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.

Harnessing genomic information for livestock improvement

Abstract

The world demand for animal-based food products is anticipated to increase by 70% by 2050. Meeting this demand in a way that has a minimal impact on the environment will require the implementation of advanced technologies, and methods to improve the genetic quality of livestock are expected to play a large part. Over the past 10 years, genomic selection has been introduced in several major livestock species and has more than doubled genetic progress in some. However, additional improvements are required. Genomic information of increasing complexity (including genomic, epigenomic, transcriptomic and microbiome data), combined with technological advances for its cost-effective collection and use, will make a major contribution.

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: Identification of mutations and genes causing monogenic defects in livestock.
Fig. 2: Selection procedure of elite dairy sires and cows.
Fig. 3: Identifying cows with subclinical mastitis by bulk genotyping of tank milk.

References

  1. Thornton, P. K. Livestock production: recent trends, future prospects. Phil. Trans. R. Soc. B 365, 2853–2867 (2010).

    PubMed  PubMed Central  Google Scholar 

  2. Lynch, M. & Walsh, B. Genetics and Analysis of Quantitative Traits (Sinauer Associates, Sunderland, MA, 1998).

    Google Scholar 

  3. Havenstein, G. B., Ferket, P. R. & Qureshi, M. A. Growth liability and feed conversion of 1957 versus 2001 broilers when fed representative 1957 and 2001 broiler diets. Poult. Sci. 82, 1500–1508 (2003).

    CAS  PubMed  Google Scholar 

  4. Sigel, P. B. Evolution of the modern broiler and feed efficiency. Annu. Rev. Anim. Biosci. 2, 375–385 (2014).

    Google Scholar 

  5. Capper, J. L. & Bauman, D. E. The role of productivity in improving the environmental sustainability of ruminant production systems. Annu. Rev. Anim. Biosci. 1, 469–489 (2013).

    PubMed  Google Scholar 

  6. Pryce, J. E., Royal, M. D., Garnsworthy, P. C. & Mao, I. L. Fertility in the high-producing dairy cow. Livestock Prod. Sci. 86, 125–135 (2004).

    Google Scholar 

  7. Goddard, M. E. & Hayes, B. J. Mapping genes for complex traits in domestic animals and their use in breeding programs. Nat. Rev. Genet. 10, 381–391 (2009).

    CAS  PubMed  Google Scholar 

  8. Meuwissen, T. H., Hayes, B. J. & Goddard, M. E. Prediction of total genetic value using genome-wide dense marker maps. Genetics 157, 1819–1829 (2001). This is a landmark paper that triggered the adoption of GS by the breeding industry.

    CAS  PubMed  PubMed Central  Google Scholar 

  9. Hill, W. G., Goddard, M. E. & Visscher, P. Data and theory point to mainly additive genetic variance for complex traits. PLOS Genet. 4, e1000008 (2008).

    PubMed  PubMed Central  Google Scholar 

  10. Wiggans, G. R., Cole, J. B., Hubbard, S. M. & Sonstegard, T. S. Genomic selection in dairy cattle: the USDA experience. Annu. Rev. Anim. Biosci. 5, 309–327 (2017).

    PubMed  Google Scholar 

  11. Van Eenennaam, A. L., Weigel, K. A., Young, A. E., Cleveland, M. A. & Dekkers, J. C. Applied animal genomics: results from the field. Annu. Rev. Anim. Biosci. 2, 105–139 (2014).

    PubMed  Google Scholar 

  12. Crossa, J. et al. Genomics selection in plant breeding: methods, models and perspectives. Trends Plant Sci. 22, 9661–9975 (2017).

    Google Scholar 

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

    CAS  PubMed  PubMed Central  Google Scholar 

  14. Gusev, A. et al. Quantifying missing heritability at known GWAS loci. PLOS Genet. 9, e1003993 (2013).

    PubMed  PubMed Central  Google Scholar 

  15. Wray, N. R. et al. Pitfalls of predicting complex traits from SNPs. Nat. Rev. Genet. 14, 507–515 (2013).

    CAS  PubMed  PubMed Central  Google Scholar 

  16. Herrero, M. et al. Livestock and the environment: what have we learned in the past decade? Annu. Rev. Environ. Resour. 40, 177–202 (2015).

    Google Scholar 

  17. International Chicken Genome Sequencing Consortium. Sequence and comparative analysis of the chicken genome provide unique perspectives on vertebrate evolution. Nature 432, 695–716 (2004).

    Google Scholar 

  18. Bovine Genome Sequencing and Analysis Consortium. The genome sequence of taurine cattle: a window to ruminant biology and evolution. Science 324, 522–528 (2009).

    Google Scholar 

  19. Swine Genome Sequencing Consortium. Analysis of pig genomes provide insight into porcine demography and evolution. Nature 491, 393–398 (2012).

    Google Scholar 

  20. Dong, Y. et al. Sequencing and automated whole-genome optical mapping of the genome of a domestic goat (Capra hircus). Nat. Biotechnol. 31, 135–141 (2013).

    CAS  PubMed  Google Scholar 

  21. Jiang, Y. et al. The sheep genome illuminates biology of the rumen and lipid metabolism. Science 344, 1168–1173 (2014).

    CAS  PubMed  PubMed Central  Google Scholar 

  22. Lien, S. et al. The Atlantic salmon genome provides insights into rediploidization. Nature 533, 200–205 (2016).

    CAS  PubMed  PubMed Central  Google Scholar 

  23. Green, E. D. Strategies for the systematic sequencing of complex genomes. Nat. Rev. Genet. 2, 573–583 (2001).

    CAS  PubMed  Google Scholar 

  24. Lindblad-Toh, K. et al. A high-resolution map of human evolutionary constraint using 29 mammals. Nature 478, 476–482 (2011).

    CAS  PubMed  PubMed Central  Google Scholar 

  25. Daetwyler, H. D. et al. Whole-genome sequencing of 234 bulls facilitates mapping of monogenic and complex traits in cattle. Nat. Genet. 8, 858–865 (2014). This is a report by the 1,000 Bull Genomes Project consortium that introduces a community resource to accelerate the detection and use of causative variants in cattle.

    Google Scholar 

  26. Worley, K. C. A golden goat genome. Nat. Genet. 49, 485–486 (2017).

    CAS  PubMed  Google Scholar 

  27. Bickhart, D. M. et al. Single-molecule sequencing and chromatin conformation capture enable de novo reference assembly of the domestic goat genome. Nat. Genet. 49, 643–650 (2017). This paper brilliantly illustrates the utility of novel scaffolding techniques in dramatically improving the quality of reference genome sequences in an affordable way.

    CAS  PubMed  PubMed Central  Google Scholar 

  28. International Chicken Polymorphism Map Consortium. A genetic variation map for chicken with 2.8 million single-nucleotide polymorphisms. Nature 432, 717–722 (2004).

    PubMed Central  Google Scholar 

  29. Bovine HapMap Consortium. Genome-wide survey of SNP variation uncovers the genetic structure of cattle breeds. Science 324, 528–532 (2009).

    Google Scholar 

  30. Charlier, C. et al. NGS-based reverse genetic screen for common embryonic lethal mutations compromising fertility in livestock. Genome Res. 26, 1–9 (2016).

    Google Scholar 

  31. The 1000 Genomes Project Consortium. A global reference for human genetic variation. Nature 526, 68–74 (2015).

    Google Scholar 

  32. MacLeod, I. M. et al. Exploiting biological priors and sequence variants enhances QTL discovery and genomic prediction of complex traits. BMC Genomics 17, 144 (2016).

    CAS  PubMed  PubMed Central  Google Scholar 

  33. Bourneuf, E. et al. Rapid discovery of de novo deleterious mutations in cattle enhances the value of livestock as model species. Sci. Rep. 7, 11466–11485 (2017).

    CAS  PubMed  PubMed Central  Google Scholar 

  34. Kadri, N. K. et al. Coding and non-coding variants in HFM1, MLH3, MSH4, MSH5, RNF212 and RNF212B affect recombination rate in cattle. Genome Res. 26, 1323–1332 (2016).

    CAS  PubMed  PubMed Central  Google Scholar 

  35. Schaub, M. A., Boyle, A. P., Kundaje, A., Batzoglou, S. & Snyder, M. Linking disease associations with regulatory information in the human genome. Genome Res. 22, 1748–1759 (2012).

    CAS  PubMed  PubMed Central  Google Scholar 

  36. Huang, H. et al. Association mapping of inflammatory bowel disease loci to single variant resolution. Nature 547, 173–178 (2017).

    CAS  PubMed  PubMed Central  Google Scholar 

  37. Andersson, L. et al. Coordinated international action to accelerate genome-to-phenome with FAANG, the Functional Annotation of Animal Genomes project. Genome Biol. 16, 57–63 (2015).

    PubMed  PubMed Central  Google Scholar 

  38. Tuggle, C. K. et al. GO-FAANG meeting: a gathering on functional annotation of animal genomes. Anim. Genet. 47, 528–533 (2016).

    PubMed  PubMed Central  Google Scholar 

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

    CAS  PubMed  PubMed Central  Google Scholar 

  40. Zhou, Y. et al. Reduced representation bisulphite sequencing of ten bovine somatic tissues reveals DNA methylation patterns and their impacts on gene expression. BMC Genomics 17, 779 (2016).

    PubMed  PubMed Central  Google Scholar 

  41. Littlejohn, M. D. et al. Expression variants of the lipogenic AGPAT6 gene affect diverse milk composition phenotypes in Bos taurus. PLOS ONE 9, e85757 (2014).

    PubMed  PubMed Central  Google Scholar 

  42. Littlejohn, M. D. et al. Sequence-based association analysis reveals an MGST1 eQTL with pleiotropic effects on bovine milk composition. Sci. Rep. 6, 25376–25390 (2016).

    CAS  PubMed  PubMed Central  Google Scholar 

  43. Kemper, K. E. et al. Leveraging genetically simple traits to identify small-effect variants for complex phenotypes. BMC Genomics 17, 858–867 (2016).

    CAS  PubMed  PubMed Central  Google Scholar 

  44. Brand, B. et al. Adrenal cortex expression quantitative trait loci in a German Holstein x Charolais cross. BMC Genetics 17, 135–146 (2016).

    PubMed  PubMed Central  Google Scholar 

  45. Lopdell, T. J. et al. DNA and RNA-sequence based GWAS highlights membrane-transport genes as key modulators of milk lactose content. BMC Genomics 18, 968 (2017).

    PubMed  PubMed Central  Google Scholar 

  46. Ponsuksili, S., Murani, E., Brand, B., Schwerin, M. & Wimmers, K. Integrating expression profiling and whole-genome association for dissection of fat traits in a porcine model. J. Lipid Res. 52, 6668–6678 (2011).

    Google Scholar 

  47. Liaubet, L. et al. Genetic variability of transcript abundance in pig peri-mortem skeletal muscle: eQTL localized genes involved in stress response, cell death, muscle disorders and metabolism. BMC Genomics 12, 548–565 (2011).

    CAS  PubMed  PubMed Central  Google Scholar 

  48. Ernst, C. W. & Steibel, J. P. Molecular advances in QTL discovery and application in pig breeding. Trends Genet. 29, 215–224 (2013).

    CAS  PubMed  Google Scholar 

  49. Heidt, H. et al. A genetical genomics approach reveals new candidates and confirms known candidate genes for drip loss in a porcine resource population. Mamm. Genome 24, 416–426 (2013).

    PubMed  Google Scholar 

  50. Chen, C. et al. Genetic dissection of blood lipid traits by integrating genome-wide association study and gene expression profiling in a porcine model. BMC Genomics 14, 848–859 (2013).

    PubMed  PubMed Central  Google Scholar 

  51. Ponsuksili, S., Murani, E., Trakooljul, N., Schwerin, M. & Wimmers, K. Discovery of candidate genes for muscle traits based on GWAS supported by eQTL-analysis. Int. J. Biol. Sci. 10, 327–337 (2014).

    PubMed  PubMed Central  Google Scholar 

  52. Reiner, G. et al. Pathway deregulation and expression QTLs in response to Actinobacillus pleuropneumoniae infection in swine. Mamm. Genome 25, 600–617 (2014).

    CAS  PubMed  Google Scholar 

  53. Ma, J. et al. A splice mutation in the PHKG1 gene causes high glycogen content and low meat quality in pig skeletal muscle. PLOS Genet. 10, e1004710 (2014).

    PubMed  PubMed Central  Google Scholar 

  54. Kogelman, L. J. et al. An integrative systems genetics approach reveals potential causal genes and pathways related to obesity. Genome Med. 7, 105–120 (2015).

    PubMed  PubMed Central  Google Scholar 

  55. Martinez-Montes, A. M. et al. Deciphering the regulation of porcine genes influencing growth, fatness and yield-related traits through genetical genomics. Mamm. Genome 28, 130–142 (2017).

    CAS  PubMed  Google Scholar 

  56. Gonzalez-Prendes, R. et al. Joint QTL mapping and gene expression analysis identify positional candidate genes influencing pork quality traits. Sci. Rep. 7, 39830–39839 (2017).

    CAS  PubMed  PubMed Central  Google Scholar 

  57. Maroilley, T. et al. Deciphering the genetic regulation of peripheral blood transcriptome in pigs through expression genome-wide association study and allele-specific expression analysis. BMC Genomics 18, 967 (2017).

    CAS  PubMed  PubMed Central  Google Scholar 

  58. Blum, Y. et al. Complex trait subtypes identification using transcriptome profiling reveals an interaction between two QTL affecting adiposity in chicken. BMC Genomics 12, 567–575 (2011).

    CAS  PubMed  PubMed Central  Google Scholar 

  59. Johnsson, M., Jonsson, K. B., Andersson, L., Jensen, P. & Wright, D. Quantitative trait locus and genetical genomics analysis identifies putatively causal genes for fecundity and brooding in the chicken. G3 6, 311–319 (2015).

    PubMed  PubMed Central  Google Scholar 

  60. Johnsson, M., Williams, M. J., Jensen, P. & Wright, D. Genetical genomics of behavior: a novel chicken genomic model for anxiety behavior. Genetics 202, 327–340 (2016).

    CAS  PubMed  Google Scholar 

  61. Fallahsharoudi, A. et al. QTL mapping of stress related gene expression in cross between domesticated chickens and ancestral red junglefowl. Mol. Cell. Endocrinol. 446, 52–58 (2017).

    CAS  PubMed  Google Scholar 

  62. GTEx Consortium. Genetic effects on gene expression across human tissues. Nature 550, 204–213 (2017).

    PubMed Central  Google Scholar 

  63. Dunne, J. et al. First dairying in green Saharan Africa in the fifth millennium BC. Nature 486, 390–394 (2012).

    CAS  PubMed  Google Scholar 

  64. Andersson, L. S. et al. Mutations in DMRT3 affect locomotion in horses and spinal circuit function in mice. Nature 488, 642–646 (2012). This paper demonstrates the value of domestic animals in uncovering functions of mammalian genes by studying unique selected phenotypes, in this case ‘ambling’.

    CAS  PubMed  PubMed Central  Google Scholar 

  65. Andersson, L. Molecular consequences of animal breeding. Curr. Opin. Genet. Dev. 23, 295–301 (2013).

    CAS  PubMed  Google Scholar 

  66. Durkin, K. et al. Serial translocations by means of circular intermediates underlies color sidedness in cattle. Nature 482, 81–84 (2012). This study identified a novel CNV-generating mechanism that may underlie exon shuffling by studying coat-colour variation in cattle.

    CAS  PubMed  Google Scholar 

  67. Clop, A. et al. A mutation creating a potential illegitimate microRNA target site in the myostatin gene affects muscularity in sheep. Nat. Genet. 38, 813–818 (2006). This paper describes one of the most convincing examples in mammals of a phenotype resulting from perturbed microRNA-mediated gene regulation.

    CAS  PubMed  Google Scholar 

  68. Georges, M. et al. in Epigenetics and Complex Traits (eds Naoumova, A. K. & Greenwood, C. M. T.) 89–106 (Springer, New York, NY, 2013). This paper reviews the current molecular understanding of the unique phenomenon of polar overdominance at the ovine callipyge locus.

  69. MacArthur, D. G. et al. A systematic survey of loss-of-function variants in human protein-coding genes. Science 335, 823–828 (2012).

    CAS  PubMed  PubMed Central  Google Scholar 

  70. Bittles, A. H. & Neel, J. V. The costs of human inbreeding and their implications for variations at the DNA level. Nat. Genet. 8, 117–121 (1994).

    CAS  PubMed  Google Scholar 

  71. Simmons, M. J. & Crow, J. F. Mutations affecting fitness in Drosophila populations. Annu. Rev. Genet. 11, 49–78 (1977).

    CAS  PubMed  Google Scholar 

  72. Lek, M. et al. Analysis of protein-coding genetic variation in 60,706 humans. Nature 536, 285–291 (2016).

    CAS  PubMed  PubMed Central  Google Scholar 

  73. Baker, R. D., Snider, G. W., Leipold, H. E. & Johnson, J. L. Embryo transfer tests for bovine syndactyly. Theriogenology 13, 87 (1980).

    CAS  PubMed  Google Scholar 

  74. Charlier, C. et al. Highly effective SNP-based association mapping and management of recessive defects in livestock. Nat. Genet. 40, 449–454 (2008). This study is one of the first illustrations of how genome-wide SNP arrays accelerated the identification of mutations causing genetic defects in domestic animals.

    CAS  PubMed  Google Scholar 

  75. Littlejohn, M. D. et al. Functionally reciprocal mutations of the prolactin signalling pathway define hairy and slick cattle. Nat. Commun. 5, 5861–5869 (2014).

    CAS  PubMed  Google Scholar 

  76. Agerholm, J. S. et al. A de novo missense mutation of FGFR2 causes facial dysplasia syndrome in Holstein cattle. BMC Genetics 18, 74–83 (2017).

    PubMed  PubMed Central  Google Scholar 

  77. Harland, C. et al. Frequency of mosaicism points towards mutation-prone early cleavage cell divisions. Preprint at bioRxiv http://biorxiv.org/content/early/2016/10/09/079863 (2016).

  78. VanRaden, P. M., Olson, K. M., Null, D. J. & Hutchison, J. L. Harmful recessive effects on fertility detected by absence of homozygous haplotypes. J. Dairy Sci. 94, 6153–6161 (2011). This study is one of the first demonstrations of how the availability of genome-wide SNP genotypes for large populations can be used to effectively identify haplotypes harbouring EL mutations affecting fertility.

    CAS  PubMed  Google Scholar 

  79. Adams, H. A. et al. Identification of a nonsense mutation in APAF1 that is likely causal for a decrease in reproductive efficiency in Holstein dairy cattle. J. Dairy Sci. 99, 6693–6701 (2016).

    CAS  PubMed  Google Scholar 

  80. Kadri, N. K. et al. A 660-Kb deletion with antagonistic effects on fertility and milk production segregates at high frequency in Nordic Red cattle: additional evidence for the common occurrence of balancing selection in livestock. PLOS Genet. 10, e1004049 (2014).

    PubMed  PubMed Central  Google Scholar 

  81. Sahana, G., Nielsen, U. S., Aamand, G. P., Lund, M. P. & Guldbrandtsen, B. Novel harmful recessive haplotypes identified for fertility traits in Nordic Holstein cattle. PLOS ONE 8, e82909 (2013).

    PubMed  PubMed Central  Google Scholar 

  82. Fritz, S. et al. Detection of haplotypes associated with prenatal death in dairy cattle and identification of deleterious mutations in GART, SHBG and SLC37A2. PLOS ONE 8, e65550 (2013).

    CAS  PubMed  PubMed Central  Google Scholar 

  83. Pausch, H. et al. Homozygous haplotype deficiency reveals deleterious mutations compromising reproductive and rearing success in cattle. BMC Genomics 16, 312 (2015).

    PubMed  PubMed Central  Google Scholar 

  84. Häggman, J. & Uimari, P. Novel harmful recessive haplotypes for reproductive traits in pigs. J. Anim. Breed. Genet. 134, 129–135 (2017).

    PubMed  Google Scholar 

  85. Derks, M. F. L. et al. A systematic survey to identify lethal recessive variation in highly managed pig populations. BMC Genomics 18, 858–870 (2017).

    PubMed  PubMed Central  Google Scholar 

  86. McClure, M. C. et al. Bovine exome sequence analysis and targeted SNP genotyping of recessive fertility defects BH1, HH2, and HH3 reveal a putative causative mutation in SMC2 for HH3. PLOS ONE 9, e92769 (2014).

    PubMed  PubMed Central  Google Scholar 

  87. Sonstegard, T. S. et al. Identification of a nonsense mutation in CWC15 associated with decreased reproductive efficiency in Jersey cattle. PLOS ONE 8, e54872 (2013).

    CAS  PubMed  PubMed Central  Google Scholar 

  88. Fritz, S. et al. An initiator codon mutation in SDE2 causes recessive embryonic lethality in Holstein cattle. J. Dairy Sci. 101, 6220–6231 (2018).

    CAS  PubMed  Google Scholar 

  89. Martinez, V., Bünger, L. & Hill, W. G. Analysis of response to 20 generations of selection for body composition in mice: fit to infinitesimal model assumptions. Genet. Sel. Evol. 32, 3 (2000).

    CAS  PubMed  PubMed Central  Google Scholar 

  90. Henderson, C. R. Sire evaluation and genetic trends. J. Anim. Sci. 1973, 10–41 (1973).

    Google Scholar 

  91. Wright, S. Coefficients of inbreeding and relationship. Am. Nat. 56, 330–338 (1922).

    Google Scholar 

  92. Malecot, G. Les mathématiques de l’hérédité (Masson, Paris, 1948).

    Google Scholar 

  93. Ritland, K. Estimators of pairwise relatedness and individual inbreeding coefficients. Genet. Res. 67, 175–185 (1996).

    Google Scholar 

  94. Toro, M. et al. Estimation of coancestry in Iberian pigs using molecular markers. Conserv. Genet. 3, 309–320 (2002).

    CAS  Google Scholar 

  95. Garant, D. & Kruuk, L. E. B. How to use molecular marker data to measure evolutionary parameters in wild populations. Mol. Ecol. 14, 1843–1859 (2005).

    CAS  PubMed  Google Scholar 

  96. Aguilar, I. et al. A unified approach to utilize phenotypic, full pedigree, and genomic information for genetic evaluation of Holstein final score. J. Dairy Sci. 93, 743–752 (2010). This study is one of the first demonstrations that pedigree, genomic and phenotypic information from millions of animals could be combined in a single-step approach to predict GEBVs.

    CAS  PubMed  Google Scholar 

  97. Christensen, O. F. & Lund, M. S. Genomic prediction when some animals are not genotyped. Genet. Sel. Evol. 42, 2 (2010).

    PubMed  PubMed Central  Google Scholar 

  98. Liu, Z., Goddard, M. E., Reinhardt, F. & Reents, R. A single-step genomic model with direct estimation of marker effects. J. Dairy Sci. 97, 5833–5850 (2014).

    CAS  PubMed  Google Scholar 

  99. Fernando, R. L., Dekkers, J. C. & Garrick, D. J. A class of Bayesian methods to combine large numbers of genotyped and non-genotyped animals for whole-genome analyses. Genet. Sel. Evol. 46, 50 (2014).

    PubMed  PubMed Central  Google Scholar 

  100. Fernando, R. L., Cheng, H., Golden, B. L. & Garrick, D. J. Computational strategies for alternative single-step Bayesian regression models with large numbers of genotyped and non-genotyped animals. Genet. Sel. Evol. 48, 96 (2016).

    PubMed  PubMed Central  Google Scholar 

  101. Misztal, I. & Legarra, A. Invited review: efficient computation strategies in genomic selection. Animal 11, 731–736 (2017).

    CAS  PubMed  Google Scholar 

  102. García-Ruiz, A. et al. Changes in genetic selection differentials and generation intervals in US Holstein dairy cattle as a result of genomic selection. Proc. Natl Acad. Sci. USA 113, 3995–4004 (2016). This study demonstrates the impact of GS on genetic gain in the US dairy industry.

    Google Scholar 

  103. Van Eenennaam, A. L., van der Werf, J. H. & Goddard, M. E. The value of using DNA markers for beef bull selection in the seedstock sector. J. Anim. Sci. 89, 307–320 (2011).

    PubMed  Google Scholar 

  104. Lourenco, D. A. et al. Genetic evaluation using single-step genomic best linear unbiased predictor in American Angus. J. Anim. Sci. 93, 2653–26662 (2015).

    CAS  PubMed  Google Scholar 

  105. Abell, C. E., Dekkers, J. C. M., Rothschild, M. F., Mabry, J. W. & Stalder, K. J. Total cost estimation for implementing genome-enabled selection in a multi-level swine production system. Genet. Sel. Evol. 46, 32 (2014).

    PubMed  PubMed Central  Google Scholar 

  106. Shumbusho, F. et al. Economic evaluation of genomic selection in small ruminants: a sheep meat breeding program. Animal 10, 1033–1041 (2016).

    CAS  PubMed  Google Scholar 

  107. Brito, L. F. et al. Prediction of genomic breeding values for growth, carcass and meat quality traits in a multi-breed sheep population using a HD SNP chip. BMC Genetics 18, 7 (2017).

    PubMed  PubMed Central  Google Scholar 

  108. Santos, B. F. S., van der Werf, J. H. J., Gibson, J. P., Byrne, T. J. & Amer, P. R. Assessment of the genetic and economic impact of performance recording and genotyping in Australian commercial sheep operations. J. Anim. Breed. Genet. 135, 221–237 (2018).

    CAS  PubMed  Google Scholar 

  109. Wolc, A. et al. Response and inbreeding from a genomic selection experiment in layer chicken. Genet. Sel. Evol. 47, 59 (2015).

    PubMed  PubMed Central  Google Scholar 

  110. Gonzalez-Recio, O., Pryce, J. E., Haile-Mariam, M. & Hayes, B. J. Incorporating heifer feed efficiency in the Australian selection index using genomic selection. J. Dairy Sci. 97, 3883 (2014).

    CAS  PubMed  Google Scholar 

  111. Goddard, M. E. Genomic selection: prediction of accuracy and maximisation of long term response. Genetica 136, 245–257 (2009).

    PubMed  Google Scholar 

  112. Grobet, L. et al. A deletion in the bovine myostatin gene causes the double-muscled phenotype in cattle. Nat. Genet. 17, 71–74 (1997).

    CAS  PubMed  Google Scholar 

  113. Kambadur, R., Sharma, M., Smith, T. P. & Bass, J. J. Mutations in myostatin (GDF8) in double-muscled Belgian Blue and Piedmontese cattle. Genome Res. 7, 910–916 (1997).

    CAS  PubMed  Google Scholar 

  114. McPherron, A. C. & Lee, S. J. Double muscling in cattle due to mutations in the myostatin gene. Proc. Natl Acad. Sci. USA 94, 12457–12461 (1997).

    CAS  PubMed  PubMed Central  Google Scholar 

  115. Grobet, L. et al. Molecular definition of an allelic series of mutations disrupting the myostatin function and causing double-muscling in cattle. Mamm. Genome 9, 210–213 (1998).

    CAS  PubMed  Google Scholar 

  116. Fujii, J. et al. Identification of a mutation in the porcine ryanodine receptor associated with malignant hyperthermia. Science 253, 448–451 (1991).

    CAS  PubMed  Google Scholar 

  117. Milan, D. et al. A mutation in PRKAG3 associated with excess glycogen content in pig skeletal muscle. Science 288, 1248–1251 (2000).

    CAS  PubMed  Google Scholar 

  118. Van Laere, A. S. et al. A regulatory mutation in IGF2 causes a major QTL effect on muscle growth in the pig. Nature 425, 832–836 (2003).

    PubMed  Google Scholar 

  119. 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).

    CAS  PubMed  Google Scholar 

  120. Blott, S. et al. Molecular dissection of a QTL: 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 1663, 253–2666 (2003).

    Google Scholar 

  121. 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).

    CAS  PubMed  PubMed Central  Google Scholar 

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

    CAS  PubMed  Google Scholar 

  123. Bouwman, A. C. 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).

    CAS  PubMed  Google Scholar 

  124. Habier, D., Fernando, R. L., Kizilkaya, K. & Garrick, D. J. Extension of the Bayesian alphabet for genomic selection. BMC Bioinformatics 12, 186 (2011).

    PubMed  PubMed Central  Google Scholar 

  125. 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).

    CAS  PubMed  Google Scholar 

  126. Druet, T. et al. Selection in action: dissecting the molecular underpinnings of the increasing muscle mass in Belgian Blue Cattle. BMC Genomics 15, 796 (2014).

    PubMed  PubMed Central  Google Scholar 

  127. Hedrick, P. W. Heterozygote advantage: the effect of artificial selection in livestock and pets. J. Hered. 106, 141–154 (2015).

    CAS  PubMed  Google Scholar 

  128. Fasquelle, C. et al. Balancing selection of a frame-shift mutation in the MRC2 gene accounts for the outbreak of the Crooked Tail Syndrome in Belgian Blue Cattle. PLOS Genet. 5, e1000666 (2009).

    PubMed  PubMed Central  Google Scholar 

  129. Sartelet, A. et al. A splice site variant in the bovine RNF11 gene compromises growth and regulation of the inflammatory response. PLOS Genet. 8, e1002581 (2012).

    CAS  PubMed  PubMed Central  Google Scholar 

  130. Sartelet, A. et al. Allelic heterogeneity of Crooked Tail Syndrome fits the balancing selection hypothesis. Anim. Genet. 43, 591–594 (2012).

    Google Scholar 

  131. Galloway, S. M. et al. Mutations in an oocyte-derived growth factor gene (BMP15) cause increased ovulation rate and infertility in a dosage-sensitive manner. Nat. Genet. 25, 279–283 (2000).

    CAS  PubMed  Google Scholar 

  132. Hanrahan, J. P. et al. Mutations in the genes for oocyte-derived growth factors GDF9 and BMP15 are associated with both increased ovulation rate and sterility in Cambridge and Belcaler sheep (Ovis aries). Biol. Reprod. 70, 900–909 (2004).

    CAS  PubMed  Google Scholar 

  133. Rupp, R. et al. A point mutation in suppressor of cytokine signalling 2 (SOCS2) increases the susceptibility to inflammation of the mammary gland while associated with higher body weight and size and higher milk production in a sheep model. PLOS Genet. 11, e1005629 (2015).

    PubMed  PubMed Central  Google Scholar 

  134. Johnston, S. E. et al. Life history trade-offs at a single locus maintains sexually selected genetic variation. Nature 502, 93–95 (2013).

    CAS  PubMed  Google Scholar 

  135. Cockett, N. E. et al. Polar overdominance at the ovine callipyge locus. Science 273, 236–238 (1996).

    CAS  PubMed  Google Scholar 

  136. Freking, B. A. et al. Identification of the single base change causing the callipyge muscle hypertrophy phenotype, the only known example of polar overdominance in mammals. Genome Res. 12, 14966–11506 (2002).

    Google Scholar 

  137. Barson, N. J. et al. Sex-dependent dominance at a single locus maintains variation in age at maturity in salmon. Nature 528, 405–498 (2015). This paper describes a remarkable example of balancing selection occurring in a natural population as a result of antagonistic selection in males and females.

    CAS  PubMed  Google Scholar 

  138. Ayllon, F. et al. The vgll3 locus controls age at maturity in wild and domesticated atlantic salmon (Salmo salar L.) males. PLOS Genet. 11, e1005628 (2015).

    PubMed  PubMed Central  Google Scholar 

  139. Utsunomiya, Y. T. et al. A PLAG1 mutation contributed to stature recovery of modern cattle. Sci. Rep. 7, 17140 (2017).

    PubMed  PubMed Central  Google Scholar 

  140. Fortes, M. R. S. et al. Evidence for pleiotropism and recent selection in the PLAG1 region in Australian Beef cattle. Anim. Genet. 44, 6636–6647 (2013).

    Google Scholar 

  141. Pszczola, M. & Calus, M. P. Updating the reference population to achieve constant genomic prediction reliability across generations. Animal 10, 1018–1024 (2016).

    CAS  PubMed  Google Scholar 

  142. Calus, M. P. Right-hand-side updating for fast computing of genomic breeding values. Genet. Sel. Evol. 46, 24 (2014).

    PubMed  PubMed Central  Google Scholar 

  143. Calus, M. P., Bouwman, A. C., Schrooten, C. & Veerkamp, R. F. Efficient genomic prediction based on whole-genome sequence data using split-and-merge Bayesian variable selection. Genet. Sel. Evol. 48, 49 (2016).

    PubMed  PubMed Central  Google Scholar 

  144. Wang, T. et al. Application of a Bayesian non-linear model hybrid scheme to sequence data for genomic prediction and QTL mapping. BMC Genomics 18, 618 (2017).

    PubMed  PubMed Central  Google Scholar 

  145. van den Berg, I. et al. Multi-breed genomic prediction using Bayes R with sequence data and dropping variants with a small effect. Genet. Sel. Evol. 49, 70 (2017).

    PubMed  PubMed Central  Google Scholar 

  146. Brøndum, R. F. et al. Quantitative trait loci markers derived from whole genome sequence data increases the reliability of genomic prediction. J. Dairy Sci. 98, 4107–4116 (2015).

    PubMed  Google Scholar 

  147. Veerkamp, R. F., Bouwman, A. C., Schrooten, C. & Calus, M. P. Genomic prediction using preselected DNA variants from a GWAS with whole-genome sequence data in Holstein-Friesian cattle. Genet. Sel. Evol. 48, 95 (2016).

    PubMed  PubMed Central  Google Scholar 

  148. VanRaden, P. M., Tooker, M. E., O’Connell, J. R., Cole, J. B. & Bickhart, D. M. Selecting sequence variants to improve genomic predictions for dairy cattle. Genet. Sel. Evol. 49, 32 (2017).

    PubMed  PubMed Central  Google Scholar 

  149. Long, H. K., Prescott, S. L. & Wysocka, J. Ever-changing landscapes: transcriptional enhancers in development and evolution. Cell 167, 1170–1187 (2016).

    CAS  PubMed  PubMed Central  Google Scholar 

  150. Wragg, D. et al. Genome-wide analysis reveals the extent of EAV-HP integration in domestic chicken. BMC Genomics 16, 784–794 (2015).

    PubMed  PubMed Central  Google Scholar 

  151. Kemper, K. E., Jayes, B. J., Daetwyler, H. D. & Goddard, M. E. How old are quantitative trait loci and how widely do they segregate? J. Anim. Breed. Genet. 132, 121–134 (2015).

    CAS  PubMed  Google Scholar 

  152. Yokota, S. et al. Contributions of FASN and SCD gene polymorphisms on fatty acid composition in muscle from Japanese Black cattle. Anim. Genet. 43, 790–792 (2012).

    CAS  PubMed  Google Scholar 

  153. Zhang, W. et al. Genome-wide association studies for fatty acid metabolic traits in five divergent pig populations. Sci. Rep. 6, 24718 (2016).

    CAS  PubMed  PubMed Central  Google Scholar 

  154. Bolormaa, S. et al. Detailed phenotyping identifies genes with pleiotropic effects on body composition. BMC Genomics 17, 224 (2016).

    PubMed  PubMed Central  Google Scholar 

  155. Momozawa, Y. et al. IBD risk loci are enriched in multigenic regulatory modules encompassing putative causative genes. Nat. Commun. 9, 2427 (2018).

    PubMed  PubMed Central  Google Scholar 

  156. Fang, L. et al. Use of biological priors enhances understanding of genetic architecture and genomic prediction of complex traits within and between dairy cattle breeds. BMC Genomics 18, 604 (2017).

    PubMed  PubMed Central  Google Scholar 

  157. Fragomeni, B. O., Lourenco, D. A. L., Masuda, Y., Legarra, A. & Misztal, I. Incorporation of causative quantitative trait nucleotides in single-step GBLUP. Genet. Sel. Evol. 49, 59 (2017).

    PubMed  PubMed Central  Google Scholar 

  158. Hayes, B. J., Bowman, P. J., Chamberlain, A. C., Verbyla, K. & Goddard, M. E. Accuracy of genomic breeding values in multi-breed dairy cattle populations. Genet. Sel. Evol. 41, 51 (2009).

    PubMed  PubMed Central  Google Scholar 

  159. Bolormaa, S. et al. Accuracy of prediction of genomic breeding values for residual feed intake and carcass and meat quality traits in Bos taurus, Bos indicus, and composite beef cattle. J. Anim. Sci. 91, 3088–3104 (2013).

    CAS  PubMed  Google Scholar 

  160. Rolf, M. M. et al. Comparison of Bayesian models to estimate direct genomic values in multi-breed commercial beef cattle. Genet. Sel. Evol. 47, 23 (2015).

    PubMed  PubMed Central  Google Scholar 

  161. Kemper, K. E. et al. Improved precision of QTL mapping using a nonlinear Bayesian method in multibreed population leads to greater accuracy of across-breed genomic predictions. Genet. Sel. Evol. 47, 29 (2015).

    PubMed  PubMed Central  Google Scholar 

  162. Lu, D. et al. Accuracy of genomic predictions for feed efficiency traits of beef cattle using 50K and imputed HD genotypes. J. Anim. Sci. 94, 1342–1353 (2016).

    CAS  PubMed  Google Scholar 

  163. Hamidi Hay, E. & Roberts, A. Genomic prediction and genome-wide association analysis of female longevity in a composite beef cattle breed. J. Anim. Sci. 95, 1467–1471 (2017).

    CAS  PubMed  Google Scholar 

  164. 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).

    PubMed  PubMed Central  Google Scholar 

  165. Bosse, M. et al. Artificial selection on introduced Asian haplotypes shaped the genetic architecture in European commercial pigs. Proc. Biol. Sci. 282, 20152019 (2015).

    PubMed  PubMed Central  Google Scholar 

  166. Sonesson, A. K., Woolliams, J. A. & Meuwissen, T. H. Genomic selection requires genomic control of inbreeding. Genet. Sel. Evol. 44, 27 (2012).

    PubMed  PubMed Central  Google Scholar 

  167. Sun, C., VanRaden, P. M., O’Connell, J. R., Weigel, K. A. & Gianola, D. Mating programs including genomic relationships and dominance effects. J. Dairy Sci. 96, 8014–8023 (2013).

    CAS  PubMed  Google Scholar 

  168. Pryce, J. E., Hayes, B. J. & Goddard, M. E. Novel strategies to minimize progeny inbreeding while maximizing genetic gain using genomic information. J. Dairy Sci. 95, 377–388 (2012).

    CAS  PubMed  Google Scholar 

  169. Palmiter, R. D. et al. Dramatic growth of mice that develop from eggs microinjected with metallothionein-growth hormone fusion genes. Nature 300, 611–615 (1982).

    CAS  PubMed  PubMed Central  Google Scholar 

  170. Hammer, R. E. et al. Production of transgenic rabbits, sheep and pigs by microinjection. Nature 315, 680–683 (1985).

    CAS  PubMed  Google Scholar 

  171. Wilmut, I., Schnieke, A. E., McWhir, J., Kind, A. J. & Campbell, K. H. Viable offspring derived from fetal and adult mammalian cells. Nature 385, 810–813 (1997).

    CAS  PubMed  Google Scholar 

  172. Tan, W., Proudfoot, C., Lillico, S. G. & Whitelaw, C. B. Gene targeting genome editing: from Dolly to editors. Transgenic Res. 25, 273–287 (2016).

    CAS  PubMed  PubMed Central  Google Scholar 

  173. Kim, H. & Kim, J.-S. A guide to genome engineering with programmable nucleases. Nat. Rev. Genet. 15, 321–334 (2014).

    CAS  PubMed  Google Scholar 

  174. Komor, A. C., Badran, A. H. & Liu, D. R. CRISPR-based technologies for the manipulation of eukaryotic genomes. Cell 168, 1–17 (2017).

    Google Scholar 

  175. Van Eenennaam, A. L. Genetic modification of food animals. Curr. Opin. Biotechnol. 44, 27–34 (2017).

    PubMed  Google Scholar 

  176. Sakuma, T., Nakade, S., Sakane, Y., Suzuki, K. T. & Yamamoto, T. MMEJ-assisted gene knock-in using TALENs and CRISP-Cas9 with the PITCh systems. Nat. Protoc. 11, 118–133 (2016).

    CAS  PubMed  Google Scholar 

  177. Suzuki, K. et al. In vivo genome editing via CRISP/Cas9 mediated homology-independent targeted integration. Nature 540, 144–149 (2016).

    CAS  PubMed  PubMed Central  Google Scholar 

  178. Laible, G., Wei, J. & Wagner, S. Improving livestock for agriculture — technological progress from random transgenesis to precision genome editing heralds a new era. Biotechnol. J. 10, 109–120 (2015).

    CAS  PubMed  Google Scholar 

  179. Lievens, A., Petrillo, M. & Querci Patak, M. A. Genetically modified animals: options and issues for traceability and enforcement. Trends Food Sci. Technol. 44, 159–176 (2015).

    CAS  Google Scholar 

  180. Pirottin, D. et al. Transgenic engineering of male-specific muscular hypertrophy. Proc. Natl Acad. Sci. USA 102, 6413–66418 (2005).

    CAS  PubMed  PubMed Central  Google Scholar 

  181. Wang, H. et al. One-step generation of mice carrying mutations in multiple genes by CRISPR/Cas-mediated genome engineering. Cell 153, 910–918 (2013).

    CAS  PubMed  PubMed Central  Google Scholar 

  182. Cong, L. et al. Multiplex genome engineering using CRISP/Cas systems. Science 339, 819–823 (2013).

    CAS  PubMed  PubMed Central  Google Scholar 

  183. Jenko, J. et al. Potential of promotion of alleles by genome editing to improve quantitative traits in livestock breeding programs. Genet. Sel. Evol. 47, 55 (2015). This paper describes a strategy to combine genome editing and GS to accelerate genetic gains in livestock.

    PubMed  PubMed Central  Google Scholar 

  184. Kasinathan, P. et al. Acceleration of genetic gain in cattle by reduction of generation interval. Sci. Rep. 5, 8674–86766 (2015).

    CAS  PubMed  PubMed Central  Google Scholar 

  185. Blard, G., Zhang, Z., Coppieters, W. & Georges, M. Identifying cows with subclinical mastitis by bulk SNP genotyping of tank milk. J. Dairy Sci. 95, 4109–4113 (2012). This paper describes a method to identify cows with subclinical mastitis by SNP genotyping tank milk.

    CAS  PubMed  Google Scholar 

  186. Hogeveen, H., Huijps, K. & Lam, T. J. Economic aspects of mastitis: new developments. NZ Vet. J. 59, 16–23 (2011).

    CAS  Google Scholar 

  187. Nicoloso, L., Crepaldi, P., Mazza, R., Ajmone-Marsan, P. & Negrini, R. Recent advance in DNA-based traceability and authentication of livestock meat PDO and PGI products. Recent Pat. Food Nutr. Agric. 5, 9–18 (2013).

    CAS  PubMed  Google Scholar 

  188. Ross, E. M., Moate, P. J., Marett, L. C., Cocks, B. G. & Hayes, B. J. Metagenomic predictions: from microbiome to complex health and environmental phenotypes in human and cattle. PLOS ONE 8, e73056 (2013).

    CAS  PubMed  PubMed Central  Google Scholar 

  189. Kittelmann, S. et al. Two different bacterial community types are linked with the low-methane emission trait in sheep. PLOS ONE 9, e103171 (2014).

    PubMed  PubMed Central  Google Scholar 

  190. Wang, M., Pryce, J. E., Savin, K. & Hayes, B. J. Prediction of residual feed intake from genome & metagenome profiles in first lactation Holstein-Friesian dairy cattle. Proc. Assoc. Adv. Breed. Genet. 21, 89–92 (2015).

    Google Scholar 

  191. Frantz, L. A. F. et al. Evidence of long-term gene flow and selection during domestication from analyses of Eurasian wild and domestic pig genomes. Nat. Genet. 47, 1141–1148 (2015). This paper describes a thorough analysis of more than 100 whole-genome sequences of pigs, which strongly suggests that long-term gene flow between wild and domestic pigs counteracted by recurrent selection for domestic traits creates ‘islands of domestication’ in the genome.

    CAS  PubMed  Google Scholar 

  192. Wade, C. M. et al. The mosaic structure of variation in the laboratory mouse genome. Nature 420, 574–578 (2002).

    CAS  PubMed  Google Scholar 

  193. Sankararaman, S. et al. The genomic landscape of Neanderthal ancestry in present-day humans. Nature 507, 354–357 (2014).

    CAS  PubMed  PubMed Central  Google Scholar 

  194. Berry, S. et al. A triad of highly divergent polymeric immunoglobulin receptor (PIGR) haplotytpes with major effect on IgA concentration in cow’s milk. PLOS ONE 8, e57219 (2013).

    CAS  PubMed  PubMed Central  Google Scholar 

  195. Ai, H. et al. Adaptation and possible ancient interspecies introgression in pigs identified by whole-genome sequencing. Nat. Genet. 47, 217–225 (2015).

    CAS  PubMed  Google Scholar 

  196. Rubin, C.-J. et al. Whole genome resequencing reveals loci under selection during chicken domestication. Nature 464, 587–591 (2010). This paper reports the identification of convincing signatures of selective sweeps following poultry domestication.

    CAS  PubMed  Google Scholar 

  197. Carneiro, M. et al. Rabbit genome analysis reveals a polygenic basis for phenotypic change during domestication. Science 345, 1074–1079 (2014). This study provides strong evidence that polygenic adaptation plays a major role in shaping the phenotype of domestic animals in response to human needs.

    CAS  PubMed  PubMed Central  Google Scholar 

  198. Dong, Y. et al. Reference genome of wild goat (Capra aegagrus) and sequencing of goat breeds provide insight into genic basis of goat domestication. BMC Genomics 16, 431–442 (2015).

    PubMed  PubMed Central  Google Scholar 

  199. Park, S. D. E. et al. Genome sequencing of the extinct Eurasian wild aurochs, Bos primigenius, illuminates the phylogeography and evolution of cattle. Genome Biol. 16, 234–249 (2015).

    PubMed  PubMed Central  Google Scholar 

  200. Field, Y. et al. Detection of human adaptation during the past 2,000 years. Science 354, 760–764 (2016).

    CAS  PubMed  PubMed Central  Google Scholar 

  201. Nejati-Javaremi, A., Smith, C. & Gibson, J. Effect of total allelic relationship on accuracy of evaluation and response to selection. J. Anim. Sci. 75, 1738–1745 (1997).

    CAS  PubMed  Google Scholar 

  202. VanRaden, P. M. Efficient methods to compute genomic predictions. J. Dairy Sci. 91, 4414–4423 (2008).

    CAS  PubMed  Google Scholar 

  203. Daetwyler, H. D., Villanueva, B. & Wooliams, J. A. Accuracy of predicting the genetic risk of disease using a genome-wide approach. PLOS ONE 3, e3395 (2008).

    PubMed  PubMed Central  Google Scholar 

  204. Hayes, B. J., Visscher, P. M. & Goddard, M. E. Increased accuracy of artificial selection by using the realized relationship matrix. Genet. Res. 91, 47–60 (2009).

    CAS  Google Scholar 

  205. Pryce, J. E. et al. Genomic selection using a multi-breed, across-country reference population. J. Dairy Sci. 94, 2625–26630 (2011).

    CAS  PubMed  Google Scholar 

  206. Kijas, J. W. et al. Genome-wide analysis of the world’s sheep breeds reveals high levels of historic mixture and strong recent selection. PLOS Biol. 10, e1001258 (2012).

    CAS  PubMed  PubMed Central  Google Scholar 

  207. Moghaddar, N., Gore, K. P., Daetwyler, H. D., Hayes, B. J. & van der Werf, J. H. J. Accuracy of genotype imputation based on random and selected reference sets in purebred and crossbred sheep populations and its effect on accuracy of genomic prediction. Genet. Sel. Evol. 47, 97 (2015).

    PubMed  PubMed Central  Google Scholar 

  208. Lee, S. H., Weerasinghe, W. M., Wray, N. R., Goddard, M. E. & van der Werf, J. H. Using information of relatives in genomic prediction to apply effective stratified medicine. Sci. Rep. 7, 42091 (2017).

    CAS  PubMed  PubMed Central  Google Scholar 

  209. Hill, W. G. & Weir, B. S. Variation in actual relationship as a consequence of Mendelian sampling and linkage. Genet. Res. 93, 47–64 (2011).

    CAS  Google Scholar 

  210. Pursel, V. G. et al. Genetic engineering of livestock. Science 24, 1281–1288 (1989).

    Google Scholar 

  211. Pursel, V. G., Hammer, R. E., Bolt, D. J., Palmiter, R. D. & Brinster, R. L. Integration, expression and germ-line transmission of growth-related genes in pigs. J. Reprod. Fertil. Suppl. 41, 77–87 (1990).

    CAS  PubMed  Google Scholar 

  212. Rexroad, C. E. et al. Production of transgenic sheep with growth-regulating genes. Mol. Reprod. Dev. 1, 164–169 (1989).

    CAS  PubMed  Google Scholar 

  213. Nottle, M. B. et al. in Transgenic Animals in Agriculture (eds Murray, J. D., Anderson, G. B., Oberbauer, A. M. & McGloughin, M. M.) 145–156 (CABI Publishing, Oxon, UK, 1999).

  214. Adams, N. R., Briegel, J. R. & Ward, K. A. The impact of a transgene for ovine growth hormone on the performance of two breeds of sheep. J. Anim. Sci. 80, 2325–2333 (2002).

    CAS  PubMed  Google Scholar 

  215. Pursel, V. G. et al. in Transgenic Animals in Agriculture (eds Murray, J. D., Anderson, G. B., Oberbauer, A. M. & McGloughin, M. M.) 131–144 (CABI Publishing, Oxon, UK, 1999).

  216. Sutrave, P., Kelly, A. M. & Hughes, S. H. Ski can cause selective growth of skeletal muscle in transgenic mice. Genes Dev. 4, 1462–1472 (1990).

    CAS  PubMed  Google Scholar 

  217. Grobet, L. et al. Modulating skeletal muscle mass by postnatal, muscle-specific inactivation of the myostatin gene. Genesis 35, 227–238 (2003).

    CAS  PubMed  Google Scholar 

  218. Luo, J. et al. Efficient generation of myostatin (MSTN) biallelic mutations in cattle using zinc finger nucleases. PLOS ONE 9, e95225 (2014).

    PubMed  PubMed Central  Google Scholar 

  219. Proudfoot, C. et al. Genome edited sheep and cattle. Transgenic Res. 24, 147–153 (2015).

    CAS  PubMed  Google Scholar 

  220. Ni, W. et al. Efficient gene knockout in goats using CRISPR/Cas9 system. PLOS ONE 9, e106718 (2014).

    PubMed  PubMed Central  Google Scholar 

  221. Han, H. et al. One-step generation of myostatin gene knockout sheep via the CRISPR/Cas9 system. Front. Agric. Sci. Eng. 1, 2–5 (2014).

    Google Scholar 

  222. Tessanne, K. et al. Production of transgenic calves expressing an shRNA targeting myostatin. Mol. Reprod. Dev. 79, 176–185 (2012).

    CAS  PubMed  Google Scholar 

  223. Lee, S. J. Quadrupling muscle mass in mice by targeting TGF-beta signaling pathways. PLOS ONE 2, e789 (2007).

    PubMed  PubMed Central  Google Scholar 

  224. Saeki, K. et al. Functional expression of a delta 12 fatty acid desaturase transgene from spinach in transgenic pigs. Proc. Natl Acad. Sci. USA 101, 6361–6366 (2004).

    CAS  PubMed  PubMed Central  Google Scholar 

  225. Lai, L. et al. Generation of cloned transgenic pigs rich in omega-3 fatty acids. Nat. Biotechnol. 24, 345–436 (2006).

    Google Scholar 

  226. Wu, X. et al. Production of cloned transgenic cow expressing omega-3 fatty acids. Transgenic Res. 21, 537–543 (2012).

    CAS  PubMed  Google Scholar 

  227. Zhang, P. et al. Handmade cloned transgenic sheep rich in omega-3 fatty acids. PLOS ONE 8, e55941 (2013).

    CAS  PubMed  PubMed Central  Google Scholar 

  228. Zheng, Q. et al. Reconstitution of UCP1 using CRISP/Cas9 in the white adipose tissue of pigs decreases fat deposition and improves thermogenic capacity. Proc. Natl Acad. Sci. USA 114, E9474–E9482 (2017).

    CAS  PubMed  PubMed Central  Google Scholar 

  229. Berg, F., Gustafson, U. & Andersson, L. The uncoupling protein 1 gene (UCP1) is disrupted in the pig lineage: a genetic explanation for poor thermoregulation in piglets. PLOS Genet. 2, e129 (2006).

    PubMed  PubMed Central  Google Scholar 

  230. Brophy, B. et al. Cloned transgenic cattle produce milk with higher levels of beta-casein and kappa-casein. Nat. Biotechnol. 21, 157–162 (2003).

    CAS  PubMed  Google Scholar 

  231. Martin, P., Szymanowska, M., Zwierzchowski, L. & Leroux, C. The impact of genetic polymorphisms on the protein composition of ruminant milks. Reprod. Nutr. Dev. 42, 433–459 (2002).

    CAS  PubMed  Google Scholar 

  232. Yu, S. et al. Highly efficient modification of beta-lactoglobulin (BLG) gene via zinc-finger nucleases in cattle. Cell Res. 21, 1638–1640 (2011).

    CAS  PubMed  PubMed Central  Google Scholar 

  233. Jabed, A., Wagner, S., McCracken, J., Wells, D. N. & Labile, G. Targeted microRNA expression in dairy cattle directs production of beta-lactoglobulin-free, high-casein milk. Proc. Natl Acad. Sci. USA 109, 16811–16816 (2012).

    CAS  PubMed  PubMed Central  Google Scholar 

  234. Cui, C. et al. Gene targeting by TALEN-induced homologous recombination in goats directs production of beta-lactoglobulin-free, high human lactoferrin milk. Sci. Rep. 5, 10482 (2015).

    PubMed  PubMed Central  Google Scholar 

  235. Zhu, H. et al. Generation of beta-lactoglobulin-modified transgenic goats by homologous recombination. FEBS J. 282, 4600–4613 (2016).

    Google Scholar 

  236. Wheeler, M. B., Bleck, G. T. & Donovan, S. M. Transgenic alteration of sow milk to improve piglet growth and health. Reprod. Suppl. 58, 313–324 (2001).

    CAS  PubMed  Google Scholar 

  237. Wang, J. et al. Expression and characterization of bioactive recombinant human alpha-lactalbumin in the milk of transgenic cloned cows. J. Dairy Sci. 91, 4466–4476 (2008).

    CAS  PubMed  Google Scholar 

  238. Jost, B., Vilotte, J. L., Duluc, I., Rodeau, J. L. & Freund, J. N. Production of low-lactose milk by ectopic expression of intestinal lactase in the mouse mammary gland. Nat. Biotechnol. 17, 160–164 (1999).

    CAS  PubMed  Google Scholar 

  239. Reh, W. A. et al. Using a stearoyl-CoA desaturase transgene to alter milk fatty acid composition. J. Dairy Sci. 87, 3510–3514 (2004).

    CAS  PubMed  Google Scholar 

  240. Damak, S., Su, H., Jay, N. P. & Bullock, D. W. Improved wool production in transgenic sheep expressing insulin-like growth factor 1. Biotechnology 14, 185–188 (1996).

    CAS  PubMed  Google Scholar 

  241. Bawden, C. S., Powell, B. C., Walker, S. K. & Rogers, G. E. Expression of a wool intermediate filament keratin transgene in sheep fibre alters structure. Transgenic Res. 7, 273–287 (1998).

    CAS  PubMed  Google Scholar 

  242. Bawden, C. S. et al. Expression of bacterial cysteine biosynthesis genes in transgenic mice and sheep: toward a new in vivo amino acid biosynthesis pathway and improved wool growth. Transgenic Res. 4, 87–104 (1995).

    CAS  PubMed  Google Scholar 

  243. Maga, E. A. et al. Production and processing of milk from transgenic goats expressing human lysozyme in the mammary gland. J. Dairy Sci. 89, 518–524 (2006).

    CAS  PubMed  Google Scholar 

  244. Liu, X. et al. Generation of mastitis resistance in cows by targeting human lysozyme gene to beta-casein locus using zinc-finger nucleases. Proc. R. Soc. B 281, 20133368 (2014).

    PubMed  PubMed Central  Google Scholar 

  245. Wall, R. J. et al. Genetically enhanced cows resist intramammary Staphylococcus aureus infection. Nat. Biotechnol. 23, 445–451 (2005).

    CAS  PubMed  Google Scholar 

  246. Liu, X. et al. Zinc-finger nickase-mediated insertion of the lysostaphin gene into the beta-casein locus in cloned cows. Nat. Commun. 4, 2565 (2013).

    PubMed  Google Scholar 

  247. Dunham, R. A. et al. Enhanced bacterial disease resistance of transgenic channel catfish Ictalurus punctatus possessing cecropin genes. Mar. Biotechnol. 4, 338–344 (2002).

    CAS  Google Scholar 

  248. Su, F. et al. Generation of transgenic cattle expressing human beta-defensin 3 as an approach to reducing susceptibility to Mycobacterium bovis infection. FEBS J. 283, 776–790 (2016).

    CAS  PubMed  Google Scholar 

  249. Yang, X. et al. Overexpression of porcine beta-defensin 2 enhances resistance to Actinobacillus pleuropneumoniae infection in pigs. Infect. Immun. 83, 2836–2843 (2015).

    CAS  PubMed  PubMed Central  Google Scholar 

  250. Wu, H. et al. TALE nickase-mediated SP110 knockin endows cattle with increased resistance to tuberculosis. Proc. Natl Acad. Sci. USA 112, 1530–1539 (2015).

    Google Scholar 

  251. Pan, H. et al. Ipr1 gene mediates innate immunity to tuberculosis. Nature 434, 767–772 (2005).

    CAS  PubMed  PubMed Central  Google Scholar 

  252. Tosh, K. et al. Variants in the SP110 gene are associated with genetic susceptibility to tuberculosis in West Africa. Proc. Natl Acad. Sci. USA 103, 10364–10368 (2006).

    CAS  PubMed  PubMed Central  Google Scholar 

  253. Fox, G. J. et al. Polymorphisms of SP110 are associated with bothpulmonary and extra-pulmonary tuberculosis among the Vietnamese. PLOS ONE 9, e99496 (2014).

    PubMed  PubMed Central  Google Scholar 

  254. Hu, W. et al. Significant resistance to the infection of foot-and-mouth disease virus in shRNA transgenic pig. Transgenic Res. 21, 901–925 (2012).

    Google Scholar 

  255. Muller, M., Brenig, B., Winnacker, E. L. & Brem, G. Transgenic pigs carrying cDNA copies encoding the murine Mx1 protein which confers resistance to influenza virus infection. Gene 121, 263–270 (1992).

    CAS  PubMed  Google Scholar 

  256. Yan, Q. et al. Production of transgenic pigs overexpressing the antiviral gene Mx1. Cell Regen. 3, 11–22 (2014).

    PubMed  PubMed Central  Google Scholar 

  257. Lyall, J. et al. Suppression of avian influenza transmission in genetically modified chickens. Science 331, 223–226 (2011).

    CAS  PubMed  Google Scholar 

  258. Clements, J. E. et al. Development of transgenic sheep that express the visan virus envelope gene. Virology 200, 370–380 (1994).

    CAS  PubMed  Google Scholar 

  259. Crittenden, L. B. & Salter, D. W. A transgene, alv6, that expresses the envelope of subgroup A avian leucosis virus reduces the rate of congenital transmission of a field strain of avian leucosis virus. Poult. Sci. 71, 799–806 (1992).

    CAS  PubMed  Google Scholar 

  260. Whitworth, K. M. et al. Use of CRISPR/Cas9 system to produce genetically engineered pigs from in vitro derived oocytes and embryos. Biol. Reprod. 91, 1–13 (2014).

    Google Scholar 

  261. Burkard, C. et al. Precision engineering for PRRSV resistance in pigs: macrophages from genome edited pigs lacking CD163 SRCR5 domain are fully resistant to both PRRSV genotypes while maintaining biological function. PLOS Pathog. 13, e1006206 (2017).

    PubMed  PubMed Central  Google Scholar 

  262. Lillico, S. G. et al. Live pigs produced from genome edited zygotes. Sci. Rep. 3, 2847–2851 (2013).

    PubMed  PubMed Central  Google Scholar 

  263. Lillico, S. G. et al. Mammalian interspecies substitution of immune modulatory alleles by genome editing. Sci. Rep. 6, 21645–21650 (2016).

    CAS  PubMed  PubMed Central  Google Scholar 

  264. Richt, J. A. et al. Production of cattle lacking prion protein. Nat. Biotechnol. 25, 132–138 (2007).

    CAS  PubMed  Google Scholar 

  265. Yu, G. et al. Functional disruption of the prion protein gene in cloned goats. J. Gen. Virol. 87, 1019–1027 (2006).

    CAS  PubMed  Google Scholar 

  266. Denning, C. et al. Deletion of the alpha(1,3)galactosyl transferase (GGTA1) gene and the prion protein (PrP) gene in sheep. Nat. Biotechnol. 19, 559–562 (2001).

    CAS  PubMed  Google Scholar 

  267. Golding, M. C., Long, C. R., Carmell, M. A., Hannon, G. J. & Westhusin, M. E. Suppression of prion protein in livestock by RNA interference. Proc. Natl Acad. Sci. USA 103, 5285–5290 (2006).

    CAS  PubMed  PubMed Central  Google Scholar 

  268. Wongsrikeao, P. et al. Combination of the somatic cell nuclear transfer method and RNAi technology for the production of a prion gene-knockdown calf using plasmid vectors harbouring the U6 or tRNA promoter. Prion 5, 39–46 (2011).

    CAS  PubMed  PubMed Central  Google Scholar 

  269. Benestad, S. L., Anstbø, L., Tranulis, M. A., Espenes, A. & Olsaker, I. Healthy goats naturally devoid of prion protein. Vet. Res. 43, 87–91 (2012).

    CAS  PubMed  PubMed Central  Google Scholar 

  270. Willyard, C. Putting sleeping sickness to bed. Nat. Med. 17, 14–17 (2011).

    CAS  PubMed  Google Scholar 

  271. Genovese, G. et al. Association of trypanolytic ApoL1 variants with kidney disease in African Americans. Science 329, 841–845 (2010).

    CAS  PubMed  PubMed Central  Google Scholar 

  272. Hew, C. L., Davies, P. L. & Fletcher, G. Antifreeze protein gene transfer in Atlantic salmon. Mol. Mar. Biol. Biotechnol. 1, 309–317 (1992).

    CAS  PubMed  Google Scholar 

  273. Davies, P. L. & Hew, C. L. Biochemistry of fish antifreeze proteins. FASEB J. 4, 2460–2468 (1990).

    CAS  PubMed  Google Scholar 

  274. Carlson, D. F. et al. Production of hornless dairy cattle from genome-edited cell lines. Nat. Biotechnol. 34, 479–481 (2016). This paper illustrates the engineering of a desirable phenotype in livestock by TALEN-mediated allele swapping and SCNT.

    CAS  PubMed  Google Scholar 

  275. Medugorac, I. et al. Bovine polledness — an autosomal dominant trait with allelic heterogeneity. PLOS ONE 7, e39477 (2012).

    CAS  PubMed  PubMed Central  Google Scholar 

  276. Allais-Bonnet, A. et al. Novel insights into the bovine polled phenotype and horn ontogenesis in Bovidae. PLOS ONE 8, e63512 (2013).

    PubMed  PubMed Central  Google Scholar 

  277. Rothammer, S. et al. The 80-Kb DNA duplication on BTA1 is the only remaining candidate mutation for the polled phenotype of Friesian origin. Genet. Sel. Evol. 46, 44 (2014).

    PubMed  PubMed Central  Google Scholar 

  278. Golovan, S. P. et al. Pigs expressing salivary phytase produce low-phosphorus manure. Nat. Biotechnol. 19, 741–745 (2001).

    CAS  PubMed  Google Scholar 

Download references

Acknowledgements

C.C. is Senior Research Associate of the Fonds de la Recherche Scientifique (FRS-FNRS). Research in animal genomics conducted by the authors is funded by the European Research Council (ERC) Advanced DAMONA, H2020 GpE and WALInnov CAUSEL grants to M.G. and the DGARNE Rilouke and ULiège RetroBlue grants to C.C. The authors are grateful to T. Druet for fruitful discussions and comments on the manuscript and to M. Goddard for excellent discussions over the years.

Reviewer information

Nature Reviews Genetics thanks D. Garrick and A. Legarra for their contribution to the peer review of this work.

Author information

Authors and Affiliations

Authors

Contributions

All authors researched data for the article, made substantial contributions to discussions of the content and reviewed and/or edited the manuscript before submission. M.G. and B.H. wrote the article.

Corresponding author

Correspondence to Michel Georges.

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.

Related links

dbSNP: https://www.ncbi.nlm.nih.gov/projects/SNP

International Mouse Phenotype Consortium: http://www.mousephenotype.org

NCBI Genome database: https://www.ncbi.nlm.nih.gov/genome

Online Mendelian Inheritance in Animals: http://www.omia.org/home

Online Mendelian Inheritance in Man: https://www.omim.org

Supplementary information

Glossary

Within-breed selection

A process by which sires and dams that have above average breeding values are selected as parents to produce the next generation of animals.

Genetic gains

Differences in the average breeding values of the population before and after selection. Genetic gain is a function of the amount of genetic variance, the accuracy of selection, the intensity of selection and the generation interval.

Quantitative trait loci

(QTL). Regions in the genome that encompass genetic variants with an effect on a quantitative trait of interest.

Genome-wide association studies

(GWAS). Scan of the entire genome to identify genetic variants for which variation in genotype is associated with variation for one or more phenotypes of interest.

Genomic selection

(GS). An ensemble of methods to estimate the breeding values of individual animals on the basis of genome-wide single-nucleotide polymorphism genotype information.

Single-nucleotide polymorphism arrays

(SNP arrays). Microarrays used to determine the genotype of individuals for hundreds to millions of SNPs at once.

Progeny testing

(PT). An approach by which the breeding value of an animal is estimated from phenotypic measures made on its progeny.

Genetic architecture

The description of the number, location and effects of the genetic variants that affect a phenotype of interest.

Genotype imputation

The in silico prediction of the genotype of an individual for ungenotyped variants on the basis of known genotypes at neighbouring variants and a reference population with genotype information for all variants. Imputation exploits the nonrandom association of alleles at neighbouring variants, referred to as linkage disequilibrium.

Soft sweeps

The process by which the frequency of a favourable old variant rapidly increases in the population by positive selection until eventual fixation. Soft sweeps are not associated with the concomitant fixation of one predominant haplotype, as the variant has been distributed over multiple haplotypes by recombination before selection. Old variants that are substrates for new selection constitute the standing variation in the population.

Epigenome

The combination of chemical modifications of the DNA sequence (such as cytosine methylation) or nucleosomes (such as methylation of Lys 27 of histone H3) that mark functionally distinct segments of the genome (such as active enhancers) and are inherited mitotically and/or meiotically.

ChIP-Seq

A combination of chromatin immunoprecipitation and next-generation sequencing for genome-wide mapping of binding sites occupied by specific DNA-binding proteins or chromatin regions enriched in specific histone modifications.

DNase-Seq

A method based on next-generation sequencing for genome-wide detection of gene-switch components on the basis of their open chromatin conformation and resulting hypersensitivity to digestion by DNase I.

ATAC-Seq

An assay based on next-generation sequencing for genome-wide detection of gene-switch components on the basis of their open chromatin conformation and resulting increased accessibility to transposase Tn5.

Expression quantitative trait loci

(eQTL). Quantitative trait loci that influence the transcript levels of specific genes. Cis-eQTL are due to regulatory variants that control the levels of RNA molecules transcribed from gene copies located on the same DNA molecule as the variant. Trans-eQTL are due to regulatory variants that can also control the levels of RNA molecules transcribed from gene copies located on different DNA molecules to the variant (homologous or other chromosomes).

Pleiotropy

The ability of a genetic variant to affect more than one phenotype.

Hypomorphic

Pertaining to an allele with partial loss of function when compared with the wild-type allele.

Overdominance

The phenotypic superiority (for example, on a quantitative scale) of heterozygotes (‘Aa’) over both homozygous classes (‘AA’ and ‘aa’).

Haplosufficient

Pertaining to genes for which one functional copy is sufficient to ensure normal development and function.

Compound heterozygosity

Pertaining to the inheritance of two distinct mutations in different alleles of the same gene, one from each parent.

Autozygosity mapping

Mapping of a recessive mutation on the basis that all affected individuals will be homozygous for the same (autozygous) haplotype. Typically applied in genetically isolated populations in which the hypothesis of allelic homogeneity is reasonable.

Modifier locus

A locus with variants that may (depending on the genotype of the individual) affect the phenotypic expression conferred by specific variants at another locus. The effects of modifier loci include suppression and epistasis.

Reverse genetic screens

Process aimed at completing the phenotype–genotype map by sorting individuals according to their genotype at a variant with unknown function and searching for shared phenotypes, as opposed to forward genetics, which consists of sorting individuals according to a phenotype and searching for shared variants.

Haplotypes

A combination of alleles at multiple variant positions transmitted by a gamete. The term is often used to describe variants that are located close to each other in the genome.

Linkage disequilibrium

(LD). The nonrandom association of alleles at two or more loci, which is manifest by the over-representation of specific haplotypes and the concomitant under-representation of others.

Selection index

A weighted sum of breeding values for several traits, each weighted by economic or perceived relevance.

Kinship coefficient

A measure of genetic relatedness between two individuals. The kinship coefficient corresponds to the probability that two alleles (one from each individual) drawn at random from the two possible alleles (maternal and paternal) for each individual for a randomly selected locus in the genome are identical by descent. The kinship coefficient between two individuals corresponds to the expected inbreeding coefficient of their putative offspring.

Hard sweeps

The process by which the frequency of a favourable new variant rapidly increases in the population by positive selection until eventual fixation of the variant and the haplotype upon which it occurred.

Balancing selection

A selective force on a locus that leads to a steady state whereby multiple alleles are simultaneously maintained in the population, rather than one allele becoming fixed at the expense of the others.

Non-synonymous

Variants that cause a change in the amino acid sequence of a protein. By contrast, synonymous variants are variants in the open reading frame of a protein-coding gene that do not change the amino acid sequence. Most non-synonymous variants affect the first and second codon positions, while most synonymous variants affect the third codon position.

Intermediate phenotypes

Phenotypes that mediate the link between a causative variant and the end-point disease or agricultural phenotype of interest — includes transcript, protein and metabolite levels.

Gene flow

The passage of alleles between populations as a result of migration or interbreeding.

Polygenic adaptation

The process by which a phenotype caused by many genes evolves in a population under selection, not by massive changes in the frequency of a few variants with major effects on the phenotype (hard and soft sweeps) but by very small changes in the frequency of many variants with minor effects on the phenotype.

Mosaicism

The occurrence of mutations in some but not all cells of an organism that is entirely derived from a single zygote.

Gartner hype cycle

A model first proposed by the Gartner firm to explain the phases of maturation, adoption and social application of new technologies.

Rights and permissions

Reprints and permissions

About this article

Check for updates. Verify currency and authenticity via CrossMark

Cite this article

Georges, M., Charlier, C. & Hayes, B. Harnessing genomic information for livestock improvement. Nat Rev Genet 20, 135–156 (2019). https://doi.org/10.1038/s41576-018-0082-2

Download citation

  • Published:

  • Issue Date:

  • DOI: https://doi.org/10.1038/s41576-018-0082-2

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