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.

  • Review Article
  • Published:

Epistasis and quantitative traits: using model organisms to study gene–gene interactions

Key Points

  • A major challenge of contemporary biology is to understand how naturally occurring genetic variation causes phenotypic variation in quantitative traits. Despite the biological plausibility that genetic variation affects nonlinear networks at multiple levels of biological organization, most efforts to explain the relationship between genetic and phenotypic variation concentrate on additive effects of individual loci.

  • Mapping gene–gene interactions (that is, epistasis) is challenging experimentally, statistically and computationally owing to the large number of interactions to be evaluated. This number is of the order of the square of the number of single-locus tests for pairwise interactions.

  • Epistatic interactions for quantitative traits result in a change of either the magnitude or the direction of allelic effects at one locus, depending on the genotype at the interacting locus. With epistasis, the additive effect (that is, the main effect) of a locus changes with the allele frequency of the interacting locus, such that estimates of effects at a single interacting locus will differ between populations with different allele frequencies.

  • Epistasis generates mostly additive variance for quantitative traits; therefore, the observation that most genetic variance for quantitative traits is additive is not inconsistent with an underlying epistatic genetic architecture. Experimental designs that are only possible in model organisms allow the exploration of the gene–gene interaction space, and the results of these analyses indicate that epistasis is pervasive.

  • Genetic interaction networks are derived by assessing quantitative trait phenotypes of wild-type, single-mutant and double-mutant genotypes. These networks show scale-free and small-world properties, such that the major features of network topology may be inferred by focusing on major hub genes and on interactions among the genes with which they interact. Combining genomics with mutant-interaction screens may aid the identification of network hubs.

  • Taking advantage of multifactorial perturbations in quantitative trait locus (QTL)-mapping populations is less laborious than constructing all pairwise combinations of mutant alleles, and the ability to construct chromosome substitution lines, introgression lines and near-isogenic lines in model organisms maximizes power to detect interactions. Epistasis is commonly observed, even between loci without significant main effects, but there are only a few cases in which the actual interacting variants have been identified.

  • Natural populations harbour hidden reservoirs of cryptic genetic variation that can be revealed by introducing mutations onto wild-derived backgrounds. When this approach is implemented in a QTL-mapping population, it is a powerful experimental design for identifying naturally occurring variants that either enhance or suppress the mutant phenotype.

  • Observations of cryptic genetic variation and less-than-additive epistatic interactions between QTLs suggest that natural populations have evolved suppressing epistatic interactions as homeostatic (that is, canalizing) mechanisms for quantitative traits. Pervasive epistasis has consequences for plant and animal breeding, evolutionary biology and human genetics.

  • In the future, assessment of the pleiotropic effects of genetic interactions on transcriptional, metabolic and protein–protein interaction networks will provide insights into the mechanistic basis of epistasis for organismal phenotypes.

Abstract

The role of epistasis in the genetic architecture of quantitative traits is controversial, despite the biological plausibility that nonlinear molecular interactions underpin the genotype–phenotype map. This controversy arises because most genetic variation for quantitative traits is additive. However, additive variance is consistent with pervasive epistasis. In this Review, I discuss experimental designs to detect the contribution of epistasis to quantitative trait phenotypes in model organisms. These studies indicate that epistasis is common, and that additivity can be an emergent property of underlying genetic interaction networks. Epistasis causes hidden quantitative genetic variation in natural populations and could be responsible for the small additive effects, missing heritability and the lack of replication that are typically observed for human complex traits.

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

Access options

Buy this article

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

Figure 1: Two-locus genotypic effects.
Figure 2: Quantitative genetics of additive-by- additive interactions.
Figure 3: Genotypes for mapping QTLs between two genetically divergent lines.
Figure 4: Two-dimensional search for epistatic interactions.
Figure 5: Epistasis between naturally occurring variation and mutations in D. melanogaster.

Similar content being viewed by others

References

  1. Kauffman, S. A. The Origins of Order (Oxford Univ. Press,1993).

    Google Scholar 

  2. Phillips, P. C. Epistasis — the essential role of gene interactions in the structure and evolution of genetic systems. Nature Rev. Genet. 9, 855–867 (2008). This is a comprehensive review that describes the importance of studying genetic interactions with respect to dissecting regulatory pathways, mapping the genetic basis of complex traits and understanding both the structure and the evolution of complex systems.

    Article  CAS  PubMed  Google Scholar 

  3. Waddington, C. H. Canalization of development and the inheritance of acquired characters. Nature 150, 563–565 (1942).

    Article  Google Scholar 

  4. Waddington, C. H. The Strategy of Genes (George Allen and Unwin,1957).

    Google Scholar 

  5. Dobzhansky, T. Genetics and the Origin of Species (Columbia Univ. Press,1937).

    Google Scholar 

  6. Muller, H. J. in The New Systematics (ed. Huxley, J. S.) 185–268 (Clarendon, 1940).

    Google Scholar 

  7. Carlborg, O. & Haley, C. S. Epistasis: too often neglected in complex trait studies? Nature Rev. Genet. 5, 618–625 (2004).

    Article  CAS  PubMed  Google Scholar 

  8. Cordell, H. J. Detecting gene–gene interactions that underlie human diseases. Nature Rev. Genet. 10, 392–404 (2009).

    Article  CAS  PubMed  Google Scholar 

  9. Cheverud, J. M. & Routman, E. J. Epistasis and its contribution to genetic variance components. Genetics 139, 1455–1461 (1995). This is the first paper to clearly articulate the difference between the effect of epistasis on individual genotypes — which is independent of allele frequency — and the contribution of epistasis to epistatic variance for a quantitative trait, which does depend on allele frequency.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  10. Falconer, D. S. & Mackay, T. F. C. Introduction to Quantitative Genetics (Longman, 1996).

    Google Scholar 

  11. Lynch, M. & Walsh, J. B. Genetics and Analysis of Quantitative Traits (Sinauer Associates, 1998).

    Google Scholar 

  12. Fisher, R. A. The Genetical Theory of Natural Selection (Clarendon, 1930).

    Book  Google Scholar 

  13. Wright, S. Evolution in Mendelian populations. Genetics 16, 97–159 (1931).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

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

    Article  PubMed  PubMed Central  CAS  Google Scholar 

  15. Alvarez-Castro, J. M. & Carlborg, O. A unified model for functional and statistical epistasis and its application in quantitative trait loci analysis. Genetics 176, 1151–1167 (2007).

    Article  PubMed  PubMed Central  Google Scholar 

  16. 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).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  17. Hayes, B. J., Lewin, H. A. & Goddard, M. E. The future of livestock breeding: genomic selection for efficiency, reduced emissions intensity, and adaptation. Trends Genet. 29, 206–214 (2013).

    Article  CAS  PubMed  Google Scholar 

  18. Elena, S. F. & Lenski, R. E. Test of synergistic interactions among deleterious mutations in bacteria. Nature 390, 395–398 (1997). This is one of the first studies to show epistasis for new mutations that affect fitness, using the E. coli model system.

    Article  CAS  PubMed  Google Scholar 

  19. Clark, A. G. & Wang, L. Epistasis in measured genotypes: Drosophila P-element insertions. Genetics 147, 157–163 (1997).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  20. Flint, J. & Mackay, T. F. C. Genetic architecture of quantitative traits in mice, flies, and humans. Genome Res. 19, 723–733 (2009).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  21. Mackay, T. F. C., Stone, E. A. & Ayroles, J. F. The genetics of quantitative traits: challenges and prospects. Nature Rev. Genet. 10, 565–577 (2009).

    Article  CAS  PubMed  Google Scholar 

  22. Magwire, M. M. et al. Quantitative and molecular genetic analyses of mutations increasing Drosophila life span. PLoS Genet. 6, e1001037 (2010).

    Article  PubMed  PubMed Central  CAS  Google Scholar 

  23. Zwarts, L. et al. Complex genetic architecture of Drosophila aggressive behavior. Proc. Natl Acad. Sci. USA 108, 17070–17075 (2011).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  24. Giaever, G. et al. Functional profiling of the Saccharomyces cerevisiae genome. Nature 418, 387–391 (2002).

    Article  CAS  PubMed  Google Scholar 

  25. Boone, C., Bussey, H. & Andrews, B. J. Exploring genetic interactions and networks with yeast. Nature Rev. Genet. 8, 437–449 (2007).

    Article  CAS  PubMed  Google Scholar 

  26. Tong, A. H. et al. Global mapping of the yeast genetic interaction network. Science 303, 808–813 (2004). This paper describes the first high-throughput analysis of synthetic lethal genetic interaction mapping in yeast, from which the first large-scale genetic interaction network was derived.

    Article  CAS  PubMed  Google Scholar 

  27. Collins, S. R., Schuldiner, M., Krogan, N. & Weissman, J. S. A strategy for extracting and analyzing large-scale quantitative epistatic interaction data. Genome Biol. 7, R63 (2006).

    Article  PubMed  PubMed Central  CAS  Google Scholar 

  28. St Onge, R. P. et al. Systematic pathway analysis using high-resolution fitness profiling of combinatorial gene deletions. Nature Genet. 39, 199–206 (2007).

    Article  CAS  PubMed  Google Scholar 

  29. Costanzo, M. et al. The genetic landscape of a cell. Science 327, 425–431 (2010).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  30. Szappanos, B. et al. An integrated approach to characterize genetic interaction networks in yeast metabolism. Nature Genet. 43, 656–662 (2011).

    Article  CAS  PubMed  Google Scholar 

  31. Lehner, B., Crombie, C., Tischler, J., Fortunato, A. & Fraser, A. G. Systematic mapping of genetic interactions in Caenorhabditis elegans identifies common modifiers of diverse signaling pathways. Nature Genet. 38, 896–903 (2006). This study describes the first large-scale mapping of genetic interaction networks in a metazoan, which was done by feeding hypomorphic C. elegans mutants with arrays of bacteria that expressed double-stranded RNAi molecules which target specific signalling pathways.

    Article  CAS  PubMed  Google Scholar 

  32. Byrne, A. B. et al. A global analysis of genetic interactions in Caenorhabditis elegans. J. Biol. 6, 8 (2007).

    Article  PubMed  PubMed Central  Google Scholar 

  33. Horn, T. et al. Mapping of signaling networks through synthetic genetic interaction analysis by RNAi. Nature Methods 8, 341–346 (2011).

    Article  CAS  PubMed  Google Scholar 

  34. Van Driessche, N. et al. Epistasis analysis with global transcriptional phenotypes. Nature Genet. 37, 471–477 (2005).

    Article  CAS  PubMed  Google Scholar 

  35. Aylor, D. L. & Zeng, Z.-B. From classical genetics to quantitative genetics to systems biology: modeling epistasis. PLoS Genet. 4, e1000029 (2008).

    Article  PubMed  PubMed Central  CAS  Google Scholar 

  36. Carter, G. W. et al. Prediction of phenotype and gene expression for combinations of mutations. Mol. Syst. Biol. 3, 96 (2007).

    Article  PubMed  PubMed Central  Google Scholar 

  37. Bellen, H. J. et al. The Drosophila gene disruption project: progress using transposons with distinctive site specificities. Genetics 188, 731–743 (2011).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  38. Dietzl, G. et al. A genome-wide transgenic RNAi library for conditional gene inactivation in Drosophila. Nature 448, 151–156 (2007).

    Article  CAS  PubMed  Google Scholar 

  39. Sabrautzki, S. et al. New mouse models for metabolic bone diseases generated by genome-wide ENU mutagenesis. Mamm. Genome 23, 416–430 (2012).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  40. O'Rourke, E. J., Conery, A. L. & Moy, T. I. Whole-animal high-throughput screens: the C. elegans model. Methods Mol. Biol. 486, 57–75 (2009).

    Article  CAS  PubMed  Google Scholar 

  41. O'Malley, R. C., Alonso, J. M., Kim, C. J., Leisse, T. J. & Ecker, J. R. An adapter ligation-mediated PCR method for high-throughput mapping of T-DNA inserts in the Arabidopsis genome. Nature Protoc. 2, 2910–2917 (2007).

    Article  CAS  Google Scholar 

  42. Anholt, R. R. H. et al. The genetic architecture of odor-guided behavior in Drosophila: epistasis and the transcriptome. Nature Genet. 35, 180–184 (2003). This is the first study to show that genes that are transcriptionally co-regulated in a mutant background themselves affect the same trait as the focal mutations, and that mutations in these genes epistatically interact with the focal mutation.

    Article  CAS  PubMed  Google Scholar 

  43. Deutschbauer, A. M. & Davis, R. W. Quantitative trait loci mapped to single-nucleotide resolution in yeast. Nature Genet. 37, 1333–1340 (2005).

    Article  CAS  PubMed  Google Scholar 

  44. Gerke, J., Lorenz, K. & Cohen, B. Genetic interactions between transcription factors cause natural variation in yeast. Science 323, 498–501 (2009).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  45. Brem, R. B., Storey, J. D., Whittle, J. & Kruglyak, L. Genetic interactions between polymorphisms that affect gene expression in yeast. Nature 436, 701–703 (2005).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  46. Gaertner, B. E., Parmenter, M. D., Rockman, M. V., Kruglyak, L. & Phillips, P. C. More than the sum of its parts: a complex epistatic network underlies natural variation in thermal preference behavior in Caenorhabditis elegans. Genetics 192, 1533–1542 (2012).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  47. Cheverud, J. M. et al. Genetic architecture of adiposity in the cross of LG/J and SM/J inbred mice. Mamm. Genome 12, 3–12 (2001).

    Article  CAS  PubMed  Google Scholar 

  48. Stylianou, I. M. et al. Quantitative trait locus analysis for obesity reveals multiple networks of interacting loci. Mamm. Genome 17, 22–36 (2006).

    Article  PubMed  Google Scholar 

  49. Jarvis, J. P. & Cheverud, J. M. Mapping the epistatic network underlying murine reproductive fatpad variation. Genetics 187, 597–610 (2011).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  50. Leamy, L. J., Gordon, R. R. & Pomp, D. Sex-, diet-, and cancer-dependent epistatic effects on complex traits in mice. Front. Genet. 2, 71 (2011).

    Article  PubMed  PubMed Central  Google Scholar 

  51. Peripato, A. C. et al. Epistasis affecting litter size in mice. J. Evol. Biol. 17, 593–602 (2004).

    Article  CAS  PubMed  Google Scholar 

  52. Hanlon, P. et al. Three-locus and four-locus QTL interactions influence mouse insulin-like growth factor-I. Physiol. Genom. 26, 46–54 (2006).

    Article  CAS  Google Scholar 

  53. Carlborg, O., Jacobsson, L., Ahgren, P., Siegel, P. & Andersson, L. Epistasis and the release of genetic variation during long-term selection. Nature Genet. 38, 418–420 (2006). This study documents the importance of epistatic interactions that govern long-term response to artificial selection for growth rate in chickens.

    Article  CAS  PubMed  Google Scholar 

  54. Pettersson, M., Besnier, F., Siegel, P. B. & Carlborg, O. Replication and explorations of high-order epistasis using a large advanced intercross line pedigree. PLoS Genet. 7, e1002180 (2011).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  55. Kroymann, J. & Mitchell-Olds, T. Epistasis and balanced polymorphism influencing complex trait variation. Nature 435, 95–98 (2005). This paper reports the surprising observation that a small chromosome interval that has no effect on growth rate of A. thaliana contains two epistatically interacting QTLs that affect growth.

    Article  CAS  PubMed  Google Scholar 

  56. Wentzel, A. M. et al. Linking metabolic QTLs with network and cis-eQTLs controlling biosynthetic pathways. PLoS Genet. 3, 1687–1701 (2007).

    Article  CAS  Google Scholar 

  57. Rowe, H. C., Hansen, B. G., Halkier, B. A. & Kliebenstein, D. J. Biochemical networks and epistasis shape the Arabidopsis thaliana metabolome. Plant Cell 20, 1199–1216 (2008).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  58. Doebley, J., Stec, A. & Gustus, C. teosinte branched1 and the origin of maize: evidence for epistasis and the evolution of dominance. Genetics 141, 333–346 (1995).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  59. Stam, L. F. & Laurie, C. C. Molecular dissection of a major gene effect on a quantitative trait: the level of alcohol dehydrogenase expression in Drosophila melanogaster. Genetics 144, 1559–1564 (1996). This classic paper uses P-element transformation to create all possible combinations of three segments of the Adh gene and showed that each of the three segments affects Adh activity, and that there is, surprisingly, epistasis between two different regions within the gene.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  60. Shao, H. et al. Genetic architecture of complex traits: large phenotypic effects and pervasive epistasis. Proc. Natl Acad. Sci. USA 105, 19910–19914 (2008).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  61. Gale, G. D. et al. A genome-wide panel of congenic mice reveals widespread epistasis of behavior quantitative trait loci. Mol. Psychiatry 14, 631–645 (2009).

    Article  CAS  PubMed  Google Scholar 

  62. Spiezio, S. H., Takada, T., Shiroishi, T. & Nadeau, J. H. Genetic divergence and the genetic architecture of complex traits in chromosome substitution strains of mice. BMC Genet. 13, 38 (2012).

    Article  PubMed  PubMed Central  Google Scholar 

  63. Edwards, A. C. & Mackay, T. F. C. Quantitative trait loci for aggressive behavior in Drosophila melanogaster. Genetics 182, 889–897 (2009).

    Article  PubMed  PubMed Central  Google Scholar 

  64. Eshed, Y. & Zamir, D. Less-than-additive epistatic interactions of quantitative trait loci in tomato. Genetics 143, 1807–1817 (1996).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  65. Causse, M., Chaïb, J., Lecomte, L., Buret, M. & Hospital, F. Both additivity and epistasis control the genetic variation for fruit quality traits in tomato. Theor. Appl. Genet. 115, 429–442 (2007).

    Article  CAS  PubMed  Google Scholar 

  66. Greene, C. S., Penrod, N. M., Williams, S. M. & Moore, J. H. Failure to replicate a genetic association may provide important clues about genetic architecture. PLoS ONE 4, e5639 (2009).

    Article  PubMed  PubMed Central  CAS  Google Scholar 

  67. Mackay, T. F. C. et al. The Drosophila melanogaster Genetic Reference Panel. Nature 482, 173–178 (2012).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  68. Huang, W. et al. Epistasis dominates the genetic architecture of Drosophila quantitative traits. Proc. Natl Acad. Sci. USA 109, 15553–15559 (2012).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  69. Ehrenreich, I. M. et al. Dissection of genetically complex traits with extremely large pools of yeast segregants. Nature 464, 1039–1042 (2010).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  70. Rendel, J. M. Canalization of the scute phenotype of Drosophila. Evolution 13, 425–439 (1959).

    Article  Google Scholar 

  71. Gibson, G. & van Helden, S. Is function of the Drosophila homeotic gene Ultrabithorax canalized? Genetics 147, 1155–1168 (1997).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  72. Gibson, G., Wemple, M. & van Helden, S. Potential variance affecting homeotic Ultrabithorax and Antennapedia phenotypes in Drosophila melanogaster. Genetics 151, 1081–1091 (1999).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  73. Polaczyk, P. J., Gasperini, R. & Gibson, G. Naturally occurring genetic variation affects Drosophila photoreceptor determination. Dev. Genes Evol. 207, 462–470 (1998).

    Article  CAS  PubMed  Google Scholar 

  74. Dworkin, I. et al. Genomic consequences of background effects on scalloped mutant expressivity in the wing of Drosophila melanogaster. Genetics 181, 1065–1076 (2009).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  75. Dworkin, I., Palsson, A., Birdsall, K. & Gibson, G. Evidence that Egfr contributes to cryptic genetic variation for photoreceptor determination in natural populations of Drosophila melanogaster. Curr. Biol. 13, 1888–1893 (2003).

    Article  CAS  PubMed  Google Scholar 

  76. Rutherford, S. L. & Lindquist, S. Hsp90 as a capacitor for morphological evolution. Nature 396, 336–342 (1998).

    Article  CAS  PubMed  Google Scholar 

  77. Gibson, G. & Dworkin, I. Uncovering cryptic genetic variation. Nature Rev. Genet. 5, 681–690 (2004). This review documents the presence of cryptic genetic variation for complex traits and argues that such variation is important in understanding the genetic bases of common diseases in humans, artificial selection response in livestock and crops, and evolutionary responses to new mutations.

    Article  CAS  PubMed  Google Scholar 

  78. Spencer, C. C., Howell, C. E., Wright, A. R. & Promislow, D. E. Testing an 'aging gene' in long-lived Drosophila strains: increased longevity depends on sex and genetic background. Aging Cell 2, 123–130 (2003).

    Article  CAS  PubMed  Google Scholar 

  79. Yamamoto, A., Anholt, R. R. H. & Mackay, T. F. C. Epistatic interactions attenuate mutations affecting startle behaviour in Drosophila melanogaster. Genet. Res. 91, 373–382 (2009).

    Article  CAS  Google Scholar 

  80. Swarup, S. et al. Extensive epistasis for olfactory behavior, sleep and waking activity in Drosophila melanogaster. Genet. Res. 94, 9–20 (2012).

    Article  CAS  Google Scholar 

  81. Cheng, Y. et al. Mapping genetic loci that interact with myostatin to affect growth traits. Heredity 107, 565–573 (2011).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  82. Chaikam, V. et al. Use of Mutant-Assisted Gene Identification and Characterization (MAGIC) to identify novel genetic loci that modify the maize hypersensitive response. Theor. Appl. Genet. 123, 985–997 (2011).

    Article  PubMed  Google Scholar 

  83. Sangster, T. A. et al. HSP90-buffered genetic variation is common in Arabidopsis thaliana. Proc. Natl Acad. Sci. USA 105, 2969–2974 (2008).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  84. Manolio, T. A. et al. Finding the missing heritability of complex diseases. Nature 461, 747–753 (2009).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  85. Valdar, W. et al. Genome-wide genetic association of complex traits in heterogeneous stock mice. Nature Genet. 38, 879–887 (2006).

    Article  CAS  PubMed  Google Scholar 

  86. Buckler, E. S. et al. The genetic architecture of maize flowering time. Science 325, 714–718 (2009).

    Article  CAS  PubMed  Google Scholar 

  87. Barton, N. H. & Turelli, M. Evolutionary quantitative genetics: how little do we know? Annu. Rev. Genet. 23, 337–370 (1989).

    Article  CAS  PubMed  Google Scholar 

  88. Barton, N. H. & Keightley, P. D. Understanding quantitative genetic variation. Nature Rev. Genet. 3, 11–21 (2002).

    Article  CAS  PubMed  Google Scholar 

  89. Johnson, T. & Barton, N. Theoretical models of selection and mutation on quantitative traits. Phil. Trans. R. Soc. B 360, 1411–1425 (2005).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  90. Houle, D., Morikawa, B. & Lynch, M. Comparing mutational variabilities. Genetics 143, 1467–1483 (1996).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  91. Zhang, X.-S. & Hill, W. G. Genetic variability under mutation selection balance. Trends Ecol. Evol. 20, 468–470 (2005).

    Article  PubMed  Google Scholar 

  92. Carson, H. L. & Templeton, A. R. Genetic revolutions in relation to speciation phenomena: the founding of new populations. Annu. Rev. Ecol. Syst. 15, 97–131 (1984).

    Article  Google Scholar 

  93. Goodnight, C. J. On the effect of founder events on epistatic genetic variance. Evolution 41, 80–91 (1987).

    Article  PubMed  Google Scholar 

  94. Tachida, H. & Cockerham, C. C. A building block model for quantitative genetics. Genetics 121, 839–844 (1989).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

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

    Article  CAS  PubMed  Google Scholar 

  96. Makowsky, R. et al. Beyond missing heritability: prediction of complex traits. PLoS Genet. 7, e1002051 (2011).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  97. Benyamin, B. et al. Childhood intelligence is heritable, highly polygenic and associated with FNBP1L. Mol. Psychiatry http://dx.doi.org/10.1038/mp.2012.184 (2013).

  98. Hu, Z. et al. Genomic value prediction for quantitative traits under the epistatic model. BMC Genet. 12, 15 (2011).

    Article  PubMed  PubMed Central  Google Scholar 

  99. Long, N., Gianola, D., Rosa, G. J. & Weigel, K. A. Marker-assisted prediction of non-additive genetic values. Genetica 139, 843–854 (2011).

    Article  PubMed  Google Scholar 

  100. Bulmer, M. G. The Mathematical Theory of Quantitative Genetics (Clarendon,1985).

    Google Scholar 

  101. Anholt, R. R. H. & Mackay, T. F. C. Principles of Behavioral Genetics (Elsevier, 2010).

    Google Scholar 

  102. Mackay, T. F. C. et al. Genetics and genomics of Drosophila mating behavior. Proc. Natl Acad. Sci. USA 102 (Suppl. 1), 6622–6629 (2005).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  103. Ayroles, J. F. et al. Systems genetics of complex traits in Drosophila melanogaster. Nature Genet. 41, 299–307 (2009).

    Article  CAS  PubMed  Google Scholar 

  104. Jordan, K. W., Carbone, M. A., Yamamoto, A., Morgan, T. J. & Mackay, T. F. C. Quantitative genomics of locomotor behavior in Drosophila melanogaster. Genome Biol. 8, R172 (2007).

    Article  PubMed  PubMed Central  CAS  Google Scholar 

  105. Edwards, A. C., Rollmann, S. M., Morgan, T. J. & Mackay, T. F. C. Quantitative genomics of aggressive behavior in Drosophila melanogaster. PLoS Genet. 2, e154 (2006).

    Article  PubMed  PubMed Central  CAS  Google Scholar 

  106. Morozova, T. V., Anholt, R. R. & Mackay, T. F. C. Phenotypic and transcriptional response to selection for alcohol sensitivity in Drosophila melanogaster. Genome Biol. 8, R231 (2007).

    Article  PubMed  PubMed Central  CAS  Google Scholar 

  107. Leips, J. & Mackay, T. F. C. Quantitative trait loci for lifespan in Drosophila melanogaster: interactions with genetic background and larval density. Genetics 155, 1773–1788 (2000).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

Download references

Acknowledgements

The author thanks S. Zhou for helping with figures 2 and 5b, and R. Anholt for comments on the manuscript. Work in the Mackay laboratory is supported by the US National Institutes of Health grants R01 GM45146, R01 GM076083, R01 GM59469 and R01 AA016560.

Author information

Authors and Affiliations

Authors

Corresponding author

Correspondence to Trudy F. C. Mackay.

Ethics declarations

Competing interests

The author declares no competing financial interests.

PowerPoint slides

Glossary

Main effect

The effect of a variable averaged over all other variables; also known as marginal effect.

Heterosis

The phenomenon whereby the mean value of a quantitative trait in the F1 progeny of two inbred lines exceeds, in the direction of increased fitness, either the mean value of the parental lines (that is, mid-parent heterosis) or the mean value of the best parent (that is, high parent heterosis); also known as hybrid vigour.

Missing heritability

The phenomenon whereby the fraction of total phenotypic variance that is explained by all individually significant loci in human genome-wide association analyses for common diseases and quantitative traits is typically much less than the heritability that is estimated from relationships among relatives.

Di-hybrid cross

A cross between parental lines that are fixed for alternative alleles at two unlinked loci (for example, A1A1B2B2 × A2A2B1B1, where A and B denote the loci and the subscripts represent the alleles) in which nine genotypes segregate in the F2 generation.

Dominance effects

Differences between the genotypic values of the heterozygous genotypes and the average genotypic values of the homozygous genotypes at loci that affect quantitative traits.

Standing variation

Allelic variation that is currently segregating within a population, as opposed to alleles that appear as the result of new mutation events.

Introgression

The substitution of a genomic region from one strain with that of another, typically by repeated backcrosses.

Diallel cross

A class of experimental designs that are used to estimate both additive and non-additive variance components for a quantitative trait from all possible crosses among a population of inbred lines. Full diallel designs include reciprocal crosses, whereas half-diallel designs do not; parental lines can be included or excluded in either case.

Synthetic enhancement

A type of epistatic interaction whereby the phenotype of a double mutant is more severe than that predicted from the additive effects of the single mutants.

Multiple testing penalty

The downward adjustment of the significance threshold for individual statistical tests that is required when multiple hypothesis tests are carried out on a single data set; for n independent tests, the Bonferroni-adjusted 5% significance threshold is 0.05/n.

Minor allele frequency

The frequency of the less common allele at a bi-allelic locus.

Founder-effect speciation models

A class of models for the evolution of reproductive isolation that is based on changes in selection pressures and on allele frequencies of epistatically interacting loci, which result from the establishment of a new population in a new environment from a small number of individuals.

Dobzhansky–Muller incompatibilities

Substitutions that occur during divergence of two lineages; these substitutions are neutral in the respective genetic backgrounds of the two lineages but cause a reduction in fertility and/or viability in hybrids between the two lineages.

Genomic prediction methods

Models that are derived from a discovery sample which consists of individuals with measured phenotypes and genome-wide marker data; these models are used to predict individual phenotypes in an independent sample from the same population using only genome-wide marker data.

Rights and permissions

Reprints and permissions

About this article

Cite this article

Mackay, T. Epistasis and quantitative traits: using model organisms to study gene–gene interactions. Nat Rev Genet 15, 22–33 (2014). https://doi.org/10.1038/nrg3627

Download citation

  • Published:

  • Issue Date:

  • DOI: https://doi.org/10.1038/nrg3627

This article is cited by

Search

Quick links

Nature Briefing

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

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