hillemanns, P.
Guénel, P.
Truong, T.
Menegaux, F.
Sanchez, M.
Radice, P.
Peterlongo, P.
Manoukian, S.
Pensotti, V.
Hopper, J.L.
Tsimiklis, H.
Apicella, C.
Southey, M.C.
Brauch, H.
Brüning, T.
Ko, Y.-.
Sigurdson, A.J.
Doody, M.M.
Hamann, U.
Torres, D.
Ulmer, H.-.
Försti, A.
Sawyer, E.J.
Tomlinson, I.
Kerin, M.J.
Miller, N.
Andrulis, I.L.
Knight, J.A.
Glendon, G.
Marie Mulligan, A.
Chenevix-Trench, G.
Balleine, R.
Giles, G.G.
Milne, R.L.
McLean, C.
Lindblom, A.
Margolin, S.
Haiman, C.A.
Henderson, B.E.
Schumacher, F.
Le Marchand, L.
Eilber, U.
Wang-Gohrke, S.
Hooning, M.J.
Hollestelle, A.
van den Ouweland, A.M.
Koppert, L.B.
Carpenter, J.
Clarke, C.
Scott, R.
Mannermaa, A.
Kataja, V.
Kosma, V.-.
Hartikainen, J.M.
Brenner, H.
Arndt, V.
Stegmaier, C.
Karina Dieffenbach, A.
Winqvist, R.
Pylkäs, K.
Jukkola-Vuorinen, A.
Grip, M.
Offit, K.
Vijai, J.
Robson, M.
Rau-Murthy, R.
Dwek, M.
Swann, R.
Annie Perkins, K.
Goldberg, M.S.
Labrèche, F.
Dumont, M.
Eccles, D.M.
Tapper, W.J.
Rafiq, S.
John, E.M.
Whittemore, A.S.
Slager, S.
Yannoukakos, D.
Toland, A.E.
Yao, S.
Zheng, W.
Halverson, S.L.
González-Neira, A.
Pita, G.
Rosario Alonso, M.
Álvarez, N.
Herrero, D.
Tessier, D.C.
Vincent, D.
Bacot, F.
Luccarini, C.
Baynes, C.
Ahmed, S.
Maranian, M.
Healey, C.S.
Simard, J.
Hall, P.
Easton, D.F.
Garcia-Closas, M.
(2015). Prediction of breast cancer risk based on profiling with common genetic variants. J natl cancer inst,
Vol.107
(5).
show abstract
BACKGROUND: Data for multiple common susceptibility alleles for breast cancer may be combined to identify women at different levels of breast cancer risk. Such stratification could guide preventive and screening strategies. However, empirical evidence for genetic risk stratification is lacking. METHODS: We investigated the value of using 77 breast cancer-associated single nucleotide polymorphisms (SNPs) for risk stratification, in a study of 33 673 breast cancer cases and 33 381 control women of European origin. We tested all possible pair-wise multiplicative interactions and constructed a 77-SNP polygenic risk score (PRS) for breast cancer overall and by estrogen receptor (ER) status. Absolute risks of breast cancer by PRS were derived from relative risk estimates and UK incidence and mortality rates. RESULTS: There was no strong evidence for departure from a multiplicative model for any SNP pair. Women in the highest 1% of the PRS had a three-fold increased risk of developing breast cancer compared with women in the middle quintile (odds ratio [OR] = 3.36, 95% confidence interval [CI] = 2.95 to 3.83). The ORs for ER-positive and ER-negative disease were 3.73 (95% CI = 3.24 to 4.30) and 2.80 (95% CI = 2.26 to 3.46), respectively. Lifetime risk of breast cancer for women in the lowest and highest quintiles of the PRS were 5.2% and 16.6% for a woman without family history, and 8.6% and 24.4% for a woman with a first-degree family history of breast cancer. CONCLUSIONS: The PRS stratifies breast cancer risk in women both with and without a family history of breast cancer. The observed level of risk discrimination could inform targeted screening and prevention strategies. Further discrimination may be achievable through combining the PRS with lifestyle/environmental factors, although these were not considered in this report..
Litchfield, K.
Thomsen, H.
Mitchell, J.S.
Sundquist, J.
Houlston, R.S.
Hemminki, K.
Turnbull, C.
(2015). Quantifying the heritability of testicular germ cell tumour using both population-based and genomic approaches. Sci rep,
Vol.5,
p. 13889.
show abstract
A sizable fraction of testicular germ cell tumour (TGCT) risk is expected to be explained by heritable factors. Recent genome-wide association studies (GWAS) have successfully identified a number of common SNPs associated with TGCT. It is however, unclear how much common variation there is left to be accounted for by other, yet to be identified, common SNPs and what contribution common genetic variation makes to the heritable risk of TGCT. We approached this question using two complimentary analytical techniques. We undertook a population-based analysis of the Swedish family-cancer database, through which we estimated that the heritability of TGCT at 48.9% (CI:47.2%-52.3%). We also applied Genome-Wide Complex Trait Analysis to 922 cases and 4,842 controls to estimate the heritability of TGCT. The heritability explained by known common risk SNPs identified by GWAS was 9.1%, whereas the heritability explained by all common SNPs was 37.4% (CI:27.6%-47.2%). These complementary findings indicate that the known TGCT SNPs only explain a small proportion of the heritability and many additional common SNPs remain to be identified. The data also suggests that a fraction of the heritability of TGCT is likely to be explained by other classes of genetic variation, such as rare disease-causing alleles..
Kristiansen, W.
Karlsson, R.
Rounge, T.B.
Whitington, T.
Andreassen, B.K.
Magnusson, P.K.
Fosså, S.D.
Adami, H.-.
Turnbull, C.
Haugen, T.B.
Grotmol, T.
Wiklund, F.
(2015). Two new loci and gene sets related to sex determination and cancer progression are associated with susceptibility to testicular germ cell tumor. Hum mol genet,
Vol.24
(14),
pp. 4138-4146.
show abstract
Genome-wide association (GWA) studies have reported 19 distinct susceptibility loci for testicular germ cell tumor (TGCT). A GWA study for TGCT was performed by genotyping 610 240 single-nucleotide polymorphisms (SNPs) in 1326 cases and 6687 controls from Sweden and Norway. No novel genome-wide significant associations were observed in this discovery stage. We put forward 27 SNPs from 15 novel regions and 12 SNPs previously reported, for replication in 710 case-parent triads and 289 cases and 290 controls. Predefined biological pathways and processes, in addition to a custom-built sex-determination gene set, were subject to enrichment analyses using Meta-Analysis Gene Set Enrichment of Variant Associations (M) and Improved Gene Set Enrichment Analysis for Genome-wide Association Study (I). In the combined meta-analysis, we observed genome-wide significant association for rs7501939 on chromosome 17q12 (OR = 0.78, 95% CI = 0.72-0.84, P = 1.1 × 10(-9)) and rs2195987 on chromosome 19p12 (OR = 0.76, 95% CI: 0.69-0.84, P = 3.2 × 10(-8)). The marker rs7501939 on chromosome 17q12 is located in an intron of the HNF1B gene, encoding a member of the homeodomain-containing superfamily of transcription factors. The sex-determination gene set (false discovery rate, FDRM < 0.001, FDRI < 0.001) and pathways related to NF-κB, glycerophospholipid and ether lipid metabolism, as well as cancer and apoptosis, was associated with TGCT (FDR < 0.1). In addition to revealing two new TGCT susceptibility loci, our results continue to support the notion that genes governing normal germ cell development in utero are implicated in the development of TGCT..
Litchfield, K.
Sultana, R.
Renwick, A.
Dudakia, D.
Seal, S.
Ramsay, E.
Powell, S.
Elliott, A.
Warren-Perry, M.
Eeles, R.
Peto, J.
Kote-Jarai, Z.
Muir, K.
Nsengimana, J.
UKTCC,
Stratton, M.R.
Easton, D.F.
Bishop, D.T.
Huddart, R.A.
Rahman, N.
Turnbull, C.
UKTCC,
(2015). Multi-stage genome-wide association study identifies new susceptibility locus for testicular germ cell tumour on chromosome 3q25. Hum mol genet,
Vol.24
(4),
pp. 1169-1176.
show abstract
Recent genome-wide association studies (GWAS) and subsequent meta-analyses have identified over 25 SNPs at 18 loci, together accounting for >15% of the genetic susceptibility to testicular germ cell tumour (TGCT). To identify further common SNPs associated with TGCT, here we report a three-stage experiment, involving 4098 cases and 18 972 controls. Stage 1 comprised previously published GWAS analysis of 307 291 SNPs in 986 cases and 4946 controls. In Stage 2, we used previously published customised Illumina iSelect genotyping array (iCOGs) data across 694 SNPs in 1064 cases and 10 082 controls. Here, we report new genotyping of eight SNPs showing some evidence of association in combined analysis of Stage 1 and Stage 2 in an additional 2048 cases of TGCT and 3944 controls (Stage 3). Through fixed-effects meta-analysis across three stages, we identified a novel locus at 3q25.31 (rs1510272) demonstrating association with TGCT [per-allele odds ratio (OR) = 1.16, 95% confidence interval (CI) = 1.06-1.27; P = 1.2 × 10(-9)]..
Litchfield, K.
Mitchell, J.S.
Shipley, J.
Huddart, R.
Rajpert-De Meyts, E.
Skakkebæk, N.E.
Houlston, R.S.
Turnbull, C.
(2015). Polygenic susceptibility to testicular cancer: implications for personalised health care. Br j cancer,
Vol.113
(10),
pp. 1512-1518.
show abstract
BACKGROUND: The increasing incidence of testicular germ cell tumour (TGCT) combined with its strong heritable basis suggests that stratified screening for the early detection of TGCT may be clinically useful. We modelled the efficiency of such a personalised screening approach, based on genetic risk profiling in combination with other diagnostic tools. METHODS: We compared the number of cases potentially detectable in the population under a number of screening models. The polygenic risk scoring (PRS) model was assumed to have a log-normal relative risk distribution across the 19 currently known TGCT susceptibility variants. The diagnostic performance of testicular biopsy and non-invasive semen analysis was also assessed, within a simulated combined screening programme. RESULTS: The area under the curve for the TGCT PRS model was 0.72 with individuals in the top 1% of the PRS having a nine-fold increased TGCT risk compared with the population median. Results from population-screening simulations only achieved a maximal positive predictive value (PPV) of 60%, highlighting broader clinical factors that challenge such strategies, not least the rare nature of TGCT. In terms of future improvements, heritability estimates suggest that a significant number of additional genetic risk factors for TGCT remain to be discovered, identification of which would potentially yield improvement of the PPV to 80-90%. CONCLUSIONS: While personalised screening models may offer enhanced TGCT risk discrimination, presently the case for population-level testing is not compelling. However, future advances, such as more routine generation of whole genome data is likely to alter the landscape. More targeted screening programs may plausibly then offer clinical benefit, particularly given the significant survivorship issues associated with the successful treatment of TGCT..
Litchfield, K.
Summersgill, B.
Yost, S.
Sultana, R.
Labreche, K.
Dudakia, D.
Renwick, A.
Seal, S.
Al-Saadi, R.
Broderick, P.
Turner, N.C.
Houlston, R.S.
Huddart, R.
Shipley, J.
Turnbull, C.
(2015). Whole-exome sequencing reveals the mutational spectrum of testicular germ cell tumours. Nat commun,
Vol.6,
p. 5973.
show abstract
full text
Testicular germ cell tumours (TGCTs) are the most common cancer in young men. Here we perform whole-exome sequencing (WES) of 42 TGCTs to comprehensively study the cancer's mutational profile. The mutation rate is uniformly low in all of the tumours (mean 0.5 mutations per Mb) as compared with common cancers, consistent with the embryological origin of TGCT. In addition to expected copy number gain of chromosome 12p and mutation of KIT, we identify recurrent mutations in the tumour suppressor gene CDC27 (11.9%). Copy number analysis reveals recurring amplification of the spermatocyte development gene FSIP2 (15.3%) and a 0.4 Mb region at Xq28 (15.3%). Two treatment-refractory patients are shown to harbour XRCC2 mutations, a gene strongly implicated in defining cisplatin resistance. Our findings provide further insights into genes involved in the development and progression of TGCT..
Litchfield, K.
Holroyd, A.
Lloyd, A.
Broderick, P.
Nsengimana, J.
Eeles, R.
Easton, D.F.
Dudakia, D.
Bishop, D.T.
Reid, A.
Huddart, R.A.
Grotmol, T.
Wiklund, F.
Shipley, J.
Houlston, R.S.
Turnbull, C.
(2015). Identification of four new susceptibility loci for testicular germ cell tumour. Nat commun,
Vol.6,
p. 8690.
show abstract
Genome-wide association studies (GWAS) have identified multiple risk loci for testicular germ cell tumour (TGCT), revealing a polygenic model of disease susceptibility strongly influenced by common variation. To identify additional single-nucleotide polymorphisms (SNPs) associated with TGCT, we conducted a multistage GWAS with a combined data set of >25,000 individuals (6,059 cases and 19,094 controls). We identified new risk loci for TGCT at 3q23 (rs11705932, TFDP2, P=1.5 × 10(-9)), 11q14.1 (rs7107174, GAB2, P=9.7 × 10(-11)), 16p13.13 (rs4561483, GSPT1, P=1.6 × 10(-8)) and 16q24.2 (rs55637647, ZFPM1, P=3.4 × 10(-9)). We additionally present detailed functional analysis of these loci, identifying a statistically significant relationship between rs4561483 risk genotype and increased GSPT1 expression in TGCT patient samples. These findings provide additional support for a polygenic model of TGCT risk and further insight into the biological basis of disease development..
Slade, I.
Riddell, D.
Turnbull, C.
Hanson, H.
Rahman, N.
MCG programme,
(2015). Development of cancer genetic services in the UK: A national consultation. Genome med,
Vol.7
(1),
p. 18.
show abstract
full text
BACKGROUND: Technological advances in DNA sequencing have made gene testing fast and affordable, but there are challenges to the translation of these improvements for patient benefit. The Mainstreaming Cancer Genetics (MCG) programme is exploiting advances in DNA sequencing to develop the infrastructure, processes and capabilities required for cancer gene testing to become routinely available to all those that can benefit. METHODS: The MCG programme held a consultation day to discuss the development of cancer genetics with senior representation from all 24 UK cancer genetic centres. The current service landscape and capacity for expansion was assessed through structured questionnaires. Workshop discussion addressed the opportunities and challenges to increasing cancer gene testing in the National Health Service (NHS). RESULTS: Services vary with respect to population served and models of service delivery, and with respect to methods and thresholds for determining risk and testing eligibility. Almost all centres want to offer more cancer gene testing (82%) and reported increasing demand for testing from non-genetic clinical colleagues (92%). Reported challenges to increasing testing include the complexity of interpreting the resulting genetic data (79%), the level of funding and complexity of commissioning (67%), the limited capacity of current processes and cross-disciplinary relationships (38%), and workforce education (29%). CONCLUSIONS: Priorities to address include the development and evaluation of models of increasing access to gene testing, the optimal process for interpretation of large-scale genetic data, implementation of appropriate commissioning and funding processes, and achieving national consistency. The UK cancer genetics community have high expertise and strong commitment to maximising scientific advances for improved patient benefit and should be pivotally involved in the implementation of increased cancer gene testing..
Ahsan, H.
Halpern, J.
Kibriya, M.G.
Pierce, B.L.
Tong, L.
Gamazon, E.
McGuire, V.
Felberg, A.
Shi, J.
Jasmine, F.
Roy, S.
Brutus, R.
Argos, M.
Melkonian, S.
Chang-Claude, J.
Andrulis, I.
Hopper, J.L.
John, E.M.
Malone, K.
Ursin, G.
Gammon, M.D.
Thomas, D.C.
Seminara, D.
Casey, G.
Knight, J.A.
Southey, M.C.
Giles, G.G.
Santella, R.M.
Lee, E.
Conti, D.
Duggan, D.
Gallinger, S.
Haile, R.
Jenkins, M.
Lindor, N.M.
Newcomb, P.
Michailidou, K.
Apicella, C.
Park, D.J.
Peto, J.
Fletcher, O.
dos Santos Silva, I.
Lathrop, M.
Hunter, D.J.
Chanock, S.J.
Meindl, A.
Schmutzler, R.K.
Müller-Myhsok, B.
Lochmann, M.
Beckmann, L.
Hein, R.
Makalic, E.
Schmidt, D.F.
Bui, Q.M.
Stone, J.
Flesch-Janys, D.
Dahmen, N.
Nevanlinna, H.
Aittomäki, K.
Blomqvist, C.
Hall, P.
Czene, K.
Irwanto, A.
Liu, J.
Rahman, N.
Turnbull, C.
Familial Breast Cancer Study,
Dunning, A.M.
Pharoah, P.
Waisfisz, Q.
Meijers-Heijboer, H.
Uitterlinden, A.G.
Rivadeneira, F.
Nicolae, D.
Easton, D.F.
Cox, N.J.
Whittemore, A.S.
(2014). A genome-wide association study of early-onset breast cancer identifies PFKM as a novel breast cancer gene and supports a common genetic spectrum for breast cancer at any age. Cancer epidemiol biomarkers prev,
Vol.23
(4),
pp. 658-669.
show abstract
Early-onset breast cancer (EOBC) causes substantial loss of life and productivity, creating a major burden among women worldwide. We analyzed 1,265,548 Hapmap3 single-nucleotide polymorphisms (SNP) among a discovery set of 3,523 EOBC incident cases and 2,702 population control women ages ≤ 51 years. The SNPs with smallest P values were examined in a replication set of 3,470 EOBC cases and 5,475 control women. We also tested EOBC association with 19,684 genes by annotating each gene with putative functional SNPs, and then combining their P values to obtain a gene-based P value. We examined the gene with smallest P value for replication in 1,145 breast cancer cases and 1,142 control women. The combined discovery and replication sets identified 72 new SNPs associated with EOBC (P < 4 × 10(-8)) located in six genomic regions previously reported to contain SNPs associated largely with later-onset breast cancer (LOBC). SNP rs2229882 and 10 other SNPs on chromosome 5q11.2 remained associated (P < 6 × 10(-4)) after adjustment for the strongest published SNPs in the region. Thirty-two of the 82 currently known LOBC SNPs were associated with EOBC (P < 0.05). Low power is likely responsible for the remaining 50 unassociated known LOBC SNPs. The gene-based analysis identified an association between breast cancer and the phosphofructokinase-muscle (PFKM) gene on chromosome 12q13.11 that met the genome-wide gene-based threshold of 2.5 × 10(-6). In conclusion, EOBC and LOBC seem to have similar genetic etiologies; the 5q11.2 region may contain multiple distinct breast cancer loci; and the PFKM gene region is worthy of further investigation. These findings should enhance our understanding of the etiology of breast cancer..
Antoniou, A.C.
Casadei, S.
Heikkinen, T.
Barrowdale, D.
Pylkäs, K.
Roberts, J.
Lee, A.
Subramanian, D.
De Leeneer, K.
Fostira, F.
Tomiak, E.
Neuhausen, S.L.
Teo, Z.L.
Khan, S.
Aittomäki, K.
Moilanen, J.S.
Turnbull, C.
Seal, S.
Mannermaa, A.
Kallioniemi, A.
Lindeman, G.J.
Buys, S.S.
Andrulis, I.L.
Radice, P.
Tondini, C.
Manoukian, S.
Toland, A.E.
Miron, P.
Weitzel, J.N.
Domchek, S.M.
Poppe, B.
Claes, K.B.
Yannoukakos, D.
Concannon, P.
Bernstein, J.L.
James, P.A.
Easton, D.F.
Goldgar, D.E.
Hopper, J.L.
Rahman, N.
Peterlongo, P.
Nevanlinna, H.
King, M.-.
Couch, F.J.
Southey, M.C.
Winqvist, R.
Foulkes, W.D.
Tischkowitz, M.
(2014). Breast-cancer risk in families with mutations in PALB2. N engl j med,
Vol.371
(6),
pp. 497-506.
show abstract
BACKGROUND: Germline loss-of-function mutations in PALB2 are known to confer a predisposition to breast cancer. However, the lifetime risk of breast cancer that is conferred by such mutations remains unknown. METHODS: We analyzed the risk of breast cancer among 362 members of 154 families who had deleterious truncating, splice, or deletion mutations in PALB2. The age-specific breast-cancer risk for mutation carriers was estimated with the use of a modified segregation-analysis approach that allowed for the effects of PALB2 genotype and residual familial aggregation. RESULTS: The risk of breast cancer for female PALB2 mutation carriers, as compared with the general population, was eight to nine times as high among those younger than 40 years of age, six to eight times as high among those 40 to 60 years of age, and five times as high among those older than 60 years of age. The estimated cumulative risk of breast cancer among female mutation carriers was 14% (95% confidence interval [CI], 9 to 20) by 50 years of age and 35% (95% CI, 26 to 46) by 70 years of age. Breast-cancer risk was also significantly influenced by birth cohort (P<0.001) and by other familial factors (P=0.04). The absolute breast-cancer risk for PALB2 female mutation carriers by 70 years of age ranged from 33% (95% CI, 25 to 44) for those with no family history of breast cancer to 58% (95% CI, 50 to 66) for those with two or more first-degree relatives with breast cancer at 50 years of age. CONCLUSIONS: Loss-of-function mutations in PALB2 are an important cause of hereditary breast cancer, with respect both to the frequency of cancer-predisposing mutations and to the risk associated with them. Our data suggest the breast-cancer risk for PALB2 mutation carriers may overlap with that for BRCA2 mutation carriers. (Funded by the European Research Council and others.)..
Antoniou, A.C.
Casadei, S.
Heikkinen, T.
Barrowdale, D.
Pylkas, K.
Roberts, J.
Lee, A.
Subramanian, D.
de Leeneer, K.
Fostira, F.
Tomiak, E.
Neuhausen, S.L.
Teo, Z.L.
Khan, S.
Aittomaki, K.
Moilanen, J.S.
Turnbull, C.
Seal, S.
Mannermaa, A.
Kallioniemi, A.
Lindeman, G.J.
Buys, S.S.
Andrulis, I.L.
Radice, P.
Tondini, C.
Manoukian, S.
Toland, A.E.
Miron, P.
Weitzel, J.N.
Domchek, S.M.
Poppe, B.
Claes, K.B.
Yannoukakos, D.
Concannon, P.
Bernstein, J.L.
James, P.A.
Easton, D.F.
Goldgar, D.E.
Hopper, J.L.
Rahman, N.
Peterlongo, P.
Nevanlinna, H.
King, M.-.
Couch, F.J.
Southey, M.C.
Winqvist, R.
Foulkes, W.D.
Tischkowitz, M.
(2014). Breast-Cancer Risk in Families With Mutations in PALB2 EDITORIAL COMMENT. Obstetrical & gynecological survey,
Vol.69
(11),
pp. 659-660.
Koster, R.
Mitra, N.
D'Andrea, K.
Vardhanabhuti, S.
Chung, C.C.
Wang, Z.
Loren Erickson, R.
Vaughn, D.J.
Litchfield, K.
Rahman, N.
Greene, M.H.
McGlynn, K.A.
Turnbull, C.
Chanock, S.J.
Nathanson, K.L.
Kanetsky, P.A.
(2014). Pathway-based analysis of GWAs data identifies association of sex determination genes with susceptibility to testicular germ cell tumors. Hum mol genet,
Vol.23
(22),
pp. 6061-6068.
show abstract
Genome-wide association (GWA) studies of testicular germ cell tumor (TGCT) have identified 18 susceptibility loci, some containing genes encoding proteins important in male germ cell development. Deletions of one of these genes, DMRT1, lead to male-to-female sex reversal and are associated with development of gonadoblastoma. To further explore genetic association with TGCT, we undertook a pathway-based analysis of SNP marker associations in the Penn GWAs (349 TGCT cases and 919 controls). We analyzed a custom-built sex determination gene set consisting of 32 genes using three different methods of pathway-based analysis. The sex determination gene set ranked highly compared with canonical gene sets, and it was associated with TGCT (FDRG = 2.28 × 10(-5), FDRM = 0.014 and FDRI = 0.008 for Gene Set Analysis-SNP (GSA-SNP), Meta-Analysis Gene Set Enrichment of Variant Associations (MAGENTA) and Improved Gene Set Enrichment Analysis for Genome-wide Association Study (i-GSEA4GWAS) analysis, respectively). The association remained after removal of DMRT1 from the gene set (FDRG = 0.0002, FDRM = 0.055 and FDRI = 0.009). Using data from the NCI GWA scan (582 TGCT cases and 1056 controls) and UK scan (986 TGCT cases and 4946 controls), we replicated these findings (NCI: FDRG = 0.006, FDRM = 0.014, FDRI = 0.033, and UK: FDRG = 1.04 × 10(-6), FDRM = 0.016, FDRI = 0.025). After removal of DMRT1 from the gene set, the sex determination gene set remains associated with TGCT in the NCI (FDRG = 0.039, FDRM = 0.050 and FDRI = 0.055) and UK scans (FDRG = 3.00 × 10(-5), FDRM = 0.056 and FDRI = 0.044). With the exception of DMRT1, genes in the sex determination gene set have not previously been identified as TGCT susceptibility loci in these GWA scans, demonstrating the complementary nature of a pathway-based approach for genome-wide analysis of TGCT..
Antoniou, A.C.
Foulkes, W.D.
Tischkowitz, M.
(2014). Breast-cancer risk in families with mutations in PALB2. N engl j med,
Vol.371
(17),
pp. 1651-1652.
Antoniou, A.C.
Foulkes, W.D.
Tischkowitz, M.
(2014). Breast-Cancer Risk in Families with Mutations in PALB2 Reply. New england journal of medicine,
Vol.371
(17),
pp. 1651-2.
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.
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.
Haiman, C.A.
Henderson, B.E.
Schumacher, F.
Le Marchand, L.
Pharoah, P.D.
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.-.
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.
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.
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.-.
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.-.
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. Plos one,
Vol.9
(11),
p. e109973.
show abstract
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..
Schumacher, F.R.
Wang, Z.
Skotheim, R.I.
Koster, R.
Chung, C.C.
Hildebrandt, M.A.
Kratz, C.P.
Bakken, A.C.
Bishop, D.T.
Cook, M.B.
Erickson, R.L.
Fosså, S.D.
Greene, M.H.
Jacobs, K.B.
Kanetsky, P.A.
Kolonel, L.N.
Loud, J.T.
Korde, L.A.
Le Marchand, L.
Lewinger, J.P.
Lothe, R.A.
Pike, M.C.
Rahman, N.
Rubertone, M.V.
Schwartz, S.M.
Siegmund, K.D.
Skinner, E.C.
Turnbull, C.
Van Den Berg, D.J.
Wu, X.
Yeager, M.
Nathanson, K.L.
Chanock, S.J.
Cortessis, V.K.
McGlynn, K.A.
(2013). Testicular germ cell tumor susceptibility associated with the UCK2 locus on chromosome 1q23. Hum mol genet,
Vol.22
(13),
pp. 2748-2753.
show abstract
Genome-wide association studies (GWASs) have identified multiple common genetic variants associated with an increased risk of testicular germ cell tumors (TGCTs). A previous GWAS reported a possible TGCT susceptibility locus on chromosome 1q23 in the UCK2 gene, but failed to reach genome-wide significance following replication. We interrogated this region by conducting a meta-analysis of two independent GWASs including a total of 940 TGCT cases and 1559 controls for 122 single-nucleotide polymorphisms (SNPs) on chromosome 1q23 and followed up the most significant SNPs in an additional 2202 TGCT cases and 2386 controls from four case-control studies. We observed genome-wide significant associations for several UCK2 markers, the most significant of which was for rs3790665 (PCombined = 6.0 × 10(-9)). Additional support is provided from an independent familial study of TGCT where a significant over-transmission for rs3790665 with TGCT risk was observed (PFBAT = 2.3 × 10(-3)). Here, we provide substantial evidence for the association between UCK2 genetic variation and TGCT risk..
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.-.
Blot, W.
Kang, D.
Hartman, M.
Sangrajrang, S.
Shen, C.-.
Southey, M.C.
Park, D.J.
Hammet, F.
Stone, J.
Veer, L.J.
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.
Menéndez, P.
Müller, H.
Arndt, V.
Stegmaier, C.
Lichtner, P.
Lochmann, M.
Justenhoven, C.
Ko, Y.-.
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.-.
Hartikainen, J.M.
Balleine, R.
kConFab Investigators,
Tseng, C.-.
Berg, D.V.
Stram, D.O.
Neven, P.
Dieudonné, A.-.
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.
Cheng, C.-.
Shrubsole, M.
Long, J.
Pylkäs, K.
Jukkola-Vuorinen, A.
Kauppila, S.
Knight, J.A.
Glendon, G.
Mulligan, A.M.
Tollenaar, R.A.
Seynaeve, C.M.
Kriege, M.
Hooning, M.J.
van den Ouweland, A.M.
van Deurzen, C.H.
Lu, W.
Gao, Y.-.
Cai, H.
Balasubramanian, S.P.
Cross, S.S.
Reed, M.W.
Signorello, L.
Cai, Q.
Shah, M.
Miao, H.
Chan, C.W.
Chia, K.S.
Jakubowska, A.
Jaworska, K.
Durda, K.
Hsiung, C.-.
Wu, P.-.
Yu, J.-.
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.-.
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.
Vachon, C.
Easton, D.F.
Haiman, C.A.
Kraft, P.
(2013). Genome-wide association studies identify four ER negative-specific breast cancer risk loci. Nat genet,
Vol.45
(4),
pp. 392-398e2.
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..
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.
Breast and Ovarian Cancer Susceptibility Collaboration,
Fletcher, O.
Peto, J.
Gibson, L.
Dos Santos Silva, I.
Nevanlinna, H.
Muranen, T.A.
Aittomäki, K.
Blomqvist, C.
Czene, K.
Irwanto, A.
Liu, J.
Waisfisz, Q.
Meijers-Heijboer, H.
Adank, M.
Hereditary Breast and Ovarian Cancer Research Group Netherlands (HEBON),
van der Luijt, R.B.
Hein, R.
Dahmen, N.
Beckman, L.
Meindl, A.
Schmutzler, R.K.
Müller-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.
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.-.
Rudolph, A.
Nickels, S.
Flesch-Janys, D.
Johnson, N.
Aitken, Z.
Aaltonen, K.
Heikkinen, T.
Broeks, A.
Veer, L.J.
van der Schoot, C.E.
Guénel, P.
Truong, T.
Laurent-Puig, P.
Menegaux, F.
Marme, F.
Schneeweiss, A.
Sohn, C.
Burwinkel, B.
Zamora, M.P.
Perez, J.I.
Pita, G.
Alonso, M.R.
Cox, A.
Brock, I.W.
Cross, S.S.
Reed, M.W.
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.
kConFab Investigators,
Australian Ovarian Cancer Study Group,
Lindblom, A.
Margolin, S.
Hooning, M.J.
Hollestelle, A.
van den Ouweland, A.M.
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.
Müller, H.
Arndt, V.
Stegmaier, C.
Swerdlow, A.
Ashworth, A.
Orr, N.
Jones, M.
Figueroa, J.
Lissowska, J.
Brinton, L.
Goldberg, M.S.
Labrèche, F.
Dumont, M.
Winqvist, R.
Pylkäs, K.
Jukkola-Vuorinen, A.
Grip, M.
Brauch, H.
Hamann, U.
Brüning, T.
GENICA (Gene Environment Interaction and Breast Cancer in Germany) Network,
Radice, P.
Peterlongo, P.
Manoukian, S.
Bonanni, B.
Devilee, P.
Tollenaar, R.A.
Seynaeve, C.
van Asperen, C.J.
Jakubowska, A.
Lubinski, J.
Jaworska, K.
Durda, K.
Mannermaa, A.
Kataja, V.
Kosma, V.-.
Hartikainen, J.M.
Bogdanova, N.V.
Antonenkova, N.N.
Dörk, T.
Kristensen, V.N.
Anton-Culver, H.
Slager, S.
Toland, A.E.
Edge, S.
Fostira, F.
Kang, D.
Yoo, K.-.
Noh, D.-.
Matsuo, K.
Ito, H.
Iwata, H.
Sueta, A.
Wu, A.H.
Tseng, C.-.
Van Den Berg, D.
Stram, D.O.
Shu, X.-.
Lu, W.
Gao, Y.-.
Cai, H.
Teo, S.H.
Yip, C.H.
Phuah, S.Y.
Cornes, B.K.
Hartman, M.
Miao, H.
Lim, W.Y.
Sng, J.-.
Muir, K.
Lophatananon, A.
Stewart-Brown, S.
Siriwanarangsan, P.
Shen, C.-.
Hsiung, C.-.
Wu, P.-.
Ding, S.-.
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.
Chenevix-Trench, G.
Dunning, A.M.
Benitez, J.
Easton, D.F.
(2013). Large-scale genotyping identifies 41 new loci associated with breast cancer risk. Nat genet,
Vol.45
(4),
pp. 353-361e2.
show abstract
full text
Breast cancer is the most common cancer among women. Common variants at 27 loci have been identified as associated with susceptibility to breast cancer, and these account for ∼9% of the familial risk of the disease. We report here a meta-analysis of 9 genome-wide association studies, including 10,052 breast cancer cases and 12,575 controls of European ancestry, from which we selected 29,807 SNPs for further genotyping. These SNPs were genotyped in 45,290 cases and 41,880 controls of European ancestry from 41 studies in the Breast Cancer Association Consortium (BCAC). The SNPs were genotyped as part of a collaborative genotyping experiment involving four consortia (Collaborative Oncological Gene-environment Study, COGS) and used a custom Illumina iSelect genotyping array, iCOGS, comprising more than 200,000 SNPs. We identified SNPs at 41 new breast cancer susceptibility loci at genome-wide significance (P < 5 × 10(-8)). Further analyses suggest that more than 1,000 additional loci are involved in breast cancer susceptibility..
Ruark, E.
Seal, S.
McDonald, H.
Zhang, F.
Elliot, A.
Lau, K.
Perdeaux, E.
Rapley, E.
Eeles, R.
Peto, J.
Kote-Jarai, Z.
Muir, K.
Nsengimana, J.
Shipley, J.
UK Testicular Cancer Collaboration (UKTCC),
Bishop, D.T.
Stratton, M.R.
Easton, D.F.
Huddart, R.A.
Rahman, N.
Turnbull, C.
(2013). Identification of nine new susceptibility loci for testicular cancer, including variants near DAZL and PRDM14. Nat genet,
Vol.45
(6),
pp. 686-689.
show abstract
full text
Testicular germ cell tumor (TGCT) is the most common cancer in young men and is notable for its high familial risks. So far, six loci associated with TGCT have been reported. From genome-wide association study (GWAS) analysis of 307,291 SNPs in 986 TGCT cases and 4,946 controls, we selected for follow-up 694 SNPs, which we genotyped in a further 1,064 TGCT cases and 10,082 controls from the UK. We identified SNPs at nine new loci (1q22, 1q24.1, 3p24.3, 4q24, 5q31.1, 8q13.3, 16q12.1, 17q22 and 21q22.3) showing association with TGCT (P < 5 × 10(-8)), which together account for an additional 4-6% of the familial risk of TGCT. The loci include genes plausibly related to TGCT development. PRDM14, at 8q13.3, is essential for early germ cell specification, and DAZL, at 3p24.3, is required for the regulation of germ cell development. Furthermore, PITX1, at 5q31.1, regulates TERT expression and is the third TGCT-associated locus implicated in telomerase regulation..
Chung, C.C.
Kanetsky, P.A.
Wang, Z.
Hildebrandt, M.A.
Koster, R.
Skotheim, R.I.
Kratz, C.P.
Turnbull, C.
Cortessis, V.K.
Bakken, A.C.
Bishop, D.T.
Cook, M.B.
Erickson, R.L.
Fosså, S.D.
Jacobs, K.B.
Korde, L.A.
Kraggerud, S.M.
Lothe, R.A.
Loud, J.T.
Rahman, N.
Skinner, E.C.
Thomas, D.C.
Wu, X.
Yeager, M.
Schumacher, F.R.
Greene, M.H.
Schwartz, S.M.
McGlynn, K.A.
Chanock, S.J.
Nathanson, K.L.
(2013). Meta-analysis identifies four new loci associated with testicular germ cell tumor. Nat genet,
Vol.45
(6),
pp. 680-685.
show abstract
full text
We conducted a meta-analysis to identify new susceptibility loci for testicular germ cell tumor (TGCT). In the discovery phase, we analyzed 931 affected individuals and 1,975 controls from 3 genome-wide association studies (GWAS). We conducted replication in 6 independent sample sets comprising 3,211 affected individuals and 7,591 controls. In the combined analysis, risk of TGCT was significantly associated with markers at four previously unreported loci: 4q22.2 in HPGDS (per-allele odds ratio (OR) = 1.19, 95% confidence interval (CI) = 1.12-1.26; P = 1.11 × 10(-8)), 7p22.3 in MAD1L1 (OR = 1.21, 95% CI = 1.14-1.29; P = 5.59 × 10(-9)), 16q22.3 in RFWD3 (OR = 1.26, 95% CI = 1.18-1.34; P = 5.15 × 10(-12)) and 17q22 (rs9905704: OR = 1.27, 95% CI = 1.18-1.33; P = 4.32 × 10(-13) and rs7221274: OR = 1.20, 95% CI = 1.12-1.28; P = 4.04 × 10(-9)), a locus that includes TEX14, RAD51C and PPM1E. These new TGCT susceptibility loci contain biologically plausible genes encoding proteins important for male germ cell development, chromosomal segregation and the DNA damage response..
Ruark, E.
Snape, K.
Humburg, P.
Loveday, C.
Bajrami, I.
Brough, R.
Rodrigues, D.N.
Renwick, A.
Seal, S.
Ramsay, E.
Duarte, S.D.
Rivas, M.A.
Warren-Perry, M.
Zachariou, A.
Campion-Flora, A.
Hanks, S.
Murray, A.
Ansari Pour, N.
Douglas, J.
Gregory, L.
Rimmer, A.
Walker, N.M.
Yang, T.-.
Adlard, J.W.
Barwell, J.
Berg, J.
Brady, A.F.
Brewer, C.
Brice, G.
Chapman, C.
Cook, J.
Davidson, R.
Donaldson, A.
Douglas, F.
Eccles, D.
Evans, D.G.
Greenhalgh, L.
Henderson, A.
Izatt, L.
Kumar, A.
Lalloo, F.
Miedzybrodzka, Z.
Morrison, P.J.
Paterson, J.
Porteous, M.
Rogers, M.T.
Shanley, S.
Walker, L.
Gore, M.
Houlston, R.
Brown, M.A.
Caufield, M.J.
Deloukas, P.
McCarthy, M.I.
Todd, J.A.
Breast and Ovarian Cancer Susceptibility Collaboration,
Wellcome Trust Case Control Consortium,
Turnbull, C.
Reis-Filho, J.S.
Ashworth, A.
Antoniou, A.C.
Lord, C.J.
Donnelly, P.
Rahman, N.
(2013). Mosaic PPM1D mutations are associated with predisposition to breast and ovarian cancer. Nature,
Vol.493
(7432),
pp. 406-410.
show abstract
full text
Improved sequencing technologies offer unprecedented opportunities for investigating the role of rare genetic variation in common disease. However, there are considerable challenges with respect to study design, data analysis and replication. Using pooled next-generation sequencing of 507 genes implicated in the repair of DNA in 1,150 samples, an analytical strategy focused on protein-truncating variants (PTVs) and a large-scale sequencing case-control replication experiment in 13,642 individuals, here we show that rare PTVs in the p53-inducible protein phosphatase PPM1D are associated with predisposition to breast cancer and ovarian cancer. PPM1D PTV mutations were present in 25 out of 7,781 cases versus 1 out of 5,861 controls (P = 1.12 × 10(-5)), including 18 mutations in 6,912 individuals with breast cancer (P = 2.42 × 10(-4)) and 12 mutations in 1,121 individuals with ovarian cancer (P = 3.10 × 10(-9)). Notably, all of the identified PPM1D PTVs were mosaic in lymphocyte DNA and clustered within a 370-base-pair region in the final exon of the gene, carboxy-terminal to the phosphatase catalytic domain. Functional studies demonstrate that the mutations result in enhanced suppression of p53 in response to ionizing radiation exposure, suggesting that the mutant alleles encode hyperactive PPM1D isoforms. Thus, although the mutations cause premature protein truncation, they do not result in the simple loss-of-function effect typically associated with this class of variant, but instead probably have a gain-of-function effect. Our results have implications for the detection and management of breast and ovarian cancer risk. More generally, these data provide new insights into the role of rare and of mosaic genetic variants in common conditions, and the use of sequencing in their identification..
Huertas-Vazquez, A.
Nelson, C.P.
Guo, X.
Reinier, K.
Uy-Evanado, A.
Teodorescu, C.
Ayala, J.
Jerger, K.
Chugh, H.
WTCCC+,
Braund, P.S.
Deloukas, P.
Hall, A.S.
Balmforth, A.J.
Jones, M.
Taylor, K.D.
Pulit, S.L.
Newton-Cheh, C.
Gunson, K.
Jui, J.
Rotter, J.I.
Albert, C.M.
Samani, N.J.
Chugh, S.S.
(2013). Novel loci associated with increased risk of sudden cardiac death in the context of coronary artery disease. Plos one,
Vol.8
(4),
p. e59905.
show abstract
BACKGROUND: Recent genome-wide association studies (GWAS) have identified novel loci associated with sudden cardiac death (SCD). Despite this progress, identified DNA variants account for a relatively small portion of overall SCD risk, suggesting that additional loci contributing to SCD susceptibility await discovery. The objective of this study was to identify novel DNA variation associated with SCD in the context of coronary artery disease (CAD). METHODS AND FINDINGS: Using the MetaboChip custom array we conducted a case-control association analysis of 119,117 SNPs in 948 SCD cases (with underlying CAD) from the Oregon Sudden Unexpected Death Study (Oregon-SUDS) and 3,050 controls with CAD from the Wellcome Trust Case-Control Consortium (WTCCC). Two newly identified loci were significantly associated with increased risk of SCD after correction for multiple comparisons at: rs6730157 in the RAB3GAP1 gene on chromosome 2 (P = 4.93×10(-12), OR = 1.60) and rs2077316 in the ZNF365 gene on chromosome 10 (P = 3.64×10(-8), OR = 2.41). CONCLUSIONS: Our findings suggest that RAB3GAP1 and ZNF365 are relevant candidate genes for SCD and will contribute to the mechanistic understanding of SCD susceptibility..
Jostins, L.
Ripke, S.
Weersma, R.K.
Duerr, R.H.
McGovern, D.P.
Hui, K.Y.
Lee, J.C.
Schumm, L.P.
Sharma, Y.
Anderson, C.A.
Essers, J.
Mitrovic, M.
Ning, K.
Cleynen, I.
Theatre, E.
Spain, S.L.
Raychaudhuri, S.
Goyette, P.
Wei, Z.
Abraham, C.
Achkar, J.-.
Ahmad, T.
Amininejad, L.
Ananthakrishnan, A.N.
Andersen, V.
Andrews, J.M.
Baidoo, L.
Balschun, T.
Bampton, P.A.
Bitton, A.
Boucher, G.
Brand, S.
Büning, C.
Cohain, A.
Cichon, S.
D'Amato, M.
De Jong, D.
Devaney, K.L.
Dubinsky, M.
Edwards, C.
Ellinghaus, D.
Ferguson, L.R.
Franchimont, D.
Fransen, K.
Gearry, R.
Georges, M.
Gieger, C.
Glas, J.
Haritunians, T.
Hart, A.
Hawkey, C.
Hedl, M.
Hu, X.
Karlsen, T.H.
Kupcinskas, L.
Kugathasan, S.
Latiano, A.
Laukens, D.
Lawrance, I.C.
Lees, C.W.
Louis, E.
Mahy, G.
Mansfield, J.
Morgan, A.R.
Mowat, C.
Newman, W.
Palmieri, O.
Ponsioen, C.Y.
Potocnik, U.
Prescott, N.J.
Regueiro, M.
Rotter, J.I.
Russell, R.K.
Sanderson, J.D.
Sans, M.
Satsangi, J.
Schreiber, S.
Simms, L.A.
Sventoraityte, J.
Targan, S.R.
Taylor, K.D.
Tremelling, M.
Verspaget, H.W.
De Vos, M.
Wijmenga, C.
Wilson, D.C.
Winkelmann, J.
Xavier, R.J.
Zeissig, S.
Zhang, B.
Zhang, C.K.
Zhao, H.
International IBD Genetics Consortium (IIBDGC),
Silverberg, M.S.
Annese, V.
Hakonarson, H.
Brant, S.R.
Radford-Smith, G.
Mathew, C.G.
Rioux, J.D.
Schadt, E.E.
Daly, M.J.
Franke, A.
Parkes, M.
Vermeire, S.
Barrett, J.C.
Cho, J.H.
(2012). Host-microbe interactions have shaped the genetic architecture of inflammatory bowel disease. Nature,
Vol.491
(7422),
pp. 119-124.
show abstract
full text
Crohn's disease and ulcerative colitis, the two common forms of inflammatory bowel disease (IBD), affect over 2.5 million people of European ancestry, with rising prevalence in other populations. Genome-wide association studies and subsequent meta-analyses of these two diseases as separate phenotypes have implicated previously unsuspected mechanisms, such as autophagy, in their pathogenesis and showed that some IBD loci are shared with other inflammatory diseases. Here we expand on the knowledge of relevant pathways by undertaking a meta-analysis of Crohn's disease and ulcerative colitis genome-wide association scans, followed by extensive validation of significant findings, with a combined total of more than 75,000 cases and controls. We identify 71 new associations, for a total of 163 IBD loci, that meet genome-wide significance thresholds. Most loci contribute to both phenotypes, and both directional (consistently favouring one allele over the course of human history) and balancing (favouring the retention of both alleles within populations) selection effects are evident. Many IBD loci are also implicated in other immune-mediated disorders, most notably with ankylosing spondylitis and psoriasis. We also observe considerable overlap between susceptibility loci for IBD and mycobacterial infection. Gene co-expression network analysis emphasizes this relationship, with pathways shared between host responses to mycobacteria and those predisposing to IBD..
Cooper, J.D.
Simmonds, M.J.
Walker, N.M.
Burren, O.
Brand, O.J.
Guo, H.
Wallace, C.
Stevens, H.
Coleman, G.
Wellcome Trust Case Control Consortium,
Franklyn, J.A.
Todd, J.A.
Gough, S.C.
(2012). Seven newly identified loci for autoimmune thyroid disease. Hum mol genet,
Vol.21
(23),
pp. 5202-5208.
show abstract
Autoimmune thyroid disease (AITD), including Graves' disease (GD) and Hashimoto's thyroiditis (HT), is one of the most common of the immune-mediated diseases. To further investigate the genetic determinants of AITD, we conducted an association study using a custom-made single-nucleotide polymorphism (SNP) array, the ImmunoChip. The SNP array contains all known and genotype-able SNPs across 186 distinct susceptibility loci associated with one or more immune-mediated diseases. After stringent quality control, we analysed 103 875 common SNPs (minor allele frequency >0.05) in 2285 GD and 462 HT patients and 9364 controls. We found evidence for seven new AITD risk loci (P < 1.12 × 10(-6); a permutation test derived significance threshold), five at locations previously associated and two at locations awaiting confirmation, with other immune-mediated diseases..
Scott, R.H.
Murray, A.
Baskcomb, L.
Turnbull, C.
Loveday, C.
Al-Saadi, R.
Williams, R.
Breatnach, F.
Gerrard, M.
Hale, J.
Kohler, J.
Lapunzina, P.
Levitt, G.A.
Picton, S.
Pizer, B.
Ronghe, M.D.
Traunecker, H.
Williams, D.
Kelsey, A.
Vujanic, G.M.
Sebire, N.J.
Grundy, P.
Stiller, C.A.
Pritchard-Jones, K.
Douglas, J.
Rahman, N.
(2012). Stratification of Wilms tumor by genetic and epigenetic analysis. Oncotarget,
Vol.3
(3),
pp. 327-335.
show abstract
Somatic defects at five loci, WT1, CTNNB1, WTX, TP53 and the imprinted 11p15 region, are implicated in Wilms tumor, the commonest childhood kidney cancer. In this study we analysed all five loci in 120 Wilms tumors. We identified epigenetic 11p15 abnormalities in 69% of tumors, 37% were H19 epimutations and 32% were paternal uniparental disomy (pUPD). We identified mutations of WTX in 32%, CTNNB1 in 15%, WT1 in 12% and TP53 in 5% of tumors. We identified several significant associations: between 11p15 and WTX (P=0.007), between WT1 and CTNNB1 (P less than 0.001), between WT1 and pUPD 11p15 (P=0.01), and a strong negative association between WT1 and H19 epimutation (P less than 0.001). We next used these data to stratify Wilms tumor into three molecular Groups, based on the status at 11p15 and WT1. Group 1 tumors (63%) were defined as 11p15-mutant and WT1-normal; a third also had WTX mutations. Group 2 tumors (13%) were WT1-mutant. They either had 11p15 pUPD or were 11p15-normal. Almost all had CTNNB1 mutations but none had H19 epimutation. Group 3 tumors (25%) were defined as 11p15-normal and WT1-normal and were typically normal at all five loci (P less than 0.001). We also identified a novel clinical association between H19 epimutation and bilateral disease (P less than 0.001). These data provide new insights into the pattern, order, interactions and clinical associations of molecular events in Wilms tumor..
Turnbull, C.
Loveday, C.
Ramsay, E.
Hughes, D.
Ruark, E.
Frankum, J.R.
Warren-Perry, M.
Snape, K.
Eccles, D.
Evans, D.G.
Renwick, A.
Seal, S.
Lord, C.J.
Ashworth, A.
Reis-Filho, J.S.
Antoniou, A.C.
Rahman, N.
(2012). Newly discovered high-intermediate penetrance ovarian cancer susceptibility genes: clinical testing of RAD51D and RAD51C may have significant utility. International journal of gynecological cancer,
Vol.22,
pp. S42-S42.
Snape, K.
Ruark, E.
Tarpey, P.
Renwick, A.
Turnbull, C.
Seal, S.
Murray, A.
Hanks, S.
Douglas, J.
Stratton, M.R.
Rahman, N.
(2012). Predisposition gene identification in common cancers by exome sequencing: insights from familial breast cancer. Breast cancer res treat,
Vol.134
(1),
pp. 429-433.
show abstract
full text
The genetic component of breast cancer predisposition remains largely unexplained. Candidate gene case-control resequencing has identified predisposition genes characterised by rare, protein truncating mutations that confer moderate risks of disease. In theory, exome sequencing should yield additional genes of this class. Here, we explore the feasibility and design considerations of this approach. We performed exome sequencing in 50 individuals with familial breast cancer, applying frequency and protein function filters to identify variants most likely to be pathogenic. We identified 867,378 variants that passed the call quality filters of which 1,296 variants passed the frequency and protein truncation filters. The median number of validated, rare, protein truncating variants was 10 in individuals with, and without, mutations in known genes. The functional candidacy of mutated genes was similar in both groups. Without prior knowledge, the known genes would not have been recognisable as breast cancer predisposition genes. Everyone carries multiple rare mutations that are plausibly related to disease. Exome sequencing in common conditions will therefore require intelligent sample and variant prioritisation strategies in large case-control studies to deliver robust genetic evidence of disease association..
Robertson, L.
Hanson, H.
Seal, S.
Warren-Perry, M.
Hughes, D.
Howell, I.
Turnbull, C.
Houlston, R.
Shanley, S.
Butler, S.
Evans, D.G.
Ross, G.
Eccles, D.
Tutt, A.
Rahman, N.
TNT Trial TMG,
BCSC (UK),
(2012). BRCA1 testing should be offered to individuals with triple-negative breast cancer diagnosed below 50 years. Br j cancer,
Vol.106
(6),
pp. 1234-1238.
show abstract
full text
BACKGROUND: Triple-negative (TN) tumours are the predominant breast cancer subtype in BRCA1 mutation carriers. Recently, it was proposed that all individuals below 50 years of age with TN breast cancer should be offered BRCA testing. We have evaluated the BRCA1 mutation frequency and the implications for clinical practice of undertaking genetic testing in women with TN breast cancer. METHODS: We undertook BRCA1 mutation analysis in 308 individuals with TN breast cancer, 159 individuals from unselected series of breast cancer and 149 individuals from series ascertained on the basis of young age and/or family history. RESULTS: BRCA1 mutations were present in 45 out of 308 individuals. Individuals with TN cancer <50 years had >10% likelihood of carrying a BRCA1 mutation in both the unselected (11 out of 58, 19%) and selected (26 out of 111, 23%) series. However, over a third would not have been offered testing using existing criteria. We estimate that testing all individuals with TN breast cancer <50 years would generate an extra 1200 tests annually in England. CONCLUSION: Women with TN breast cancer diagnosed below 50 years have >10% likelihood of carrying a BRCA1 mutation and are therefore eligible for testing in most centres. However, implementation may place short-term logistical and financial burdens on genetic services..
Turnbull, C.
Seal, S.
Renwick, A.
Warren-Perry, M.
Hughes, D.
Elliott, A.
Pernet, D.
Peock, S.
Adlard, J.W.
Barwell, J.
Berg, J.
Brady, A.F.
Brewer, C.
Brice, G.
Chapman, C.
Cook, J.
Davidson, R.
Donaldson, A.
Douglas, F.
Greenhalgh, L.
Henderson, A.
Izatt, L.
Kumar, A.
Lalloo, F.
Miedzybrodzka, Z.
Morrison, P.J.
Paterson, J.
Porteous, M.
Rogers, M.T.
Shanley, S.
Walker, L.
Breast Cancer Susceptibility Collaboration (UK), EMBRACE,
Ahmed, M.
Eccles, D.
Evans, D.G.
Donnelly, P.
Easton, D.F.
Stratton, M.R.
Rahman, N.
(2012). Gene-gene interactions in breast cancer susceptibility. Hum mol genet,
Vol.21
(4),
pp. 958-962.
show abstract
There have been few definitive examples of gene-gene interactions in humans. Through mutational analyses in 7325 individuals, we report four interactions (defined as departures from a multiplicative model) between mutations in the breast cancer susceptibility genes ATM and CHEK2 with BRCA1 and BRCA2 (case-only interaction between ATM and BRCA1/BRCA2 combined, P = 5.9 × 10(-4); ATM and BRCA1, P= 0.01; ATM and BRCA2, P= 0.02; CHEK2 and BRCA1/BRCA2 combined, P = 2.1 × 10(-4); CHEK2 and BRCA1, P= 0.01; CHEK2 and BRCA2, P= 0.01). The interactions are such that the resultant risk of breast cancer is lower than the multiplicative product of the constituent risks, and plausibly reflect the functional relationships of the encoded proteins in DNA repair. These findings have important implications for models of disease predisposition and clinical translation..
Siddiq, A.
Couch, F.J.
Chen, G.K.
Lindström, S.
Eccles, D.
Millikan, R.C.
Michailidou, K.
Stram, D.O.
Beckmann, L.
Rhie, S.K.
Ambrosone, C.B.
Aittomäki, K.
Amiano, P.
Apicella, C.
Australian Breast Cancer Tissue Bank Investigators,
Baglietto, L.
Bandera, E.V.
Beckmann, M.W.
Berg, C.D.
Bernstein, L.
Blomqvist, C.
Brauch, H.
Brinton, L.
Bui, Q.M.
Buring, J.E.
Buys, S.S.
Campa, D.
Carpenter, J.E.
Chasman, D.I.
Chang-Claude, J.
Chen, C.
Clavel-Chapelon, F.
Cox, A.
Cross, S.S.
Czene, K.
Deming, S.L.
Diasio, R.B.
Diver, W.R.
Dunning, A.M.
Durcan, L.
Ekici, A.B.
Fasching, P.A.
Familial Breast Cancer Study,
Feigelson, H.S.
Fejerman, L.
Figueroa, J.D.
Fletcher, O.
Flesch-Janys, D.
Gaudet, M.M.
GENICA Consortium,
Gerty, S.M.
Rodriguez-Gil, J.L.
Giles, G.G.
van Gils, C.H.
Godwin, A.K.
Graham, N.
Greco, D.
Hall, P.
Hankinson, S.E.
Hartmann, A.
Hein, R.
Heinz, J.
Hoover, R.N.
Hopper, J.L.
Hu, J.J.
Huntsman, S.
Ingles, S.A.
Irwanto, A.
Isaacs, C.
Jacobs, K.B.
John, E.M.
Justenhoven, C.
Kaaks, R.
Kolonel, L.N.
Coetzee, G.A.
Lathrop, M.
Le Marchand, L.
Lee, A.M.
Lee, I.-.
Lesnick, T.
Lichtner, P.
Liu, J.
Lund, E.
Makalic, E.
Martin, N.G.
McLean, C.A.
Meijers-Heijboer, H.
Meindl, A.
Miron, P.
Monroe, K.R.
Montgomery, G.W.
Müller-Myhsok, B.
Nickels, S.
Nyante, S.J.
Olswold, C.
Overvad, K.
Palli, D.
Park, D.J.
Palmer, J.R.
Pathak, H.
Peto, J.
Pharoah, P.
Rahman, N.
Rivadeneira, F.
Schmidt, D.F.
Schmutzler, R.K.
Slager, S.
Southey, M.C.
Stevens, K.N.
Sinn, H.-.
Press, M.F.
Ross, E.
Riboli, E.
Ridker, P.M.
Schumacher, F.R.
Severi, G.
Dos Santos Silva, I.
Stone, J.
Sund, M.
Tapper, W.J.
Thun, M.J.
Travis, R.C.
Turnbull, C.
Uitterlinden, A.G.
Waisfisz, Q.
Wang, X.
Wang, Z.
Weaver, J.
Schulz-Wendtland, R.
Wilkens, L.R.
Van Den Berg, D.
Zheng, W.
Ziegler, R.G.
Ziv, E.
Nevanlinna, H.
Easton, D.F.
Hunter, D.J.
Henderson, B.E.
Chanock, S.J.
Garcia-Closas, M.
Kraft, P.
Haiman, C.A.
Vachon, C.M.
(2012). A meta-analysis of genome-wide association studies of breast cancer identifies two novel susceptibility loci at 6q14 and 20q11. Hum mol genet,
Vol.21
(24),
pp. 5373-5384.
show abstract
Genome-wide association studies (GWAS) of breast cancer defined by hormone receptor status have revealed loci contributing to susceptibility of estrogen receptor (ER)-negative subtypes. To identify additional genetic variants for ER-negative breast cancer, we conducted the largest meta-analysis of ER-negative disease to date, comprising 4754 ER-negative cases and 31 663 controls from three GWAS: NCI Breast and Prostate Cancer Cohort Consortium (BPC3) (2188 ER-negative cases; 25 519 controls of European ancestry), Triple Negative Breast Cancer Consortium (TNBCC) (1562 triple negative cases; 3399 controls of European ancestry) and African American Breast Cancer Consortium (AABC) (1004 ER-negative cases; 2745 controls). We performed in silico replication of 86 SNPs at P ≤ 1 × 10(-5) in an additional 11 209 breast cancer cases (946 with ER-negative disease) and 16 057 controls of Japanese, Latino and European ancestry. We identified two novel loci for breast cancer at 20q11 and 6q14. SNP rs2284378 at 20q11 was associated with ER-negative breast cancer (combined two-stage OR = 1.16; P = 1.1 × 10(-8)) but showed a weaker association with overall breast cancer (OR = 1.08, P = 1.3 × 10(-6)) based on 17 869 cases and 43 745 controls and no association with ER-positive disease (OR = 1.01, P = 0.67) based on 9965 cases and 22 902 controls. Similarly, rs17530068 at 6q14 was associated with breast cancer (OR = 1.12; P = 1.1 × 10(-9)), and with both ER-positive (OR = 1.09; P = 1.5 × 10(-5)) and ER-negative (OR = 1.16, P = 2.5 × 10(-7)) disease. We also confirmed three known loci associated with ER-negative (19p13) and both ER-negative and ER-positive breast cancer (6q25 and 12p11). Our results highlight the value of large-scale collaborative studies to identify novel breast cancer risk loci..
Varghese, J.S.
Thompson, D.J.
Michailidou, K.
Lindström, S.
Turnbull, C.
Brown, J.
Leyland, J.
Warren, R.M.
Luben, R.N.
Loos, R.J.
Wareham, N.J.
Rommens, J.
Paterson, A.D.
Martin, L.J.
Vachon, C.M.
Scott, C.G.
Atkinson, E.J.
Couch, F.J.
Apicella, C.
Southey, M.C.
Stone, J.
Li, J.
Eriksson, L.
Czene, K.
Boyd, N.F.
Hall, P.
Hopper, J.L.
Tamimi, R.M.
MODE Consortium,
Rahman, N.
Easton, D.F.
(2012). Mammographic breast density and breast cancer: evidence of a shared genetic basis. Cancer res,
Vol.72
(6),
pp. 1478-1484.
show abstract
Percent mammographic breast density (PMD) is a strong heritable risk factor for breast cancer. However, the pathways through which this risk is mediated are still unclear. To explore whether PMD and breast cancer have a shared genetic basis, we identified genetic variants most strongly associated with PMD in a published meta-analysis of five genome-wide association studies (GWAS) and used these to construct risk scores for 3,628 breast cancer cases and 5,190 controls from the UK2 GWAS of breast cancer. The signed per-allele effect estimates of single-nucleotide polymorphisms (SNP) were multiplied with the respective allele counts in the individual and summed over all SNPs to derive the risk score for an individual. These scores were included as the exposure variable in a logistic regression model with breast cancer case-control status as the outcome. This analysis was repeated using 10 different cutoff points for the most significant density SNPs (1%-10% representing 5,222-50,899 SNPs). Permutation analysis was also conducted across all 10 cutoff points. The association between risk score and breast cancer was significant for all cutoff points from 3% to 10% of top density SNPs, being most significant for the 6% (2-sided P = 0.002) to 10% (P = 0.001) cutoff points (overall permutation P = 0.003). Women in the top 10% of the risk score distribution had a 31% increased risk of breast cancer [OR = 1.31; 95% confidence interval (CI), 1.08-1.59] compared with women in the bottom 10%. Together, our results show that PMD and breast cancer have a shared genetic basis that is mediated through a large number of common variants..
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.
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.-.
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.-.
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.
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.
Cox, A.
Reed, M.W.
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.-.
Noh, D.-.
Shen, C.-.
Yu, J.-.
Wu, P.-.
Hsiung, C.-.
Perkins, A.
Swann, R.
Velentzis, L.
Eccles, D.M.
Tapper, W.J.
Gerty, S.M.
Graham, N.J.
Ponder, B.A.
Chenevix-Trench, G.
Pharoah, P.D.
Lathrop, M.
Dunning, A.M.
Rahman, N.
Peto, J.
Easton, D.F.
(2012). Genome-wide association analysis identifies three new breast cancer susceptibility loci. Nat genet,
Vol.44
(3),
pp. 312-318.
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..
Loveday, C.
Turnbull, C.
Ruark, E.
Xicola, R.M.
Ramsay, E.
Hughes, D.
Warren-Perry, M.
Snape, K.
Breast Cancer Susceptibility Collaboration (UK),
Eccles, D.
Evans, D.G.
Gore, M.
Renwick, A.
Seal, S.
Antoniou, A.C.
Rahman, N.
(2012). Germline RAD51C mutations confer susceptibility to ovarian cancer. Nat genet,
Vol.44
(5),
pp. 475-476.
Turnbull, C.
Perdeaux, E.R.
Pernet, D.
Naranjo, A.
Renwick, A.
Seal, S.
Munoz-Xicola, R.M.
Hanks, S.
Slade, I.
Zachariou, A.
Warren-Perry, M.
Ruark, E.
Gerrard, M.
Hale, J.
Hewitt, M.
Kohler, J.
Lane, S.
Levitt, G.
Madi, M.
Morland, B.
Neefjes, V.
Nicholson, J.
Picton, S.
Pizer, B.
Ronghe, M.
Stevens, M.
Traunecker, H.
Stiller, C.A.
Pritchard-Jones, K.
Dome, J.
Grundy, P.
Rahman, N.
(2012). A genome-wide association study identifies susceptibility loci for Wilms tumor. Nat genet,
Vol.44
(6),
pp. 681-684.
show abstract
full text
Wilms tumor is the most common renal malignancy of childhood. To identify common variants that confer susceptibility to Wilms tumor, we conducted a genome-wide association study in 757 individuals with Wilms tumor (cases) and 1,879 controls. We evaluated ten SNPs in regions significantly associated at P < 5 × 10(-5) in two independent replication series from the UK (769 cases and 2,814 controls) and the United States (719 cases and 1,037 controls). We identified clear significant associations at 2p24 (rs3755132, P = 1.03 × 10(-14); rs807624, P = 1.32 × 10(-14)) and 11q14 (rs790356, P = 4.25 × 10(-15)). Both regions contain genes that are plausibly related to Wilms tumorigenesis. We also identified candidate association signals at 5q14, 22q12 and Xp22..
Kirchhoff, T.
Gaudet, M.M.
Antoniou, A.C.
McGuffog, L.
Humphreys, M.K.
Dunning, A.M.
Bojesen, S.E.
Nordestgaard, B.G.
Flyger, H.
Kang, D.
Yoo, K.-.
Noh, D.-.
Ahn, S.-.
Dork, T.
Schürmann, P.
Karstens, J.H.
Hillemanns, P.
Couch, F.J.
Olson, J.
Vachon, C.
Wang, X.
Cox, A.
Brock, I.
Elliott, G.
Reed, M.W.
Burwinkel, B.
Meindl, A.
Brauch, H.
Hamann, U.
Ko, Y.-.
GENICA Network,
Broeks, A.
Schmidt, M.K.
Van 't Veer, L.J.
Braaf, L.M.
Johnson, N.
Fletcher, O.
Gibson, L.
Peto, J.
Turnbull, C.
Seal, S.
Renwick, A.
Rahman, N.
Wu, P.-.
Yu, J.-.
Hsiung, C.-.
Shen, C.-.
Southey, M.C.
Hopper, J.L.
Hammet, F.
Van Dorpe, T.
Dieudonne, A.-.
Hatse, S.
Lambrechts, D.
Andrulis, I.L.
Bogdanova, N.
Antonenkova, N.
Rogov, J.I.
Prokofieva, D.
Bermisheva, M.
Khusnutdinova, E.
van Asperen, C.J.
Tollenaar, R.A.
Hooning, M.J.
Devilee, P.
Margolin, S.
Lindblom, A.
Milne, R.L.
Arias, J.I.
Zamora, M.P.
Benítez, J.
Severi, G.
Baglietto, L.
Giles, G.G.
kConFab,
AOCS Study Group,
Spurdle, A.B.
Beesley, J.
Chen, X.
Holland, H.
Healey, S.
Wang-Gohrke, S.
Chang-Claude, J.
Mannermaa, A.
Kosma, V.-.
Kauppinen, J.
Kataja, V.
Agnarsson, B.A.
Caligo, M.A.
Godwin, A.K.
Nevanlinna, H.
Heikkinen, T.
Fredericksen, Z.
Lindor, N.
Nathanson, K.L.
Domchek, S.M.
SWE-BRCA,
Loman, N.
Karlsson, P.
Stenmark Askmalm, M.
Melin, B.
von Wachenfeldt, A.
HEBON,
Hogervorst, F.B.
Verheus, M.
Rookus, M.A.
Seynaeve, C.
Oldenburg, R.A.
Ligtenberg, M.J.
Ausems, M.G.
Aalfs, C.M.
Gille, H.J.
Wijnen, J.T.
Gómez García, E.B.
EMBRACE,
Peock, S.
Cook, M.
Oliver, C.T.
Frost, D.
Luccarini, C.
Pichert, G.
Davidson, R.
Chu, C.
Eccles, D.
Ong, K.-.
Cook, J.
Douglas, F.
Hodgson, S.
Evans, D.G.
Eeles, R.
Gold, B.
Pharoah, P.D.
Offit, K.
Chenevix-Trench, G.
Easton, D.F.
BCAC/CIMBA,
(2012). Breast cancer risk and 6q22 33: combined results from Breast Cancer Association Consortium and Consortium of Investigators on Modifiers of BRCA1/2. Plos one,
Vol.7
(6),
p. e35706.
show abstract
Recently, a locus on chromosome 6q22.33 (rs2180341) was reported to be associated with increased breast cancer risk in the Ashkenazi Jewish (AJ) population, and this association was also observed in populations of non-AJ European ancestry. In the present study, we performed a large replication analysis of rs2180341 using data from 31,428 invasive breast cancer cases and 34,700 controls collected from 25 studies in the Breast Cancer Association Consortium (BCAC). In addition, we evaluated whether rs2180341 modifies breast cancer risk in 3,361 BRCA1 and 2,020 BRCA2 carriers from 11 centers in the Consortium of Investigators of Modifiers of BRCA1/2 (CIMBA). Based on the BCAC data from women of European ancestry, we found evidence for a weak association with breast cancer risk for rs2180341 (per-allele odds ratio (OR) = 1.03, 95% CI 1.00-1.06, p = 0.023). There was evidence for heterogeneity in the ORs among studies (I(2) = 49.3%; p = <0.004). In CIMBA, we observed an inverse association with the minor allele of rs2180341 and breast cancer risk in BRCA1 mutation carriers (per-allele OR = 0.89, 95%CI 0.80-1.00, p = 0.048), indicating a potential protective effect of this allele. These data suggest that that 6q22.33 confers a weak effect on breast cancer risk..
Hein, R.
Maranian, M.
Hopper, J.L.
Kapuscinski, M.K.
Southey, M.C.
Park, D.J.
Schmidt, M.K.
Broeks, A.
Hogervorst, F.B.
Bueno-de-Mesquita, H.B.
Muir, K.R.
Lophatananon, A.
Rattanamongkongul, S.
Puttawibul, P.
Fasching, P.A.
Hein, A.
Ekici, A.B.
Beckmann, M.W.
Fletcher, O.
Johnson, N.
dos Santos Silva, I.
Peto, J.
Sawyer, E.
Tomlinson, I.
Kerin, M.
Miller, N.
Marmee, F.
Schneeweiss, A.
Sohn, C.
Burwinkel, B.
Guénel, P.
Cordina-Duverger, E.
Menegaux, F.
Truong, T.
Bojesen, S.E.
Nordestgaard, B.G.
Flyger, H.
Milne, R.L.
Perez, J.I.
Zamora, M.P.
Benítez, J.
Anton-Culver, H.
Ziogas, A.
Bernstein, L.
Clarke, C.A.
Brenner, H.
Müller, H.
Arndt, V.
Stegmaier, C.
Rahman, N.
Seal, S.
Turnbull, C.
Renwick, A.
Meindl, A.
Schott, S.
Bartram, C.R.
Schmutzler, R.K.
Brauch, H.
Hamann, U.
Ko, Y.-.
GENICA Network,
Wang-Gohrke, S.
Dörk, T.
Schürmann, P.
Karstens, J.H.
Hillemanns, P.
Nevanlinna, H.
Heikkinen, T.
Aittomäki, K.
Blomqvist, C.
Bogdanova, N.V.
Zalutsky, I.V.
Antonenkova, N.N.
Bermisheva, M.
Prokovieva, D.
Farahtdinova, A.
Khusnutdinova, E.
Lindblom, A.
Margolin, S.
Mannermaa, A.
Kataja, V.
Kosma, V.-.
Hartikainen, J.
Chen, X.
Beesley, J.
Kconfab Investigators,
AOCS Group,
Lambrechts, D.
Zhao, H.
Neven, P.
Wildiers, H.
Nickels, S.
Flesch-Janys, D.
Radice, P.
Peterlongo, P.
Manoukian, S.
Barile, M.
Couch, F.J.
Olson, J.E.
Wang, X.
Fredericksen, Z.
Giles, G.G.
Baglietto, L.
McLean, C.A.
Severi, G.
Offit, K.
Robson, M.
Gaudet, M.M.
Vijai, J.
Alnæs, G.G.
Kristensen, V.
Børresen-Dale, A.-.
John, E.M.
Miron, A.
Winqvist, R.
Pylkäs, K.
Jukkola-Vuorinen, A.
Grip, M.
Andrulis, I.L.
Knight, J.A.
Glendon, G.
Mulligan, A.M.
Figueroa, J.D.
García-Closas, M.
Lissowska, J.
Sherman, M.E.
Hooning, M.
Martens, J.W.
Seynaeve, C.
Collée, M.
Hall, P.
Humpreys, K.
Czene, K.
Liu, J.
Cox, A.
Brock, I.W.
Cross, S.S.
Reed, M.W.
Ahmed, S.
Ghoussaini, M.
Pharoah, P.D.
Kang, D.
Yoo, K.-.
Noh, D.-.
Jakubowska, A.
Jaworska, K.
Durda, K.
Złowocka, E.
Sangrajrang, S.
Gaborieau, V.
Brennan, P.
McKay, J.
Shen, C.-.
Yu, J.-.
Hsu, H.-.
Hou, M.-.
Orr, N.
Schoemaker, M.
Ashworth, A.
Swerdlow, A.
Trentham-Dietz, A.
Newcomb, P.A.
Titus, L.
Egan, K.M.
Chenevix-Trench, G.
Antoniou, A.C.
Humphreys, M.K.
Morrison, J.
Chang-Claude, J.
Easton, D.F.
Dunning, A.M.
(2012). Comparison of 6q25 breast cancer hits from Asian and European Genome Wide Association Studies in the Breast Cancer Association Consortium (BCAC). Plos one,
Vol.7
(8),
p. e42380.
show abstract
The 6q25.1 locus was first identified via a genome-wide association study (GWAS) in Chinese women and marked by single nucleotide polymorphism (SNP) rs2046210, approximately 180 Kb upstream of ESR1. There have been conflicting reports about the association of this locus with breast cancer in Europeans, and a GWAS in Europeans identified a different SNP, tagged here by rs12662670. We examined the associations of both SNPs in up to 61,689 cases and 58,822 controls from forty-four studies collaborating in the Breast Cancer Association Consortium, of which four studies were of Asian and 39 of European descent. Logistic regression was used to estimate odds ratios (OR) and 95% confidence intervals (CI). Case-only analyses were used to compare SNP effects in Estrogen Receptor positive (ER+) versus negative (ER-) tumours. Models including both SNPs were fitted to investigate whether the SNP effects were independent. Both SNPs are significantly associated with breast cancer risk in both ethnic groups. Per-allele ORs are higher in Asian than in European studies [rs2046210: OR (A/G) = 1.36 (95% CI 1.26-1.48), p = 7.6 × 10(-14) in Asians and 1.09 (95% CI 1.07-1.11), p = 6.8 × 10(-18) in Europeans. rs12662670: OR (G/T) = 1.29 (95% CI 1.19-1.41), p = 1.2 × 10(-9) in Asians and 1.12 (95% CI 1.08-1.17), p = 3.8 × 10(-9) in Europeans]. SNP rs2046210 is associated with a significantly greater risk of ER- than ER+ tumours in Europeans [OR (ER-) = 1.20 (95% CI 1.15-1.25), p = 1.8 × 10(-17) versus OR (ER+) = 1.07 (95% CI 1.04-1.1), p = 1.3 × 10(-7), p(heterogeneity) = 5.1 × 10(-6)]. In these Asian studies, by contrast, there is no clear evidence of a differential association by tumour receptor status. Each SNP is associated with risk after adjustment for the other SNP. These results suggest the presence of two variants at 6q25.1 each independently associated with breast cancer risk in Asians and in Europeans. Of these two, the one tagged by rs2046210 is associated with a greater risk of ER- tumours..
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.
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.-.
Soini, Y.
Kataja, V.
Lambrechts, D.
Yesilyurt, B.T.
Chrisiaens, M.-.
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.
van den Ouweland, A.M.
Collée, J.M.
Hall, P.
Czene, K.
Cox, A.
Brock, I.W.
Reed, M.W.
Cross, S.S.
Pharoah, P.
Dunning, A.M.
Kang, D.
Yoo, K.-.
Noh, D.-.
Ahn, S.-.
Jakubowska, A.
Lubinski, J.
Jaworska, K.
Durda, K.
Sangrajrang, S.
Gaborieau, V.
Brennan, P.
McKay, J.
Shen, C.-.
Ding, S.-.
Hsu, H.-.
Yu, J.-.
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. Hum mol genet,
Vol.20
(23),
pp. 4693-4706.
show abstract
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..
Slade, I.
Bacchelli, C.
Davies, H.
Murray, A.
Abbaszadeh, F.
Hanks, S.
Barfoot, R.
Burke, A.
Chisholm, J.
Hewitt, M.
Jenkinson, H.
King, D.
Morland, B.
Pizer, B.
Prescott, K.
Saggar, A.
Side, L.
Traunecker, H.
Vaidya, S.
Ward, P.
Futreal, P.A.
Vujanic, G.
Nicholson, A.G.
Sebire, N.
Turnbull, C.
Priest, J.R.
Pritchard-Jones, K.
Houlston, R.
Stiller, C.
Stratton, M.R.
Douglas, J.
Rahman, N.
(2011). DICER1 syndrome: clarifying the diagnosis, clinical features and management implications of a pleiotropic tumour predisposition syndrome. J med genet,
Vol.48
(4),
pp. 273-278.
show abstract
BACKGROUND: Constitutional DICER1 mutations were recently reported to cause familial pleuropulmonary blastoma (PPB). AIM: To investigate the contribution and phenotypic spectrum of constitutional and somatic DICER1 mutations to cancer. METHODS AND RESULTS: The authors sequenced DICER1 in constitutional DNA from 823 unrelated patients with a variety of tumours and in 781 cancer cell lines. Constitutional DICER1 mutations were identified in 19 families including 11/14 with PPB, 2/3 with cystic nephroma, 4/7 with ovarian Sertoli-Leydig-type tumours, 1/243 with Wilms tumour (this patient also had a Sertoli-Leydig tumour), 1/1 with intraocular medulloepithelioma (this patient also had PPB), 1/86 with medulloblastoma/infratentorial primitive neuroectodermal tumour, and 1/172 with germ cell tumour. The inheritance was investigated in 17 families. DICER1 mutations were identified in 25 relatives: 17 were unaffected, one mother had ovarian Sertoli-Leydig tumour, one half-sibling had cystic nephroma, and six relatives had non-toxic thyroid cysts/goitre. Analysis of eight tumours from DICER1 mutation-positive patients showed universal retention of the wild-type allele. DICER1 truncating mutations were identified in 4/781 cancer cell lines; all were in microsatellite unstable lines and therefore unlikely to be driver mutations. CONCLUSION: Constitutional DICER1 haploinsufficiency predisposes to a broad range of tumours, making a substantial contribution to PPB, cystic nephroma and ovarian Sertoli-Leydig tumours, but a smaller contribution to other tumours. Most mutation carriers are unaffected, indicating that tumour risk is modest. The authors define the clinical contexts in which DICER1 mutation testing should be considered, the associated tumour risks, and the implications for at-risk individuals. They have termed this condition 'DICER1 syndrome'. ACCESSION NUMBERS: The cDNA Genbank accession number for the DICER1 sequence reported in this paper is NM_030621.2..
Loveday, C.
Turnbull, C.
Ramsay, E.
Hughes, D.
Ruark, E.
Frankum, J.R.
Bowden, G.
Kalmyrzaev, B.
Warren-Perry, M.
Snape, K.
Adlard, J.W.
Barwell, J.
Berg, J.
Brady, A.F.
Brewer, C.
Brice, G.
Chapman, C.
Cook, J.
Davidson, R.
Donaldson, A.
Douglas, F.
Greenhalgh, L.
Henderson, A.
Izatt, L.
Kumar, A.
Lalloo, F.
Miedzybrodzka, Z.
Morrison, P.J.
Paterson, J.
Porteous, M.
Rogers, M.T.
Shanley, S.
Walker, L.
Breast Cancer Susceptibility Collaboration (UK),
Eccles, D.
Evans, D.G.
Renwick, A.
Seal, S.
Lord, C.J.
Ashworth, A.
Reis-Filho, J.S.
Antoniou, A.C.
Rahman, N.
(2011). Germline mutations in RAD51D confer susceptibility to ovarian cancer. Nat genet,
Vol.43
(9),
pp. 879-882.
show abstract
full text
Recently, RAD51C mutations were identified in families with breast and ovarian cancer. This observation prompted us to investigate the role of RAD51D in cancer susceptibility. We identified eight inactivating RAD51D mutations in unrelated individuals from 911 breast-ovarian cancer families compared with one inactivating mutation identified in 1,060 controls (P = 0.01). The association found here was principally with ovarian cancer, with three mutations identified in the 59 pedigrees with three or more individuals with ovarian cancer (P = 0.0005). The relative risk of ovarian cancer for RAD51D mutation carriers was estimated to be 6.30 (95% CI 2.86-13.85, P = 4.8 × 10(-6)). By contrast, we estimated the relative risk of breast cancer to be 1.32 (95% CI 0.59-2.96, P = 0.50). These data indicate that RAD51D mutation testing may have clinical utility in individuals with ovarian cancer and their families. Moreover, we show that cells deficient in RAD51D are sensitive to treatment with a PARP inhibitor, suggesting a possible therapeutic approach for cancers arising in RAD51D mutation carriers..
Turnbull, C.
Rahman, N.
(2011). Genome-wide association studies provide new insights into the genetic basis of testicular germ-cell tumour. Int j androl,
Vol.34
(4 Pt 2),
pp. e86-e96.
show abstract
Testicular germ-cell tumour (TGCT) is the most common cancer in young men, and genetic epidemiological studies suggest that the disease has a strong genetic basis. Until 2009, very little of this genetic component had been explained. Genome-wide association studies have since identified eight SNPs at six loci which together account for approximately 15% of the genetic risk of TGCT and offer novel biological insights into testicular germ-cell oncogenesis. In this review, we summarize the genetic epidemiology of TGCT, detail the contribution genome-wide association studies have made to our understanding of the genetic basis of TGCT and reflect on how future technological advances may assist in revealing the remaining genetic factors underlying TGCT susceptibility..
Milne, R.L.
Goode, E.L.
García-Closas, M.
Couch, F.J.
Severi, G.
Hein, R.
Fredericksen, Z.
Malats, N.
Zamora, M.P.
Arias Pérez, J.I.
Benítez, J.
Dörk, T.
Schürmann, P.
Karstens, J.H.
Hillemanns, P.
Cox, A.
Brock, I.W.
Elliot, G.
Cross, S.S.
Seal, S.
Turnbull, C.
Renwick, A.
Rahman, N.
Shen, C.-.
Yu, J.-.
Huang, C.-.
Hou, M.-.
Nordestgaard, B.G.
Bojesen, S.E.
Lanng, C.
Grenaker Alnæs, G.
Kristensen, V.
Børrensen-Dale, A.-.
Hopper, J.L.
Dite, G.S.
Apicella, C.
Southey, M.C.
Lambrechts, D.
Yesilyurt, B.T.
Floris, G.
Leunen, K.
Sangrajrang, S.
Gaborieau, V.
Brennan, P.
McKay, J.
Chang-Claude, J.
Wang-Gohrke, S.
Radice, P.
Peterlongo, P.
Manoukian, S.
Barile, M.
Giles, G.G.
Baglietto, L.
John, E.M.
Miron, A.
Chanock, S.J.
Lissowska, J.
Sherman, M.E.
Figueroa, J.D.
Bogdanova, N.V.
Antonenkova, N.N.
Zalutsky, I.V.
Rogov, Y.I.
Fasching, P.A.
Bayer, C.M.
Ekici, A.B.
Beckmann, M.W.
Brenner, H.
Müller, H.
Arndt, V.
Stegmaier, C.
Andrulis, I.L.
Knight, J.A.
Glendon, G.
Mulligan, A.M.
Mannermaa, A.
Kataja, V.
Kosma, V.-.
Hartikainen, J.M.
Meindl, A.
Heil, J.
Bartram, C.R.
Schmutzler, R.K.
Thomas, G.D.
Hoover, R.N.
Fletcher, O.
Gibson, L.J.
dos Santos Silva, I.
Peto, J.
Nickels, S.
Flesch-Janys, D.
Anton-Culver, H.
Ziogas, A.
Sawyer, E.
Tomlinson, I.
Kerin, M.
Miller, N.
Schmidt, M.K.
Broeks, A.
Van 't Veer, L.J.
Tollenaar, R.A.
Pharoah, P.D.
Dunning, A.M.
Pooley, K.A.
Marme, F.
Schneeweiss, A.
Sohn, C.
Burwinkel, B.
Jakubowska, A.
Lubinski, J.
Jaworska, K.
Durda, K.
Kang, D.
Yoo, K.-.
Noh, D.-.
Ahn, S.-.
Hunter, D.J.
Hankinson, S.E.
Kraft, P.
Lindstrom, S.
Chen, X.
Beesley, J.
Hamann, U.
Harth, V.
Justenhoven, C.
GENICA Network,
Winqvist, R.
Pylkäs, K.
Jukkola-Vuorinen, A.
Grip, M.
Hooning, M.
Hollestelle, A.
Oldenburg, R.A.
Tilanus-Linthorst, M.
Khusnutdinova, E.
Bermisheva, M.
Prokofieva, D.
Farahtdinova, A.
Olson, J.E.
Wang, X.
Humphreys, M.K.
Wang, Q.
Chenevix-Trench, G.
kConFab Investigators,
AOCS Group,
Easton, D.F.
(2011). Confirmation of 5p12 as a susceptibility locus for progesterone-receptor-positive, lower grade breast cancer. Cancer epidemiol biomarkers prev,
Vol.20
(10),
pp. 2222-2231.
show abstract
BACKGROUND: The single-nucleotide polymorphism (SNP) 5p12-rs10941679 has been found to be associated with risk of breast cancer, particularly estrogen receptor (ER)-positive disease. We aimed to further explore this association overall, and by tumor histopathology, in the Breast Cancer Association Consortium. METHODS: Data were combined from 37 studies, including 40,972 invasive cases, 1,398 cases of ductal carcinoma in situ (DCIS), and 46,334 controls, all of white European ancestry, as well as 3,007 invasive cases and 2,337 controls of Asian ancestry. Associations overall and by tumor invasiveness and histopathology were assessed using logistic regression. RESULTS: For white Europeans, the per-allele OR associated with 5p12-rs10941679 was 1.11 (95% CI = 1.08-1.14, P = 7 × 10(-18)) for invasive breast cancer and 1.10 (95% CI = 1.01-1.21, P = 0.03) for DCIS. For Asian women, the estimated OR for invasive disease was similar (OR = 1.07, 95%CI = 0.99-1.15, P = 0.09). Further analyses suggested that the association in white Europeans was largely limited to progesterone receptor (PR)-positive disease (per-allele OR = 1.16, 95% CI = 1.12-1.20, P = 1 × 10(-18) vs. OR = 1.03, 95% CI = 0.99-1.07, P = 0.2 for PR-negative disease; P(heterogeneity) = 2 × 10(-7)); heterogeneity by ER status was not observed (P = 0.2) once PR status was accounted for. The association was also stronger for lower grade tumors [per-allele OR (95% CI) = 1.20 (1.14-1.25), 1.13 (1.09-1.16), and 1.04 (0.99-1.08) for grade 1, 2, and 3/4, respectively; P(trend) = 5 × 10(-7)]. CONCLUSION: 5p12 is a breast cancer susceptibility locus for PR-positive, lower grade breast cancer. IMPACT: Multicenter fine-mapping studies of this region are needed as a first step to identifying the causal variant or variants..
Wellcome Trust Case Control Consortium,
Craddock, N.
Hurles, M.E.
Cardin, N.
Pearson, R.D.
Plagnol, V.
Robson, S.
Vukcevic, D.
Barnes, C.
Conrad, D.F.
Giannoulatou, E.
Holmes, C.
Marchini, J.L.
Stirrups, K.
Tobin, M.D.
Wain, L.V.
Yau, C.
Aerts, J.
Ahmad, T.
Andrews, T.D.
Arbury, H.
Attwood, A.
Auton, A.
Ball, S.G.
Balmforth, A.J.
Barrett, J.C.
Barroso, I.
Barton, A.
Bennett, A.J.
Bhaskar, S.
Blaszczyk, K.
Bowes, J.
Brand, O.J.
Braund, P.S.
Bredin, F.
Breen, G.
Brown, M.J.
Bruce, I.N.
Bull, J.
Burren, O.S.
Burton, J.
Byrnes, J.
Caesar, S.
Clee, C.M.
Coffey, A.J.
Connell, J.M.
Cooper, J.D.
Dominiczak, A.F.
Downes, K.
Drummond, H.E.
Dudakia, D.
Dunham, A.
Ebbs, B.
Eccles, D.
Edkins, S.
Edwards, C.
Elliot, A.
Emery, P.
Evans, D.M.
Evans, G.
Eyre, S.
Farmer, A.
Ferrier, I.N.
Feuk, L.
Fitzgerald, T.
Flynn, E.
Forbes, A.
Forty, L.
Franklyn, J.A.
Freathy, R.M.
Gibbs, P.
Gilbert, P.
Gokumen, O.
Gordon-Smith, K.
Gray, E.
Green, E.
Groves, C.J.
Grozeva, D.
Gwilliam, R.
Hall, A.
Hammond, N.
Hardy, M.
Harrison, P.
Hassanali, N.
Hebaishi, H.
Hines, S.
Hinks, A.
Hitman, G.A.
Hocking, L.
Howard, E.
Howard, P.
Howson, J.M.
Hughes, D.
Hunt, S.
Isaacs, J.D.
Jain, M.
Jewell, D.P.
Johnson, T.
Jolley, J.D.
Jones, I.R.
Jones, L.A.
Kirov, G.
Langford, C.F.
Lango-Allen, H.
Lathrop, G.M.
Lee, J.
Lee, K.L.
Lees, C.
Lewis, K.
Lindgren, C.M.
Maisuria-Armer, M.
Maller, J.
Mansfield, J.
Martin, P.
Massey, D.C.
McArdle, W.L.
McGuffin, P.
McLay, K.E.
Mentzer, A.
Mimmack, M.L.
Morgan, A.E.
Morris, A.P.
Mowat, C.
Myers, S.
Newman, W.
Nimmo, E.R.
O'Donovan, M.C.
Onipinla, A.
Onyiah, I.
Ovington, N.R.
Owen, M.J.
Palin, K.
Parnell, K.
Pernet, D.
Perry, J.R.
Phillips, A.
Pinto, D.
Prescott, N.J.
Prokopenko, I.
Quail, M.A.
Rafelt, S.
Rayner, N.W.
Redon, R.
Reid, D.M.
Renwick,
Ring, S.M.
Robertson, N.
Russell, E.
St Clair, D.
Sambrook, J.G.
Sanderson, J.D.
Schuilenburg, H.
Scott, C.E.
Scott, R.
Seal, S.
Shaw-Hawkins, S.
Shields, B.M.
Simmonds, M.J.
Smyth, D.J.
Somaskantharajah, E.
Spanova, K.
Steer, S.
Stephens, J.
Stevens, H.E.
Stone, M.A.
Su, Z.
Symmons, D.P.
Thompson, J.R.
Thomson, W.
Travers, M.E.
Turnbull, C.
Valsesia, A.
Walker, M.
Walker, N.M.
Wallace, C.
Warren-Perry, M.
Watkins, N.A.
Webster, J.
Weedon, M.N.
Wilson, A.G.
Woodburn, M.
Wordsworth, B.P.
Young, A.H.
Zeggini, E.
Carter, N.P.
Frayling, T.M.
Lee, C.
McVean, G.
Munroe, P.B.
Palotie, A.
Sawcer, S.J.
Scherer, S.W.
Strachan, D.P.
Tyler-Smith, C.
Brown, M.A.
Burton, P.R.
Caulfield, M.J.
Compston, A.
Farrall, M.
Gough, S.C.
Hall, A.S.
Hattersley, A.T.
Hill, A.V.
Mathew, C.G.
Pembrey, M.
Satsangi, J.
Stratton, M.R.
Worthington, J.
Deloukas, P.
Duncanson, A.
Kwiatkowski, D.P.
McCarthy, M.I.
Ouwehand, W.
Parkes, M.
Rahman, N.
Todd, J.A.
Samani, N.J.
Donnelly, P.
(2010). Genome-wide association study of CNVs in 16,000 cases of eight common diseases and 3,000 shared controls. Nature,
Vol.464
(7289),
pp. 713-720.
show abstract
Copy number variants (CNVs) account for a major proportion of human genetic polymorphism and have been predicted to have an important role in genetic susceptibility to common disease. To address this we undertook a large, direct genome-wide study of association between CNVs and eight common human diseases. Using a purpose-designed array we typed approximately 19,000 individuals into distinct copy-number classes at 3,432 polymorphic CNVs, including an estimated approximately 50% of all common CNVs larger than 500 base pairs. We identified several biological artefacts that lead to false-positive associations, including systematic CNV differences between DNAs derived from blood and cell lines. Association testing and follow-up replication analyses confirmed three loci where CNVs were associated with disease-IRGM for Crohn's disease, HLA for Crohn's disease, rheumatoid arthritis and type 1 diabetes, and TSPAN8 for type 2 diabetes-although in each case the locus had previously been identified in single nucleotide polymorphism (SNP)-based studies, reflecting our observation that most common CNVs that are well-typed on our array are well tagged by SNPs and so have been indirectly explored through SNP studies. We conclude that common CNVs that can be typed on existing platforms are unlikely to contribute greatly to the genetic basis of common human diseases..
Turnbull, C.
Ahmed, S.
Morrison, J.
Pernet, D.
Renwick, A.
Maranian, M.
Seal, S.
Ghoussaini, M.
Hines, S.
Healey, C.S.
Hughes, D.
Warren-Perry, M.
Tapper, W.
Eccles, D.
Evans, D.G.
Breast Cancer Susceptibility Collaboration (UK),
Hooning, M.
Schutte, M.
van den Ouweland, A.
Houlston, R.
Ross, G.
Langford, C.
Pharoah, P.D.
Stratton, M.R.
Dunning, A.M.
Rahman, N.
Easton, D.F.
(2010). Genome-wide association study identifies five new breast cancer susceptibility loci. Nat genet,
Vol.42
(6),
pp. 504-507.
show abstract
full text
Breast cancer is the most common cancer in women in developed countries. To identify common breast cancer susceptibility alleles, we conducted a genome-wide association study in which 582,886 SNPs were genotyped in 3,659 cases with a family history of the disease and 4,897 controls. Promising associations were evaluated in a second stage, comprising 12,576 cases and 12,223 controls. We identified five new susceptibility loci, on chromosomes 9, 10 and 11 (P = 4.6 x 10(-7) to P = 3.2 x 10(-15)). We also identified SNPs in the 6q25.1 (rs3757318, P = 2.9 x 10(-6)), 8q24 (rs1562430, P = 5.8 x 10(-7)) and LSP1 (rs909116, P = 7.3 x 10(-7)) regions that showed more significant association with risk than those reported previously. Previously identified breast cancer susceptibility loci were also found to show larger effect sizes in this study of familial breast cancer cases than in previous population-based studies, consistent with polygenic susceptibility to the disease..
Turnbull, C.
Rapley, E.A.
Seal, S.
Pernet, D.
Renwick, A.
Hughes, D.
Ricketts, M.
Linger, R.
Nsengimana, J.
Deloukas, P.
Huddart, R.A.
Bishop, D.T.
Easton, D.F.
Stratton, M.R.
Rahman, N.
UK Testicular Cancer Collaboration,
(2010). Variants near DMRT1, TERT and ATF7IP are associated with testicular germ cell cancer. Nat genet,
Vol.42
(7),
pp. 604-607.
show abstract
full text
We conducted a genome-wide association study for testicular germ cell tumor, genotyping 298,782 SNPs in 979 affected individuals and 4,947 controls from the UK and replicating associations in a further 664 cases and 3,456 controls. We identified three new susceptibility loci, two of which include genes that are involved in telomere regulation. We identified two independent signals within the TERT-CLPTM1L locus on chromosome 5, which has previously been associated with multiple other cancers (rs4635969, OR=1.54, P=1.14x10(-23); rs2736100, OR=1.33, P=7.55x10(-15)). We also identified a locus on chromosome 12 (rs2900333, OR=1.27, P=6.16x10(-10)) that contains ATF7IP, a regulator of TERT expression. Finally, we identified a locus on chromosome 9 (rs755383, OR=1.37, P=1.12x10(-23)), containing the sex determination gene DMRT1, which has been linked to teratoma susceptibility in mice..
Turnbull, C.
Hines, S.
Renwick, A.
Hughes, D.
Pernet, D.
Elliott, A.
Seal, S.
Warren-Perry, M.
Gareth Evans, D.
Eccles, D.
Breast Cancer Susceptibility Collaboration UK,
Stratton, M.R.
Rahman, N.
(2010). Mutation and association analysis of GEN1 in breast cancer susceptibility. Breast cancer res treat,
Vol.124
(1),
pp. 283-288.
show abstract
full text
GEN1 was recently identified as a key Holliday junction resolvase involved in homologous recombination. Somatic truncating GEN1 mutations have been reported in two breast cancers. Together these data led to the proposition that GEN1 is a breast cancer predisposition gene. In this article we have formally investigated this hypothesis. We performed full-gene mutational analysis of GEN1 in 176 BRCA1/2-negative familial breast cancer samples and 159 controls. We genotyped six SNPs tagging the 30 common variants in the transcribed region of GEN1 in 3,750 breast cancer cases and 4,907 controls. Mutation analysis revealed one truncating variant, c.2515_2519delAAGTT, which was present in 4% of cases and 4% of controls. We identified control individuals homozygous for the deletion, demonstrating that the last 69 amino acids of GEN1 are dispensable for its function. We identified 17 other variants, but their frequency did not significantly differ between cases and controls. Analysis of 3,750 breast cancer cases and 4,907 controls demonstrated no evidence of significant association with breast cancer for six SNPs tagging the 30 common GEN1 variants. These data indicate that although it also plays a key role in double-strand DNA break repair, GEN1 does not make an appreciable contribution to breast cancer susceptibility by acting as a high- or intermediate-penetrance breast cancer predisposition gene like BRCA1, BRCA2, CHEK2, ATM, BRIP1 and PALB2 and that common GEN1 variants do not act as low-penetrance susceptibility alleles analogous to SNPs in FGFR2. Furthermore, our analyses demonstrate the importance of undertaking appropriate genetic investigations, typically full gene screening in cases and controls together with large-scale case-control association analyses, to evaluate the contribution of genes to cancer susceptibility..
Rapley, E.A.
Turnbull, C.
Al Olama, A.A.
Dermitzakis, E.T.
Linger, R.
Huddart, R.A.
Renwick, A.
Hughes, D.
Hines, S.
Seal, S.
Morrison, J.
Nsengimana, J.
Deloukas, P.
UK Testicular Cancer Collaboration,
Rahman, N.
Bishop, D.T.
Easton, D.F.
Stratton, M.R.
(2009). A genome-wide association study of testicular germ cell tumor. Nat genet,
Vol.41
(7),
pp. 807-810.
show abstract
full text
We conducted a genome-wide association study for testicular germ cell tumor (TGCT), genotyping 307,666 SNPs in 730 cases and 1,435 controls from the UK and replicating associations in a further 571 cases and 1,806 controls. We found strong evidence for susceptibility loci on chromosome 5 (per allele OR = 1.37 (95% CI = 1.19-1.58), P = 3 x 10(-13)), chromosome 6 (OR = 1.50 (95% CI = 1.28-1.75), P = 10(-13)) and chromosome 12 (OR = 2.55 (95% CI = 2.05-3.19), P = 10(-31)). KITLG, encoding the ligand for the receptor tyrosine kinase KIT, which has previously been implicated in the pathogenesis of TGCT and the biology of germ cells, may explain the association on chromosome 12..
Turnbull, C.
Ruddy, D.
Barnicoat, A.
(2009). Nasopharyngeal teratoma and diaphragmatic hernia: no longer a random association but a new syndrome?. Clin dysmorphol,
Vol.18
(3),
pp. 131-134.
show abstract
We describe a female infant with a diaphragmatic hernia and nasopharyngeal teratoma. The case is compared with two previous reports of this combination of features. We suggest that this can no longer be considered a random association but instead represents the emergence of a distinct syndrome of which this case represents the third report..
Milne, R.L.
Benítez, J.
Nevanlinna, H.
Heikkinen, T.
Aittomäki, K.
Blomqvist, C.
Arias, J.I.
Zamora, M.P.
Burwinkel, B.
Bartram, C.R.
Meindl, A.
Schmutzler, R.K.
Cox, A.
Brock, I.
Elliott, G.
Reed, M.W.
Southey, M.C.
Smith, L.
Spurdle, A.B.
Hopper, J.L.
Couch, F.J.
Olson, J.E.
Wang, X.
Fredericksen, Z.
Schürmann, P.
Bremer, M.
Hillemanns, P.
Dörk, T.
Devilee, P.
van Asperen, C.J.
Tollenaar, R.A.
Seynaeve, C.
Hall, P.
Czene, K.
Liu, J.
Li, Y.
Ahmed, S.
Dunning, A.M.
Maranian, M.
Pharoah, P.D.
Chenevix-Trench, G.
Beesley, J.
kConFab Investigators,
AOCS Group,
Bogdanova, N.V.
Antonenkova, N.N.
Zalutsky, I.V.
Anton-Culver, H.
Ziogas, A.
Brauch, H.
Justenhoven, C.
Ko, Y.-.
Haas, S.
Fasching, P.A.
Strick, R.
Ekici, A.B.
Beckmann, M.W.
Giles, G.G.
Severi, G.
Baglietto, L.
English, D.R.
Fletcher, O.
Johnson, N.
dos Santos Silva, I.
Peto, J.
Turnbull, C.
Hines, S.
Renwick, A.
Rahman, N.
Nordestgaard, B.G.
Bojesen, S.E.
Flyger, H.
Kang, D.
Yoo, K.-.
Noh, D.-.
Mannermaa, A.
Kataja, V.
Kosma, V.-.
García-Closas, M.
Chanock, S.
Lissowska, J.
Brinton, L.A.
Chang-Claude, J.
Wang-Gohrke, S.
Shen, C.-.
Wang, H.-.
Yu, J.-.
Chen, S.-.
Bermisheva, M.
Nikolaeva, T.
Khusnutdinova, E.
Humphreys, M.K.
Morrison, J.
Platte, R.
Easton, D.F.
Breast Cancer Association Consortium,
(2009). Risk of estrogen receptor-positive and -negative breast cancer and single-nucleotide polymorphism 2q35-rs13387042. J natl cancer inst,
Vol.101
(14),
pp. 1012-1018.
show abstract
BACKGROUND: A recent genome-wide association study identified single-nucleotide polymorphism (SNP) 2q35-rs13387042 as a marker of susceptibility to estrogen receptor (ER)-positive breast cancer. We attempted to confirm this association using the Breast Cancer Association Consortium. METHODS: 2q35-rs13387042 SNP was genotyped for 31 510 women with invasive breast cancer, 1101 women with ductal carcinoma in situ, and 35 969 female control subjects from 25 studies. Odds ratios (ORs) were estimated by logistic regression, adjusted for study. Heterogeneity in odds ratios by each of age, ethnicity, and study was assessed by fitting interaction terms. Heterogeneity by each of invasiveness, family history, bilaterality, and hormone receptor status was assessed by subclassifying case patients and applying polytomous logistic regression. All statistical tests were two-sided. RESULTS: We found strong evidence of association between rs13387042 and breast cancer in white women of European origin (per-allele OR = 1.12, 95% confidence interval [CI] = 1.09 to 1.15; P(trend) = 1.0 x 10(-19)). The odds ratio was lower than that previously reported (P = .02) and did not vary by age or ethnicity (all P > or = .2). However, it was higher when the analysis was restricted to case patients who were selected for a strong family history (P = .02). An association was observed for both ER-positive (OR = 1.14, 95% CI = 1.10 to 1.17; P = 10(-15)) and ER-negative disease (OR = 1.10, 95% CI = 1.04 to 1.15; P = .0003) and both progesterone receptor (PR)-positive (OR = 1.15, 95% CI = 1.11 to 1.19; P = 5 x 10(-14)) and PR-negative disease (OR = 1.10, 95% CI = 1.06 to 1.15; P = .00002). CONCLUSION: The rs13387042 is associated with both ER-positive and ER-negative breast cancer in European women..
Song, H.
Koessler, T.
Ahmed, S.
Ramus, S.J.
Kjaer, S.K.
Dicioccio, R.A.
Wozniak, E.
Hogdall, E.
Whittemore, A.S.
McGuire, V.
Ponder, B.A.
Turnbull, C.
Hines, S.
Rahman, N.
Breast Cancer Susceptibility Collaboration UK,
Eeles, R.A.
Easton, D.F.
Gayther, S.A.
Dunning, A.M.
Pharoah, P.D.
(2008). Association study of prostate cancer susceptibility variants with risks of invasive ovarian, breast, and colorectal cancer. Cancer res,
Vol.68
(21),
pp. 8837-8842.
show abstract
Several prostate cancer susceptibility loci have recently been identified by genome-wide association studies. These loci are candidates for susceptibility to other epithelial cancers. The aim of this study was to test these tag single nucleotide polymorphisms (SNP) for association with invasive ovarian, colorectal, and breast cancer. Twelve prostate cancer-associated tag SNPs were genotyped in ovarian (2,087 cases/3,491 controls), colorectal (2,148 cases/2,265 controls) and breast (first set, 4,339 cases/4,552 controls; second set, 3,800 cases/3,995 controls) case-control studies. The primary test of association was a comparison of genotype frequencies between cases and controls, and a test for trend stratified by study where appropriate. Genotype-specific odds ratios (OR) were estimated by logistic regression. SNP rs2660753 (chromosome 3p12) showed evidence of association with ovarian cancer [per minor allele OR, 1.19; 95% confidence interval (95% CI), 1.04-1.37; P(trend) = 0.012]. This association was stronger for the serous histologic subtype (OR, 1.29; 95% CI, 1.09-1.53; P = 0.003). SNP rs7931342 (chromosome 11q13) showed some evidence of association with breast cancer (per minor allele OR, 0.95; 95% CI, 0.91-0.99; P(trend) = 0.028). This association was somewhat stronger for estrogen receptor-positive tumors (OR, 0.92; 95% CI, 0.87-0.98; P = 0.011). None of these tag SNPs were associated with risk of colorectal cancer. In conclusion, loci associated with risk of prostate cancer may also be associated with ovarian and breast cancer susceptibility. However, the effects are modest and warrant replication in larger studies..
Turnbull, C.
Rahman, N.
(2008). Genetic predisposition to breast cancer: past, present, and future. Annu rev genomics hum genet,
Vol.9,
pp. 321-345.
show abstract
In recent years, our understanding of genetic predisposition to breast cancer has advanced significantly. Three classes of predisposition factors, categorized by their associated risks of breast cancer, are currently known. BRCA1 and BRCA2 are high-penetrance breast cancer predisposition genes identified by genome-wide linkage analysis and positional cloning. Mutational screening of genes functionally related to BRCA1 and/or BRCA2 has revealed four genes, CHEK2, ATM, BRIP1, and PALB2; mutations in these genes are rare and confer an intermediate risk of breast cancer. Association studies have further identified eight common variants associated with low-penetrance breast cancer predisposition. Despite these discoveries, most of the familial risk of breast cancer remains unexplained. In this review, we describe the known genetic predisposition factors, expound on the methods by which they were identified, and consider how further technological and intellectual advances may assist in identifying the remaining genetic factors underlying breast cancer susceptibility..
David, A.L.
Turnbull, C.
Scott, R.
Freeman, J.
Bilardo, C.M.
van Maarle, M.
Chitty, L.S.
(2007). Diagnosis of Apert syndrome in the second-trimester using 2D and 3D ultrasound. Prenat diagn,
Vol.27
(7),
pp. 629-632.
show abstract
OBJECTIVES: To illustrate how Apert syndrome, a rare autosomal dominant genetic syndrome, can be detected in the second-trimester of pregnancy using 2D ultrasound, and how 3D ultrasound examination may provide parents with a better understanding of the structural defects affecting their baby. METHODS: Fetal Medicine Unit database searches to identify Apert syndrome cases. RESULTS: Five cases of Apert syndrome were suspected in the second-trimester when sonography showed abnormal extremities including syndactyly, and an abnormal skull shape. In 1 case there was increased nuchal translucency with a normal fetal karyotype in the first-trimester. In all cases, a mutation of the FGFR2 gene confirmed the diagnosis of Apert syndrome. 3D ultrasound was used to show parents the extent of the abnormalities of the skull, face and extremities. Parents were counseled by craniofacial surgeons and geneticists. CONCLUSION: Apert syndrome can be accurately suspected in the second-trimester by careful ultrasound examination of the fetus including the extremities and skull shape. 3D ultrasound can be a useful adjunct to 2D examination for parental counseling..
Siega-Riz, A.M.
Hartzema, A.G.
Turnbull, C.
Thorp, J.
McDonald, T.
Cogswell, M.E.
(2006). The effects of prophylactic iron given in prenatal supplements on iron status and birth outcomes: a randomized controlled trial. Am j obstet gynecol,
Vol.194
(2),
pp. 512-519.
show abstract
OBJECTIVE: The hypothesis that daily use of a prenatal supplement with iron from enrollment to third trimester to initially iron-replete, nonanemic pregnant women would reduce third-trimester anemia and improve birth outcomes was tested. STUDY DESIGN: Eight hundred sixty-seven women in Raleigh, North Carolina, who were at < 20 weeks of gestation were enrolled; 429 of these women had hemoglobin levels of > or = 110 g/L and ferritin levels of > or = 40 microg/L and were assigned randomly to receive prenatal supplements with 30 mg of iron as ferrous sulfate (n = 218 women) or 0 mg of iron (n = 211 women) until 26 to 29 weeks of gestation. Intent-to-treat analysis was used for the outcomes of third-trimester iron status, birth weight, preterm birth, and small-for-gestational age. RESULTS: Mean birth weight was higher by 108 g (P = .03), and the incidence of preterm delivery was lower (8% vs 14%; P = .05) in the 30-mg group compared with the control group, respectively. Iron supplementation did not affect the prevalence of small-for-gestational age infants or third-trimester iron status. CONCLUSION: Prophylactic iron supplementation that is begun early in pregnancy among low income women in the United States may have benefits beyond the reduction of iron deficiency anemia during pregnancy..
Turnbull, C.
Mirugaesu, N.
Eeles, R.
(2006). Radiotherapy and genetic predisposition to breast cancer. Clin oncol (r coll radiol),
Vol.18
(3),
pp. 257-267.
show abstract
Cancer genetics is becoming increasingly integrated into oncological care; particularly in breast cancer management. The recognition of monogenic breast cancer predisposition syndromes, such as BRCA, is critical as there is also a risk of other cancers in addition to a markedly elevated risk of contralateral breast cancer. In individuals with breast cancer due to some predisposition genes, radiation should be avoided, e.g. the ATM and TP53 genes, but there is still controversy as to whether radiation should be used in BRCA1/2 mutation carriers and more follow-up is needed. There are some radiation-sensitive genetic conditions where, if breast cancer occurs (as it may occur in any individual, not necessarily due to an increased risk associated with the condition), radiation should be avoided. These conditions are often associated with dysmorphic features. If such features are noticed then the advice of a geneticist should be sought urgently prior to giving radiation..
Turnbull, C.
Lees, M.
Chitty, L.S.
(2006). Prenatal sonographic diagnosis of Malpuech syndrome. Prenat diagn,
Vol.26
(12),
pp. 1121-1123.
show abstract
Malpuech syndrome (MS) is a rare autosomal recessive syndrome featuring pre- and post-natal growth deficiency, mental retardation, facial dysmorphism, cleft lip and palate (typically midline or bilateral), caudal appendage, renal malformations and male genital abnormalities. A prenatal diagnosis of MS was made in this fetus based on the family history and a combination of conventional and 3D prenatal ultrasound findings. The family were consanguineous with an affected first child. Prenatal ultrasound in the second pregnancy demonstrated bilateral cleft lip and palate in association with intrauterine growth retardation on serial prenatal ultrasound scans. Dysmorphic facial features and a small penis consistent with the diagnosis were confirmed on 3D scanning. Post-natal examination of the neonate confirmed the diagnosis of MS. To the best of our knowledge, this is the first prenatal diagnosis of this syndrome..
Ousey, J.C.
Rossdale, P.D.
Fowden, A.L.
Palmer, L.
Turnbull, C.
Allen, W.R.
(2005). Effects of manipulating intrauterine growth on post natal adrenocortical development and other parameters of maturity in neonatal foals (vol 36, pg 616, 2004). Equine veterinary journal,
Vol.37
(1),
pp. 36-1.
Spratt, J.C.
Leslie, S.J.
White, A.
Fenn, L.
Turnbull, C.
Northridge, D.B.
(2004). Harmonic imaging improves estimation of left ventricular mass. Int j cardiovasc imaging,
Vol.20
(2),
pp. 107-111.
show abstract
OBJECTIVES: To assess the effect of tissue harmonic imaging (THI) on assessment of left ventricular mass index (LVMI) measurements by M-mode trans-thoracic echocardiography, when compared with magnetic resonance imaging (MRI). METHODS: 20 hypertensive male subjects were studied. LVMI was measured in all subjects by both gradient-echo MRI (Lscelsint Prestige 1.9 T) and by transthoracic echocardiography (ATL HDI 5000). M-mode echocardiography recordings were taken for each patient, two with fundamental imaging (FI) and two using THI in a randomised order and the images unlabelled. Recordings were analysed off-line, by a blinded observer. LVMI by MRI was calculated using Simpson's rule on serial short axis slices of 8 mm thickness. Data are expressed as mean +/- SD. RESULTS: There was a difference in LVMI measurements between FI and THI (LVMI) (79 +/- 20 vs. 93 +/- 25 g2; p < 0.001). A lower mean difference was obtained by THI, compared to FI, when compared with MRI (2 +/- 15 vs. -32 +/- 22 g2; p < 0.001) suggesting that FI underestimates LVMI. Inter-observer variability was similar between THI and FI (4.5 +/- 15 vs. 6.4 +/- 15 g2; p = 0.46). CONCLUSION: In hypertensive males, M-mode echo derived from FI underestimated LVMI. These results imply that widely accepted reference ranges for LVMI using FI are not applicable when THI is used..
Ousey, J.C.
Rossdale, P.D.
Fowden, A.L.
Palmer, L.
Turnbull, C.
Allen, W.R.
(2004). Effects of manipulating intrauterine growth on post natal adrenocortical development and other parameters of maturity in neonatal foals. Equine vet j,
Vol.36
(7),
pp. 616-621.
show abstract
REASONS FOR PERFORMING STUDY: Intrauterine growth retardation (IUGR) impairs post natal adaptive responses and is associated with increased adrenocortical activity in many species. OBJECTIVES: To determine whether a restricted or enhanced intrauterine environment affects neonatal adaptation and adrenocortical function in horses. METHODS: Embryos from large (577 kg) Thoroughbred (TB) mares were transferred to smaller (343 kg) pony (P) mares and vice versa, to create a restricted (TB-in-P, n = 11) or enhanced (P-in-TB, n = 8) intrauterine environment. Control groups (TB-in-TB, n = 8; P-in-P, n = 7) were also included. RESULTS: Thirty foals were born live at full term (range 314-348 days) and 4 (3 TB-in-P, 1 P-in-TB) were stillborn between 275 and 335 days. TB-in-P foals were significantly (P<0.05) lighter than TB-in-TB, but heavier than P-in-P foals. TB-in-P foals took longer to first stand and suck and some had fetlock hyperextension and low (<4 g/l) plasma immunoglobulin G concentrations. Other foal groups showed normal behavioural responses. Haematological parameters were normal in all 4 groups of foals. Plasma ACTH levels were high at birth and plasma cortisol concentrations increased after delivery and returned to baseline within 6 h post partum in all but the TB-in-P foals, which had elevated levels until 48 h post partum. Plasma cortisol concentrations increased in all groups following exogenous ACTH administered on Days 1 and 5 postpartum. CONCLUSIONS: The TB-in-P foals showed IUGR and impaired post natal adaptive responses with basal hypercortisolaemia. POTENTIAL RELEVANCE: Foals born following IUGR may require clinical assistance in the early post natal period, but appear mature with respect to adrenocortical function..
Sorensen, J.S.
Turnbull, C.A.
Dearing, M.D.
(2004). A specialist herbivore (Neotoma stephensi) absorbs fewer plant toxins than does a generalist (Neotoma albigula). Physiol biochem zool,
Vol.77
(1),
pp. 139-148.
show abstract
Detoxification capacity of enzymes in the liver is thought to be the primary factor governing dietary toxin intake by mammalian herbivores. Recently, toxin absorption in the gut was proposed as an alternative process that also influences toxin intake. We examined the role of the gut in regulating toxin absorption by quantifying excretion of a plant secondary compound in the feces. We hypothesized that specialists have a greater capacity to reduce intestinal absorption of toxins than do generalists. To test this hypothesis, we compared fecal excretion of alpha-pinene in specialist (Neotoma stephensi) and generalist (Neotoma albigula) woodrats. Alpha-pinene is the most abundant monoterpene in Juniperus monosperma, which occurs in the natural diet of both woodrat species. Woodrats were fed alpha-pinene in diets containing juniper foliage for 3 wk and, in a separate experiment, were given a single oral dose of alpha-pinene. Feces were collected from animals at the end of each experiment and analyzed for alpha-pinene concentration using gas chromatography. Both woodrat species excreted unchanged alpha-pinene in the feces. However, specialist woodrats excreted 40% more alpha-pinene per unit ingested from a juniper diet and excreted nearly four times a greater percentage of an oral dose of alpha-pinene compared with generalists..
Spears, I.R.
Cummins, N.K.
Brenchley, Z.
Donohue, C.
Turnbull, C.
Burton, S.
Macho, G.A.
(2003). The effect of saddle design on stresses in the perineum during cycling. Med sci sports exerc,
Vol.35
(9),
pp. 1620-1625.
show abstract
PURPOSE: Repetitive internal stress in the perineum has been associated with soft-tissue trauma in bicyclists. Using an engineering approach, the purpose of this study was to quantify the amount of compression exerted in the perineum for a range of saddle widths and orientations. METHODS: Computer tomography was used to create a three-dimensional voxel-based finite element model of the right side of the male perineum-pelvis. For the creation of the saddle model, a commercially available saddle was digitized and the surface manipulated to represent a variety of saddle widths and orientations. The two models were merged, and a static downward load of 189 N was applied to the model at the region representing the sacroiliac joint. For validation purposes, external stresses along the perineum-saddle interface were compared with the results of pressure sensitive film. Good agreement was found for these external stresses. The saddles were then stretched and rotated, and the magnitude and location of maximum stresses within the perineum were both recorded. In all cases, the model of the pelvis-perineum was held in an upright position. RESULTS: Stresses within the perineum were reduced when the saddle was sufficiently wide to support both ischial tuberosities. This supporting mechanism was best achieved when the saddle was at least two times wider than the bi-ischial width of the cyclist. Stresses in the anterior of the perineum were reduced when the saddle was tilted downward, whereas stresses in the posterior were reduced when the saddle was tilted upward. CONCLUSIONS: Recommendations that saddles should be sufficiently wide to support the ischial tuberosities appear to be well founded. Recommendations that saddles be tilted downward (i.e., nose down) are supported by the model, but with caution, given the limitations of the model..
Allen, W.R.
Wilsher, S.
Turnbull, C.
Stewart, F.
Ousey, J.
Rossdale, P.D.
Fowden, A.L.
(2002). Influence of maternal size on placental, fetal and postnatal growth in the horse I Development in utero. Reproduction,
Vol.123
(3),
pp. 445-453.
show abstract
The interacting influences of maternal size and fetal genotype on placental and fetal development in the mare were assessed by comparing conventional within-breed Thoroughbred (Tb-in-Tb, n = 7) and Pony (P-in-P, n = 7) control pregnancies established by artificial insemination (AI) with between-breed (Tb-in-P, n = 8; deprived in utero condition and P-in-Tb, n = 7; luxurious in utero condition) experimental pregnancies established by embryo transfer. All foals were born spontaneously and the mean (+/- SEM) duration of gestation in the two groups of control mares was significantly different (P < 0.001) at 325 +/- 3.0 days for the P-in-P pregnancies and 339 +/- 3.0 days for the Tb-in-Tb pregnancies, whereas the durations of gestation for the two experimental groups were very similar and midway between those of the control pregnancies at 332 +/- 2.8 days for the Tb-in-P and 331 +/- 2.7 days for the P-in-Tb. Mean (+/- SEM) foal birth weight and the mean (+/- SEM) values for the mass, gross area and volume of the allantochorion were all highest in the seven Tb-in-Tb pregnancies (53.1 +/- 2.6 kg, 3.8 +/- 0.3 kg, 12.9 +/- 0.3 x 10(3) cm(2), 3.5 +/- 0.2 l, respectively) and lowest in the seven P-in-P control pregnancies (24.0 +/- 1.3 kg, 1.7 +/- 0.1 kg, 8.3 +/- 0.3 x 10(3) cm(2), 1.8 +/- 0.1 l, respectively). These parameters were higher in the seven P-in-Tb pregnancies (37.9 +/- 2.1 kg, 2.7 +/- 0.1 kg, 10.1 +/- 0.5 x 10(3) cm(2), 2.5 +/- 0.1 l, respectively) than in the eight Tb-in-P (33.0 +/- 2.4 kg, 2.3 +/- 0.2 kg, 9.0 +/- 0.5 x 10(3) cm(2), 2.1 +/- 0.1 l) experimental pregnancies. Foal birth weight was positively correlated with the mass (r = 0.84, P < 0.001), gross area (r = 0.87, P < 0.001) and volume (r = 0.91, P < 0.001) of the allantochorion, and maternal weight was also positively correlated with both the mass and gross area of the allantochorion (r = 0.64 and 0.69, respectively; both P < 0.001). Application of stereology to multiple random biopsies recovered from each placenta produced mean values for the surface density of microcotyledons on the allantochorion (S(v)). Values were higher in Thoroughbred than in Pony mares regardless of the breed of fetus being carried. Multiplication of S(v) by the volume of the allantochorion to give the total microscopic area of fetomaternal contact at the placental interface was also positively correlated with foal birth weight (r = 0.84, P < 0.001). Foal birth weight was determined by the microscopic area of fetomaternal contact of the placenta and there were no differences in foal weight per m(2) of placenta regardless of fetal or maternal genomes. Thus, the results indicate that in equids, maternal size interacts with both the maternal and fetal genotypes to control the rate and extent of fetal growth by influencing the gross area of the diffuse allantochorion, and the density, complexity and depth of the microcotyledons on its surface..
TURNBULL, C.
(1980). The Human Potential for Inhumanity. The sciences,
Vol.20
(3),
pp. 18-21.