Skip to main content

Thank you for visiting nature.com. You are using a browser version with limited support for CSS. To obtain the best experience, we recommend you use a more up to date browser (or turn off compatibility mode in Internet Explorer). In the meantime, to ensure continued support, we are displaying the site without styles and JavaScript.

  • Article
  • Published:

Genome-wide association meta-analyses combining multiple risk phenotypes provide insights into the genetic architecture of cutaneous melanoma susceptibility

Abstract

Most genetic susceptibility to cutaneous melanoma remains to be discovered. Meta-analysis genome-wide association study (GWAS) of 36,760 cases of melanoma (67% newly genotyped) and 375,188 controls identified 54 significant (P < 5 × 10−8) loci with 68 independent single nucleotide polymorphisms. Analysis of risk estimates across geographical regions and host factors suggests the acral melanoma subtype is uniquely unrelated to pigmentation. Combining this meta-analysis with GWAS of nevus count and hair color, and transcriptome association approaches, uncovered 31 potential secondary loci for a total of 85 cutaneous melanoma susceptibility loci. These findings provide insights into cutaneous melanoma genetic architecture, reinforcing the importance of nevogenesis, pigmentation and telomere maintenance, together with identifying potential new pathways for cutaneous melanoma pathogenesis.

This is a preview of subscription content, access via your institution

Access options

Rent or buy this article

Prices vary by article type

from$1.95

to$39.95

Prices may be subject to local taxes which are calculated during checkout

Fig. 1: Manhattan plot for the total cutaneous melanoma meta-analysis.
Fig. 2: Overlap of loci identified by primary and secondary analyses.

Similar content being viewed by others

Data availability

Genome-wide summary statistics for the confirmed meta-analysis have been made publicly available at dbGaP (phs001868.v1.p1), with the exclusion of self-reported data from 23andMe and UK Biobank. Results for SNPs with a fixed or random P < 5 × 10−7 from the total meta-analysis are reported in Supplementary Table 7. The total meta-analysis includes self-reported cutaneous melanoma GWAS data from the UK Biobank and 23andMe. The raw genetic and phenotypic UK Biobank data used in this study, which were used under license, are available from: http://www.ukbiobank.ac.uk/. The genome-wide summary statistics from 23andMe data were obtained under a data transfer agreement. Further information about obtaining access to the 23andMe summary statistics is available from https://research.23andme.com/collaborate/. Source data for Fig. 2, Extended Data Figs. 4–6 and Supplementary Figs. 2 and 3 are available with the paper.

References

  1. Karimkhani, C. et al. The global burden of melanoma: results from the global burden of disease study 2015. Br. J. Dermatol. 177, 134–140 (2017).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  2. Secretan, B. et al. WHO International agency for research on cancer monograph working group a review of human carcinogens—Part E: tobacco, areca nut, alcohol, coal smoke, and salted fish. Lancet Oncol. 10, 1033–1034 (2009).

    Article  PubMed  Google Scholar 

  3. Ford, D. et al. Risk of cutaneous melanoma associated with a family history of the disease. Int. J. Cancer 62, 377–381 (1995).

    Article  CAS  PubMed  Google Scholar 

  4. Olsen, C. M., Carroll, H. J. & Whiteman, D. C. Familial melanoma: a meta-analysis and estimates of attributable fraction. Cancer Epidemiol. Biomark. Prev. 19, 65–73 (2010).

    Article  Google Scholar 

  5. Olsen, C. M., Carroll, H. J. & Whiteman, D. C. Estimating the attributable fraction for melanoma: a meta-analysis of pigmentary characteristics and freckling. Int. J. Cancer 127, 2430–2445 (2010).

    Article  CAS  PubMed  Google Scholar 

  6. Chang, Y. M. et al. A pooled analysis of melanocytic nevus phenotype and the risk of cutaneous melanoma at different latitudes. Int. J. Cancer 124, 420–428 (2009).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  7. Olsen, C. M., Carroll, H. J. & Whiteman, D. C. Estimating the attributable fraction for cancer: a meta-analysis of nevi and melanoma. Cancer Prev. Res. 3, 233–245 (2010).

    Article  Google Scholar 

  8. Bataille, V. et al. Nevus size and number are associated with telomere length and represent potential markers of a decreased senescence in vivo. Cancer Epidemiol. Biomark. Prev. 16, 1499–1502 (2007).

    Article  CAS  Google Scholar 

  9. Han, J. et al. A prospective study of telomere length and the risk of skin cancer. J. Invest. Dermatol. 129, 415–421 (2009).

    Article  CAS  PubMed  Google Scholar 

  10. Green, A. C. & Olsen, C. M. Increased risk of melanoma in organ transplant recipients: systematic review and meta-analysis of cohort studies. Acta Derm. Venereol. 95, 923–927 (2015).

    Article  CAS  PubMed  Google Scholar 

  11. Kamb, A. et al. Analysis of the p16 gene (CDKN2) as a candidate for the chromosome 9p melanoma susceptibility locus. Nat. Genet. 8, 23–26 (1994).

    Article  CAS  PubMed  Google Scholar 

  12. Berwick, M. et al. The prevalence of CDKN2A germ-line mutations and relative risk for cutaneous malignant melanoma: an international population-based study. Cancer Epidemiol. Biomark. Prev. 15, 1520–1525 (2006).

    Article  CAS  Google Scholar 

  13. Robles-Espinoza, C. D. et al. POT1 loss-of-function variants predispose to familial melanoma. Nat. Genet. 46, 478–481 (2014).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  14. Shi, J. et al. Rare missense variants in POT1 predispose to familial cutaneous malignant melanoma. Nat. Genet. 46, 482–486 (2014).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  15. Palmer, J. S. et al. Melanocortin-1 receptor polymorphisms and risk of melanoma: is the association explained solely by pigmentation phenotype? Am. J. Hum. Genet. 66, 176–186 (2000).

    Article  CAS  PubMed  Google Scholar 

  16. Landi, M. T. et al. MC1R, ASIP, and DNA repair in sporadic and familial melanoma in a Mediterranean population. J. Natl. Cancer Inst. 98, 144–145 (2005).

    Google Scholar 

  17. Brown, K. M. et al. Common sequence variants on 20q11.22 confer melanoma susceptibility. Nat. Genet. 40, 838–840 (2008).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  18. Bishop, D. T. et al. Genome-wide association study identifies three loci associated with melanoma risk. Nat. Genet. 41, 920–925 (2009).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  19. Amos, C. I. et al. Genome-wide association study identifies novel loci predisposing to cutaneous melanoma. Hum. Mol. Genet. 20, 5012–5023 (2011).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  20. Barrett, J. H. et al. Genome-wide association study identifies three new melanoma susceptibility loci. Nat. Genet. 43, 1108–1113 (2011).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  21. Macgregor, S. et al. Genome-wide association study identifies a new melanoma susceptibility locus at 1q21.3. Nat. Genet. 43, 1114–1118 (2011).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  22. Iles, M. M. et al. A variant in FTO shows association with melanoma risk not due to BMI. Nat. Genet. 45, 428–432 (2013).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  23. Law, M. H. et al. Genome-wide meta-analysis identifies five new susceptibility loci for cutaneous malignant melanoma. Nat. Genet. 47, 987–995 (2015).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  24. Ransohoff, K. J. et al. Two-stage genome-wide association study identifies a novel susceptibility locus associated with melanoma. Oncotarget 8, 17586–17592 (2017).

    Article  PubMed  PubMed Central  Google Scholar 

  25. Yokoyama, S. et al. A novel recurrent mutation in MITF predisposes to familial and sporadic melanoma. Nature 480, 99–103 (2011).

  26. Bertolotto, C. et al. A SUMOylation-defective MITF germline mutation predisposes to melanoma and renal carcinoma. Nature 480, 94–98 (2011).

    Article  CAS  PubMed  Google Scholar 

  27. Duffy, D. L. et al. Novel pleiotropic risk loci for melanoma and nevus density implicate multiple biological pathways. Nat. Commun. 9, 4774 (2018).

    Article  PubMed  PubMed Central  CAS  Google Scholar 

  28. Zhang, T. et al. Cell-type-specific eQTL of primary melanocytes facilitates identification of melanoma susceptibility genes. Genome Res. 28, 1621–1635 (2018).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  29. Elder, D. E., Massi, D., Willemze, R. & Scolyer, R. WHO Classification of Skin Tumours (International Agency for Research on Cancer, 2018).

  30. Higgins, J. P. T. & Thompson, S. G. Quantifying heterogeneity in a meta-analysis. Stat. Med. 21, 1539–1558 (2002).

  31. Bulik-Sullivan, B. et al. An atlas of genetic correlations across human diseases and traits. Nat. Genet. 47, 1236–1241 (2015).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  32. Zhang, Y., Qi, G., Park, J.-H. & Chatterjee, N. Estimation of complex effect-size distributions using summary-level statistics from genome-wide association studies across 32 complex traits. Nat. Genet. 50, 1318–1326 (2018).

    Article  CAS  PubMed  Google Scholar 

  33. Yang, J. et al. Conditional and joint multiple-SNP analysis of GWAS summary statistics identifies additional variants influencing complex traits. Nat. Genet. 44, 369–375 (2012).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  34. Duffy, D. L. et al. Multiple pigmentation gene polymorphisms account for a substantial proportion of risk of cutaneous malignant melanoma. J. Invest. Dermatol. 130, 520–528 (2010).

    Article  CAS  PubMed  Google Scholar 

  35. Duffy, D. L. et al. IRF4 variants have age-specific effects on nevus count and predispose to melanoma. Am. J. Hum. Genet. 87, 6–16 (2010).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  36. Delgado, D. A. et al. Genome-wide association study of telomere length among South Asians identifies a second RTEL1 association signal. J. Med. Genet. 55, 64–71 (2018).

    Article  CAS  PubMed  Google Scholar 

  37. Iles, M. M. et al. The effect on melanoma risk of genes previously associated with telomere length. J. Natl. Cancer Inst. 106, dju267 (2014).

    Article  PubMed  PubMed Central  CAS  Google Scholar 

  38. Pickrell, J. K. et al. Detection and interpretation of shared genetic influences on 42 human traits. Nat. Genet. 48, 709–717 (2016).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  39. GTEx Consortium Human genomics: the genotype-tissue expression (GTEx) pilot analysis: multitissue gene regulation in humans. Science 348, 648–660 (2015).

    Article  PubMed Central  CAS  Google Scholar 

  40. Gamazon, E. R. et al. A gene-based association method for mapping traits using reference transcriptome data. Nat. Genet. 47, 1091–1098 (2015).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  41. Gusev, A. et al. Integrative approaches for large-scale transcriptome-wide association studies. Nat. Genet. 48, 245–252 (2016).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  42. Finucane, H. K. et al. Partitioning heritability by functional annotation using genome-wide association summary statistics. Nat. Genet. 47, 1228–1235 (2015).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  43. Visconti, A. et al. Genome-wide association study in 176,678 Europeans reveals genetic loci for tanning response to sun exposure. Nat. Commun. 9, 1684 (2018).

    Article  PubMed  PubMed Central  CAS  Google Scholar 

  44. Choi, J. et al. A common intronic variant of PARP1 confers melanoma risk and mediates melanocyte growth via regulation of MITF. Nat. Genet. 49, 1326–1335 (2017).

    Article  CAS  PubMed  Google Scholar 

  45. Li, F. P. & Fraumeni, J. F. Jr. Soft-tissue sarcomas, breast cancer, and other neoplasms. A familial syndrome? Ann. Intern. Med. 71, 747–752 (1969).

    Article  CAS  PubMed  Google Scholar 

  46. Curiel-Lewandrowski, C., Speetzen, L. S., Cranmer, L., Warneke, J. A. & Loescher, L. J. Multiple primary cutaneous melanomas in Li–Fraumeni syndrome. Arch. Dermatol. 147, 248–250 (2011).

    Article  PubMed  Google Scholar 

  47. Beausejour, C. M. Reversal of human cellular senescence: roles of the p53 and p16 pathways. EMBO J. 22, 4212–4222 (2003).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  48. Kuilman, T., Michaloglou, C., Mooi, W. J. & Peeper, D. S. The essence of senescence. Genes Dev. 24, 2463–2479 (2010).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  49. Aoude, L. G. et al. Nonsense mutations in the shelterin complex genes ACD and TERF2IP in familial melanoma. J. Natl. Cancer Inst. 107, dju408 (2015).

    Article  PubMed  Google Scholar 

  50. Rafnar, T. et al. Sequence variants at the TERT-CLPTM1L locus associate with many cancer types. Nat. Genet. 41, 221–227 (2009).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  51. Rachakonda, S. et al. Telomere length, telomerase reverse transcriptase promoter mutations, and melanoma risk. Genes Chromosomes Cancer 57, 564–572 (2018).

    Article  CAS  PubMed  Google Scholar 

  52. Derheimer, F. A. & Kastan, M. B. Multiple roles of ATM in monitoring and maintaining DNA integrity. FEBS Lett. 584, 3675–3681 (2010).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  53. Demenais, F. et al. A linkage study between HLA and cutaneous malignant melanoma or precursor lesions or both. J. Med. Genet. 21, 429–435 (1984).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  54. Bale, S. J. et al. Hereditary malignant melanoma is not linked to the HLA complex on chromosome 6. Int. J. Cancer 36, 439–443 (1985).

    Article  CAS  PubMed  Google Scholar 

  55. Holland, E. A., Beaton, S. C., Kefford, R. F. & Mann, G. J. Linkage analysis of familial melanoma and chromosome 6 in 14 Australian kindreds. Genes Chromosomes Cancer 19, 241–249 (1997).

    Article  CAS  PubMed  Google Scholar 

  56. Barger, B. O., Acton, R. T., Soong, S. J., Roseman, J. & Balch, C. Increase of HLA-DR4 in melanoma patients from Alabama. Cancer Res. 42, 4276–4279 (1982).

    CAS  PubMed  Google Scholar 

  57. Rovini, D., Sacchini, V., Codazzi, V., Vaglini, M. & Illeni, M. T. HLA antigen frequencies in malignant melanoma patients. A second study. Tumori 70, 29–33 (1984).

    Article  CAS  PubMed  Google Scholar 

  58. Hors, J. et al. in Histocompatibility Testing 1984 (ed. Albert, E. D.) 407–410 (Springer, 1984).

  59. Lee, J. E., Reveille, J. D., Ross, M. I. & Platsoucas, C. D. HLA-DQB1* 0301 association with increased cutaneous melanoma risk. Int. J. Cancer 59, 510–513 (1994).

    Article  CAS  PubMed  Google Scholar 

  60. Muto, M. et al. HLA class I polymorphism and the susceptibility to malignant melanoma. Tissue Antigens 47, 447–449 (1996).

    CAS  PubMed  Google Scholar 

  61. Kageshita, T. et al. Molecular genetic analysis of HLA class II alleles in Japanese patients with melanoma. Tissue Antigens 49, 466–470 (1997).

    Article  CAS  PubMed  Google Scholar 

  62. Bateman, A. C., Turner, S. J., Theaker, J. M. & Howell, W. M. HLA-DQB1*0303 and *0301 alleles influence susceptibility to and prognosis in cutaneous malignant melanoma in the British Caucasian population. Tissue Antigens 52, 67–73 (1998).

    Article  CAS  PubMed  Google Scholar 

  63. Lombardi, M. L. et al. Molecular analysis of HLA DRB1 and DQB1 polymorphism in Italian melanoma patients. J. Immunother. 21, 435–439 (1998).

    Article  CAS  PubMed  Google Scholar 

  64. Luongo, V. et al. HLA allele frequency and clinical outcome in Italian patients with cutaneous melanoma. Tissue Antigens 64, 84–87 (2004).

    Article  CAS  PubMed  Google Scholar 

  65. Campillo, J. A. et al. HLA class I and class II frequencies in patients with cutaneous malignant melanoma from southeastern Spain: the role of HLA-C in disease prognosis. Immunogenetics 57, 926–933 (2006).

    Article  CAS  PubMed  Google Scholar 

  66. Planelles, D. et al. HLA class II polymorphisms in Spanish melanoma patients: homozygosity for HLA-DQA1 locus can be a potential melanoma risk factor. Br. J. Dermatol. 154, 261–266 (2006).

    Article  CAS  PubMed  Google Scholar 

  67. Jin, Y. et al. Genome-wide association studies of autoimmune vitiligo identify 23 new risk loci and highlight key pathways and regulatory variants. Nat. Genet. 48, 1418–1424 (2016).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  68. Jin, Y. et al. Genome-wide association analyses identify 13 new susceptibility loci for generalized vitiligo. Nat. Genet. 44, 676–680 (2012).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  69. Curran, K. et al. Interplay between Foxd3 and Mitf regulates cell fate plasticity in the zebrafish neural crest. Dev. Biol. 344, 107–118 (2010).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  70. Thomas, A. J. & Erickson, C. A. FOXD3 regulates the lineage switch between neural crest-derived glial cells and pigment cells by repressing MITF through a non-canonical mechanism. Development 136, 1849–1858 (2009).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  71. Kumano, K. et al. Both Notch1 and Notch2 contribute to the regulation of melanocyte homeostasis. Pigment Cell Melanoma Res. 21, 70–78 (2008).

    Article  CAS  PubMed  Google Scholar 

  72. Schouwey, K., Larue, L., Radtke, F., Delmas, V. & Beermann, F. Transgenic expression of notch in melanocytes demonstrates RBP-Jkappa-dependent signaling. Pigment Cell Melanoma Res. 23, 134–136 (2010).

    Article  CAS  PubMed  Google Scholar 

  73. Zabierowski, S. E. et al. Direct reprogramming of melanocytes to neural crest stem-like cells by one defined factor. Stem Cells 29, 1752–1762 (2011).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  74. Falchi, M. et al. Genome-wide association study identifies variants at 9p21 and 22q13 associated with development of cutaneous nevi. Nat. Genet. 41, 915–919 (2009).

    Article  CAS  PubMed  Google Scholar 

  75. Garraway, L. A. et al. Integrative genomic analyses identify MITF as a lineage survival oncogene amplified in malignant melanoma. Nature 436, 117–122 (2005).

    Article  CAS  PubMed  Google Scholar 

  76. Abel, E. V. & Aplin, A. E. FOXD3 is a mutant B-RAF-regulated inhibitor of G(1)-S progression in melanoma cells. Cancer Res. 70, 2891–2900 (2010).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  77. Weiss, M. B., Abel, E. V., Dadpey, N. & Aplin, A. E. FOXD3 modulates migration through direct transcriptional repression of TWIST1 in melanoma. Mol. Cancer Res. 12, 1314–1323 (2014).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  78. Golan, T. et al. Interactions of melanoma cells with distal keratinocytes trigger metastasis via notch signaling inhibition of MITF. Mol. Cell 59, 664–676 (2015).

    Article  CAS  PubMed  Google Scholar 

  79. Cronin, J. C. et al. SOX10 ablation arrests cell cycle, induces senescence, and suppresses melanomagenesis. Cancer Res. 73, 5709–5718 (2013).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  80. Guilford, P. et al. E-cadherin germline mutations in familial gastric cancer. Nature 392, 402–405 (1998).

    Article  CAS  PubMed  Google Scholar 

  81. Hansford, S. et al. Hereditary diffuse gastric cancer syndrome: CDH1 mutations and beyond. JAMA Oncol. 1, 23–32 (2015).

    Article  PubMed  Google Scholar 

  82. Kim, H. C. et al. The E-cadherin gene (CDH1) variants T340A and L599V in gastric and colorectal cancer patients in Korea. Gut 47, 262–267 (2000).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  83. Tang, A. et al. E-cadherin is the major mediator of human melanocyte adhesion to keratinocytes in vitro. J. Cell Sci. 107(Pt 4), 983–992 (1994).

    Article  CAS  PubMed  Google Scholar 

  84. Hsu, M. Y., Wheelock, M. J., Johnson, K. R. & Herlyn, M. Shifts in cadherin profiles between human normal melanocytes and melanomas. J. Investig. Dermatol. Symp. Proc. 1, 188–194 (1996).

    CAS  PubMed  Google Scholar 

  85. Study, C. et al. Meta-analysis of genome-wide association data identifies four new susceptibility loci for colorectal cancer. Nat. Genet. 40, 1426–1435 (2008).

    Article  CAS  Google Scholar 

  86. Wagner, R. Y. et al. Altered E-cadherin levels and distribution in melanocytes precede clinical manifestations of vitiligo. J. Invest. Dermatol. 135, 1810–1819 (2015).

    Article  CAS  PubMed  Google Scholar 

  87. Padmanaban, V. et al. E-cadherin is required for metastasis in multiple models of breast cancer. Nature 573, 439–444 (2019).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  88. Peña-Chilet, M. et al. Genetic variants in PARP1 (rs3219090) and IRF4 (rs12203592) genes associated with melanoma susceptibility in a Spanish population. BMC Cancer 13, 160 (2013).

    Article  PubMed  PubMed Central  CAS  Google Scholar 

  89. Law, M. H. et al. Meta-analysis combining new and existing data sets confirms that the TERT-CLPTM1L locus influences melanoma risk. J. Invest. Dermatol. 132, 485–487 (2012).

    Article  CAS  PubMed  Google Scholar 

  90. Antonopoulou, K. et al. Updated field synopsis and systematic meta-analyses of genetic association studies in cutaneous melanoma: the MelGene database. J. Invest. Dermatol. 135, 1074–1079 (2015).

    Article  CAS  PubMed  Google Scholar 

  91. McCarthy, S. et al. A reference panel of 64,976 haplotypes for genotype imputation. Nat. Genet. 48, 1279–1283 (2016).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  92. Purcell, S. et al. PLINK: a tool set for whole-genome association and population-based linkage analyses. Am. J. Hum. Genet. 81, 559–575 (2007).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  93. Chang, C. C. et al. Second-generation PLINK: rising to the challenge of larger and richer datasets. Gigascience 4, 7 (2015).

    Article  PubMed  PubMed Central  CAS  Google Scholar 

  94. Das, S. et al. Next-generation genotype imputation service and methods. Nat. Genet. 48, 1284–1287 (2016).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  95. DerSimonian, R. & Laird, N. Meta-analysis in clinical trials. Control. Clin. Trials 7, 177–188 (1986).

    Article  CAS  PubMed  Google Scholar 

  96. Machiela, M. J. & Chanock, S. J. LDassoc: an online tool for interactively exploring genome-wide association study results and prioritizing variants for functional investigation. Bioinformatics 34, 887–889 (2018).

    Article  CAS  PubMed  Google Scholar 

  97. Willer, C. J., Li, Y. & Abecasis, G. R. METAL: fast and efficient meta-analysis of genomewide association scans. Bioinformatics 26, 2190–2191 (2010).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  98. Watanabe, K., Taskesen, E., Van Bochoven, A. & Posthuma, D. Functional mapping and annotation of genetic associations with FUMA. Nat. Commun. 8, 1826 (2017).

    Article  PubMed  PubMed Central  CAS  Google Scholar 

  99. Bulik-Sullivan, B. K. et al. LD Score regression distinguishes confounding from polygenicity in genome-wide association studies. Nat. Genet. 47, 291–295 (2015).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  100. DeLuca, D. S. et al. RNA-SeQC: RNA-seq metrics for quality control and process optimization. Bioinformatics 28, 1530–1532 (2012).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  101. Finucane, H. K. et al. Heritability enrichment of specifically expressed genes identifies disease-relevant tissues and cell types. Nat. Genet. 50, 621–629 (2018).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  102. Hormozdiari, F. et al. Colocalization of GWAS and eQTL signals detects target genes. Am. J. Hum. Genet. 99, 1245–1260 (2016).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  103. Stacey, S. N. et al. A germline variant in the TP53 polyadenylation signal confers cancer susceptibility. Nat. Genet. 43, 1098–1103 (2011).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  104. Chahal, H. S. et al. Genome-wide association study identifies 14 novel risk alleles associated with basal cell carcinoma. Nat. Commun. 7, 12510 (2016).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  105. Ostrom, Q. T. et al. Sex-specific glioma genome-wide association study identifies new risk locus at 3p21.31 in females, and finds sex-differences in risk at 8q24.21. Sci. Rep. 8, 7352 (2018).

    Article  PubMed  PubMed Central  CAS  Google Scholar 

  106. Melin, B. S. et al. Genome-wide association study of glioma subtypes identifies specific differences in genetic susceptibility to glioblastoma and non-glioblastoma tumors. Nat. Genet. 49, 789–794 (2017).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  107. Zhou, Y., Wu, H., Zhao, M., Chang, C. & Lu, Q. The Bach family of transcription factors: a comprehensive review. Clin. Rev. Allergy Immunol. 50, 345–356 (2016).

    Article  CAS  PubMed  Google Scholar 

  108. Milovic-Holm, K., Krieghoff, E., Jensen, K., Will, H. & Hofmann, T. G. FLASH links the CD95 signaling pathway to the cell nucleus and nuclear bodies. EMBO J. 26, 391–401 (2007).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

Download references

Acknowledgements

NCI: This study was supported by the Intramural Research Program of the Division of Cancer Epidemiology and Genetics, National Cancer Institute (NCI), National Institutes of Health (NIH) and Department of Health and Human Services (DHHS). AOCS/OCAC/SEARCH: AOCS/OCAC/SEARCH is accessible via European Genome–Phenome Archive. We acknowledge their support and data, and the contribution of the study nurses, research assistants and all clinical and scientific collaborators in generation of these data. We also acknowledge their funding sources: OCAC (NIH grant no. U19CA148112), SEARCH team (Cancer Research UK grant no.C490/A16561), AOCS (US Army Medical Research and Material Command under grant no. DAMD17‐01‐1‐0729, The Cancer Council Victoria, Queensland Cancer Fund, The Cancer Council New South Wales, The Cancer Council South Australia, The Cancer Foundation of Western Australia, The Cancer Council Tasmania and the National Health and Medical Research Council of Australia (NHMRC) (grant nos. ID400413 and ID400281, as well as support from S. Boldeman, the Agar family, Ovarian Cancer Action (UK), Ovarian Cancer Australia and the Peter MacCallum Foundation). MelaNostrum Consortium: We thank the participants of the MelaNostrum Consortium from Italy (Genoa, L’Aquila, Rome, Padua, Milan, Florence and Bergamo), Spain (Valencia and Barcelona), Greece (Athens) and Cyprus (Nicosia) who provided data and biospecimens for this study. The Consortium is partially supported by the Intramural Research Program of the Division of Cancer Epidemiology and Genetics, NCI, NIH, DHHS. Funding for the University of Genoa and Genetics of Rare Cancers, Ospedale Policlinico San Martino came from Italian Ministry of Health 5 × 1000 per la Ricerca Corrente to Ospedale Policlinico San Martino and AIRC IG 15460. The research at the Melanoma Unit in Barcelona was supported by the Spanish Fondo de Investigaciones Sanitarias grant nos. PI15/00716 and PI15/00956 cofinanced by FEDER ‘Una manera de hacer Europa’; CIBER de Enfermedades Raras of the Instituto de Salud Carlos III, Spain, cofinanced by European Development Regional Fund ‘A way to achieve Europe’ ERDF; AGAUR 2014_SGR_603 of the Catalan Government, Spain; European Commission, contract no. LSHC-CT-2006-018702 (GenoMEL) and by the European Commission under the 7th Framework Programme, Diagnostics; ‘Fundació La Marató de TV3’ grant no. 201331-30, Catalonia, Spain; ‘Fundación Científica de la Asociación Española Contra el Cáncer’ grant no. GCB15152978SOEN, Spain, and CERCA Programme/Generalitat de Catalunya. Melanoma research at the Department of Dermatology, University of L’Aquila, Italy was supported by the Italian Ministry of the University and Scientific Research (PRIN-2012 grant no. 2012JJX494). Q-MEGA/QTWIN: The Q-MEGA/QTWIN study was supported by the Melanoma Research Alliance, the NIH NCI (grant nos. CA88363, CA83115, CA122838, CA87969, CA055075, CA100264, CA133996 and CA49449), the NHMRC (grant nos. 200071, 241944, 339462, 380385, 389927, 389875, 389891, 389892,389938, 443036, 442915, 442981, 496610, 496675, 496739, 552485, 552498 and APP1049894), the Cancer Councils New South Wales, Victoria and Queensland, the Cancer Institute New South Wales, the Cooperative Research Centre for Discovery of Genes for Common Human Diseases, Cerylid Biosciences (Melbourne), the Australian Cancer Research Foundation, The Wellcome Trust (grant no. WT084766/Z/08/Z) and donations from N. and S. Hawkins. S. MacGregor acknowledges fellowship support from the Australian National Health and Medical Research Council and from the Australian Research Council.

Please see the Supplementary Note for additional acknowledgments.

Author information

Authors and Affiliations

Authors

Consortia

Contributions

M.T.L., M.M.I. and M.H.L. conceptualized and designed the project. D.T.B., S. MacGregor, M.T.L. and S.J.C. provided funding support. M.T.L., D.T.B., S. MacGregor, M.J.M., J.S., M.M.I. and M.H.L. interpreted the results and supervized the study. M.J.M., M.T.L., M.M.I., K.B., J.C. and M.H.L. wrote the manuscript. J.S., M.M.I., K.B., T.Z., J.C., D.L.D. and M.H.L. analyzed the data. A.J. Stratigos, P. Ghiorzo, S.P. and E.N. coordinated the study and collected the data. M.T.L., D.T.B., S. MacGregor, M.J.M., A.J. Stratigos, P. Ghiorzo, M.B., D.C., J.C., M.C.F., T.Z., M.R., A.J.T., C.M., J. Martinez, A. Hadjisavvas, L.S., I.S., R.S., X.R.Y., A.M.G., M.P., K.P.K., L.P., P.Q., C.P., L.C., M.Z., P. Gimenez-Xavier, A.R., L.E., S. Manoukian, L.R., B.H.S., M.A.L., L.D.R., D.M., M. Mandala, K.K., L.A.A., C.I.A., P.A.A., M.A., E.A., H.P.S., V.B., B.D., L.M.B., K.P.B., W.V.C., V.C., J.E.C., T.D., M.F., S.F., E.F., S.S., P. Galan, Z.G., E.M.G., S.G., A.G., N.A.G., J. Hansson, M. Harland, J. Harris, P.H., A. Henders, M. Hočevar, V.H., D.H., C.I., R.K., J. Lang, G.M.L., J.E.L., X.L., J. Lubiński, R.M.M., M. Malt, J. Malvehy, K.M., H.M., A.M., E.K.M., R.E.N., S.N., D.R.N., H.O., N.O., L.G.F., J.A.P., A.A.Q., G.L.R., J.R., C. Requena, C. Rowe, N.J.S., M. Sanna, D.S., H.S., L.A.S., M. Smithers, F.S., A.J. Swerdlow, N.V.D.S., N.A.K., A. Visconti, L. Wallace, S.V.W., L. Wheeler, R.A.S., A. Hutchinson, K.J., M. Malasky, A. Vogt, W.Z., K.A.P., D.E.E., J. Han, B.H., N.K.H., P.A.K., C.B., G.W.M., C.M.O., C.H., A.M.D., N.G.M., E.E., G.J.M., G.L., P.D.P.P., D.F.E., J.H.B., A.E.C., G.A., D.L.D., D.C.W., H.G., A.D.N., M.A.T., J.A.N.B., K.P., S.J.C., K.M.B., F.D., S.P., E.N., J.S., M.M.I. and M.H.L. participated in data collection, results interpretation and manuscript review.

Corresponding authors

Correspondence to Maria Teresa Landi, Mark M. Iles or Matthew H. Law.

Ethics declarations

Competing interests

The authors declare no competing interests.

Additional information

Publisher’s note Springer Nature remains neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Extended data

Extended Data Fig. 1 Quantile-Quantile plot of total CM meta-analysis.

Quantile-quantile plots of negative log10 two-sided P value derived from a fixed-effects inverse-variance weighted meta-analysis of log(OR) effect-sizes derived from the logistic regression GWAS listed in Supplementary Table 1. All confirmed and self-report cases are included, with a total sample size of 36,760 melanoma cases and 375,188 controls.

Extended Data Fig. 2 Manhattan plots of melanoma risk loci from total and confirmed-only GWAS meta-analyses.

Negative log10 two-sided P value derived from a fixed-effects inverse-variance weighted meta-analysis of log(OR) effect-sizes derived from the logistic regression GWAS (y-axis) are plotted by their chromosome position. The confirmed-only analysis included 30,134 cases with histopathologically confirmed CM, and 81,415 controls. The total CM meta-analysis includes all confirmed and self-report cases, with a total sample size of 36,760 CM cases and 375,188 controls. Multiple-testing corrected genome-wide significance threshold was P<5×10−8. We display in order the total CM meta-analysis without limiting the y-axis; the pathologically confirmed CM cases only meta-analysis with the y-axis limited to 1×10−25 and without a limit to more clearly display loci other than MC1R.

Extended Data Fig. 3 Quantile-Quantile plot of confirmed-only CM meta-analysis.

Quantile-quantile plots of negative log10 two-sided P value derived from a fixed-effects inverse-variance weighted meta-analysis of log(OR) effect-sizes derived from the logistic regression GWAS listed in Supplementary Table 1. Only cases with histopathologically confirmed CM are included, with a total sample size of 30,134 melanoma cases and 81,415 controls.

Extended Data Fig. 4 Distribution of pigmentation polygenic risk scores across melanoma histological subtypes.

The figure shows whether PRS defined based on SNPs associated with hair color differ across CM histological types (Methods; SSM: superficial spreading melanoma; NM: nodular melanoma; LM: lentigo melanoma; Acral: acral lentiginous melanoma). The higher the PRS the lighter the hair color. When comparing subtype 1 vs. subtype 2, we report the effect size for the linear regression of PRS on subtype 1, including study and principal components as covariates to control for population stratification. The regression coefficient, 95% confidence interval, and statistical significance are shown. The positive beta indicates the PRS is higher in subtype 2 (for example, nonacral melanomas). This analysis included 9828 SSM, 2137 NM, 900 LM, 353 acral melanoma cases and 44676 controls. Two-sided t-statistic was used for testing significance. P values reported were not adjusted for multiple comparison.

Source data

Extended Data Fig. 5 LD score regression plots.

LD score regression was performed for the top 4000 (A) 2000 (B) and 1000 (C) tissue-specific genes from melanocyte and GTEx tissue types (v7 datasets), to assess the enrichment of melanoma heritability in these genomic regions using summary statistics from Total CM GWAS meta-analysis. The level of enrichment and P values are shown, with an FDR = 0.05 cutoff marked as a dashed horizontal line (See Methods for statistical test). Tissue categories are color-coded, and a subset of top individual tissue types are shown on the plot. Tissue types from “Skin” category including melanocytes are highlighted in magenta.

Source data

Extended Data Fig. 6 Effect sizes for confirmed-only meta-analysis versus UKBB self-report set.

For each independent genome-wide significant (P<5×10−8) lead SNP from the confirmed-only meta-analysis (30,134 melanoma cases and 81,415 controls), we plot on the Y-axis UK Biobank self-report GWAS (UKBB SR) log(OR) and standard error from a logistic regression GWAS (1,802 self-report CM cases and 7,208 controls) and on the X-axis we plot the log(OR) and standard error from a fixed-effects inverse-variance weighted meta-analysis of log(OR) effect-sizes derived from the logistic regression GWAS for confirmed melanoma cases listed in Supplementary Table 1. We also report the r2 correlation from the linear regression of UKBB SR log(OR) on the confirmed met-analysis estimates, weighted by their standard error.

Source data

Supplementary information

Supplementary Information

Supplementary Note and Figs. 1–9

Reporting Summary

Supplementary Tables 1–18

Supplementary Data 1

P value and regional LD for each lead SNP from Supplementary Table 3. Regional association P values and LD patterns for identified CM susceptibility regions. Top panel plots –log10 association P values from the total CM fixed-effects meta-analysis. Blue dot is the most significant regional variant and points above blue line indicate association P < 5× 10−8. Darker shades of red indicate higher LD with most significant regional variant. Bottom panel displays nearby genes. Legend applies to all subsequent plots.

Source data

Source Data Fig. 2

Statistical Source Data for Figure 2: Overlap of loci identified by primary and secondary analyses. Positions for each locus and 0,1 value for detected in the listed analyses as per Fig. 2.

Source Data Extended Data Fig. 4

Statistical source data

Source Data Extended Data Fig. 5

Statistical source data

Source Data Extended Data Fig. 6

Statistical source data

Source Data Supplementary Fig. 2

Statistical source data. A1 is the effect allele for the listed BETA etc.

Source Data Supplementary Fig. 3

Statistical source data

Rights and permissions

Reprints and permissions

About this article

Check for updates. Verify currency and authenticity via CrossMark

Cite this article

Landi, M.T., Bishop, D.T., MacGregor, S. et al. Genome-wide association meta-analyses combining multiple risk phenotypes provide insights into the genetic architecture of cutaneous melanoma susceptibility. Nat Genet 52, 494–504 (2020). https://doi.org/10.1038/s41588-020-0611-8

Download citation

  • Received:

  • Accepted:

  • Published:

  • Issue Date:

  • DOI: https://doi.org/10.1038/s41588-020-0611-8

This article is cited by

Search

Quick links

Nature Briefing

Sign up for the Nature Briefing newsletter — what matters in science, free to your inbox daily.

Get the most important science stories of the day, free in your inbox. Sign up for Nature Briefing