Skip to main content Accessibility help
×
Hostname: page-component-76fb5796d-9pm4c Total loading time: 0 Render date: 2024-04-25T08:33:04.523Z Has data issue: false hasContentIssue false

2 - Genome-wide association studies of cancer predisposition

from Part 1.1 - Analytical techniques: analysis of DNA

Published online by Cambridge University Press:  05 February 2015

Zsofia K. Stadler
Affiliation:
Department of Medicine, Clinical Genetics Service, Memorial Sloan-Kettering Cancer Center, New York, NY, USA
Sohela Shah
Affiliation:
Department of Medicine, Clinical Genetics Service, Memorial Sloan-Kettering Cancer Center, New York, NY, USA
Kenneth Offit
Affiliation:
Department of Medicine, Clinical Genetics Service, Memorial Sloan-Kettering Cancer Center, New York, NY, USA
Edward P. Gelmann
Affiliation:
Columbia University, New York
Charles L. Sawyers
Affiliation:
Memorial Sloan-Kettering Cancer Center, New York
Frank J. Rauscher, III
Affiliation:
The Wistar Institute Cancer Centre, Philadelphia
Get access

Summary

Introduction

Until relatively recently the field of medical genetics has focused on the identification and treatment of rare, single-gene disorders that are usually associated with a high risk of a particular disease or trait. However, such high-penetrance susceptibility loci only account for a fraction of the familial aggregation of common diseases, and the genetic architecture of complex diseases, such as cancer, is probably better characterized by “polygenic” inheritance, wherein heritability is determined by the joint action of multiple genes. Since 2007, genome-wide association studies (GWAS) have resulted in a paradigm shift in the discovery of gene–disease associations and helped to identify multiple low-penetrance germline genetic variants for most common cancer types (1,2). Through a hypothesis-neutral genome-based approach, GWAS compare frequencies of common DNA variations, usually in the form of single-nucleotide polymorphisms (SNPs), in a large set of unrelated cases and controls to identify genetic variants associated with disease risk. GWAS have resulted in the mapping of susceptibility loci for many human diseases, including the identification of over 100 loci for cancer, most of which were not previously implicated in carcinogenesis.

This unprecedented rapid amassing of new genetic risk variants associated with cancer risk has generated hope that these germline markers may prove useful for cancer prevention, improve our understanding of cancer pathogenesis, and possibly lead to a new era of personalized genomics (3). However, as the majority of genetic variants confer only a modest risk of disease with effect size under 1.5 (Figure 2.1) and the biological mechanisms underpinning most associations are unknown, significant scientific barriers must be overcome before GWAS results can be meaningfully translated into patient care.

Type
Chapter
Information
Molecular Oncology
Causes of Cancer and Targets for Treatment
, pp. 10 - 20
Publisher: Cambridge University Press
Print publication year: 2013

Access options

Get access to the full version of this content by using one of the access options below. (Log in options will check for institutional or personal access. Content may require purchase if you do not have access.)

References

Stadler, ZK, Thom, P, Robson, ME, et al. Genome-wide association studies of cancer. Journal of Clinical Oncology 2010;28:4255–67.CrossRefGoogle Scholar
Stadler, ZK, Vijai, J, Thom, P, et al. Genome-wide association studies of cancer predisposition. Hematology/Oncology Clinics of North America 2010;24:973–96.CrossRef
Offit, K. Personalized medicine: new genomics, old lessons. Human Genetics 2011;130:3–14.CrossRef
Lichtenstein, P, Holm, NV, Verkasalo, PK, et al. Environmental and heritable factors in the causation of cancer-analyses of cohorts of twins from Sweden, Denmark, and Finland. New England Journal of Medicine. 2000;343:78–85.CrossRefGoogle Scholar
Goldgar, DE, Easton, DF, Cannon-Albright, LA, Skolnick, MH. Systematic population-based assessment of cancer risk in first-degree relatives of cancer probands. Journal of the National Cancer Institute 1994;86:1600–8.CrossRefGoogle ScholarPubMed
Athma, P, Rappaport, R, Swift, M. Molecular genotyping shows that ataxia-telangiectasia heterozygotes are predisposed to breast cancer. Cancer Genetics and Cytogenetics. 1996;92:130–4.CrossRef
Bernstein, JL, Bernstein, L, Thompson, WD, et al. ATM variants 7271T<G and IVS10–6T<G among women with unilateral and bilateral breast cancer. British Journal of Cancer 2003;89:1513–16.CrossRefGoogle ScholarPubMed
Bretsky, P, Haiman, CA, Gilad, S, et al. The relationship between twenty missense ATM variants and breast cancer risk: the Multiethnic Cohort. Cancer Epidemiology, Biomarkers and Prevention 2003;12:733–8.
The CHEK2 Breast Cancer Case-Control Consortium CHEK2*1100delC and susceptibility to breast cancer: a collaborative analysis involving 10,860 breast cancer cases and 9,065 controls from 10 studies. American Journal of Human Genetics 2004;74:1175–82.CrossRef
Meijers-Heijboer, H, van den Ouweland, A, Klijn, J, et al. Low-penetrance susceptibility to breast cancer due to CHEK2(*)1100delC in noncarriers of BRCA1 or BRCA2 mutations. Nature Genetics 2002;31:55–9.CrossRef
Thompson, D, Seal, S, Schutte, M, et al. A multicenter study of cancer incidence in CHEK2 1100delC mutation carriers. Cancer Epidemiology, Biomarkers and Prevention 2006;15:2542–5.CrossRef
Seal, S, Thompson, D, Renwick, A, et al. Truncating mutations in the Fanconi anemia J gene BRIP1 are low-penetrance breast cancer susceptibility alleles. Nature Genetics 2006;38:1239–41.CrossRef
Rahman, N, Seal, S, Thompson, D, et al. PALB2, which encodes a BRCA2-interacting protein, is a breast cancer susceptibility gene. Nature Genetics 2007;39:165–7.CrossRef
Easton, DF. How many more breast cancer predisposition genes are there? Breast Cancer Research 1999;1:14–17.
Lander, ES. The new genomics: global views of biology. Science 1996;274:536–9.CrossRef
Xiong, M, Guo, SW. Fine-scale genetic mapping based on linkage disequilibrium: theory and applications. American Journal of Human Genetics 1997;60:1513–31.CrossRefGoogle ScholarPubMed
Frazer, KA, Ballinger, DG, Cox, DR, et al. A second generation human haplotype map of over 3.1 million SNPs. Nature. 2007;449:851–61.
Kruglyak, L. The road to genome-wide association studies. Nature Reviews Genetics 2008;9:314–18.CrossRef
Satagopan, JM, Elston, RC. Optimal two-stage genotyping in population-based association studies. Genetic Epidemiology 2003;25:149–57.CrossRef
Wang, WY, Barratt, BJ, Clayton, DG, Todd, JA. Genome-wide association studies: theoretical and practical concerns. Nature Reviews Genetics 2005;6:109–18.CrossRef
Garcia-Closas, M, Chanock, S. Genetic susceptibility loci for breast cancer by estrogen receptor status. Clinical Cancer Research 2008;14:8000–9.CrossRef
Ardlie, KG, Lunetta, KL, Seielstad, M. Testing for population subdivision and association in four case-control studies. American Journal of Human Genetics 2002;71:304–11.CrossRefGoogle ScholarPubMed
Wacholder, S, Rothman, N, Caporaso, N. Counterpoint: bias from population stratification is not a major threat to the validity of conclusions from epidemiological studies of common polymorphisms and cancer. Cancer Epidemiology, Biomarkers and Prevention 2002;11:513–20.
Marchini, J, Cardon, LR, Phillips, MS, Donnelly, P. The effects of human population structure on large genetic association studies. Nature Genetics 2004;36:512–17.CrossRef
Freedman, ML, Reich, D, Penney, KL, et al. Assessing the impact of population stratification on genetic association studies. Nature Genetics 2004;36:388–93.CrossRef
Little, J, Higgins, JP, Ioannidis, JP, et al. STrengthening the REporting of Genetic Association studies (STREGA): an extension of the STROBE Statement. Annals of Internal Medicine 2009;150:206–15.CrossRef
Garner, C. Upward bias in odds ratio estimates from genome-wide association studies. Genetic Epidemiology 2007;31:288–95.CrossRef
Price, AL, Patterson, NJ, Plenge, RM, et al. Principal components analysis corrects for stratification in genome-wide association studies. Nature Genetics 2006;38:904–9.CrossRef
Easton, DF, Pooley, KA, Dunning, AM, et al. Genome-wide association study identifies novel breast cancer susceptibility loci. Nature. 2007;447:1087–93.CrossRef
Garcia-Closas, M, Hall, P, Nevanlinna, H, et al. Heterogeneity of breast cancer associations with five susceptibility loci by clinical and pathological characteristics. PLoS Genetics 2008;4:e1000054.
Ahmed, S, Thomas, G, Ghoussaini, M, et al. Newly discovered breast cancer susceptibility loci on 3p24 and 17q23.2. Nature Genetics 2009;41:585–90.
Hunter, DJ, Kraft, P, Jacobs, KB, et al. A genome-wide association study identifies alleles in FGFR2 associated with risk of sporadic postmenopausal breast cancer. Nature Genetics 2007;39:870–4.CrossRef
Thomas, G, Jacobs, KB, Kraft, P, et al. A multistage genome-wide association study in breast cancer identifies two new risk alleles at 1p11.2 and 14q24.1 (RAD51L1). Nature Genetics 2009;41:579–84.
Stacey, SN, Manolescu, A, Sulem, P, et al. Common variants on chromosomes 2q35 and 16q12 confer susceptibility to estrogen receptor-positive breast cancer. Nature Genetics 2007;39:865–9.CrossRef
Stacey, SN, Manolescu, A, Sulem, P, et al. Common variants on chromosome 5p12 confer susceptibility to estrogen receptor-positive breast cancer. Nature Genetics 2008;40:703–6.CrossRef
Gold, B, Kirchhoff, T, Stefanov, S, et al. Genome-wide association study provides evidence for a breast cancer risk locus at 6q22.33. Proceedings of the National Academy of Sciences USA 2008;105:4340–5.
Zheng, W, Long, J, Gao, YT, et al. Genome-wide association study identifies a new breast cancer susceptibility locus at 6q25.1. Nature Genetics 2009;41:324–8.
Eeles, RA, Kote-Jarai, Z, Giles, GG, et al. Multiple newly identified loci associated with prostate cancer susceptibility. Nature Genetics 2008;40:316–21.CrossRef
Thomas, G, Jacobs, KB, Yeager, M, et al. Multiple loci identified in a genome-wide association study of prostate cancer. Nature Genetics 2008;40:310–15.CrossRef
Yeager, M, Orr, N, Hayes, RB, et al. Genome-wide association study of prostate cancer identifies a second risk locus at 8q24. Nature Genetics 2007;39:645–9.CrossRef
Yeager, M, Chatterjee, N, Ciampa, J, et al. Identification of a new prostate cancer susceptibility locus on chromosome 8q24. Nature Genetics 2009;41:1055–7.CrossRef
Eeles, RA, Kote-Jarai, Z, Al Olama, AA, et al. Identification of seven new prostate cancer susceptibility loci through a genome-wide association study. Nature Genetics 2009;41:1116–21.CrossRef
Gudmundsson, J, Sulem, P, Manolescu, A, et al. Genome-wide association study identifies a second prostate cancer susceptibility variant at 8q24. Nature Genetics 2007;39:631–7.CrossRef
Gudmundsson, J, Sulem, P, Rafnar, T, et al. Common sequence variants on 2p15 and Xp11.22 confer susceptibility to prostate cancer. Nature Genetics 2008;40:281–3.CrossRef
Gudmundsson, J, Sulem, P, Gudbjartsson, DF, et al. Genome-wide association and replication studies identify four variants associated with prostate cancer susceptibility. Nature Genetics 2009;41:1122–6.CrossRef
Amundadottir, LT, Sulem, P, Gudmundsson, J, et al. A common variant associated with prostate cancer in European and African populations. Nature Genetics 2006;38:652–8.CrossRef
Freedman, ML, Haiman, CA, Patterson, N, et al. Admixture mapping identifies 8q24 as a prostate cancer risk locus in African-American men. Proceedings of the National Academy of Sciences USA 2006;103:14 068–73.
Al Olama, AA, Kote-Jarai, Z, Giles, GG, et al. Multiple loci on 8q24 associated with prostate cancer susceptibility. Nature Genetics 2009;41:1058–60.CrossRef
Sun, J, Zheng, SL, Wiklund, F, et al. Evidence for two independent prostate cancer risk-associated loci in the HNF1B gene at 17q12. Nature Genetics 2008;40:1153–5.CrossRef
Haiman, CA, Chen, GK, Blot, WJ, et al. Genome-wide association study of prostate cancer in men of African ancestry identifies a susceptibility locus at 17q21. Nature Genetics 2011;43:570–3.CrossRef
Kote-Jarai, Z, Olama, AA, Giles, GG, et al. Seven prostate cancer susceptibility loci identified by a multi-stage genome-wide association study. Nature Genetics 2011;43:785–91.CrossRef
Schumacher, FR, Berndt, SI, Siddiq, A, et al. Genome-wide association study identifies new prostate cancer susceptibility loci. Human Molecular Genetics. 2011;20:3867–75.CrossRef
Haiman, CA, Patterson, N, Freedman, ML, et al. Multiple regions within 8q24 independently affect risk for prostate cancer. Nature Genetics 2007;39:638–44.CrossRef
Tomlinson, I, Webb, E, Carvajal-Carmona, L, et al. A genome-wide association scan of tag SNPs identifies a susceptibility variant for colorectal cancer at 8q24.21. Nature Genetics 2007;39:984–8.
Zanke, BW, Greenwood, CM, Rangrej, J, et al. Genome-wide association scan identifies a colorectal cancer susceptibility locus on chromosome 8q24. Nature Genetics 2007;39:989–94.CrossRef
Tomlinson, IP, Webb, E, Carvajal-Carmona, L, et al. A genome-wide association study identifies colorectal cancer susceptibility loci on chromosomes 10p14 and 8q23.3. Nature Genetics 2008;40:623–30.
Houlston, RS, Webb, E, Broderick, P, et al. Meta-analysis of genome-wide association data identifies four new susceptibility loci for colorectal cancer. Nature Genetics 2008;40:1426–35.CrossRef
Tenesa, A, Farrington, SM, Prendergast, JG, et al. Genome-wide association scan identifies a colorectal cancer susceptibility locus on 11q23 and replicates risk loci at 8q24 and 18q21. Nature Genetics 2008;40:631–7.CrossRef
Jaeger, E, Webb, E, Howarth, K, et al. Common genetic variants at the CRAC1 (HMPS) locus on chromosome 15q13.3 influence colorectal cancer risk. Nature Genetics 2008;40:26–8.
Broderick, P, Carvajal-Carmona, L, Pittman, AM, et al. A genome-wide association study shows that common alleles of Smad7 influence colorectal cancer risk. Nature Genetics 2007;39:1315–17.CrossRef
Houlston, RS, Cheadle, J, Dobbins, SE, et al. Meta-analysis of three genome-wide association studies identifies susceptibility loci for colorectal cancer at 1q41, 3q26.2, 12q13.13 and 20q13.33. Nature Genetics 2010;42:973–7.
Cui, R, Okada, Y, Jang, SG, et al. Common variant in 6q26-q27 is associated with distal colon cancer in an Asian population. Gut 2011;60:799–805.CrossRef
Wang, Y, Broderick, P, Webb, E, et al. Common 5p15.33 and 6p21.33 variants influence lung cancer risk. Nature Genetics 2008;40:1407–9.CrossRef
McKay, JD, Hung, RJ, Gaborieau, V, et al. Lung cancer susceptibility locus at 5p15.33. Nature Genetics 2008;40:1404–6.
Amos, CI, Wu, X, Broderick, P, et al. Genome-wide association scan of tag SNPs identifies a susceptibility locus for lung cancer at 15q25.1. Nature Genetics 2008;40:616–22.
Hung, RJ, McKay, JD, Gaborieau, V, et al. A susceptibility locus for lung cancer maps to nicotinic acetylcholine receptor subunit genes on 15q25. Nature. 2008;452:633–7.CrossRef
Liu, P, Vikis, HG, Wang, D, et al. Familial aggregation of common sequence variants on 15q24–25.1 in lung cancer. Journal of the National Cancer Institute 2008;100:1326–30.CrossRefGoogle ScholarPubMed
Li, Y, Sheu, CC, Ye, Y, et al. Genetic variants and risk of lung cancer in never smokers: a genome-wide association study. Lancet Oncology 2010;11:321–30.CrossRef
Miki, D, Kubo, M, Takahashi, A, et al. Variation in TP63 is associated with lung adenocarcinoma susceptibility in Japanese and Korean populations. Nature Genetics 2010;42:893–6.CrossRef
Hsiung, CA, Lan, Q, Hong, YC, et al. The 5p15.33 locus is associated with risk of lung adenocarcinoma in never-smoking females in Asia. PLoS Genetics 2010;6:e1001051.
Hu, Z, Wu, C, Shi, Y, et al. A genome-wide association study identifies two new lung cancer susceptibility loci at 13q12.12 and 22q12.2 in Han Chinese. Nature Genetics 2011;43:792–6.
Amundadottir, L, Kraft, P, Stolzenberg-Solomon, RZ, et al. Genome-wide association study identifies variants in the ABO locus associated with susceptibility to pancreatic cancer. Nature Genetics 2009;41:986–90.CrossRef
Low, SK, Kuchiba, A, Zembutsu, H, et al. Genome-wide association study of pancreatic cancer in Japanese population. PLoS One. 2010;5:e11824.
Sakamoto, H, Yoshimura, K, Saeki, N, et al. Genetic variation in PSCA is associated with susceptibility to diffuse-type gastric cancer. Nature Genetics 2008;40:730–40.CrossRef
Bass, AJ, Meyerson, M. Genome-wide association study in esophageal squamous cell carcinoma. Gastroenterology 2009;137:1573–6.CrossRef
Wu, C, Hu, Z, He, Z, et al. Genome-wide association study identifies three new susceptibility loci for esophageal squamous-cell carcinoma in Chinese populations. Nature Genetics 2011;43:679–84.CrossRef
Kiemeney, LA, Thorlacius, S, Sulem, P, et al. Sequence variant on 8q24 confers susceptibility to urinary bladder cancer. Nature Genetics 2008;40:1307–12.CrossRef
Wu, X, Ye, Y, Kiemeney, LA, et al. Genetic variation in the prostate stem cell antigen gene PSCA confers susceptibility to urinary bladder cancer. Nature Genetics 2009;41:991–5.CrossRef
Kiemeney, LA, Sulem, P, Besenbacher, S, et al. A sequence variant at 4p16.3 confers susceptibility to urinary bladder cancer. Nature Genetics 2010;42:415–19.CrossRef
Rapley, EA, Turnbull, C, Al Olama, AA, et al. A genome-wide association study of testicular germ cell tumor. Nature Genetics 2009;41:807–10.CrossRef
Kanetsky, PA, Mitra, N, Vardhanabhuti, S, et al. Common variation in KITLG and at 5q31.3 predisposes to testicular germ cell cancer. Nature Genetics 2009;41:811–15.CrossRef
Kanetsky, PA, Mitra, N, Vardhanabhuti, S, et al. A second independent locus within DMRT1 is associated with testicular germ cell tumor susceptibility. Human Molecular Genetics 2011;20:3109–17.CrossRef
Turnbull, C, Rapley, EA, Seal, S, et al. Variants near DMRT1, TERT and ATF7IP are associated with testicular germ cell cancer. Nature Genetics 2010;42:604–7.CrossRef
Song, H, Ramus, SJ, Tyrer, J, et al. A genome-wide association study identifies a new ovarian cancer susceptibility locus on 9p22.2. Nature Genetics 2009;41:996–1000.
Bolton, KL, Tyrer, J, Song, H, et al. Common variants at 19p13 are associated with susceptibility to ovarian cancer. Nature Genetics 2010;42:880–4.CrossRef
Goode, EL, Chenevix-Trench, G, Song, H, et al. A genome-wide association study identifies susceptibility loci for ovarian cancer at 2q31 and 8q24. Nature Genetics 2010;42:874–9.CrossRef
Spurdle, AB, Thompson, DJ, Ahmed, S, et al. Genome-wide association study identifies a common variant associated with risk of endometrial cancer. Nature Genetics 2011;43:451–4.CrossRef
Kumar, V, Kato, N, Urabe, Y, et al. Genome-wide association study identifies a susceptibility locus for HCV-induced hepatocellular carcinoma. Nature Genetics 2011;43:455–8.CrossRef
Clifford, RJ, Zhang, J, Meerzaman, DM, et al. Genetic variations at loci involved in the immune response are risk factors for hepatocellular carcinoma. Hepatology 2010;52:2034–43.CrossRef
Zhang, H, Zhai, Y, Hu, Z, et al. Genome-wide association study identifies 1p36.22 as a new susceptibility locus for hepatocellular carcinoma in chronic hepatitis B virus carriers. Nature Genetics 2010;42:755–8.CrossRef
Bei, JX, Li, Y, Jia, WH, et al. A genome-wide association study of nasopharyngeal carcinoma identifies three new susceptibility loci. Nature Genetics 2010;42:599–603.CrossRef
Tse, KP, Su, WH, Chang, KP, et al. Genome-wide association study reveals multiple nasopharyngeal carcinoma-associated loci within the HLA region at chromosome 6p21.3. American Journal of Human Genetics 2009;85:194–203.CrossRefGoogle ScholarPubMed
Ng, CC, Yew, PY, Puah, SM, et al. A genome-wide association study identifies ITGA9 conferring risk of nasopharyngeal carcinoma. Journal of Human Genetics 2009;54:392–7.CrossRefGoogle ScholarPubMed
Purdue, MP, Johansson, M, Zelenika, D, et al. Genome-wide association study of renal cell carcinoma identifies two susceptibility loci on 2p21 and 11q13.3. Nature Genetics 2011;43:60–5.
Stacey, SN, Sulem, P, Masson, G, et al. New common variants affecting susceptibility to basal cell carcinoma. Nature Genetics 2009;41:909–14.CrossRef
Stacey, SN, Gudbjartsson, DF, Sulem, P, et al. Common variants on 1p36 and 1q42 are associated with cutaneous basal cell carcinoma but not with melanoma or pigmentation traits. Nature Genetics 2008;40:1313–18.CrossRef
Bishop, DT, Demenais, F, Iles, MM, et al. Genome-wide association study identifies three loci associated with melanoma risk. Nature Genetics 2009;41:920–5.CrossRef
Gudbjartsson, DF, Sulem, P, Stacey, SN, et al. ASIP and TYR pigmentation variants associate with cutaneous melanoma and basal cell carcinoma. Nature Genetics 2008;40:886–91.CrossRef
Brown, KM, Macgregor, S, Montgomery, GW, et al. Common sequence variants on 20q11.22 confer melanoma susceptibility. Nature Genetics 2008;40:838–40.CrossRef
Gudmundsson, J, Sulem, P, Gudbjartsson, DF, et al. Common variants on 9q22.33 and 14q13.3 predispose to thyroid cancer in European populations. Nature Genetics 2009;41:460–4.CrossRef
Capasso, M, Devoto, M, Hou, C, et al. Common variations in BARD1 influence susceptibility to high-risk neuroblastoma. Nature Genetics 2009;41:718–23.CrossRef
Maris, JM, Mosse, YP, Bradfield, JP, et al. Chromosome 6p22 locus associated with clinically aggressive neuroblastoma. New England Journal of Medicine 2008;358:2585–93.CrossRefGoogle ScholarPubMed
Shete, S, Hosking, FJ, Robertson, LB, et al. Genome-wide association study identifies five susceptibility loci for glioma. Nature Genetics 2009;41:899–904.CrossRef
Wrensch, M, Jenkins, RB, Chang, JS, et al. Variants in the CDKN2B and RTEL1 regions are associated with high-grade glioma susceptibility. Nature Genetics 2009;41:905–8.CrossRef
Wang, K, Diskin, SJ, Zhang, H, et al. Integrative genomics identifies LMO1 as a neuroblastoma oncogene. Nature 2011;469:216–20.CrossRef
Papaemmanuil, E, Hosking, FJ, Vijayakrishnan, J, et al. Loci on 7p12.2, 10q21.2 and 14q11.2 are associated with risk of childhood acute lymphoblastic leukemia. Nature Genetics 2009;41:1006–10.CrossRef
Trevino, LR, Yang, W, French, D, et al. Germline genomic variants associated with childhood acute lymphoblastic leukemia. Nature Genetics 2009;41:1001–5.CrossRef
Di Bernardo, MC, Crowther-Swanepoel, D, Broderick, P, et al. A genome-wide association study identifies six susceptibility loci for chronic lymphocytic leukemia. Nature Genetics 2008;40:1204–10.CrossRef
Skibola, CF, Bracci, PM, Halperin, E, et al. Genetic variants at 6p21.33 are associated with susceptibility to follicular lymphoma. Nature Genetics 2009;41:873–5.CrossRef
Kilpivaara, O, Mukherjee, S, Schram, AM, et al. A germline JAK2 SNP is associated with predisposition to the development of JAK2(V617F)-positive myeloproliferative neoplasms. Nature Genetics 2009;41:455–9.CrossRef
Slager, SL, Rabe, KG, Achenbach, SJ, et al. Genome-wide association study identifies a novel susceptibility locus at 6p21.3 among familial CLL. Blood 2011;117:1911–16.
Kim, DH, Lee, ST, Won, HH, et al. A genome-wide association study identifies novel loci associated with susceptibility to chronic myeloid leukemia. Blood 2011;117:6906–11.CrossRef
Enciso-Mora, V, Broderick, P, Ma, Y, et al. A genome-wide association study of Hodgkin's lymphoma identifies new susceptibility loci at 2p16.1 (REL), 8q24.21 and 10p14 (GATA3). Nature Genetics 2010;42:1126–30.
Kumar, V, Matsuo, K, Takahashi, A, et al. Common variants on 14q32 and 13q12 are associated with DLBCL susceptibility. Journal of Human Genetics 2011;56:436–9.CrossRefGoogle ScholarPubMed
Hindorff, LA. A Catalog of Published Genome-Wide Association Studies. . Accessed July 15, 2011.
Mahakali Zama, A, Hudson, FP, 3rd, Bedell, MA. Analysis of hypomorphic KitlSl mutants suggests different requirements for KITL in proliferation and migration of mouse primordial germ cells. Biology of Reproduction 2005;73:639–47.CrossRef
Crowther-Swanepoel, D, Broderick, P, Di Bernardo, MC, et al. Common variants at 2q37.3, 8q24.21, 15q21.3 and 16q24.1 influence chronic lymphocytic leukemia risk. Nature Genetics 2010;42:132–6.CrossRef
Grisanzio, C, Freedman, ML. Chromosome 8q24-associated cancers and MYC. Genes and Cancer 2010;1:555–9.CrossRef
Polakis, P. Wnt signaling and cancer. Genes and Development 2000;14:1837–51.
Sole, X, Hernandez, P, de Heredia, ML, et al. Genetic and genomic analysis modeling of germline c-MYC overexpression and cancer susceptibility. BMC Genomics 2008;9:12.CrossRef
Pomerantz, MM, Ahmadiyeh, N, Jia, L, et al. The 8q24 cancer risk variant rs6983267 shows long-range interaction with MYC in colorectal cancer. Nature Genetics 2009;41:882–4.CrossRef
Tuupanen, S, Turunen, M, Lehtonen, R, et al. The common colorectal cancer predisposition SNP rs6983267 at chromosome 8q24 confers potential to enhanced Wnt signaling. Nature Genetics 2009;41:885–90.CrossRef
Sotelo, J, Esposito, D, Duhagon, MA, et al. Long-range enhancers on 8q24 regulate c-Myc. Proceedings of the National Academy of Sciences USA 2010;107:3001–5.CrossRef
Wright, JB, Brown, SJ, Cole, MD. Upregulation of c-MYC in cis through a large chromatin loop linked to a cancer risk-associated single-nucleotide polymorphism in colorectal cancer cells. Molecular and Cellular Biology 2010;30:1411–20.CrossRef
Ahmadiyeh, N, Pomerantz, MM, Grisanzio, C, et al. 8q24 prostate, breast, and colon cancer risk loci show tissue-specific long-range interaction with MYC. Proceedings of the National Academy of Sciences USA 2010;107:9742–6.CrossRef
Janssens, AC, Gwinn, M, Bradley, LA, et al. A critical appraisal of the scientific basis of commercial genomic profiles used to assess health risks and personalize health interventions. American Journal of Human Genetics 2008;82:593–9.CrossRefGoogle ScholarPubMed
Grosse, SD, Khoury, MJ. What is the clinical utility of genetic testing? Genetics in Medicine 2006;8:448–50.
Burke, W. Clinical validity and clinical utility of genetic tests. Current Protocols in Human Genetics 2009;Chapter 9:Unit 9 15.
Gail, MH. Value of adding single-nucleotide polymorphism genotypes to a breast cancer risk model. Journal of the National Cancer Institute 2009;101:959–63.CrossRefGoogle ScholarPubMed
Gail, MH. Discriminatory accuracy from single-nucleotide polymorphisms in models to predict breast cancer risk. Journal of the National Cancer Institute 2008;100:1037–41.CrossRefGoogle ScholarPubMed
Wacholder, S, Hartge, P, Prentice, R, et al. Performance of common genetic variants in breast-cancer risk models. New England Journal of Medicine. 2010;362:986–93.CrossRefGoogle ScholarPubMed
Michailidou, K, Hall, P, Gonzalez-Neira, A, et al. Large-scale genotyping identifies 41 new loci associated with breast cancer risk. Nature Genetics 2013;45:353–61.CrossRefGoogle ScholarPubMed
Garcia-Closas, M, Couch, FJ, Lindstrom, S, et al. Genome-wide association studies identify four ER negative-specific breast cancer risk loci. Nature Genetics 2013;45:392–8.CrossRefGoogle ScholarPubMed
Bojesen, SE, Pooley, KA, Johnatty, SE, et al. Multiple independent variants at the TERT locus are associated with telomere length and risks of breast and ovarian cancer. Nature Genetics 2013;45:371–84.CrossRefGoogle ScholarPubMed
Pharoah, PD, Tsai, YY, Ramus, SJ, et al. GWAS meta-analysis and replication identifies three new susceptibility loci for ovarian cancer. Nature Genetics 2013;45:362–70.CrossRefGoogle ScholarPubMed
Shen, H, Fridley, BL, Song, H, et al. Epigenetic analysis leads to identification of HNF1B as a subtype-specific susceptibility gene for ovarian cancer. Nature Communications 2013;4:1628.CrossRefGoogle ScholarPubMed
Permuth-Wey, J, Lawrenson, K, Shen, HC, et al. Identification and molecular characterization of a new ovarian cancer susceptibility locus at 17q21.31. Nature Communications 2013;4:1627.CrossRefGoogle ScholarPubMed
Eeles, RA, Olama, AA, Benlloch, S, et al. Identification of 23 new prostate cancer susceptibility loci using the iCOGS custom genotyping array. Nature Genetics 2013;45:385–91.CrossRefGoogle ScholarPubMed
Kote-Jarai, Z, Saunders, EJ, Leongamornlert, DA, et al. Fine-mapping identifies multiple prostate cancer risk loci at 5p15, one of which associates with TERT expression. Human Molecular Genetics 2013.CrossRefGoogle ScholarPubMed
Burton, H, Chowdhury, S, Dent, T, et al. Public health implications from COGS and potential for risk stratification and screening. Nature Genetics 2013;45:349–51.CrossRefGoogle ScholarPubMed

Save book to Kindle

To save this book to your Kindle, first ensure coreplatform@cambridge.org is added to your Approved Personal Document E-mail List under your Personal Document Settings on the Manage Your Content and Devices page of your Amazon account. Then enter the ‘name’ part of your Kindle email address below. Find out more about saving to your Kindle.

Note you can select to save to either the @free.kindle.com or @kindle.com variations. ‘@free.kindle.com’ emails are free but can only be saved to your device when it is connected to wi-fi. ‘@kindle.com’ emails can be delivered even when you are not connected to wi-fi, but note that service fees apply.

Find out more about the Kindle Personal Document Service.

Available formats
×

Save book to Dropbox

To save content items to your account, please confirm that you agree to abide by our usage policies. If this is the first time you use this feature, you will be asked to authorise Cambridge Core to connect with your account. Find out more about saving content to Dropbox.

Available formats
×

Save book to Google Drive

To save content items to your account, please confirm that you agree to abide by our usage policies. If this is the first time you use this feature, you will be asked to authorise Cambridge Core to connect with your account. Find out more about saving content to Google Drive.

Available formats
×