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.

Meta-analysis of genome-wide association studies for cattle stature identifies common genes that regulate body size in mammals

Subjects

Abstract

Stature is affected by many polymorphisms of small effect in humans1. In contrast, variation in dogs, even within breeds, has been suggested to be largely due to variants in a small number of genes2,3. Here we use data from cattle to compare the genetic architecture of stature to those in humans and dogs. We conducted a meta-analysis for stature using 58,265 cattle from 17 populations with 25.4 million imputed whole-genome sequence variants. Results showed that the genetic architecture of stature in cattle is similar to that in humans, as the lead variants in 163 significantly associated genomic regions (P < 5 × 10−8) explained at most 13.8% of the phenotypic variance. Most of these variants were noncoding, including variants that were also expression quantitative trait loci (eQTLs) and in ChIP–seq peaks. There was significant overlap in loci for stature with humans and dogs, suggesting that a set of common genes regulates body size in mammals.

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 meta-analysis of bovine stature with n = 58,265 animals.
Fig. 2: Validation of lead variants.
Fig. 3: Haplotype diversity for 15 cattle breeds in two genomic regions (NCAPGLCORL and PLAG1) where selection signatures match segregation of QTLs for stature.

Similar content being viewed by others

References

  1. Wood, A. R. et al. Defining the role of common variation in the genomic and biological architecture of adult human height. Nat. Genet. 46, 1173–1186 (2014).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  2. Rimbault, M. et al. Derived variants at six genes explain nearly half of size reduction in dog breeds. Genome Res. 23, 1985–1995 (2013).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  3. Hayward, J. J. et al. Complex disease and phenotype mapping in the domestic dog. Nat. Commun. 7, 10460 (2016).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  4. Daetwyler, H. D. et al. Whole-genome sequencing of 234 bulls facilitates mapping of monogenic and complex traits in cattle. Nat. Genet. 46, 858–865 (2014).

    Article  CAS  PubMed  Google Scholar 

  5. Kang, H. M. et al. Variance component model to account for sample structure in genome-wide association studies. Nat. Genet. 42, 348–354 (2010).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  6. Yang, J. et al. Common SNPs explain a large proportion of the heritability for human height. Nat. Genet. 42, 565–569 (2010).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  7. Karim, L. et al. Variants modulating the expression of a chromosome domain encompassing PLAG1 influence bovine stature. Nat. Genet. 43, 405–413 (2011).

    Article  CAS  PubMed  Google Scholar 

  8. Pryce, J. E., Hayes, B. J., Bolormaa, S. & Goddard, M. E. Polymorphic regions affecting human height also control stature in cattle. Genetics 187, 981–984 (2011).

    Article  PubMed  PubMed Central  Google Scholar 

  9. Fortes, M. R. et al. Evidence for pleiotropism and recent selection in the PLAG1 region in Australian Beef cattle. Anim. Genet. 44, 636–647 (2013).

    Article  CAS  PubMed  Google Scholar 

  10. Pausch, H. et al. Meta-analysis of sequence-based association studies across three cattle breeds reveals 25 QTL for fat and protein percentages in milk at nucleotide resolution. BMC Genomics 18, 853 (2017).

  11. Li, Z. et al. An HMGA2–IGF2BP2 axis regulates myoblast proliferation and myogenesis. Dev. Cell 23, 1176–1188 (2012).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  12. Fernandes, I. et al. Ligand-dependent nuclear receptor corepressor LCoR functions by histone deacetylase–dependent and –independent mechanisms. Mol. Cell 11, 139–150 (2003).

    Article  CAS  PubMed  Google Scholar 

  13. Calderon, M. R. et al. Ligand-dependent corepressor (LCoR) recruitment by Kruppel-like factor 6 (KLF6) regulates expression of the cyclin-dependent kinase inhibitor CDKN1A gene. J. Biol. Chem. 287, 8662–8674 (2012).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  14. Calderon, M. R. et al. Ligand-dependent corepressor contributes to transcriptional repression by C2H2 zinc-finger transcription factor ZBRK1 through association with KRAB-associated protein-1. Nucleic Acids Res. 42, 7012–7027 (2014).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  15. Kemper, K. E., Visscher, P. M. & Goddard, M. E. Genetic architecture of body size in mammals. Genome Biol. 13, 244 (2012).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  16. 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 

  17. Villar, D. et al. Enhancer evolution across 20 mammalian species. Cell 160, 554–566 (2015).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  18. GTEx Consortium. The Genotype-Tissue Expression (GTEx) pilot analysis: multitissue gene regulation in humans. Science 348, 648–660 (2015).

    Article  PubMed Central  Google Scholar 

  19. Zhu, Z. et al. Integration of summary data from GWAS and eQTL studies predicts complex trait gene targets. Nat. Genet. 48, 481–487 (2016).

    Article  CAS  PubMed  Google Scholar 

  20. Akhtar, M. et al. Cell type and context-specific function of PLAG1 for IGF2 P3 promoter activity. Int. J. Oncol. 41, 1959–1966 (2012).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  21. DeChiara, T. M., Efstratiadis, A. & Robertson, E. J. A growth-deficiency phenotype in heterozygous mice carrying an insulin-like growth factor II gene disrupted by targeting. Nature 345, 78–80 (1990).

    Article  CAS  PubMed  Google Scholar 

  22. Voz, M. L., Agten, N. S., Van de Ven, W. J. & Kas, K. PLAG1, the main translocation target in pleomorphic adenoma of the salivary glands, is a positive regulator of IGF-II. Cancer Res. 60, 106–113 (2000).

    CAS  PubMed  Google Scholar 

  23. Nielsen, J. et al. A family of insulin-like growth factor II mRNA-binding proteins represses translation in late development. Mol. Cell. Biol. 19, 1262–1270 (1999).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  24. Reik, W. et al. Igf2 imprinting in development and disease. Int. J. Dev. Biol. 44, 145–150 (2000).

    CAS  PubMed  Google Scholar 

  25. Sutter, N. B. et al. A single IGF1 allele is a major determinant of small size in dogs. Science 316, 112–115 (2007).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  26. Park, S. D. et al. Genome sequencing of the extinct Eurasian wild aurochs, Bos primigenius, illuminates the phylogeography and evolution of cattle. Genome Biol. 16, 234 (2015).

    Article  PubMed  PubMed Central  Google Scholar 

  27. Clutton-Brock, J. A Natural History of Domesticated Mammals (Cambridge University Press, Cambridge, UK, 1987).

    Google Scholar 

  28. Vretemark, M. From Bones to Livestock (Stockholm University, City, 1997).

    Google Scholar 

  29. Manning, K., Timpson, A., Shennan, S. & Crema, E. Size reduction in early European domestic cattle relates to intensification of Neolithic herding strategies. PLoS One 10, e0141873 (2015).

    Article  PubMed  PubMed Central  Google Scholar 

  30. Svensson, E. M. et al. Tracing genetic change over time using nuclear SNPs in ancient and modern cattle. Anim. Genet. 38, 378–383 (2007).

    Article  CAS  PubMed  Google Scholar 

  31. Krogmeier, D. Zusammenhänge zwischen Nutzungsdauer und Körpergröße unter besonderer Berücksichtigung des Stallsystems bei Braunvieh und Fleckvieh. Zuchtungskunde 81, 328–340 (2009).

    Google Scholar 

  32. Beavers, L. & Van Doormaal, B. A closer look at stature (CDN Report) (Publisher, City, 2016).

  33. Laumay, A. & le Mezec, P. Bilan de l’indexation des races bovines laitieres. Resultats de la campagne 2014 (INRA Report 0015202017) (Publisher, City, 2015).

  34. Bonhomme, M. et al. Detecting selection in population trees: the Lewontin and Krakauer test extended. Genetics 186, 241–262 (2010).

    Article  PubMed  PubMed Central  Google Scholar 

  35. Fariello, M. I., Boitard, S., Naya, H., SanCristobal, M. & Servin, B. Detecting signatures of selection through haplotype differentiation among hierarchically structured populations. Genetics 193, 929–941 (2013).

    Article  PubMed  PubMed Central  Google Scholar 

  36. Goddard, M. E., Kemper, K. E., MacLeod, I. M., Chamberlain, A. J. & Hayes, B. J. Genetics of complex traits: prediction of phenotype, identification of causal polymorphisms and genetic architecture. Proc. Biol. Sci. 283, 1835 (2016).

    Article  Google Scholar 

  37. 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 

  38. Arthur, P. F., Parnell, P. F. & Richardson, E. C. Correlated responses in calf body weight and size to divergent selection for yearling growth rate in Angus cattle. Livest. Prod. Sci. 49, 305–312 (1997).

    Article  Google Scholar 

  39. Arango, J. A., Cundiff, L. V. & Van Vleck, L. D. Comparisons of Angus-, Braunvieh-, Chianina-, Hereford-, Gelbvieh-, Maine Anjou–, and Red Poll–sired cows for weight, weight adjusted for body condition score, height, and body condition score. J. Anim. Sci. 80, 3133–3141 (2002).

    Article  CAS  PubMed  Google Scholar 

  40. Arango, J. A., Cundiff, L. V. & Van Vleck, L. D. Breed comparisons of Angus, Charolais, Hereford, Jersey, Limousin, Simmental, and South Devon for weight, weight adjusted for body condition score, height, and body condition score of cows. J. Anim. Sci. 80, 3123–3132 (2002).

    Article  CAS  PubMed  Google Scholar 

  41. Arango, J. A., Cundiff, L. V. & Van Vleck, L. D. Comparisons of Angus, Charolais, Galloway, Hereford, Longhorn, Nellore, Piedmontese, Salers, and Shorthorn breeds for weight, weight adjusted for condition score, height, and condition score of cows. J. Anim. Sci. 82, 74–84 (2004).

    Article  CAS  PubMed  Google Scholar 

  42. Erbe, M. et al. Improving accuracy of genomic predictions within and between dairy cattle breeds with imputed high-density single nucleotide polymorphism panels. J. Dairy Sci. 95, 4114–4129 (2012).

    Article  CAS  PubMed  Google Scholar 

  43. Browning, B. L. & Browning, S. R. Genotype imputation with millions of reference samples. Am. J. Hum. Genet. 98, 116–126 (2016).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  44. Li, Y., Willer, C. J., Ding, J., Scheet, P. & Abecasis, G. R. MaCH: using sequence and genotype data to estimate haplotypes and unobserved genotypes. Genet. Epidemiol. 34, 816–834 (2010).

    Article  PubMed  PubMed Central  Google Scholar 

  45. Sargolzaei, M., Chesnais, J. P. & Schenkel, F. S. A new approach for efficient genotype imputation using information from relatives. BMC Genomics 15, 478 (2014).

    Article  PubMed  PubMed Central  Google Scholar 

  46. Harris, R.S. Improved Pairwise Alignment of GenomicDNA. PhD thesis, Pennsylvania State Univ. (2007).

  47. Rocha, D., Billerey, C., Samson, F., Boichard, D. & Boussaha, M. Identification of the putative ancestral allele of bovine single-nucleotide polymorphisms. J. Anim. Breed. Genet. 131, 483–486 (2014).

    Article  CAS  PubMed  Google Scholar 

  48. Zimin, A. V. et al. A whole-genome assembly of the domestic cow, Bos taurus. Genome Biol. 10, R42 (2009).

    Article  PubMed  PubMed Central  Google Scholar 

  49. Kim, D. et al. TopHat2: accurate alignment of transcriptomes in the presence of insertions, deletions and gene fusions. Genome Biol. 14, R36 (2013).

    Article  PubMed  PubMed Central  Google Scholar 

  50. Anders, S., Pyl, P. T. & Huber, W. HTSeq—a Python framework to work with high-throughput sequencing data. Bioinformatics 31, 166–169 (2015).

    Article  CAS  PubMed  Google Scholar 

  51. Anders, S. & Huber, W. Differential expression analysis for sequence count data. Genome Biol. 11, R106 (2010).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  52. Gilmour, A. R., Gogel, B., Cullis, B., Thompson, R. & Butler, D. ASReml User Guide Release 3.0. (Hemel Hempstead: VSN International, Stockholm, 2009).

    Google Scholar 

  53. Sun, L., Craiu, R. V., Paterson, A. D. & Bull, S. B. Stratified false discovery control for large-scale hypothesis testing with application to genome-wide association studies. Genet. Epidemiol. 30, 519–530 (2006).

    Article  PubMed  Google Scholar 

  54. Blott, S. et al. Molecular dissection of a quantitative trait locus: a phenylalanine-to-tyrosine substitution in the transmembrane domain of the bovine growth hormone receptor is associated with a major effect on milk yield and composition. Genetics 163, 253–266 (2003).

    CAS  PubMed  PubMed Central  Google Scholar 

  55. Grisart, B. et al. Positional candidate cloning of a QTL in dairy cattle: identification of a missense mutation in the bovine DGAT1 gene with major effect on milk yield and composition. Genome Res. 12, 222–231 (2002).

    Article  CAS  PubMed  Google Scholar 

  56. Cohen-Zinder, M. et al. Identification of a missense mutation in the bovine ABCG2 gene with a major effect on the QTL on chromosome 6 affecting milk yield and composition in Holstein cattle. Genome Res. 15, 936–944 (2005).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

Download references

Acknowledgements

H.D.D., A.J.C., P.J.B. and B.J.H. would like to acknowledge the Dairy Futures Cooperative Research Centre for funding. H.P. and R.F. acknowledge funding from the German Federal Ministry of Education and Research (BMBF) within the AgroClustEr ‘Synbreed—Synergistic Plant and Animal Breeding’ (grant 0315527B). H.P., R.F., R.E. and K.-U.G. acknowledge the Arbeitsgemeinschaft Süddeutscher Rinderzüchter, the Arbeitsgemeinschaft Österreichischer Fleckviehzüchter and ZuchtData EDV Dienstleistungen for providing genotype data. A. Bagnato acknowledges the European Union (EU) Collaborative Project LowInputBreeds (grant agreement 222623) for providing Brown Swiss genotypes. Braunvieh Schweiz is acknowledged for providing Brown Swiss phenotypes. H.P. and R.F. acknowledge the German Holstein Association (DHV) and the Confederación de Asociaciones de Frisona Española (CONCAFE) for sharing genotype data. H.P. was financially supported by a postdoctoral fellowship from the Deutsche Forschungsgemeinschaft (DFG) (grant PA 2789/1-1). D.B. and D.C.P. acknowledge funding from the Research Stimulus Fund (11/S/112) and Science Foundation Ireland (14/IA/2576). M.S. and F.S.S. acknowledge the Canadian Dairy Network (CDN) for providing the Holstein genotypes. P.S. acknowledges funding from the Genome Canada project entitled ‘Whole Genome Selection through Genome Wide Imputation in Beef Cattle’ and acknowledges WestGrid and Compute/Calcul Canada for providing computing resources. J.F.T. was supported by the National Institute of Food and Agriculture, US Department of Agriculture, under awards 2013-68004-20364 and 2015-67015-23183. A. Bagnato, F.P., M.D. and J.W. acknowledge EU Collaborative Project Quantomics (grant 516 agreement 222664) for providing Brown Swiss and Finnish Ayrshire sequences and genotypes. A.C.B. and R.F.V. acknowledge funding from the public–private partnership ‘Breed4Food’ (code BO-22.04-011-001-ASG-LR) and EU FP7 IRSES SEQSEL (grant 317697). A.C.B. and R.F.V. acknowledge CRV (Arnhem, the Netherlands) for providing data on Dutch and New Zealand Holstein and Jersey bulls.

Author information

Authors and Affiliations

Authors

Contributions

A.C.B. conducted the meta-analysis and contributed to writing the manuscript. H.D.D., A.J.C. and C.V.J. ran the 1000 Bull Genomes pipeline and extracted sequence variants, and A.J.C. and C.V.J. performed the eQTL analysis. C.H.P. sourced samples for miniature cattle and generated whole-genome sequence alignments for these. M.S., D.P.B., P.J.B. and F.S.S. contributed to genotype imputation and writing of the manuscript. M.S., F.S.S., G.S., D.C.P., H.P., J.V., B. Gredler, J.J.C., J.L.H. and R.F.B. performed GWAS analysis. S.B., B.S. and M.D. performed selection signature analysis. R.E. and K.-U.G. prepared daughter yield deviations and yield deviaitons of Fleckvieh animals, and the Intergenomics Consortium contributed genotypes. A.G.-G., C.H., M.-P.S., A.C., T.T., A. Bieber, P.C. and A. Barbat prepared phenotypes and genotypes for French cattle and ran GWAS. M.F., I.R. and J.S. prepared phenotypes and genotypes for Swiss and Austrian cattle and ran GWAS. A.A.E.V. contributed to across-species identification of stature-related genes. M.B., M.W., P.S., D.R., V.J. and R.D.S. performed variant annotation. B.J.H., D.J.G., J.F.T., C.B., J.R., A. Bagnato, F.P., B.T., L.-E.H., C.D., R.F., C.P.V.T., R.F.V., D.B., P.S., M.E.G., B. Guldbrandtsen and M.S.L. conceived the experimental design, analyzed stature data for contributed breeds and wrote the manuscript.

Corresponding author

Correspondence to Ben J. Hayes.

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.

Integrated supplementary information

Supplementary Figure 1 Accuracy of imputing sequence variants in cattle populations.

The accuracy of imputing sequence variant genotypes was assessed in 61 Jersey, 59 Brown Swiss, 187 Fleckvieh, 93 Angus and 312 Holstein animals from 1000 Bull Genomes Run4. a, A principal-components analysis of the genomic relationship matrix that was built using 1,147 animals and 20,211,021 autosomal sequence variants with minor allele frequency greater than 0.01 clearly separated the animals by breed. The genomic relationship matrix was built using PLINK1.9, and GCTA6 was used to perform the principal-components analysis. (b) The overall accuracy of imputing sequence variant genotypes was assessed for 2,231,805, 1,036,588 and 622,011 sequence variants, respectively, on chromosomes 1, 20 and 25 using FImpute45 and Minimac44. The accuracy of imputing genotypes was calculated by 15-fold cross-validation, where the sequence genotypes of 25 randomly selected animals per breed from 1000 Bull Genomes Run4 were masked to those on the Illumina BovineHD BeadChip and all the sequence variants were then imputed using FImpute or Minimac with all other sequences (n = 1,122) as a reference. The random selection of animals was performed 15 times for each chromosome and breed analysis. The squared correlation (r2) between the real sequence variant called genotypes and the imputed variant genotypes was taken as the accuracy of imputation. Box plots show the overall accuracy of imputing sequence variants per breed using FImpute or Minimac. The interquartile range defines the height of the box, and whiskers extend to 1.5 times the interquartile range. In each box, the median is represented by the horizontal black bar. c,d, Accuracy of imputing sequence variant genotypes using Minimac (c) or FImpute (d) by minor allele frequency. The dotted and solid lines in c represent the squared correlation between true genotypes (0, 1 or 2) and imputed dosages and best-guess genotypes, respectively.

Supplementary Figure 2 The accuracy of imputing sequence variant genotypes assessed in Montebeliarde, Normande and Danish Red cattle using 1000 Bull Genomes Run4.

a, Accuracy was assessed for sequence variants on chromosomes 1, 20 and 25 using FImpute. The accuracy of imputing genotypes was calculated by masking the sequence genotypes of 14 randomly selected animals per breed from 1000 Bull Genomes Run4 to those on the Illumina BovineHD BeadChip and then imputing all sequence variants using FImpute45 with all other sequences (n = 1,133) as a reference. The squared correlation (r2) between the real sequence variant called genotypes and the imputed variant genotypes was taken as the accuracy of imputation. The accuracy of imputing sequence variant genotypes is plotted by minor allele frequency. b, The accuracy of imputation to sequence variants in Holstein and Jersey cattle was compared for FImpute and Beagle. The accuracy was assessed for sequence variants on chromosome 14. The accuracy of imputing genotypes was calculated by masking the sequence genotypes of 25 randomly selected animals per breed from 1000 Bull Genomes Run4 to those on the Illumina BovineHD BeadChip and then imputing all sequence variants using FImpute and Beagle with all other sequences (n = 1,122) as a reference.

Supplementary Figure 3 Association testing with imputed sequence variants in the region of the GHR gene.

ac, Sequence variant genotypes were imputed in 6,777 Fleckvieh, 5,204 Holstein and 1,646 Brown Swiss animals using the 1000 Bull Genomes Run4 multi-breed reference population with Minimac44. Association tests were performed between imputed sequence variant genotypes on chromosome 20 and daughter-derived values for protein percentage. Association testing was carried out with EMMAX using the -Z flag to consider predicted allele dosages for the imputed sequence variants. a, In Fleckvieh, the GHR p.Y279F mutation (rs385640152) was the second most strongly associated marker. b,c, The association testing revealed the causal GHR p.Y279F mutation (rs385640152)55 (in the growth hormone receptor gene) to be the most significantly associated variant in Holstein (b) and Brown Swiss (c) cattle. The frequency of the minor allele was 0.16, 0.08 and 0.12 in Holstein, Fleckvieh and Brown Swiss cattle, respectively. Data were used from Pausch et al.10.

Supplementary Figure 4 Association testing with imputed sequence variants in the region of the DGAT1 gene.

a,b, Sequence variant genotypes were imputed in 6,777 Fleckvieh and 5,204 Holstein animals using the 1000 Bull Genomes Run4 multi-breed reference population with Minimac44. Association tests were performed between imputed sequence variant genotypes on chromosome 14 and daughter-derived values for fat percentage. Association testing was carried out with EMMAX5 using the -Z flag to consider predicted allele dosages for the imputed sequence variants. The association testing revealed the causal mutation (p.A232K) in the DGAT1 gene56 to be the most significantly associated variant in Fleckvieh (a) and Holstein (b) cattle, as has been demonstrated in an earlier run of the 1000 Bull Genomes project4. The frequency of the minor allele was 0.31, 0.07 and 0.12 in Holstein and Fleckvieh cattle, respectively. Data were used from Pausch et al10.

Supplementary Figure 5 Association testing with imputed sequence variants in the region of the ABCG2 gene.

Sequence variant genotypes were imputed into 5,204 Holstein animals using the 1000 Bull Genomes Run4 multi-breed reference population with Minimac44. Association tests were performed between imputed sequence variant genotypes on chromosome 6 and daughter-derived values for protein percentage. Association testing was carried out with EMMAX5 using the -Z flag to consider predicted allele dosages for the imputed sequence variants. The association testing revealed the causal p.Y581S variant in ABCG256 to be the most significantly associated variant in Holstein cattle. The frequency of the minor allele was 0.013. Data were used from Pausch et al10.

Supplementary Figure 6

The number of breeds that each variant segregates in plotted against the size of the variant effect estimated in the meta-analysis for the 163 lead variants.

Supplementary Figure 7 Selection signature analysis.

Signatures of selection based on haplotype (hapFLK statistical test35; top) or single-SNP (FLK statistical test36; bottom) tests for differentiation (–log10P values), with n = 380 animals. Darker colors indicate significant hits at an FDR of 5%. Vertical red lines highlight the positions of 163 lead variants from the stature GWAS meta-analysis.

Supplementary Figure 8

The distance of the most significant SNP from the gene for 659 genes with eQTLs.

Supplementary information

Supplementary Text and Figures

Supplementary Figures 1–8 and Supplementary Tables 1, 3, 5 and 7–9.

Life Sciences Reporting Summary

Supplementary Table 2

Positions and effects of the most significant SNPs, confidence intervals, overlap with human genes for stature, overlap with eQTLs from whole blood and indication of the ancestral allele.

Supplementary Table 4

Validation of lead SNPs in ten populations encompassing eight breeds.

Supplementary Table 6

Proportion of bootstrap samples in which the lead variant from the original meta-analysis remained the lead variant in the bootstrap sample.

Rights and permissions

Reprints and permissions

About this article

Check for updates. Verify currency and authenticity via CrossMark

Cite this article

Bouwman, A.C., Daetwyler, H.D., Chamberlain, A.J. et al. Meta-analysis of genome-wide association studies for cattle stature identifies common genes that regulate body size in mammals. Nat Genet 50, 362–367 (2018). https://doi.org/10.1038/s41588-018-0056-5

Download citation

  • Received:

  • Accepted:

  • Published:

  • Issue Date:

  • DOI: https://doi.org/10.1038/s41588-018-0056-5

This article is cited by

Search

Quick links

Nature Briefing: Translational Research

Sign up for the Nature Briefing: Translational Research newsletter — top stories in biotechnology, drug discovery and pharma.

Get what matters in translational research, free to your inbox weekly. Sign up for Nature Briefing: Translational Research