Shared heritability and functional enrichment across six solid cancers

Abstract

Quantifying the genetic correlation between cancers can provide important insights into the mechanisms driving cancer etiology. Using genome-wide association study summary statistics across six cancer types based on a total of 296,215 cases and 301,319 controls of European ancestry, here we estimate the pair-wise genetic correlations between breast, colorectal, head/neck, lung, ovary and prostate cancer, and between cancers and 38 other diseases. We observed statistically significant genetic correlations between lung and head/neck cancer (rg = 0.57, p = 4.6 × 10-8), breast and ovarian cancer (rg = 0.24, p = 7 × 10-5), breast and lung cancer (rg = 0.18, p =1.5 × 10-6) and breast and colorectal cancer (rg = 0.15, p = 1.1 × 10-4). We also found that multiple cancers are genetically correlated with non-cancer traits including smoking, psychiatric diseases and metabolic characteristics. Functional enrichment analysis revealed a significant excess contribution of conserved and regulatory regions to cancer heritability. Our comprehensive analysis of cross-cancer heritability suggests that solid tumors arising across tissues share in part a common germline genetic basis.

Conflict of interest statement

The authors declare no competing interests.

Figures

Fig. 1
Fig. 1
Estimates of SNP-heritability (hg2) and cross-cancer heritability (rg) for the six cancer types. SNP-heritability and cross-cancer heritability are calculated based on HapMap3 SNPs using LD score regression (LDSC). a The solid bar represents overall SNP hg2 on the liability scale, calculated based on all HapMap3 SNPs. The dark green bar represents hg2 calculated based on non-significant SNPs—the remaining SNPs after excluding genome-wide significant hits (p < 5 × 10−8) ± 500 kb. The black bar with density texture indicates proportion of hg2 (as reflected by the percentages displayed on top of each bar) that could be explained by top hits ±500 kb surrounded areas. The orange error bars represent 95% confidence intervals. b The solid blue bar represents overall SNP hg2 in liability scale (no SNP exclusion), with black error bars indicating 95% confidence intervals. The red short lines correspond to classical estimates of h2 measured in a twin study of Scandinavian countries (Mucci et al.). c Genetic correlations between cancers. Estimates withstood Bonferroni corrections (p < 0.05/15) are marked with double asterisk (**), and nominal significant results (p < 0.05) are marked with single asterisk (*)
Fig. 2
Fig. 2
Local genetic correlation between breast, lung and prostate cancer. The region-specific p-values for the local genetic covariance for breast and prostate cancer are shown in a, and for lung and prostate cancer in b. Each dot presents a specific genomic region. In the QQ plots, red color indicates significance after multiple corrections (p < 0.05/1703 regions compared), and blue color indicates nominal significance (p < 0.05/15 pairs of cancers compared). Manhattan-style plots showing the estimates of local genetic covariance for breast and prostate cancer (c), and for lung and prostate cancer (d). Although breast and prostate cancer only show modest genome-wide genetic correlation, two loci exhibit significant local genetic covariance. Similarly, albeit the negligible overall genetic correlation for lung and prostate cancer, three loci present significant local genetic covariance. In the Manhattan plots, red color indicates even number chromosomes and blue color indicates odd number chromosomes
Fig. 3
Fig. 3
Cross-trait genetic correlation (rg) analysis between cancers and non-cancer traits. The traits were divided into four categories: a Common phenotypes, b Metabolic or cardiovascular related traits, c Psychiatric traits, d Autoimmune inflammatory diseases. Pair-wise genetic correlations withstood Bonferroni corrections (228 tests) are marked with double asterisk (**), with estimates of correlation shown in the cells. Pair-wise genetic correlations with significance at p < 0.01 are marked with a single asterisk (*). The color of cells represents the magnitude of correlation
Fig. 4
Fig. 4
Putative directional relationships between cancers and traits. For each cancer–trait pair identified as candidates to be related in a causal manner, the plots show trait-specific effect sizes (beta coefficients) of the included genetic variants. Gray lines represent the relevant standard errors. a HDL and breast cancer. Trait-specific effect sizes for HDL and breast cancer are shown for SNPs associated with HDL levels (left) and breast cancer (right). b Schizophrenia and breast cancer. Trait-specific effect sizes for schizophrenia and breast cancer are shown for SNPs associated with schizophrenia (left) and breast cancer (right). c Age at natural menopause and breast cancer. Trait-specific effect sizes for age at natural menopause and breast cancer are shown for SNPs associated with age at natural menopause (left) and breast cancer (right). d Lupus and prostate cancer. Trait-specific effect sizes for lupus and prostate cancer are shown for SNPs associated with lupus (left) and prostate cancer (right)
Fig. 5
Fig. 5
Enrichment p-values of 24 non-cell-type-specific functional categories over six cancer types. The x-axis represents each of the 24 functional categories, y-axis represents log-transformed p-values of enrichment. Annotations with statistical significance after Bonferroni corrections (p < 0.05/24) were plotted in orange, otherwise blue. The horizontal gray dash line indicates p-threshold of 0.05; horizontal red dash line indicates p-threshold of 0.05/24. From top to bottom are six panels representing six cancers: breast cancer, colorectal cancer, head/neck cancer, lung cancer, ovarian cancer, and prostate cancer. TSS transcription start site, UTR untranslated region, TFBS transcription factor binding sites, DHS DNase I hypersensitive sites, DGF digital genomic foot printing, CTCF CCCTC-binding factor

References

    1. Lichtenstein P, et al. Environmental and heritable factors in the causation of cancer–analyses of cohorts of twins from Sweden, Denmark, and Finland. N. Engl. J. Med. 2000;343:78–85. doi: 10.1056/NEJM200007133430201.
    1. Mucci LA, et al. Familial risk and heritability of cancer among twins in nordic countries. JAMA. 2016;315:68. doi: 10.1001/jama.2015.17703.
    1. Polderman TJC, et al. Meta-analysis of the heritability of human traits based on fifty years of twin studies. Nat. Genet. 2015;47:702–709. doi: 10.1038/ng.3285.
    1. Amundadottir LT, et al. Cancer as a complex phenotype: pattern of cancer distribution within and beyond the nuclear family. PLoS Med. 2004;1:e65. doi: 10.1371/journal.pmed.0010065.
    1. Yu H, Frank C, Sundquist J, Hemminki A, Hemminki K. Common cancers share familial susceptibility: implications for cancer genetics and counselling. J. Med. Genet. 2017;54:248–253. doi: 10.1136/jmedgenet-2016-103932.
    1. Frank C, Sundquist J, Yu H, Hemminki A, Hemminki K. Concordant and discordant familial cancer: familial risks, proportions and population impact. Int. J. Cancer. 2017;140:1510–1516. doi: 10.1002/ijc.30583.
    1. Fehringer G, et al. Cross-cancer genome-wide analysis of lung, ovary, breast, prostate, and colorectal cancer reveals novel pleiotropic associations. Cancer Res. 2016;76:5103–5114. doi: 10.1158/0008-5472.CAN-15-2980.
    1. Kar SP, et al. Genome-wide meta-analyses of breast, ovarian, and prostate cancer association studies identify multiple new susceptibility loci shared by at least two cancer types. Cancer Discov. 2016;6:1052–1067. doi: 10.1158/-15-1227.
    1. Sampson JN, et al. Analysis of heritability and shared heritability based on genome-wide association studies for thirteen cancer types. J. Natl. Cancer Inst. 2015;107:djv279. doi: 10.1093/jnci/djv279.
    1. Gusev A, et al. Atlas of prostate cancer heritability in European and African-American men pinpoints tissue-specific regulation. Nat. Commun. 2016;7:10979. doi: 10.1038/ncomms10979.
    1. Jiao S, et al. Estimating the heritability of colorectal cancer. Hum. Mol. Genet. 2014;23:3898–3905. doi: 10.1093/hmg/ddu087.
    1. Lu Y, et al. Most common ‘sporadic’ cancers have a significant germline genetic component. Hum. Mol. Genet. 2014;23:6112–6118. doi: 10.1093/hmg/ddu312.
    1. Yang J, Lee SH, Goddard ME, Visscher PM. GCTA: a tool for genome-wide complex trait analysis. Am. J. Hum. Genet. 2011;88:76–82. doi: 10.1016/j.ajhg.2010.11.011.
    1. Finucane HK, et al. Partitioning heritability by functional annotation using genome-wide association summary statistics. Nat. Genet. 2015;47:1228–1235. doi: 10.1038/ng.3404.
    1. Bulik-Sullivan BK, et al. LD Score regression distinguishes confounding from polygenicity in genome-wide association studies. Nat. Genet. 2015;47:291–295. doi: 10.1038/ng.3211.
    1. Lindström S, et al. Quantifying the genetic correlation between multiple cancer types. Cancer Epidemiol. Biomark. Prev. 2017;26:1427–1435. doi: 10.1158/1055-9965.EPI-17-0211.
    1. Yang J, et al. Common SNPs explain a large proportion of the heritability for human height. Nat. Genet. 2010;42:565–569. doi: 10.1038/ng.608.
    1. Amos CI, et al. The OncoArray Consortium: a network for understanding the genetic architecture of common cancers. Cancer Epidemiol. Biomark. Prev. 2017;26:126–135. doi: 10.1158/1055-9965.EPI-16-0106.
    1. SMOKING and health. Joint report of the Study Group on Smoking and Health. Science. 1957;125:1129–1133. doi: 10.1126/science.125.3258.1129.
    1. Shaw R, Beasley N. Aetiology and risk factors for head and neck cancer: United Kingdom National Multidisciplinary Guidelines. J. Laryngol. Otol. 2016;130:S9–S12. doi: 10.1017/S0022215116000360.
    1. Koene RJ, Prizment AE, Blaes A, Konety SH. Shared risk factors in cardiovascular disease and cancer. Circulation. 2016;133:1104–1114. doi: 10.1161/CIRCULATIONAHA.115.020406.
    1. Thompson CL, et al. Short duration of sleep increases risk of colorectal adenoma. Cancer. 2011;117:841–847. doi: 10.1002/cncr.25507.
    1. Sigurdardottir LG, et al. Sleep disruption among older men and risk of prostate cancer. Cancer Epidemiol. Prev. Biomark. 2013;22:872–879. doi: 10.1158/1055-9965.EPI-12-1227-T.
    1. Gazal S, et al. Linkage disequilibrium-dependent architecture of human complex traits shows action of negative selection. Nat. Genet. 2017;49:1421–1427. doi: 10.1038/ng.3954.
    1. Field RW, Withers BL. Occupational and environmental causes of lung cancer. Clin. Chest Med. 2012;33:681–703. doi: 10.1016/j.ccm.2012.07.001.
    1. Hulka BS. Epidemiologic analysis of breast and gynecologic cancers. Prog. Clin. Biol. Res. 1997;396:17–29.
    1. Gaudet MM, et al. Pooled analysis of active cigarette smoking and invasive breast cancer risk in 14 cohort studies. Int. J. Epidemiol. 2017;46:881–893.
    1. Cancer Genome Atlas Network. Comprehensive molecular portraits of human breast tumours. Nature. 2012;490:61–70. doi: 10.1038/nature11412.
    1. Maas P, et al. Breast cancer risk from modifiable and nonmodifiable risk factors among white women in the United States. JAMA Oncol. 2016;2:1295–1302. doi: 10.1001/jamaoncol.2016.1025.
    1. Nakaya N, et al. Personality traits and cancer risk and survival based on finnish and swedish registry data. Am. J. Epidemiol. 2010;172:377–385. doi: 10.1093/aje/kwq046.
    1. Oksbjerg Dalton S, Munk Laursen T, Mellemkjaer L, Johansen C, Mortensen PB. Schizophrenia and the risk for breast cancer. Schizophr. Res. 2003;62:89–92. doi: 10.1016/S0920-9964(02)00430-9.
    1. Hung RJ, et al. A susceptibility locus for lung cancer maps to nicotinic acetylcholine receptor subunit genes on 15q25. Nature. 2008;452:633–637. doi: 10.1038/nature06885.
    1. Amos CI, et al. Genome-wide association scan of tag SNPs identifies a susceptibility locus for lung cancer at 15q25.1. Nat. Genet. 2008;40:616–622. doi: 10.1038/ng.109.
    1. Gao C, et al. Mendelian randomization study of adiposity-related traits and risk of breast, ovarian, prostate, lung and colorectal cancer. Int. J. Epidemiol. 2016;45:896–908. doi: 10.1093/ije/dyw129.
    1. Collaborative Group on Hormonal Factors in Breast Cancer. Menarche, menopause, and breast cancer risk: individual participant meta-analysis, including 118 964 women with breast cancer from 117 epidemiological studies. Lancet Oncol. 2012;13:1141–1151. doi: 10.1016/S1470-2045(12)70425-4.
    1. Day FR, et al. Large-scale genomic analyses link reproductive aging to hypothalamic signaling, breast cancer susceptibility and BRCA1-mediated DNA repair. Nat. Genet. 2015;47:1294–1303. doi: 10.1038/ng.3412.
    1. Day FR, et al. Genomic analyses identify hundreds of variants associated with age at menarche and support a role for puberty timing in cancer risk. Nat. Genet. 2017;49:834–841. doi: 10.1038/ng.3841.
    1. Zack TI, et al. Pan-cancer patterns of somatic copy number alteration. Nat. Genet. 2013;45:1134–1140. doi: 10.1038/ng.2760.
    1. Ciriello G, et al. Emerging landscape of oncogenic signatures across human cancers. Nat. Genet. 2013;45:1127–1133. doi: 10.1038/ng.2762.
    1. Lindblad-Toh K, et al. A high-resolution map of human evolutionary constraint using 29 mammals. Nature. 2011;478:476–482. doi: 10.1038/nature10530.
    1. Calin GA, et al. Ultraconserved regions encoding ncRNAs are altered in human leukemias and carcinomas. Cancer Cell. 2007;12:215–229. doi: 10.1016/j.ccr.2007.07.027.
    1. Peng JC, Shen J, Ran ZH. Transcribed ultraconserved region in human cancers. RNA Biol. 2013;10:1771–1777. doi: 10.4161/rna.26995.
    1. Michailidou K, et al. Association analysis identifies 65 new breast cancer risk loci. Nature. 2017;551:92–94. doi: 10.1038/nature24284.
    1. Milne RL, et al. Identification of ten variants associated with risk of estrogen-receptor-negative breast cancer. Nat. Genet. 2017;49:1767–1778. doi: 10.1038/ng.3785.
    1. McKay JD, et al. Large-scale association analysis identifies new lung cancer susceptibility loci and heterogeneity in genetic susceptibility across histological subtypes. Nat. Genet. 2017;49:1126–1132. doi: 10.1038/ng.3892.
    1. Schmit, S. L. et al. Novel Common Genetic Susceptibility Loci for Colorectal Cancer. J. Natl. Cancer Inst. 111, djy099 (2019).
    1. Phelan CM, et al. Identification of 12 new susceptibility loci for different histotypes of epithelial ovarian cancer. Nat. Genet. 2017;49:680–691. doi: 10.1038/ng.3826.
    1. Lesseur C, et al. Genome-wide association analyses identify new susceptibility loci for oral cavity and pharyngeal cancer. Nat. Genet. 2016;48:1544–1550. doi: 10.1038/ng.3685.
    1. Schumacher FR, et al. Association analyses of more than 140,000 men identify 63 new prostate cancer susceptibility loci. Nat. Genet. 2018;50:928–936. doi: 10.1038/s41588-018-0142-8.
    1. Shi H, Mancuso N, Spendlove S, Pasaniuc B. Local genetic correlation gives insights into the shared genetic architecture of complex traits. Am. J. Hum. Genet. 2017;101:737–751. doi: 10.1016/j.ajhg.2017.09.022.
    1. Sakoda LC, Jorgenson E, Witte JS. Turning of COGS moves forward findings for hormonally mediated cancers. Nat. Genet. 2013;45:345–348. doi: 10.1038/ng.2587.
    1. Pickrell JK, et al. Detection and interpretation of shared genetic influences on 42 human traits. Nat. Genet. 2016;48:709–717. doi: 10.1038/ng.3570.
    1. Pickrell JK. Joint analysis of functional genomic data and genome-wide association studies of 18 human traits. Am. J. Hum. Genet. 2014;94:559–573. doi: 10.1016/j.ajhg.2014.03.004.
    1. Bowden J, Davey Smith G, Burgess S. Mendelian randomization with invalid instruments: effect estimation and bias detection through Egger regression. Int. J. Epidemiol. 2015;44:512–525. doi: 10.1093/ije/dyv080.
    1. Gusev A, et al. Partitioning heritability of regulatory and cell-type-specific variants across 11 common diseases. Am. J. Hum. Genet. 2014;95:535–552. doi: 10.1016/j.ajhg.2014.10.004.
    1. Roadmap Epigenomics Consortium. et al. Integrative analysis of 111 reference human epigenomes. Nature. 2015;518:317–330. doi: 10.1038/nature14248.
    1. ENCODE Project Consortium. An integrated encyclopedia of DNA elements in the human genome. Nature. 2012;489:57–74. doi: 10.1038/nature11247.
    1. Trynka G, et al. Chromatin marks identify critical cell types for fine mapping complex trait variants. Nat. Genet. 2013;45:124–130. doi: 10.1038/ng.2504.
    1. Hnisz D, et al. Super-enhancers in the control of cell identity and disease. Cell. 2013;155:934–947. doi: 10.1016/j.cell.2013.09.053.
    1. Schizophrenia Working Group of the Psychiatric Genomics Consortium. Biological insights from 108 schizophrenia-associated genetic loci. Nature. 2014;511:421–427. doi: 10.1038/nature13595.
    1. Hoffman MM, et al. Integrative annotation of chromatin elements from ENCODE data. Nucleic Acids Res. 2013;41:827–841. doi: 10.1093/nar/gks1284.
    1. Ward LD, Kellis M. Evidence of abundant purifying selection in humans for recently acquired regulatory functions. Science. 2012;337:1675–1678. doi: 10.1126/science.1225057.
    1. Andersson R, et al. An atlas of active enhancers across human cell types and tissues. Nature. 2014;507:455–461. doi: 10.1038/nature12787.

Source: PubMed

3
Subskrybuj