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:

Molecular spandrels: tests of adaptation at the genetic level

A Corrigendum to this article was published on 29 November 2011

This article has been updated

Key Points

  • It is now possible to identify genes that underlie fitness-related traits and to detect molecular evidence that they have been affected by natural selection. However, neither of these approaches functionally connects genotype, phenotype and fitness, and thus they do not provide direct evidence that specific alleles are adaptive.

  • Demonstrating that an allele is adaptive requires evidence of the direct effects of alternative alleles on an ecologically relevant phenotype, and hence organismal performance, as well as the effects of these alternative alleles on fitness through known functional mechanisms.

  • Experimental tests of selection on the genes underlying phenotypic traits will help to avoid being led astray by intuitively appealing but potentially incomplete or incorrect adaptive stories ('molecular spandrels').

  • Pioneering studies have used a variety of approaches to isolate the fitness effects of specific alleles from the confounding influence of the surrounding genome, typically by using populations in which the alleles of interest are segregating within either a randomized genetic background or a homogenous genetic background.

  • The use of next-generation sequencing in field experiments represents a promising avenue for carrying out comprehensive tests of adaptation. Incorporating genomics into field experiments should help to: differentiate between the effects of pleiotropy and linkage; characterize the role of epistasis; and allow for fine-scale discrimination of the mechanisms and targets of selection.

Abstract

Although much progress has been made in identifying the genes (and, in rare cases, mutations) that contribute to phenotypic variation, less is known about the effects that these genes have on fitness. Nonetheless, genes are commonly labelled as 'adaptive' if an allele has been shown to affect a phenotype with known or suspected functional importance or if patterns of nucleotide variation at the locus are consistent with positive selection. In these cases, the 'adaptive' designation may be premature and may lead to incorrect conclusions about the relationships between gene function and fitness. Experiments to test targets and agents of natural selection within a genomic context are necessary for identifying the adaptive consequences of individual alleles.

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: Connections between various approaches for studying the genetics of adaptation.
Figure 2: Examples of using complementary ecological and molecular approaches to investigate the genetic basis of adaptation.
Figure 3: How understanding genetic architecture can inform us about selection.

Similar content being viewed by others

Change history

  • 29 November 2011

    In the above article, Table 1 and Supplementary information S1 (table) incorrectly stated that estimates of selection had not been calculated for the Ace1 gene, a variant of which confers insecticide resistance in mosquitoes. The tables also contained a spelling mistake: the species name Culex pipiens was originally given as Culex pipens in the 'Phenotypic effect' column. The revised tables now include four new papers (listed as references 150 to 153 in the reference list) that discuss the relevant selection studies, and the misspelling of Culex pipiens has been corrected. Additionally, Box 3 contained a typographical error: the time for the carbonaria morph of Biston betularia to reach 98% frequency was 47 years and not 7 years as stated in the Review. The authors apologize for these errors.

References

  1. Gould, S. & Lewontin, R. The spandrels of San Marco and the Panglossian paradigm: a critique of the adaptationist programme. Proc. R. Soc. Lond. B 205, 581–598 (1979). This is a much-cited paper that criticized evolutionary biologists for adaptationist 'story-telling' in the absence of rigorous tests of alternative possibilities.

    Article  CAS  PubMed  Google Scholar 

  2. Arnold, S. Morphology, performance and fitness. Am. Zool. 23, 347–361 (1983).

    Article  Google Scholar 

  3. Lande, R. Quantitative genetic analysis of multivariate evolution, applied to brain: body size allometry. Evolution 33, 402–416 (1979).

    Article  PubMed  Google Scholar 

  4. Lande, R. & Arnold, S. J. The measurement of selection on correlated characters. Evolution 37, 1210–1226 (1983).

    Article  PubMed  Google Scholar 

  5. Kingsolver, J. G. et al. The strength of phenotypic selection in natural populations. Am. Nat. 157, 245–261 (2001).

    Article  CAS  PubMed  Google Scholar 

  6. Morjan, C. L. & Rieseberg, L. H. How species evolve collectively: implications of gene flow and selection for the spread of advantageous alleles. Mol. Ecol. 13, 1341–1356 (2004). This is an innovative literature review that combines estimates of selection on phenotypic traits with the trait variance explained by underlying QTLs to obtain a rough measure of the selection acting on individual loci.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  7. Siepielski, A., Dibattista, J. & Carlson, S. It's about time: the temporal dynamics of phenotypic selection in the wild. Ecol. Lett. 12, 1261–1276 (2009).

    Article  PubMed  Google Scholar 

  8. Nielsen, R. Adaptationism — 30 years after Gould and Lewontin. Evolution 63, 2487–2490 (2009).

    Article  PubMed  Google Scholar 

  9. Thornton, K. R., Jensen, J. D., Becquet, C. & Andolfatto, P. Progress and prospects in mapping recent selection in the genome. Heredity 98, 340–348 (2007).

    Article  CAS  PubMed  Google Scholar 

  10. Reznick, D. N. & Ghalambor, C. K. Selection in nature: experimental manipulations of natural populations. Integr. Comp. Biol. 45, 456–462 (2005).

    Article  PubMed  Google Scholar 

  11. Tauber, E. et al. Natural selection favors a newly derived timeless allele in Drosophila melanogaster. Science 316, 1895–1898 (2007).

    Article  CAS  PubMed  Google Scholar 

  12. Sezgin, E. Single-locus latitudinal clines and their relationship to temperate adaptation in metabolic genes and derived alleles in Drosophila melanogaster. Genetics 168, 923–931 (2004).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  13. Umina, P., Weeks, A., Kearney, M., McKechnie, S. & Hoffmann, A. A. A rapid shift in a classic clinal pattern in Drosophila reflecting climate change. Science 308, 691–693 (2005).

    Article  CAS  PubMed  Google Scholar 

  14. Beaumont, M. A. Adaptation and speciation: what can Fst tell us? Trends Ecol. Evol. 20, 435–440 (2005).

    Article  PubMed  Google Scholar 

  15. Beaumont, M. A. & Balding, D. J. Identifying adaptive genetic divergence among populations from genome scans. Mol. Ecol. 13, 969–980 (2004).

    Article  CAS  PubMed  Google Scholar 

  16. Vasemagi, A. The adaptive hypothesis of clinal variation revisited: single-locus clines as a result of spatially restricted gene flow. Genetics 173, 2411–2414 (2006).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  17. Jensen, J. & Bachtrog, D. Characterizing recurrent positive selection at fast evolving genes in Drosophila miranda and Drosophila pseudoobscura. Genom. Biol. Evol. 2, 371–378 (2010).

    Article  CAS  Google Scholar 

  18. Andres, A. et al. Targets of balancing selection in the human genome. Mol. Biol. Evol. 26, 2755–2764 (2009).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  19. Nielsen, R. et al. Darwinian and demographic forces affecting human protein coding genes. Genome Res. 19, 838–849 (2009).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  20. Bazin, E., Dawson, K. J. & Beaumont, M. A. Likelihood-free inference of population structure and local adaptation in a Bayesian hierarchical model. Genetics 185, 587–602 (2011).

    Article  CAS  Google Scholar 

  21. Thornton, K. R. & Jensen, J. D. Controlling the false-positive rate in multilocus genome scans for selection. Genetics 175, 737–750 (2007).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  22. Przeworski, M., Coop, G. & Wall, J. D. The signature of positive selection on standing genetic variation. Evolution 59, 2312–2323 (2005).

    Article  PubMed  Google Scholar 

  23. Teshima, K. M., Coop, G. & Przeworski, M. How reliable are empirical genomic scans for selective sweeps? Genome Res. 16, 702–712 (2006).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  24. Baird, N. A. et al. Rapid SNP discovery and genetic mapping using sequenced RAD markers. PLoS ONE 3, e3376 (2008).

    Article  PubMed  PubMed Central  CAS  Google Scholar 

  25. Andolfatto, P. et al. Multiplexed shotgun genotyping for rapid and efficient genetic mapping. Genome Res. 21, 610–617 (2011).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  26. Xie, W. et al. Parent-independent genotyping for constructing an ultra high-density linkage map based on population sequencing. Proc. Nat. Acad. Sci. USA 107, 10578–10583 (2010).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  27. 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 

  28. Hoekstra, H. E., Hirschmann, R. J., Bundey, R. A., Insel, P. A. & Crossland, J. P. A single amino acid mutation contributes to adaptive beach mouse color pattern. Science 313, 101–104 (2006).

    Article  CAS  PubMed  Google Scholar 

  29. Colosimo, P. F. et al. Widespread parallel evolution in sticklebacks by repeated fixation of ectodysplasin alleles. Science 307, 1928–1933 (2005).

    Article  CAS  PubMed  Google Scholar 

  30. Gratten, J. et al. A localized negative genetic correlation constrains microevolution of coat color in wild sheep. Science 319, 318–320 (2008). This study demonstrates the difficulty of predicting the microevolutionary consequences of selection in the absence of knowledge about the genetic basis of fitness variation.

    Article  CAS  PubMed  Google Scholar 

  31. Steiner, C. C., Weber, J. N. & Hoekstra, H. E. Adaptive variation in beach mice produced by two interacting pigmentation genes. PLoS Biol. 5, e219 (2007).

    Article  PubMed  PubMed Central  CAS  Google Scholar 

  32. Rogers, S. M. & Bernatchez, L. The genetic architecture of ecological speciation and the association with signatures of selection in natural lake whitefish (Coregonus sp. Salmonidae) species pairs. Mol. Biol. Evol. 24, 1423–1438 (2007).

    Article  CAS  PubMed  Google Scholar 

  33. Nosil, P., Egan, S. P. & Funk, D. J. Heterogeneous genomic differentiation between walking-stick ecotypes: “isolation by adaptation” and multiple roles for divergent selection. Evolution 62, 316–336 (2008).

    Article  PubMed  Google Scholar 

  34. Via, S. J. W. The genetic mosaic suggests a new role for hitchhiking in ecological speciation. Mol. Ecol. 17, 4334–4345 (2008).

    Article  PubMed  Google Scholar 

  35. Todesco, M. et al. Natural allelic variation underlying a major fitness trade-off in Arabidopsis thaliana. Nature 465, 632–636 (2010).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  36. Linnen, C. R., Kingsley, E. P., Jensen, J. D. & Hoekstra, H. E. On the origin and spread of an adaptive allele in deer mice. Science 325, 1095–1098 (2009).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  37. Rokyta, D., Joyce, P., Caudle, S. & Wichman, H. An empirical test of the mutational landscape model of adaptation using a single-stranded DNA virus. Nature Genet. 37, 441–444 (2005).

    Article  CAS  PubMed  Google Scholar 

  38. Araya, C. L., Payen, C., Dunham, M. J. & Fields, S. Whole-genome sequencing of a laboratory-evolved yeast strain. BMC Genom. 11, 88 (2010).

    Article  CAS  Google Scholar 

  39. Betancourt, A. J. Genome-wide patterns of substitution in adaptively evolving populations of the RNA bacteriophage MS2. Genetics 181, 1535–1544 (2009).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  40. Maclean, R. C., Hall, A. R., Perron, G. G. & Buckling, A. The population genetics of antibiotic resistance: integrating molecular mechanisms and treatment contexts. Nature Rev. Genet. 11, 405–414 (2010).

    Article  CAS  PubMed  Google Scholar 

  41. Sandgren, A. et al. Tuberculosis drug resistance mutation database. PLoS Med. 6, e1000002 (2009).

    Article  PubMed Central  CAS  Google Scholar 

  42. Stanek, M., Cooper, T. & Lenski, R. E. Identification and dynamics of a beneficial mutation in a long-term evolution experiment with Escherichia coli. BMC Evol. Biol. 9, 302 (2009).

    Article  PubMed  PubMed Central  CAS  Google Scholar 

  43. Barrick, J. E. et al. Genome evolution and adaptation in a long-term experiment with Escherichia coli. Nature 461, 1243–1247 (2009). This was one of the first studies to investigate the genome-wide consequences of adaptation in experimentally evolving populations.

    Article  CAS  PubMed  Google Scholar 

  44. Herring, C. et al. Comparative genome sequencing of Escherichia coli allows observation of bacterial evolution on a laboratory timescale. Nature Genet. 38, 1406–1412 (2006).

    Article  CAS  PubMed  Google Scholar 

  45. Paterson, S. et al. Antagonistic coevolution accelerates molecular evolution. Nature 464, 275–278 (2010).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  46. Burke, M. K. et al. Genome-wide analysis of a long-term evolution experiment with Drosophila. Nature 467, 587–590 (2010).

    Article  CAS  PubMed  Google Scholar 

  47. Falconer, D. S. & Mackay, T. F. C. Introduction to Quantitative Genetics (Addison Wesley Longman, Harlow, UK, 1996).

    Google Scholar 

  48. Thompson, J. N. The Geographic Mosaic of Coevolution. (Univ. Chicago Press, Chicago, 2005).

    Book  Google Scholar 

  49. Spencer, C. C. & Promislow, D. E. L. Genes, culture, and aging flies—what the lab can and cannot tell us about natural genetic variation for senescence. Sci. Aging Knowl. Environ. 2002, pe6 (2002).

    Google Scholar 

  50. Endler, J. A. Natural Selection in the Wild (Princeton Univ. Press, Princeton, 1986).

    Google Scholar 

  51. Boag, T. & Grant, P. R. Intense natural selection in a population of Darwin's finches (Geospizinae) in the Galapagos. Science 214, 82–85 (1981). This important study addresses the challenge raised by Gould & Lewontin's 'spandrel' paper (reference 1) by showing that natural selection can be measured in action in the wild. This work served as a catalyst for other field researchers to estimate selection coefficients in natural populations.

    Article  CAS  PubMed  Google Scholar 

  52. Price, T., Grant, P. R. & Gibbs, H. L. Recurrent patterns of natural selection in a population of Darwin's finches. Nature 309, 787–789 (1984).

    Article  CAS  PubMed  Google Scholar 

  53. Reznick, D. N., Shaw, F. H., Rodd, F. H. & Shaw, R. G. Evaluation of the rate of evolution in natural populations of guppies (Poecilia reticulata). Science 275, 1934–1937 (1997).

    Article  CAS  PubMed  Google Scholar 

  54. Reznick, D. N. & Bryga, H. Life-history evolution in guppies (Poecilia reticulata).1. Phenotypic and genetic changes in an introduction experiment. Evolution 41, 1370–1385 (1987).

    PubMed  Google Scholar 

  55. Losos, J. B., Warheit, K. I. & Schoener, T. W. Adaptive differentation following experimental island colonization in Anolis lizards. Nature 387, 70–73 (1997).

    Article  CAS  Google Scholar 

  56. Schoener, T. W., Spiller, D. A. & Losos, J. B. Predators increase the risk of catastrophic exinction of prey populations. Nature 412, 183–186 (2001).

    Article  CAS  PubMed  Google Scholar 

  57. Nosil, P., Crespi, B. & Sandoval, C. Host-plant adaptation drives the parallel evolution of reproductive isolation. Nature 417, 440–443 (2002).

    Article  CAS  PubMed  Google Scholar 

  58. Nosil, P. & Crespi, B. Experimental evidence that predation promotes divergence in adaptive radiation. Proc. Nat. Acad. Sci. USA 103, 9090–9095 (2006).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  59. Schluter, D. Experimental evidence that competition promotes divergence in adaptive radiation. Science 266, 798–801 (1994).

    Article  CAS  PubMed  Google Scholar 

  60. Rundle, H. D., Vamosi, S. M. & Schluter, D. Experimental test of predation's effect on divergent selection during character displacement in sticklebacks. Proc. Nat. Acad. Sci. USA 100, 14943–14948 (2003).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  61. Mackay, T. F. C. The genetic architecture of quantitative traits. Ann. Rev. Genet. 35, 303–339 (2001).

    Article  CAS  PubMed  Google Scholar 

  62. Hall, M., Lowry, D. & Willis, J. Is local adaptation in Mimulus guttatus caused by trade-offs at individual loci? Mol. Ecol. 19, 2739–2753 (2010). This paper discusses a detailed ecological transplant experiment that demonstrated that local adaptation is largely controlled by non-overlapping loci.

    Article  CAS  PubMed  Google Scholar 

  63. Lowry, D., Hall, M., Salt, D. & Willis, J. Genetic and physiological basis of adaptive salt tolerance divergence between coastal and inland Mimulus guttatus. New Phyt. 183, 776–788 (2009).

    Article  Google Scholar 

  64. Anderson, J., Lee, C. & Mitchell-Olds, T. Life history QTLs and natural selection on flowering time in Boechera stricta, a perennial relative of Arabidopsis. Evolution 65, 771–787 (2010).

    Article  PubMed  PubMed Central  Google Scholar 

  65. Gardner, K. M. & Latta, R. G. Identifying loci under selection across contrasting environments in Avena barbata using quantitative trait locus mapping. Mol. Ecol. 15, 1321–1333 (2006).

    Article  CAS  PubMed  Google Scholar 

  66. Verhoeven, K. J. F., Poorter, H., Nevo, E. & Biere, A. Habitat-specific natural selection at a flowering-time QTL is a main driver of local adaptation in two wild barley populations. Mol. Ecol. 17, 3416–3424 (2008).

    CAS  PubMed  Google Scholar 

  67. Ungerer, M., Linder, C. & Rieseberg, L. Effects of genetic background on response to selection in experimental populations of Arabidopsis thaliana. Genetics 163, 277–286 (2003). This was one of the first experimental studies to measure natural selection on phenotypic traits that are connected to specific QTLs.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  68. Ungerer, M. & Rieseberg, L. Genetic architecture of a selection response in Arabidopsis thaliana. Evolution 57, 2531–2539 (2003).

    Article  CAS  PubMed  Google Scholar 

  69. Beraldi, D. et al. Mapping quantitative trait loci underlying fitness-related traits in a free-living sheep population. Evolution 61, 1403–1416 (2007).

    Article  PubMed  Google Scholar 

  70. Beavis, W. D. in Molecular Dissection of Complex Traits (ed. Paterson, A. H.) 145–162 (CRC Press, New York, 1998).

    Google Scholar 

  71. Otto, S. P. & Jones, C. D. Detecting the undetected: estimating the total number of loci underlying a quantitative trait. Genetics 156, 2093–2107 (2000).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  72. Weinig, C. et al. Heterogeneous selection at specific loci in natural environments in Arabidopsis thaliana. Genetics 165, 321–329 (2003).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  73. Lexer, C., Welch, M. E., Durphy, J. L. & Rieseberg, L. H. Natural selection for salt tolerance quantitative trait loci (QTLs) in wild sunflower hybrids: implications for the origin of Helianthus paradoxus, a diploid hybrid species. Mol. Ecol. 12, 1225–1235 (2003).

    Article  CAS  PubMed  Google Scholar 

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

    Article  CAS  PubMed  Google Scholar 

  75. Korves, T. et al. Fitness effects associated with the major flowering time gene FRIGIDA in Arabidopsis thaliana in the field. Am. Nat. 169, 141–157 (2007).

    Article  Google Scholar 

  76. Dalziel, A., Rogers, S. M. & Schulte, P. Linking genotypes to phenotypes and fitness: how mechanistic biology can inform molecular ecology. Mol. Ecol. 18, 4997–5017 (2009).

    Article  CAS  PubMed  Google Scholar 

  77. Barrett, R., Rogers, S. M. & Schluter, D. Natural selection on a major armor gene in threespine stickleback. Science 322, 255–257 (2008).

    Article  CAS  PubMed  Google Scholar 

  78. Hudson, M. Sequencing breakthroughs for genomic ecology and evolutionary biology. Mol. Ecol. Res. 8, 3–17 (2008).

    Article  CAS  Google Scholar 

  79. Metzker, M. L. Sequencing technologies: the next generation. Nature Rev. Genet. 11, 31–46 (2009).

    Article  CAS  PubMed  Google Scholar 

  80. Stapley, J. et al. Adaptation genomics: the next generation. Trends Ecol. Evol. 25, 705–712 (2010).

    Article  PubMed  Google Scholar 

  81. Storz, J. & Wheat, C. Integrating evolutionary and functional approaches to infer adaptation at specific loci. Evolution 64, 2489–2509 (2010). This excellent perspective piece called for a synthesis of insights from indirect population genetic analyses and direct experiments on molecular function in making inferences about adaptation at the genetic level.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  82. Ungerer, M., Johnson, L. & Herman, M. Ecological genomics: understanding gene and genome function in the natural environment. Heredity 100, 178–183 (2008).

    Article  CAS  PubMed  Google Scholar 

  83. Excoffier, L., Hofer, T. & Foll, M. Detecting loci under selection in a hierarchically structured population. Heredity 103, 285–298 (2009).

    Article  CAS  PubMed  Google Scholar 

  84. Ferguson, L. et al. Characterization of a hotspot for mimicry: assembly of a butterfly wing transcriptome to genomic sequence at the HmYb/Sb locus. Mol. Ecol. 19, 240–254 (2010).

    Article  PubMed  Google Scholar 

  85. Renaut, S., Nolte, A. & Bernatchez, L. Mining transcriptome sequences towards identifying adaptive single nucleotide polymorphisms in lake whitefish species pairs (Coregonus sp. Salmonidae). Mol. Ecol. 19, 115–131 (2010).

    Article  CAS  PubMed  Google Scholar 

  86. Korbel, J. et al. The current excitement about copy-number variation: how it relates to gene duplications and protein families. Curr. Op. Struct. Biol. 18, 366–374 (2008).

    Article  CAS  Google Scholar 

  87. Christians, J. K. & Senger, L. K. Fine mapping dissects pleiotropic growth quantitative trait locus into linked loci. Mamm. Genom. 18, 240–245 (2007).

    Article  CAS  Google Scholar 

  88. Lawniczak, M. K. N., Emrich, S. J. & Holloway, A. K. Widespread divergence between incipient Anopheles gambiae species revealed by whole genome sequences. Science 330, 512–514 (2010).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  89. Houle, D., Govindaraju, D. R. & Omholt, S. Phenomics: the next challenge. Nature Rev. Genet. 11, 855–866 (2010).

    Article  CAS  PubMed  Google Scholar 

  90. Nosil, P., Funk, D. & Ortiz-Barrientos, D. Divergent selection and heterogeneous genomic divergence. Mol. Ecol. 18, 375–402 (2009).

    Article  PubMed  Google Scholar 

  91. Michel, A. et al. Widespread genomic divergence during sympatric speciation. Proc. Nat. Acad. Sci. USA 107, 9724–9729 (2010). This novel study used a combination of laboratory-based selection experiments, genetic mapping, LD analyses and field surveys to show genomic patterns of divergent natural selection.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  92. 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).

    Article  CAS  PubMed  Google Scholar 

  93. Cadieu, E. et al. Coat variation in the domestic dog is governed by variants in three genes. Science 326, 150–153 (2009).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  94. Dean, A. M. & Thornton, J. W. Mechanistic approaches to the study of evolution: the functional synthesis. Nature Rev. Genet. 8, 675–688 (2007).

    Article  CAS  PubMed  Google Scholar 

  95. Kimura, M. Evolutionary rate at the molecular level. Nature 217, 624–636 (1968).

    Article  CAS  PubMed  Google Scholar 

  96. Kimura, M. The Neutral Theory of Molecular Evolution (Cambridge Univ. Press, Cambridge, UK, 1983).

    Book  Google Scholar 

  97. Lynch, M. The Origins of Genomic Architecture (Sunderland, Massachusetts, 2007).

    Google Scholar 

  98. Ohta, T. Slightly deleterious mutant substitutions in evolution. Nature 246, 96–98 (1973).

    Article  CAS  PubMed  Google Scholar 

  99. Lewontin, R. C. & Hubby, J. L. A molecular approach to the study of genic heterozygosity in natural populations. II. Amount of variation and degree of heterozygosity in natural populations of Drosophila pseudoobscura. Genetics 54, 595–609 (1966).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  100. Nei, M., Suzuki, Y. & Nozawa, M. The neutral theory of molecular evolution in the genomic era. Ann. Rev. Genom. Hum. Genet. 11, 265–289 (2010). This recent review provides a comprehensive critique of the methods used to detect selection in genome scans.

    Article  CAS  Google Scholar 

  101. Hermisson, J. Who believes in whole-genome scans for selection? Heredity 103, 283–284 (2009).

    Article  CAS  PubMed  Google Scholar 

  102. Sella, G., Petrov, D., Przeworski, M. & Andolfatto, P. Pervasive natural selection in the Drosophila genome? PLoS Genet. 5, e1000495 (2009).

    Article  PubMed  PubMed Central  CAS  Google Scholar 

  103. Halligan, D. & Keightley, P. Ubiquitous selective constraints in the Drosophila genome revealed by a genome-wide interspecies comparison. Genom. Res. 16, 875–884 (2006).

    Article  CAS  Google Scholar 

  104. Shapiro, J. A. et al. Adaptive genic evolution in the Drosophila genomes. Proc. Nat. Acad. Sci. USA 104, 2271–2276 (2007).

    Article  PubMed  PubMed Central  Google Scholar 

  105. Sawyer, S. A., Parsch, J., Zhang, Z. & Hartl, D. L. Prevalence of positive selection among nearly neutral amino acid replacements in Drosophila. Proc. Nat. Acad. Sci. USA 104, 6504–6510 (2007).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  106. Macpherson, J., Sella, G., Davis, J. & Petrov, D. Genome-wide spatial correspondence between nonsynonymous divergence and neutral polymorphism reveals extensive adaptation in Drosophila. Genetics 177, 2083–2099 (2007).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  107. Asthana, S. et al. Widely distributed noncoding purifying selection in the human genome. Proc. Nat. Acad. Sci. USA 104, 12410–12415 (2007).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  108. Akey, J. M. Constructing genomic maps of positive selection in humans: where do we go from here? Genom. Res. 19, 711–722 (2009).

    Article  CAS  Google Scholar 

  109. Wright, S. & Andolfatto, P. The impact of natural selection on the genome: emerging patterns in Drosophila and Arabidopsis. Ann. Rev. Ecol. Syst. 39, 193–213 (2008).

    Article  Google Scholar 

  110. Foll, M. & Gaggiotti, O. E. A genome scan method to identify selected loci appropriate for both dominant and codominant markers: a Bayesian perspective. Genetics 180, 977–993 (2008).

    Article  PubMed  PubMed Central  Google Scholar 

  111. Andolfatto, P. Hitchhiking effects of recurrent beneficial amino acid substitutions in the Drosophila melanogaster genome. Genom. Res. 17, 1755–1762 (2007).

    Article  CAS  Google Scholar 

  112. Sattath, S., Elyashiv, E., Kolodny, O., Rinott, Y. & Sella, G. Pervasive adaptive protein evolution apparent in diversity patterns around amino acid substitutions in Drosophila simulans. PLoS Genet. 7, e1001302 (2011).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  113. Hughes, A. L. Looking for Darwin in all the wrong places: the misguided quest for positive selection at the nucleotide sequence level. Heredity 99, 364–373 (2007).

    Article  CAS  PubMed  Google Scholar 

  114. Otto, S. P. Two steps forward, one step back: the pleiotropic effects of favoured alleles. Proc. R. Soc. Lond. B. 271, 705–714 (2004).

    Article  CAS  Google Scholar 

  115. McVean, G. A. T. & Charlesworth, B. The effects of Hill–Robertson interference between weakly selected mutations on patterns of molecular evolution and variation. Genetics 155, 929–944 (2000).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  116. Coberly, L. & Rausher, M. Pleiotropic effects of an allele producing white flowers in Ipomoea purpurea. Evolution 62, 1076–1085 (2008).

    Article  PubMed  Google Scholar 

  117. Baucom, R. S., Chang, S. M., Kniskern, J. M., Rausher, M. D. & Stinchcombe, J. R. Morning glory as a powerful model in ecological genomics: tracing adaptation through both natural and artificial selection. Heredity 30 Mar 2011 (doi:10.1038/hdy.2011.25).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  118. Cook, L. The rise and fall of the carbonaria form of the peppered moth. Q. Rev. Biol. 78, 399–417 (2003).

    Article  PubMed  Google Scholar 

  119. Haldane, J. B. S. A mathmatical theory of natural and artificial selection, part 1. Trans. Camb. Phil. Soc. 23, 19–41 (1924).

    Google Scholar 

  120. Kettlewell, H. Further selection experiments on Lepidoptera. Heredity 10, 287–301 (1956).

    Article  Google Scholar 

  121. Grant, B. S., Cook, A. D., Clarke, C. A. & Owen, D. F. Geographic and temporal variation in the incidence of melanism in peppered moth populations in America and Britain. J. Hered. 89, 465–471 (1998).

    Article  Google Scholar 

  122. Majerus, M. E. N. Melanism: Evolution in Action (Oxford Univ. Press, Oxford, 1998).

    Google Scholar 

  123. Coyne, J. A. Not black and white. Nature 396, 35–36 (1998).

    Article  CAS  Google Scholar 

  124. van't Hof, A., Edmonds, N., Dalíková, M., Marec, F. & Saccheri, I. Industrial melanism in british peppered moths has a singular and recent mutational origin. Science 332, 958–960 (2011).

    Article  CAS  PubMed  Google Scholar 

  125. Hohenlohe, P. A. et al. Population genomics of parallel adaptation in threespine stickleback using sequenced RAD tags. PLoS Genet. 6, e1000862 (2010).

    Article  PubMed  PubMed Central  CAS  Google Scholar 

  126. Kingsley, E. P., Manceau, M., Wiley, C. D. & Hoekstra, H. E. Melanism in Peromyscus is caused by independent mutations in Agouti. PLoS ONE 4, e6435 (2009).

    Article  PubMed  PubMed Central  CAS  Google Scholar 

  127. Dice, L. R. Effectiveness of selection by owls of deer-mice (Peromyscus maniculatus) which contrast in color with their background. Cont. Lab. Vert. Biol. Univ. Mich. 34, 1–20 (1947).

    Google Scholar 

  128. Labbe, P. et al. Independent duplications of the Acetylcholinesterase gene conferring insecticide resistance in the mosquito Culex pipiens. Mol. Biol. Evol. 24, 1056–1067 (2007).

    Article  CAS  PubMed  Google Scholar 

  129. Habu, Y., Hisatomi, Y. & Iida, S. Molecular characterization of the mutableflaked allele for flower variegation in the common morning glory. Plant J. 16, 371–376 (1998).

    Article  CAS  PubMed  Google Scholar 

  130. Rebeiz, M., Pool, J., Kassner, V., Aquadro, C. & Carroll, S. Stepwise modification of a modular enhancer underlies adaptation in a Drosophila population. Science 326, 1663–1667 (2009).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  131. Pool, J. E. & Aquadro, C. F. The genetic basis of adaptive pigmentation variation in Drosophila melanogaster. Mol. Ecol. 16, 2844–2851 (2007).

    Article  PubMed  PubMed Central  Google Scholar 

  132. Marchinko, K. Predation's role in repeated phenotypic and genetic divergence of armor in threespine stickleback. Evolution 63, 127–138 (2009).

    Article  PubMed  Google Scholar 

  133. Reimchen, T. E. & Nosil, P. Lateral plate asymmetry, diet and parasitism in threespine stickleback. J. Evol. Biol. 14, 632–645 (2001).

    Article  Google Scholar 

  134. Le Corre, V., Roux., F. & Reboud, X. DNA polymorphism at the FRIGIDA gene in Arabidopsis thaliana: extensive nonsynonymous variation is consistent with local selection for flowering time. Mol. Biol. Evol. 19, 1261–1271 (2002).

    Article  CAS  PubMed  Google Scholar 

  135. Storz, J. & Kelly, J. Effects of spatially varying selection on nucleotide diversity and linkage disequilibrium: insights from deer mouse globin genes. Genetics 180, 367–379 (2008).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  136. Storz, J. F., Runck, A. M., Moriyama, H., Weber, R. E. & Fago, A. Genetic differences in hemoglobin function between highland and lowland deer mice. J. Exp. Biol. 213, 2565–2574 (2010).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  137. Hayes, J. P. & O'Connor, C. S. Natural selection on thermogenic capacity of high-altitude deer mice. Evolution 53, 1280–1287 (1999).

    PubMed  Google Scholar 

  138. Tishkoff, S. A. et al. Convergent adaptation of human lactase persistence in Africa and Europe. Nature Genet. 39, 31–40 (2007).

    Article  CAS  PubMed  Google Scholar 

  139. Mullen, L. M. & Hoekstra, H. E. Natural selection along an environmental gradient: A classic cline in mouse pigmentation. Evolution 62, 1555–1569 (2008).

    Article  CAS  PubMed  Google Scholar 

  140. Vignieri, S. N., Larson, J. N. & Hoekstra, H. E. The selective advantage of cryptic coloration in mice. Evolution 64, 2153–2158 (2010).

    PubMed  Google Scholar 

  141. Evans, P. et al. Microcephalin, a gene regulating brain size, continues to evolve adaptively in humans. Science 309, 1717–1720 (2005).

    Article  CAS  PubMed  Google Scholar 

  142. Jackson, A. P. et al. Identification of microcephalin, a protein implicated in determining the size of the human brain. Am. J. Hum. Genet. 71, 136–142 (2002).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  143. Geffeney, S., Brodie, E. D Jr. Ruben, P. C. & Brodie, E. D. Mechanisms of adaptation in a predator-prey arms race: TTX-resistant sodium channels. Science 297, 1336–1339 (2002).

    Article  CAS  PubMed  Google Scholar 

  144. Brodie, E. D. & Brodie, E. D Jr. Evolutionary response of predators to dangerous prey: reduction of toxicity of newts and resistance of garter snakes in island populations. Evolution 45, 221–224 (1991).

    PubMed  Google Scholar 

  145. Wheat, C. W., Watt, W. B., Pollock, D. D. & Schulte, P. M. From DNA to fitness differences: sequences and structures of adaptive variants of colias phosphoglucose isomerase (PGI). Mol. Biol. Evol. 23, 499–512 (2006).

    Article  CAS  PubMed  Google Scholar 

  146. Watt, W. B. in Butterflies: Ecology and Evolution Taking Flight (eds. Boggs, C. L., Watt, W. B. & Ehrlich, P. R.) 319–352 (Univ. Chicago Press, Chicago, 2003).

    Google Scholar 

  147. Bradshaw, H. D. & Schemske, D. W. Allele substitution at a flower colour locus produces a pollinator shift in monkeyflowers. Nature 426, 176–178 (2003).

    Article  CAS  PubMed  Google Scholar 

  148. Jackson, M., Stinchcombe, J., Korves, T. & Schmitt, J. Costs and benefits of cold tolerance in transgenic Arabidopsis thaliana. Mol . Ecol. 13, 3609–3615 (2004).

    Article  CAS  PubMed  Google Scholar 

  149. Bell, G. Experimental genomics of fitness in yeast. Proc. R. Soc. Lond. B. 277, 1459–1467 (2010).

    Google Scholar 

  150. Lenormand, T., Guillemaud, T., Bourguet, D. & Raymond, M. Evaluating gene flow using selected markers: a case study. Genetics. 149, 1383–1392 (2008).

    Article  Google Scholar 

  151. Lenormand, T., Bourguet, D., Guillemaud, T. & Raymond, M. Tracking the evolution of insecticide resistance in the mosquito Culex pipiens. Nature. 400, 861–864 (1999).

    Article  CAS  PubMed  Google Scholar 

  152. Labbé, P. et al. Forty years of erratic insecticide resistance evolution in the mosquito Culex pipiens. PLoS Genet. 3, 2190–2199 (2007).

    Article  CAS  Google Scholar 

  153. Duron, O. et al. High Wolbachia density correlates with cost of infection for insecticide resistant Culex pipiens mosquitoes. Evolution. 60, 303–314 (2006).

    Article  CAS  PubMed  Google Scholar 

Download references

Acknowledgements

We thank P. Andolfatto, A. Berry, J. Jensen, J. Losos, D. Lowry, R. Nielsen, P. Nosil, S. Otto, D. Schluter, J. Slate and M. Szulkin for useful discussion and comments on the manuscript. We thank W. Cresko and P. Hohenlohe for providing a modified figure and L. Cook, J. Mallet and I. Saccheri for providing unpublished data on Biston betularia. We are supported by National Sciences and Research Council of Canada (NSERC) Howard Alper and John Templeton Foundational Questions in Evolutionary Biology (FQEB) fellowships (R.D.H.B.) and a US National Science Foundation (NSF) grant DEB-0919190 (H.E.H.).

Author information

Authors and Affiliations

Authors

Corresponding authors

Correspondence to Rowan D. H. Barrett or Hopi E. Hoekstra.

Ethics declarations

Competing interests

The authors declare no competing financial interests.

Supplementary information

Supplementary information 1 (table)

Examples of genes commonly described as adaptive. (PDF 206 kb)

Related links

Related links

FURTHER INFORMATION

Rowan D. H. Barrett's homepage

Hopi E. Hoekstra's homepage

Nature Reviews Genetics series on Study Designs

Glossary

Spandrel

An architectural feature that is necessary in the construction of domed cathedrals, but because of its aesthetic and purposeful design it can easily be confused as the featured element rather than a by-product of an engineering constraint.

Selection coefficient

The strength of selection as measured by the difference in fitness among genotypes.

Directional selection

Natural selection that favours one end of a distribution of a quantitative trait.

Stabilizing selection

Natural selection that favours intermediate values of a quantitative trait.

Disruptive selection

Natural selection that favours extreme values of a quantitative trait over intermediate values.

Neutral theory

This theory holds that standing genetic variation is predominantly neutral, whereas most new mutations are deleterious.

Purifying selection

Natural selection that favours the current condition by removing deleterious alleles that arise in the population.

Nearly neutral theory

An extension of the neutral theory that suggests that polymorphism at functionally important sites is predominantly nearly, rather than completely, neutral.

Effective population sizes

The size of an ideal population of breeding individuals that would experience the same amount of genetic drift as the observed population.

Selective sweeps

The increase in frequency of an allele (and closely linked chromosomal segments) that is caused by positive selection for the allele. Sweeps initially reduce variation and subsequently lead to a local excess of rare alleles (and an excess of homozygosity) as new unique mutations accumulate.

Linkage disequilibrium

(LD). Nonrandom association of alleles at two or more loci. The pattern and extent of linkage disequilibrium in a genomic region is affected by mutation, recombination, genetic drift, natural selection and demographic history.

Fixation index

(FST). A measure of population subdivision, indicating the proportion of genetic diversity found between populations relative to the amount of genetic diversity found within populations.

Ascertainment bias

A consequence of collecting a nonrandom subsample with a systematic bias, so that results based on the subsample are not representative of the entire sample.

Pleiotropic

The effect of a gene on more than one phenotypic trait.

Eumelanic

Consisting of brown to black pigments (that is, eumelanin) found in vertebrate hair, feathers or scales.

Recombinant inbred lines

(RILs). A population of fully homozygous individuals obtained through the repeated self-fertilization of an F1 generation hybrid.

Backcross lines

A population of individuals obtained by crossing a hybrid with one of its parents or to an individual that is genetically similar to its parent in order to generate offspring with genetic identity closer to that of the parent.

Likelihood ratio

The ratio of how many times more likely the observed data are under one model than the other, computed for all pairwise models.

Identical-by-descent lines

A population of individuals sharing identical copies of the same ancestral allele.

Near-isogenic lines

A population of individuals that differs from its parent in only one genomic location, typically a QTL of interest. Using markers that are diagnostic for that QTL, backcrosses are made to the parent until the entire genome of the line is exactly like the parent except in the region around the marker locus.

Rights and permissions

Reprints and permissions

About this article

Cite this article

Barrett, R., Hoekstra, H. Molecular spandrels: tests of adaptation at the genetic level. Nat Rev Genet 12, 767–780 (2011). https://doi.org/10.1038/nrg3015

Download citation

  • Published:

  • Issue Date:

  • DOI: https://doi.org/10.1038/nrg3015

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