Comprehensive Physiology Wiley Online Library

Phenotypic Plasticity: Molecular Mechanisms and Adaptive Significance

Full Article on Wiley Online Library



Abstract

Phenotypic plasticity can be broadly defined as the ability of one genotype to produce more than one phenotype when exposed to different environments, as the modification of developmental events by the environment, or as the ability of an individual organism to alter its phenotype in response to changes in environmental conditions. Not surprisingly, the study of phenotypic plasticity is innately interdisciplinary and encompasses aspects of behavior, development, ecology, evolution, genetics, genomics, and multiple physiological systems at various levels of biological organization. From an ecological and evolutionary perspective, phenotypic plasticity may be a powerful means of adaptation and dramatic examples of phenotypic plasticity include predator avoidance, insect wing polymorphisms, the timing of metamorphosis in amphibians, osmoregulation in fishes, and alternative reproductive tactics in male vertebrates. From a human health perspective, documented examples of plasticity most commonly include the results of exercise, training, and/or dieting on human morphology and physiology. Regardless of the discipline, phenotypic plasticity has increasingly become the target of a plethora of investigations with the methodological approaches utilized ranging from the molecular to whole organsimal. In this article, we provide a brief historical outlook on phenotypic plasticity; examine its potential adaptive significance; emphasize recent molecular approaches that provide novel insight into underlying mechanisms, and highlight examples in fishes and insects. Finally, we highlight examples of phenotypic plasticity from a human health perspective and underscore the use of mouse models as a powerful tool in understanding the genetic architecture of phenotypic plasticity. © 2012 American Physiological Society. Compr Physiol 2:1417‐1439, 2012.

Comprehensive Physiology offers downloadable PowerPoint presentations of figures for non-profit, educational use, provided the content is not modified and full credit is given to the author and publication.

Download a PowerPoint presentation of all images


Figure 1. Figure 1.

Two genetically identical water fleas, Daphnia lumholtzi. The helmet and extended tail spine of the individual on the left were induced as a result of chemical cues from a predaceous fish and serve as protection 67. This figure is recreated, with permission, from 3 Agrawal, A. “Phenotypic plasticity in the interactions and evolution of species”, Science, October 12, 2001, 294:321‐326, Figure 1, with permission of D. Laforsch.

Figure 2. Figure 2.

Possible relationships among plasticity and genetic variation. In each panel, dots connected by lines represent the phenotypes expressed by each of three genotypes (or families), numbered 1, 2, and 3, in each of two alternative environments (A, B). These lines are the reaction norms. (A) The three genotypes differ in their phenotypes within each environment, indicating genetic variation (VG = yes). However, a given genotype expresses the same phenotype, regardless of environment; that is, the reaction norms are flat. Hence, there is no environmental effect on the phenotype (VE, or plasticity, is absent). Because the reactions norms are parallel, there is no genotype‐by‐environment interaction (GxE = no). (B) As in panel “A,” the three genotypes differ in phenotype within a given environment, but in addition, each genotype expresses a different phenotype in environment “B,” relative to that expressed in environment “A.” That is, the reaction norms are not flat; each genotype is plastic for the trait of interest. However, because the slopes of all reaction norms are parallel, there is no genotype‐by‐environment interaction. The genotypes, although plastic, are all similarly plastic (for the trait of interest). (C) Genotypes differ within environments, show plasticity, and differ in plasticity. That is, the reaction norms are not parallel, indicating genotype‐by‐environment interactions. In this case, reaction norms 2 and 3 cross, but this may not always be the case. This figure is conceptually similar to, and derived, with permission, from Figure 1.4 from Pigliucci 121; p. 15) and Figure 1.1 from reference 38; p. 4)

Figure 3. Figure 3.

The panels above (A‐D) are representations of the general requirements needed for recognizing phenotypic plasticity as an adaptation. Although the representation of each of these criteria is pictorial simplistic, we acknowledge that their conclusive demonstration is quite complex. Accordingly, under each of the four general criteria we have listed additional considerations that should be taken into account. For extensive discussion of these criteria see text (also see reference 42. (A) Environmental heterogeneity must exist, and the degree of heterogeneity may determine the evolutionary (see panel D) consequences as opposed to an alteration in population mean. Heterogeneity may be biotic (e.g., predator presence) or abiotic (e.g., temperature), and special care should be taken to consider the extent of the spatial heterogeneity, the speed of fluctuations, and how these relate to the behavior and life history of the investigated organism. (B) Organisms must be able to reliably predict heterogeneity using environmental signals and these signals must be highly correlated with future environmental conditions. These signals must be spatial and temporally reliable and the organism must have the ability to sense and respond (even if the response is imperfect, see reference 87. (C) There must be an underlying genetic architecture regulating the plastic response. Here, we have presented methodologies for the evaluation of genetic variation. Using quantitative trait locus (QTL) mapping researchers may identify plasticity regions that directly affect the reaction norm (see text and Fig. 3 for examples), or evaluate differential gene expression in different environments with microarray technology. (D) There is a measurable response to selection and the response confers a fitness benefit.

Figure 4. Figure 4.

Conceptual representation of how one would assess whether plasticity is adaptive. Imagine environmentally induced variation in coloration, from light to dark, which is associated with seasonal, latitudinal, or elevation variation in temperature. One plastic (P) and three alternative nonplastic (N) genotypes expressing phenotypes in warm and cool environments are shown, with the three nonplastic genotypes expressing dark coloration, light coloration, and one of intermediate (i.e., moderate) coloration (Ndark, Nlight, and Nmoderate, respectively). The plastic genotype expresses light coloration when it is warm (e.g., summer phenotype), but dark coloration when it is cool (e.g., spring or autumn phenotypes). Suppose that when it is warm, a light‐colored phenotype avoids overheating. Thus, in the warm environment, the fitness of the plastic genotype (P) is similar to that of the nonplastic light‐colored genotype (Nlight), both of which are more fit than either of the other two nonplastic genotypes. Suppose also that in the cool environment, the plastic genotype (which now expresses a dark phenotype) has similar fitness to that of the nonplastic dark‐colored phenotype, because both can convert the absorbed solar radiation into higher body temperatures necessary for some components of fitness. Provided all of the previously mentioned conditions are true (and making certain assumptions about the probability of encountering both environments), we would conclude that the plasticity is adaptive, because there is no single nonplastic genotype that has, on average, higher fitness. However, if there is no cost to being dark under warm conditions, such that the fitness of the Ndark genotype is as high as that of the plastic genotype under the warm conditions, then there is no fitness advantage to being plastic. Therefore, we could not conclude that the plasticity is adaptive (and indeed, if there are costs to being plastic, per se, we might expect the plastic genotype to have lower fitness than the Ndark genotype, despite that both have the same beneficial, dark‐coloration phenotype under the cool conditions.)

Figure 5. Figure 5.

Depiction of the use of molecular techniques to better understand the genetic basis of phenotypic plasticity. Gene expression data were collected using a global transcriptome analysis of Fundulus heteroclitus gill tissue during hypo‐osmotic challenge. Log2 gene expression values are plotted as a function of time of exposure and partitioned by functional category. This figure is recreated, with permission, from 184 Whitehead et al. “Functional genomics of physiological plasticity and local adaptation in killifish”, Journal of Heredity, June 25, 2010, Figure 3, with permission of Oxford University Press on behalf of the American Genetic Association. Gene names are as follows: “14.3.3.a, 14‐3‐3.a protein; APOC1, apolipoprotein C‐I; APOM, apolipoprotein M; AQP‐3, aquaporin‐3; Arpc1a, actin‐related protein 2/3 complex subunit 1A; ATP6V1E1, V‐type proton ATPase subunit E 1; BCDO2, beta‐carotene dioxygenase 2; CALM, calmodulin; CFTR, cystic fibrosis transmembrane conductance regulator; CLD3, claudin‐3; CLD4, claudin‐4; CMC1, calcium‐binding mitochondrial carrier protein Aralar1; CP24A, 1,25‐dihydroxyvitamin D(3) 24‐hydroxylase; CX32, gap junction connexin‐32.2 protein; CYLC1, cylicin‐1; Dio1; DSC1, desmocollin‐1; ECH1, delta(3,5)‐delta(2,4)‐dienoyl‐CoA isomerase; ECHB, acetyl‐CoA acyltransferase; ERG28, probable ergosterol biosynthetic protein 28; F16PA, fructose‐1,6‐bisphosphatase class 1; G6PI, glucose‐6‐phosphate isomerase; GADD45, growth arrest and DNA damage‐inducible protein GADD45 beta; GSN, gelsolin; H1, histone H1; H2AX, histone H2A.x; H2B1, histone H2B.1; H4, histone H4; HMDH, 3‐hydroxy‐3‐methylglutaryl‐coenzyme A reductase; IMPA1, inositol monophosphatase; INO1, inositol‐3‐phosphate synthase; K6PF, 6‐phosphofructokinase; KCRB, creatine kinase B‐type; KCRM, creatine kinase M‐type; KCRT, creatine kinase, testis isozyme; KRT18, keratin, type I cytoskeletal 18; Marcks, myristoylated alanine‐rich C‐kinase substrate; MYL6, myosin light polypeptide 6; MYL7, myosin regulatory light chain 2, atrial isoform; NEFL, neurofilament light polypeptide; NKA, sodium/potassium‐transporting ATPase subunit alpha‐1; NKCC2, sodium/calcium exchanger 2; ODP2, pyruvate dehydrogenase complex E2 subunit; ODPB, pyruvate dehydrogenase E1 component subunit beta; Ostf1, Oreochromis mossambicus osmotic stress transcription factor 1; OTOP1, otopetrin‐1; PGM2, phosphoglucomutase‐2; PLCD1, phospholipase C‐delta‐1; PLEC1, plectin‐1; PYGM, glycogen phosphorylase; REDD‐1, DNA damage‐inducible transcript 4 protein; RHOA, transforming protein RhoA; S10AD, S100 calcium‐binding protein A13; SAPK‐3, MAP kinase p38γ; SCMC2, small calcium‐binding mitochondrial carrier protein 2; SIAS, sialic acid synthase; TBCA, tubulin‐specific chaperone A; Tnnt3, troponin T, fast skeletal muscle; TPI1, triosephosphate isomerase; TUB1, tubulin alpha chain” (184).

Figure 6. Figure 6.

An integrative approach to the study of phenotypic plasticity utilizing a mouse model. This figure is partially recreated, with permission, from 82 Kelly et al. “Genetic architecture of voluntary exercise in an advanced intercross line of mice” Physiological Genomics, 2010, 42: 190‐200, Figures 1 and 2; and 81 Kelly et al. “Exercise, weight loss, and changes in body composition in mice: phenotypic relationships and genetic architecture,” Physiological Genomics, 2011, 43: 199‐212, Figures 1 and 5. In this integrative example, we illustrate the complex control of a single phenotype, body mass, and the intricacies of its response to an altered environment, exercise. It has been well established that the variation in body mass is partially regulated by genetics (orange panel). Additionally, as demonstrated by the presence of an altered environment (exercise, as opposed to none) there is a plastic response in body weight (blue panel). Although this change is most commonly a reduction in weight, there is substantial variation among a given population (blue panel). Additionally, this variation in the change in body weight in response to exercise has a genetic basis as indicated by the presence of a quantitative trait locus (QTL) on chromosome 11 (red panel). Furthermore, the predisposition to engage in the altered environment (exercise or not) has an underlying genetic architecture (green panel). This type of comprehensive approach demonstrates the complexities of the study of phenotypic plasticity and the power of a molecular approach using a mouse model. For further discussion of examples from this mouse model see text and Garland and Kelly 56. Additionally, for a similar theoretical approach studying bone structure and performance see Middleton et al. 103. For the genome wide QTL plots, red traces are the simple mapping output, and black traces are corrected for family structure in this fourth generation population. The solid black and gray dotted lines represent the permuted 95% [logarithm of odds (LOD) ≥ 3.9, P ≤ 0.05] and 90% (LOD ≥ 3.5, P ≤ 0.1) LOD thresholds, respectively.



Figure 1.

Two genetically identical water fleas, Daphnia lumholtzi. The helmet and extended tail spine of the individual on the left were induced as a result of chemical cues from a predaceous fish and serve as protection 67. This figure is recreated, with permission, from 3 Agrawal, A. “Phenotypic plasticity in the interactions and evolution of species”, Science, October 12, 2001, 294:321‐326, Figure 1, with permission of D. Laforsch.



Figure 2.

Possible relationships among plasticity and genetic variation. In each panel, dots connected by lines represent the phenotypes expressed by each of three genotypes (or families), numbered 1, 2, and 3, in each of two alternative environments (A, B). These lines are the reaction norms. (A) The three genotypes differ in their phenotypes within each environment, indicating genetic variation (VG = yes). However, a given genotype expresses the same phenotype, regardless of environment; that is, the reaction norms are flat. Hence, there is no environmental effect on the phenotype (VE, or plasticity, is absent). Because the reactions norms are parallel, there is no genotype‐by‐environment interaction (GxE = no). (B) As in panel “A,” the three genotypes differ in phenotype within a given environment, but in addition, each genotype expresses a different phenotype in environment “B,” relative to that expressed in environment “A.” That is, the reaction norms are not flat; each genotype is plastic for the trait of interest. However, because the slopes of all reaction norms are parallel, there is no genotype‐by‐environment interaction. The genotypes, although plastic, are all similarly plastic (for the trait of interest). (C) Genotypes differ within environments, show plasticity, and differ in plasticity. That is, the reaction norms are not parallel, indicating genotype‐by‐environment interactions. In this case, reaction norms 2 and 3 cross, but this may not always be the case. This figure is conceptually similar to, and derived, with permission, from Figure 1.4 from Pigliucci 121; p. 15) and Figure 1.1 from reference 38; p. 4)



Figure 3.

The panels above (A‐D) are representations of the general requirements needed for recognizing phenotypic plasticity as an adaptation. Although the representation of each of these criteria is pictorial simplistic, we acknowledge that their conclusive demonstration is quite complex. Accordingly, under each of the four general criteria we have listed additional considerations that should be taken into account. For extensive discussion of these criteria see text (also see reference 42. (A) Environmental heterogeneity must exist, and the degree of heterogeneity may determine the evolutionary (see panel D) consequences as opposed to an alteration in population mean. Heterogeneity may be biotic (e.g., predator presence) or abiotic (e.g., temperature), and special care should be taken to consider the extent of the spatial heterogeneity, the speed of fluctuations, and how these relate to the behavior and life history of the investigated organism. (B) Organisms must be able to reliably predict heterogeneity using environmental signals and these signals must be highly correlated with future environmental conditions. These signals must be spatial and temporally reliable and the organism must have the ability to sense and respond (even if the response is imperfect, see reference 87. (C) There must be an underlying genetic architecture regulating the plastic response. Here, we have presented methodologies for the evaluation of genetic variation. Using quantitative trait locus (QTL) mapping researchers may identify plasticity regions that directly affect the reaction norm (see text and Fig. 3 for examples), or evaluate differential gene expression in different environments with microarray technology. (D) There is a measurable response to selection and the response confers a fitness benefit.



Figure 4.

Conceptual representation of how one would assess whether plasticity is adaptive. Imagine environmentally induced variation in coloration, from light to dark, which is associated with seasonal, latitudinal, or elevation variation in temperature. One plastic (P) and three alternative nonplastic (N) genotypes expressing phenotypes in warm and cool environments are shown, with the three nonplastic genotypes expressing dark coloration, light coloration, and one of intermediate (i.e., moderate) coloration (Ndark, Nlight, and Nmoderate, respectively). The plastic genotype expresses light coloration when it is warm (e.g., summer phenotype), but dark coloration when it is cool (e.g., spring or autumn phenotypes). Suppose that when it is warm, a light‐colored phenotype avoids overheating. Thus, in the warm environment, the fitness of the plastic genotype (P) is similar to that of the nonplastic light‐colored genotype (Nlight), both of which are more fit than either of the other two nonplastic genotypes. Suppose also that in the cool environment, the plastic genotype (which now expresses a dark phenotype) has similar fitness to that of the nonplastic dark‐colored phenotype, because both can convert the absorbed solar radiation into higher body temperatures necessary for some components of fitness. Provided all of the previously mentioned conditions are true (and making certain assumptions about the probability of encountering both environments), we would conclude that the plasticity is adaptive, because there is no single nonplastic genotype that has, on average, higher fitness. However, if there is no cost to being dark under warm conditions, such that the fitness of the Ndark genotype is as high as that of the plastic genotype under the warm conditions, then there is no fitness advantage to being plastic. Therefore, we could not conclude that the plasticity is adaptive (and indeed, if there are costs to being plastic, per se, we might expect the plastic genotype to have lower fitness than the Ndark genotype, despite that both have the same beneficial, dark‐coloration phenotype under the cool conditions.)



Figure 5.

Depiction of the use of molecular techniques to better understand the genetic basis of phenotypic plasticity. Gene expression data were collected using a global transcriptome analysis of Fundulus heteroclitus gill tissue during hypo‐osmotic challenge. Log2 gene expression values are plotted as a function of time of exposure and partitioned by functional category. This figure is recreated, with permission, from 184 Whitehead et al. “Functional genomics of physiological plasticity and local adaptation in killifish”, Journal of Heredity, June 25, 2010, Figure 3, with permission of Oxford University Press on behalf of the American Genetic Association. Gene names are as follows: “14.3.3.a, 14‐3‐3.a protein; APOC1, apolipoprotein C‐I; APOM, apolipoprotein M; AQP‐3, aquaporin‐3; Arpc1a, actin‐related protein 2/3 complex subunit 1A; ATP6V1E1, V‐type proton ATPase subunit E 1; BCDO2, beta‐carotene dioxygenase 2; CALM, calmodulin; CFTR, cystic fibrosis transmembrane conductance regulator; CLD3, claudin‐3; CLD4, claudin‐4; CMC1, calcium‐binding mitochondrial carrier protein Aralar1; CP24A, 1,25‐dihydroxyvitamin D(3) 24‐hydroxylase; CX32, gap junction connexin‐32.2 protein; CYLC1, cylicin‐1; Dio1; DSC1, desmocollin‐1; ECH1, delta(3,5)‐delta(2,4)‐dienoyl‐CoA isomerase; ECHB, acetyl‐CoA acyltransferase; ERG28, probable ergosterol biosynthetic protein 28; F16PA, fructose‐1,6‐bisphosphatase class 1; G6PI, glucose‐6‐phosphate isomerase; GADD45, growth arrest and DNA damage‐inducible protein GADD45 beta; GSN, gelsolin; H1, histone H1; H2AX, histone H2A.x; H2B1, histone H2B.1; H4, histone H4; HMDH, 3‐hydroxy‐3‐methylglutaryl‐coenzyme A reductase; IMPA1, inositol monophosphatase; INO1, inositol‐3‐phosphate synthase; K6PF, 6‐phosphofructokinase; KCRB, creatine kinase B‐type; KCRM, creatine kinase M‐type; KCRT, creatine kinase, testis isozyme; KRT18, keratin, type I cytoskeletal 18; Marcks, myristoylated alanine‐rich C‐kinase substrate; MYL6, myosin light polypeptide 6; MYL7, myosin regulatory light chain 2, atrial isoform; NEFL, neurofilament light polypeptide; NKA, sodium/potassium‐transporting ATPase subunit alpha‐1; NKCC2, sodium/calcium exchanger 2; ODP2, pyruvate dehydrogenase complex E2 subunit; ODPB, pyruvate dehydrogenase E1 component subunit beta; Ostf1, Oreochromis mossambicus osmotic stress transcription factor 1; OTOP1, otopetrin‐1; PGM2, phosphoglucomutase‐2; PLCD1, phospholipase C‐delta‐1; PLEC1, plectin‐1; PYGM, glycogen phosphorylase; REDD‐1, DNA damage‐inducible transcript 4 protein; RHOA, transforming protein RhoA; S10AD, S100 calcium‐binding protein A13; SAPK‐3, MAP kinase p38γ; SCMC2, small calcium‐binding mitochondrial carrier protein 2; SIAS, sialic acid synthase; TBCA, tubulin‐specific chaperone A; Tnnt3, troponin T, fast skeletal muscle; TPI1, triosephosphate isomerase; TUB1, tubulin alpha chain” (184).



Figure 6.

An integrative approach to the study of phenotypic plasticity utilizing a mouse model. This figure is partially recreated, with permission, from 82 Kelly et al. “Genetic architecture of voluntary exercise in an advanced intercross line of mice” Physiological Genomics, 2010, 42: 190‐200, Figures 1 and 2; and 81 Kelly et al. “Exercise, weight loss, and changes in body composition in mice: phenotypic relationships and genetic architecture,” Physiological Genomics, 2011, 43: 199‐212, Figures 1 and 5. In this integrative example, we illustrate the complex control of a single phenotype, body mass, and the intricacies of its response to an altered environment, exercise. It has been well established that the variation in body mass is partially regulated by genetics (orange panel). Additionally, as demonstrated by the presence of an altered environment (exercise, as opposed to none) there is a plastic response in body weight (blue panel). Although this change is most commonly a reduction in weight, there is substantial variation among a given population (blue panel). Additionally, this variation in the change in body weight in response to exercise has a genetic basis as indicated by the presence of a quantitative trait locus (QTL) on chromosome 11 (red panel). Furthermore, the predisposition to engage in the altered environment (exercise or not) has an underlying genetic architecture (green panel). This type of comprehensive approach demonstrates the complexities of the study of phenotypic plasticity and the power of a molecular approach using a mouse model. For further discussion of examples from this mouse model see text and Garland and Kelly 56. Additionally, for a similar theoretical approach studying bone structure and performance see Middleton et al. 103. For the genome wide QTL plots, red traces are the simple mapping output, and black traces are corrected for family structure in this fourth generation population. The solid black and gray dotted lines represent the permuted 95% [logarithm of odds (LOD) ≥ 3.9, P ≤ 0.05] and 90% (LOD ≥ 3.5, P ≤ 0.1) LOD thresholds, respectively.

References
 1. Able KW, Palmer RE. Salinity effects on fertilization success and larval mortality of Fundulus heteroclitus. Copeia 2: 345‐350, 1988.
 2. Ang ET, Gomez‐Padilla F. Potential therapeutic effects of exercise to the brain. Curr Med Chem 14: 2564‐2571, 2007.
 3. Agrawal AA. Phenotypic plasticity in the interactions and evolution of species. Science 294: 321‐326, 2001.
 4. Aubin‐Horth N, Renn SCP. Genomic reaction norms: Using integrative biology to understand molecular mechanisms of phenotypic plasticity. Mol Ecol 18: 3763‐3780, 2009.
 5. Auld JR, Agrawal AA, Relyea RA. Re‐evaluating the costs and limits of adaptive phenotypic plasticity. Proc Biol Sci 277: 503‐511, 2010.
 6. Bateson P. Developmental plasticity and evolutionary biology. J Nutr 137: 1060‐1062, 2007.
 7. Bateson P, Barker D, Clutton‐Brock T, Deb D, D'Udine B, Foley RA, Gluckman P, Godfrey K, Kirkwood T, Lahr MM, Metcalf NB, Monaghan P, Spencer HG, Sultan SE. Developmental plasticity and human health. Nature 430: 419‐421, 2004.
 8. Beldade P, Mateus ARA, Keller RA. Evolution and molecular mechanisms of adaptive development plasticity. Mol Ecol 20: 1347‐1363, 2011.
 9. Bell GAC. Experimental evolution in Chlamydomonas. I. Short‐term selection in uniform and diverse environments. Heredity 78: 490‐497, 1997.
 10. Bennett AF. Adaptation and the evolution of physiological characters. In: Dantzler WH, editor. Handbook of Physiology: Comparative Physiology, Vol. I. New York: Oxford University Press, sect. 3, Chapter 1: Evolutionary Physiology 1‐16, 1997.
 11. Bennett AF. Experimental evolution and the Krogh Principle: Generating biological novelty for functional and genetic analyses. Physiol Biochem Zool 76: 1‐11, 2003.
 12. Bennett AF, Lenski RE. Evolutionary adaptation to temperature: VI. Phenotypic acclimation and its evolution in Escherichia coli. Evolution 51: 36‐44, 1997.
 13. Blair SN, Morris JN. Healthy hearts‐and the universal benefits of being physically active: Physical activity and health. Ann Epidemiol 19: 253‐256, 2009.
 14. Blundell JE, Stubbs RJ, Hughes DA, Whybrow S, King NA. Cross talk between physical activity and appetite control: Does physical activity stimulate appetite? Proc Nutr Soc 62: 651‐661, 2003.
 15. Booth FW, Chakravathy MV, Spangenburg EE. Exercise and gene expression: Physiological regulation of the human genome through physical activity. J Physiol (Lond) 543: 399‐411, 2002.
 16. Bossdorf O, Richards CL, Pigliucci M. Epigenetics for ecologists. Ecol Lett 11: 106‐115, 2008.
 17. Bouchard C, Tremblay A, Nadeau A, Dussault J, Despres JP, Theriault G, Lupien PJ, Serresse O, Boulay MR, Fournier G. Long‐term exercise training with constant energy intake. 1: Effect on body composition and selected metabolic variables. Int J Obes 14: 57‐73, 1990.
 18. Bradshaw AD. Evolutionary significance of phenotypic plasticity in plants. Adv Gen 13: 115‐155, 1965.
 19. Brakefield PM, Frankino WA. Polyphenisms in Lepidoptera: multidisciplinary approaches to studies of evolution and development. In: Whitman DW, Ananthakrishnan TN, editors. Phenotypic Plasticity of Insects: Mechanisms and Consequences. Enfield, New Hampshire, USA: Science Publishers, 281‐312, 2009.
 20. Bray MS, Hagberg JM, Perusse L, Rankinen T, Roth SM, Wolfarth B, Bouchard C. The human gene map for performance and health related fitness phenotypes: The 2006‐2007 update. Med Sci Sports Exerc 41: 34‐42, 2009.
 21. Burnett KG, Bain LJ, Baldwin WS, Callard GV, Cohen S, Di Giulio, RT, Evans DH, Gomez‐Chiarri M, Hahn ME, Hoover CA, Karchner SI, Katoh F, Maclatchy DL, Marshall WS, Meyer JN, Nacci DE, Oleksiak MF, Rees BB, Singer TD, Stegeman JJ, Towle DW, Van Veld PA, Vogelbein WK, Whitehead A, Winn RN, Crawford DL. Fundulus as the premier teleost model in environmental biology: Opportunities for new insights using genomic. Comp Biochem Physiol Part D Genomics Proteomics 2: 257‐286, 2007.
 22. Canfield M, Greene E. Phenotypic plasticity and the semantics of polyphenism: A historical review and current perspectives. In: Whitman DW, Ananthakrishnan TN, editors. Phenotypic Plasticity of Insects: Mechanisms and Consequences. Enfield, New Hampshire, USA: Science Publishers, 11‐26, 2009.
 23. Castilho PC, Buckley BA, Somero G, Block BA. Heterologous hybridization to a complementary DNA microarray reveals the effect of thermal acclimation in the endothermic bluefin tuna (Thunnus orientalis). Mol Ecol 18: 2092‐2102, 2009.
 24. Chapman LJ, Albert J, Galis F. Developmental plasticity, genetic differentiation, and hypoxia‐induced trade‐offs in an African cichlid fish. Open Evol J 2: 75‐88, 2008.
 25. Chapman LJ, Chapman CA, Nordlie FG, Rosenberger AE. Physiological refugia: Swamps, hypoxia tolerance, and maintenance of fish biodiversity in the Lake Victoria Region. Comp Biochem Physiol 133: 421‐437, 2002.
 26. Chapman LJ, Galis F, Shinn J. Phenotypic plasticity and the possible role of genetic assimilation: Hypoxia‐induced trade‐offs in the morphological traits of an African cichlid. Ecol Lett 3: 388‐393, 2000.
 27. Chapman LJ, Kaufman LS, Chapman CA, McKenzie FE. Hypoxia tolerance in twelve species of East African cichlids: Potential for low oxygen refugia in Lake Victoria. Conserv Biol 9: 1274‐1288, 1995.
 28. Chippari‐Gomes AR, Gomes LC, Lopes NP, Val AL, Almeida‐Val VM. Metabolic adjustments in two Amazonian cichlids exposed to hypoxia and anoxia. Comp Biochem Physiol B Biochem Mol Biol 141: 347‐55, 2005.
 29. Chong S, Youngson NA, Whitelaw E. Heritable germline epimutation is not the same as transgenerational epigenetic inheritance. Nat Genet 39: 574‐575, 2007.
 30. Church TS, Blair SN. When will we treat physical activity as a legitimate medical therapy even though it does not come in a pill? Br J Sports Med 43: 80‐81, 2009.
 31. Cohet Y, David JR. Control of adult reproductive potential by preimaginal thermal conditions: A study in Drosophila melanogaster. Oecologia (Berlin) 36: 295‐306, 1978.
 32. Cotman CW, Berchtold NC, Christie LA. Exercise builds brain health: Key roles of growth factor cascades and inflammation. Trends Neurosci 30: 464‐472, 2007.
 33. Cotter SC, Simpson SJ, Raubenheimer D, Wilson K. Macronutrient balance mediates trade‐offs between immune function and life history traits. Func Ecol 25: 186‐198, 2011.
 34. Cossins A, Fraser J, Hughes M, Gracey A. Post‐genomic approaches to understanding the mechanisms of environmentally induced phenotypic plasticity. J Exp Biol 209: 2328‐2336, 2006.
 35. Crispo E, Chapman LJ. Geographic variation in phenotypic plasticity in response to dissolved oxygen in an African cichlid fish. J Evol Biol 23: 2091‐103, 2010.
 36. DeKoning ABL, Picard DJ, Bond SR, Schulte PM. Stress and inter‐population variation in glycolytic enzyme expression in a teleost fish, Fundulus heteroclitus. Physiol Biochem Zool 77: 18‐26, 2004.
 37. DeLoof A, Claeys I, Simonet G, Verleyen P, Vandersmissen T, Sas F, Huybrechts J. Molecular markers of phase transition in locusts. Insect Science 13: 3‐12, 2006.
 38. DeWitt TJ, Schiener SM. Phenotypic Plasticity: Functional and Conceptual Approaches. New York: Oxford University Press, 2004.
 39. DeWitt TJ, Sih A, Wilson DS. Costs and limits of phenotypic plasticity. Trends Ecol Evol 13: 77‐81, 1998.
 40. Donnelly JE, Blair SN, Jakicic JM, Manore MM, Rankin JW, Smith BK, American College of Sports Medicine. American college of sports medicine position stand. Appropriate physical activity intervention strategies for weight loss and prevention of weight regain for adults. Med Sci Sports Exerc 41: 459‐471, 2009.
 41. Donnelly JE, Smith BK. Is exercise effective for weight loss with ad libitum diet? Energy balance, compensation, and gender differences. Exerc Sport Sci Rev 33: 169‐174, 2005.
 42. Doughty P, Reznick DN. Patterns and analysis of adaptive phenotypic plasticity in animals. In: DeWitt TJ, Schiener SM, editors. Phenotypic Plasticity: Functional and Conceptual Approaches. New York: Oxford University Press, 126‐150, 2004.
 43. Evans DH, Claiborne JB. The Physiology of Fishes. Boca Raton, FL: CRC Press, Taylor and Francis Group, 2006.
 44. Evans DH, Piermarini PM, Choe KP. The multifunctional fish gill: Dominant site of gas exchange, osmoregulation, acid‐base regulation, and excretion of nitrogenous waste. Physiol Rev 85: 97‐177, 2005.
 45. Falconer DS. Selection of mice for growth on high and low planes of nutrition. Genet Res 1: 91‐113, 1960.
 46. Falconer DS. Selection in different environments: Effects on environmental sensitivity (reaction norm) and on mean performance. Genet Res 56: 57‐70, 1990.
 47. Falconer DS, MacKay TFC. Introduction to Quantitative Genetics. New York: Pearson Education Limited, 1981.
 48. Feder ME, Bennett AF, Huey RB. Evolutionary physiology. Annu Rev Ecol Syst 31: 315‐341, 2000.
 49. Fedorka KM, Zuk M, Mousseau TA. Immune suppression and the cost of reproduction in the ground cricket, Allonemobius socius. Evolution 58: 2478‐2485, 2004.
 50. Fischer M, van Kleunen, Schmid B. Experimental life‐history evolution: Selection on growth form and its plasticity in a clonal plant. J Evol Biol 17: 331‐341, 2004.
 51. Flück M. Functional, structural and molecular plasticity of mammalian skeletal muscle in response to exercise stimuli. J Exp Biol 209: 2239‐2248, 2006.
 52. Fordyce JA. The evolutionary consequences of ecological interactions mediated through phenotypic plasticity. J Exp Biol 209: 2377‐2383, 2006.
 53. Franks PW, Ling C. Epigenetics and obesity: The devil is in the details. CMB Medicine 8: 88‐92, 2010.
 54. Garland T Jr. Selection experiments: An under‐utilized tool in biomechanics and organismal biology. In: Bels VL, Gasc JP, editors. Vertebrate Biomechanics and Evolution. Casinos A. Oxford: BIOS Scientific Publishers, 2003.
 55. Garland T Jr, Carter PA. Evolutionary physiology. Annu Rev Physiol 56: 579‐621, 1994.
 56. Garland T Jr, Kelly SA. Phenotypic plasticity and experimental evolution. J Exp Biol 209: 2344‐2361, 2006.
 57. Garland T Jr, Rose MR. Experimental Evolution: Concepts, Methods, and Applications of Selection Experiments. Berkeley CA: Univ. of California Press, 2009.
 58. Garland T Jr, Schutz H, Chappell M, Keeney BK, Meek TH, Copes LE, Acosta W, Drenowatz C, Maciel RC, van Dijk G, Kotz CM, Eisenmann JC. The biological control of voluntary exercise, spontaneous physical activity and daily energy expenditure in relation to obesity: Human and rodent perspectives. J Exp Biol 214(Pt 2): 206‐229, 2011.
 59. Gibert P, Huey RB, Gilchrist GW. Locomotor performance of Drosophila melanogaster: Interactions among developmental and adult temperatures, age, and geography. Evolution 55: 205‐209, 2001.
 60. Ghalambor C, McKay J, Carroll S, Reznick D. Adaptive versus non‐adaptive phenotypic plasticity and the potential for contemporary adaptation in new environments. Func Ecol 21: 394‐407, 2007.
 61. Gilbert SF, Epel D. Ecological Developmental Biology: Integrating Epigenetics, Medicine, and Evolution. Sunderland, MA: Sinauer, 2009.
 62. Gluckman PD, Hanson MA, Spencer HG, Bateson P. Environmental influences during development and their later consequences for health and disease: Implications for the interpretation of empirical studies. Proc Biol Sci 1564: 671‐677, 2005.
 63. Gomes FR, Rezende EL, Malisch JL, Lee SK, Rivas DA, Kelly SA, Lytle C, Yaspelkis III BB, Garland T Jr. Glycogen storage and muscle glucose transporters (GLUT‐4) of mice selectively bred for high voluntary wheel running. J Exp Biol 212: 238‐248, 2009.
 64. Gordon DM. Phenotypic plasticity. In: Keller EF, Loyd EA, editor. Keywords in Evolutionary Biology. Cambridge, MA: Harvard University Press, 1992.
 65. Grant V. Organismic Evolution. San Francisco, CA: Freeman, 1977.
 66. Grediagin A, Cody M, Rupp J, Benardot D, Shern R. Exercise does not effect body composition in untrained, moderately overfat women. J Am Diet Assoc 95: 661‐665, 1995.
 67. Green J. The distribution and variation of Daphnia lumholtz (Crustacea: Cladocera) in relation to fish predation in Lake Albert, East Africa. J Zool 151: 181‐197, 1967.
 68. Griffith RW. Environment and salinity tolerance in the genus Fundulus. Copeia 2: 319‐331, 1974.
 69. Guo W, Wang X, Ma Z, Xue L, Han J, Yu D, Kang L. CSP and Takeout genes modulate the switch between attraction and repulsion during behavioral phase change in the migratory locust. PLoS Genet 7: e1001291, 1‐13, 2011.
 70. Hanson MA, Gluckman PD. Developmental origins of health and disease: New insights. Basic Clin Pharmacol Toxicol 102: 90‐93, 2008.
 71. Huey RB, Berrigan D. Testing evolutionary hypotheses of acclimation. In: Johnston IA, Bennett AF, editors. Phenotypic and Evolutionary Adaptation to Temperature. Cambridge, MA: Cambridge University Press, 1996.
 72. Huey RB, Berrigan D, Gilchrist GW, Herron JC. Testing the adaptive significance of acclimation: A strong inference approach. Am Zool 39: 323‐336, 1999.
 73. Hughes GM, Morgan M. Structure of fish gills in relation to their respiratory function. Biol Rev 48: 419‐475, 1973.
 74. Ingleby FC, Hunt J, Hosken DJ. The role of genotype‐by‐environment interactions in sexual selection. J Evol Biol 23: 2031‐2045, 2010.
 75. Irschick DJ, Reznick DN. Field experiments, introductions, and experimental evolution: A review and practical guide. In: Garland T Jr, Rose MR, editor. Experimental Evolution: Concepts, Methods, and Applications of Selection Experiments. Berkeley CA: University of California Press, 2009.
 76. Jakicic JM, Clark K, Coleman E, Donnelly JE, Foreyt J, Melanson E, Volek J, Volpe SL, American College of Sports Medicine. American College of Sports Medicine position stand. Appropriate intervention strategies for weight loss and prevention of weight regain for adults. Med Sci Sports Exerc 33: 2145‐2156, 2001.
 77. Jakicic JM, Marcus BH, Gallagher KI, Napolitano M, Lang W. Effect of exercise duration and intensity on weight loss in overweight, sedentary women. JAMA 290: 1323‐1330, 2003.
 78. Kang L, Chen XY, Zhou Y, Liu BW, Li RQ, Wang J, Yu J. The analysis of large‐scale gene expression correlated to the phase changes of the migratory locust. Proc Natl Acad Sci U S A 101: 17611‐17615, 2004.
 79. Karavirta L, Hakkinen K, Kauhanen A, Arija‐Blazquez A, Sillanpaa E, Rinkinen N, Hakkinen A. Individual responses to combined endurance and strength training in older adults. Med Sci Sports Exerc 43: 484‐490, 2010.
 80. Karnaky KJ. Structure and function of the chloride dell of Fundulus heteroclitus and other teleosts. Am Zool 26: 209‐224, 1986.
 81. Kelly SA, Nehrenberg DL, Hua K, Garland T Jr, Pomp D. Exercise, weight loss, and changes in body composition in mice: Phenotypic relationships and genetic architecture. Physiol Genomics 43: 199‐212, 2011.
 82. Kelly SA, Nehrenberg DL, Peirce JL, Hua K, Steffy BM, Wiltshire T, Pardo‐Manuel de Villena F, Garland T Jr, Pomp D. Genetic architecture of voluntary exercise in an advanced intercross line of mice. Physiol Genomics 42: 190‐200, 2010.
 83. King NA, Hopkins M, Caudwell P, Stubbs RJ, Blundell JE. Individual variability following 12 weeks of supervised exercise: Identification and characterization of compensation for exercise‐induced weight loss. Int J Obes 32: 177‐184, 2008.
 84. King NA, Hopkins M, Caudwell P, Stubbs RJ, Blundell JE. Beneficial effects of exercise: Shifting the focus from body weight to other markers of health. Br J Sports Med 43: 924‐927, 2009.
 85. Koch LG, Green CL, Lee AD, Hornyak JE, Cicila GT, Britton SL. Test of the principle of the initial value in rat genetic models of exercise capacity. Am J Physiol Integr Comp Physiol 288: R466‐R472, 2005.
 86. Kokkinos P, Myers J. Exercise and physical activity: Clinical outcomes and applications. Circulation 122: 1637‐1648, 2010.
 87. Langerhans RB, DeWitt TJ. Plasticity constrained: Over‐generalized induction cues cause maladaptive phenotypes. Evol Ecol Res 4: 857‐870.
 88. Leamy LJ, Pomp D, Lightfoot JT. Genetic variation in the pleiotropic association between physical activity and body weight in mice. Genet Sel Evol 41: 41, 2009.
 89. Lee KP, Simpson SJ, Wilson K. Dietary protein‐quality influences melanization and immune function in an insect. Func Ecol 22: 1052‐1061, 2008.
 90. Leroi A, Bennett AF, Lenski RE. Temperature acclimation and competitive fitness: An experimental test of the beneficial acclimation assumption. Proc Natl Acad Sci U S A 91: 1917‐1921, 1994.
 91. Levin DA. Local differentiation and the breeding structure of plant populations. In: LD Gottlieb, SK Jain, editors. Plant Evolutionary Biology. New York: Chapman & Hall, 1988, p. 305‐329
 92. Losos JB, Creer DA, Glossip D, Hampton A, Roberts G, Haskell N, Taylor P, Ettling J. Evolutionary implications of phenotypic plasticity in the hindlimb of the lizard Anolis sagrei. Evolution 54: 301‐305, 2000.
 93. Ma Z, Yu J, Kang L. LocustDB: A relational database for the transcriptome and biology of the migratory locust (Locusta migratoria). BMC Genomics 7: 11, 2006.
 94. Mandic M, Todgham A, Richards J. Mechanisms and evolution of hypoxia tolerance in fish. Proc Biol Sci 276: 735‐44, 2009.
 95. Mancera JM, McCormick SD. Influence of cortisol, growth hormone, insulin‐like growth factor I and 3,3V,5‐triiodo‐l‐thyronine on hypoosmoregulatory ability in the euryhaline teleost Fundulus heteroclitus. Fish Physiol Biochem 21: 25‐33, 1999.
 96. Marshall WS. Rapid regulation of NaCl secretion by estuarine teleost fish: coping strategies for short‐duration freshwater exposures. Biochim Biophys Acta 1618: 95‐105, 2003.
 97. Marshall WS, Emberley TR, Singer TD, Bryson SE, McCormick SD. Time course of salinity adaptation in a strongly euryhaline estuarine teleost, Fundulus heteroclitus: A multivariable approach. J Exp Biol 202: 1535‐1544, 1999.
 98. Martinez ML, Chapman LJ, Rees BB. Population variation in hypoxic responses of the cichlid Pseudocrenilabrus multicolor victoriae. Canadian J Zool 87: 188‐194, 2009.
 99. McCairns RJ, Bernatchez L. Adaptive divergence between freshwater and marine sticklebacks: Insights into the role of phenotypic plasticity from an integrated analysis of candidate gene expression. Evolution 64: 1029‐47, 2010.
 100. McKean K, Nunney L. Increased sexual activity reduces male immune function in Drosophila melanogaster. Proc Natl Acad Sci U S A 98: 7904‐7909, 2001.
 101. Melzer K, Kayser B, Saris WH, Pichard C. Effects of physical activity on food intake. Clin Nutr 24: 885‐895, 2005.
 102. Meyer A. Phenotypic plasticity and heterochrony in cichlasoma‐mangaguense (Pisces, Cichlidae) and their implications for speciation in cichlid fishes. Evolution 41: 1357‐1369, 1987.
 103. Middleton KM, Kelly SA, Garland T Jr. Selective breeding as a tool to probe skeletal response to high voluntary locomotor activity in mice. Integr Comp Biol 48: 394‐410, 2008.
 104. Mikolajewski DJ, Stoks R, Rolff J, Joop G. Predators and cannibals modulate sex‐specific plasticity in life‐history and immune traits. Func Ecol 22: 114‐120, 2008.
 105. Mitchell JA, Church TS, Rankinen T, Earnest CP, Sui X, Blair SN. FTO genotype and the weight loss benefits of moderate intensity exercise. Obesity 18: 641‐643, 2010.
 106. Moczek AP. Developmental capacitance, genetic accommodation, and adaptive evolution. Evol Dev 9: 299‐305, 2007.
 107. Molteni R, Barnard RJ, Ying Z, Roberts CK, Gomez‐Pinilla F. A high‐fat sugar diet reduces hippocampal brain‐derived neurotrophic factor, neuronal plasticity, and learning. Neuroscience 112: 803‐814, 2002.
 108. Mykles DL, Ghalambor CK, Stillman JH, Tomanek L. Grand challenges in comparative physiology: Integration across disciplines and across levels of biological organization. Integr Comp Biol 50: 6‐16, 2010.
 109. Nacci D, Champlin D, Jayaraman S. Adaptation of the estuarine fish Fundulus heteroclitus (Atlantic Killifish) to polychlorinated biphenyls (PCBs). Estuar Coast 33: 853‐864, 2010.
 110. Nehrenberg DL, Hua K, Estrada‐Smith D, Garland T Jr, Pomp D. Voluntary exercise and its effects on body composition depend on genetic selection history. Obesity 17: 1402‐1409, 2009.
 111. NHLBI Obesity Task Force. Clinical guidelines on the identification, evaluation, and treatment of overweight and obesity in adults‐the evidence report. National Institutes of Health. Obes Res 2: 51S‐209S, 1998.
 112. Nijhout HF. Development and evolution of adaptive polyphenisms. Evol Dev 5: 9‐18, 2003.
 113. Nunney L, Cheung W. The effect of temperature on body size and fecundity in female Drosophila melanogaster: Evidence for adaptive plasticity. Evolution 51: 1529‐1535, 1997.
 114. Pener MP, Simpson SJ. Locust phase polyphenism: An update. Adv Insect Physiol 36: 1‐272, 2009.
 115. Pener MP, Yerushalmi Y. The physiology of locust phase polymorphism: An update. J Insect Physiol 44: 365‐377, 1998.
 116. Picard DJ, Schulte PM. Variation in gene expression in response to stress in two populations of Fundulus heteroclitus. Comp Biochem Physiol A Mol Integr Physiol 137: 205‐216, 2004.
 117. Pierce VA, Crawford DL. Variation in the glycolytic pathway: The role of evolutionary and physiological processes. Physiol Zool 69: 489‐508, 1996.
 118. Piersma T, Drent J. Phenotypic flexibility and the evolution of organismal design. Trends Ecol Evol 18: 228‐233, 2003.
 119. Piersma T, Lindstrom A. Rapid reversible changes in organ size as a component of adaptive behavior. Trends Ecol Evol 12: 134‐138, 1997.
 120. Piersma T, van Gils JA. The Flexible Phenotype: A Body Centered Integration of Ecology, Physiology, and Behaviour. New York, NY: Oxford University Press, 2011.
 121. Pigliucci M. Phenotypic Plasticity: Beyond Nature and Nurture. Baltimore, MD: The Johns Hopkins University Press, 2001.
 122. Pigliucci M. Phenotypic plasticity. In: Fox CW, Roff DA, Fairbairn DJ, editors. Evolutionary Ecology: Concepts and Case Studies. New York, New York: Oxford University Press, 2001.
 123. Pigliucci M. Studying the plasticity of phenotypic integration in a model organism. In: Pigliucci M, Preston K, editors. Phenotypic Integration: Studying The Ecology and Evolution of Complex Phenotypes. New York: Oxford University Press, 2004.
 124. Pigliucci M. Evolution of phenotypic plasticity: Where are we going now? Trends Ecol Evol 20: 481‐486, 2005.
 125. Pigliucci M. Epigenetics for ecologists. Ecol Lett 11: 106‐115, 2008.
 126. Pigliucci M, Murran CJ, Schlichting CD. Phenotypic plasticity and evolution by genetic assimilation. J Exp Biol 209: 2362‐2367, 2006.
 127. Pigliucci M, Schlichting CD. Phenotypic Evolution: A Reaction Norm Perspective. Sunderland, MA: Sinauer, 1998.
 128. Podrabsky JE, Javillonar C, Hand SC, Crawford DL. Intraspecific variation in aerobic metabolism and glycolytic enzyme expression in heart ventricles. Am J Physiol Regul Integr Comp Physiol 279: R2344‐R2348, 2000.
 129. Poulton EB. Further experiments upon the colour‐relation between certain lepidopterous larvae, pupae, cocoons, and imagines and their surroundings. Trans Royal Entomol Soc Lond 1892: 293‐487, 1892.
 130. Powers DA, Ropson I, Brown DC, Vanbeneden R, Cashon R, Gonzalez‐Villaseñor LI, Dimichele, JA. Genetic variation in Fundulus heteroclitus: Geographic distribution. Am Zool 26: 131‐144, 1986.
 131. Powers DA, Smith M, Gonzalez‐Villasenor I, DiMichele L, Crawford DL, Bernardi G, Lauerman T. A multidisciplinary approach to the selectionist/neutralist controversy using the model teleost, Fundulus heteroclitus. In: Futuyama D, Antonovics J, editors. Oxford Survey in Evolutionary Biology, Vol. 9. Oxford, UK: Oxford University Press, 43‐107, 1993.
 132. Price TD. Phenotypc plasticity, sexual selection and the evolution of colour patterns. J Exp Biol 209: 2368‐2376, 2006.
 133. Prudic KL, Jeon C, Cao H, Monteiro A. Developmental plasticity in sexual roles of butterfly species drives mutual sexual ornamentation. Science 331: 73‐75, 2011.
 134. Reboud X, Bell G. Experimental evolution of Chlamydomonas. III. Evolution of specialist and generalist types in environments that vary in space and time. Heredity 78: 507‐514, 1997.
 135. Rhodes JS, van Praag H, Jeffery S, Girard I, Mitchell GS, Garland T Jr, Gage FH. Exercise increase hippocampal neurogenesis to high levels but does not improve spatial learning in mice bred for increased voluntary wheel running. Behav Neurosci 117: 1006‐1016, 2003.
 136. Richards JG. Physiological, behavioral and biochemical adaptations in intertidal fishes to hypoxia. J Exp Biol 214: 191‐199, 2011.
 137. Ritt M, Lechleitner M. Effect of exercise intensity on body composition. JAMA 290: 3069, 2003.
 138. Rolff J, Siva‐Jothy MT. Copulation corrupts immunity: A mechanism for a cost of mating in insects. Proc Natl Acad Sci U S A 99: 9916‐9918, 2002.
 139. Rolff J, Siva‐Jothy MT. Invertebrate ecological immunity. Science 301: 472‐475, 2003.
 140. Ross R, Janssen I. Physical activity, total and regional obesity: Dose‐response considerations. Med Sci Sports Exerc 33: S521‐S527, 2001.
 141. Sarkar S. From the Reakionsnorm to the evolution of adaptive plasticity: A historical sketch, 1909‐1999. In: DeWitt TJ, Schiener SM, editors. Phenotypic Plasticity: Functional and Conceptual Approaches. New York: Oxford University Press, 2004.
 142. Schaack SR, Chapman LJ. Interdemic variation in the African cyprinid Barbus neumayeri: Correlations among hypoxia, morphology, and feeding performance. Can J Zool 81: 430‐440, 2003.
 143. Scheiner SM. Genetics and evolution of phenotypic plasticity. Annu Rev Ecol Syst 24: 35‐68, 1993.
 144. Scheiner SM. Selection experiments and the study of phenotypic plasticity. J Evol Biol 15: 889‐898, 2002.
 145. Scheiner SM, Lyman R. The genetics of phenotypic plasticity. II. Response to selection. J Evol Biol 4: 23–50, 1991.
 146. Schlichting CD, Smith H. Phenotypic plasticity: Linking molecular mechanisms with evolutionary outcomes. Evol Ecol 16: 189‐211, 2002.
 147. Schmalhausen II. Factors of Evolution: The Theory of Stabilizing Selection. University of Chicago Press, Chicago. 1949.
 148. Schmid‐Hempel P. Evolutionary ecology of insect immune defenses. Annu Rev Entomol 50: 529‐551, 2005.
 149. Scott GR, Schulte PM. Intraspecific variation in gene expression after seawater transfer in gills of the euryhaline killifish Fundulus heteroclitus. Comp Biochem Physiol A Mol Integr Physiol 141: 176‐182, 2005.
 150. Scott GR, Baker DW, Schulte PM, Wood CM. Physiological and molecular mechanisms of osmoregulatory plasticity in killifish after seawater transfer. J Exp Biol 211: 2450‐2459, 2008.
 151. Scott GR, Claiborne JB, Edwards SL, Schulte PM, Wood CM. Gene expression after freshwater transfer in gills and opercular epithelia of killifish: Insight into divergent mechanisms of ion transport. J Exp Biol 208: 2719‐2729, 2005.
 152. Scott GR, Richards JG, Forbush B, Isenring P, Schulte PM. Changes in gene expression in gills of the euryhaline killifish Fundulus heteroclitus after abrupt salinity transfer. Am J Physiol Cell Physiol 287: 300‐309, 2004.
 153. Scott GR, Rogers JT, Richards JG, Wood CM, Schulte PM. Intraspecific divergence of ionoregulatory physiology in the euryhaline teleost Fundulus heteroclitus: Possible mechanisms of freshwater adaptation. J Exp Biol 207: 3399‐3410, 2004.
 154. Shaw JR, Gabor KH, Hand E, Lankowski A, Durant L, Thibodeau R, Stanton CR, Barnaby R, Coutermarsh B, Karlson KH, Sato JD, Hamilton JW, Stanton BA. Role of glucocorticoid receptor in acclimation of killifish (Fundulus heteroclitus) to seawater and effects of arsenic. Am J Physiol 292: R1052‐R1060, 2007.
 155. Simpson GG. The Baldwin effect. Evolution 7: 110‐117, 1953.
 156. Simpson SJ, Sword GA. Locusts. Curr Biol 18: R364‐R366, 2008
 157. Simpson SJ, Sword GA. Phase polyphenism in locusts: Mechanisms, population consequences, adaptive significance and evolution. In: Whitman DW, Ananthakrishnan TN, editors. Phenotypic Plasticity of Insects: Mechanisms and Consequences. Enfield, New Hampshire, USA: Science Publishers, 2009.
 158. Simpson SJ, Sword GA, DeLoof A. Advances, controversies and consensus in locust phase polyphenism research. J Orthopteran Res 14: 213‐222, 2005.
 159. Simpson SJ, Despland E, Hagele BF, Dodgson T. Gregarious behavior in desert locusts is evoked by touching their back legs. Proc Natl Acad Sci U S A 98: 3895‐3897, 2001.
 160. Simpson SJ, Raubenheimer D, Behmer ST, Whitworth A, Wright GA. A comparison of nutritional regulation in solitarious‐ and gregarious‐phase nymphs of the desert locust Schistocerca gregaria. J Exp Biol 205: 121‐129, 2002.
 161. Singer TD, Keir KR, Hinton M, Scott GR, McKinley RS, Schulte PM. Structure and regulation of the cystic fibrosis transmembrane conductance regulator (CFTR) gene in killifish: A comparative genomics approach. Comp Biochem and Phys Part D 3: 172‐185, 2008.
 162. Siva‐Jothy MT, Moret Y, Rolff J. Insect immunity: An evolutionary ecology perspective. Adv Insect Physiol 32: 1‐48, 2005.
 163. Siva‐Jothy MT, Thompson JJW. Short‐term nutrient deprivation affects immune function. Physiol Entomol 27: 206‐212, 2002.
 164. Siva‐Jothy MT, Tsubaki Y, Hooper RE. Decreased immune response as a proximate cost of copulation and oviposition in a damselfly. Physiol Entomol 23: 274–277, 1998.
 165. Song H, Wenzel JW. Phylogeny of bird‐grasshopper subfamily Cyrtacanthacridinae (Orthoptera: Acrididae) and the evolution of locust phase polyphenism. Cladistics 24: 515‐542, 2008.
 166. Stearns SC. The evolutionary significance of phenotypic plasticity. BioScience 39: 436‐445, 1989.
 167. Storz JF, Scott GR, Cheviron ZA. Phenotypic plasticity and genetic adaptation to high‐altitude hypoxia in vertebrates. J Exp Biol 213: 4125‐4136, 2010.
 168. Swallow JG, Carter PA, Garland T Jr. Artificial selection for increased wheel‐running behavior in house mice. Behav Genet 28: 227‐237, 1998.
 169. Swallow JG, Garland T Jr. Selection experiments as a tool in evolutionary and comparative physiology: Insights into complex traits ‐ an introduction to the symposium. Integr Comp Biol 45: 387‐390, 2005.
 170. Swallow JG, Rhodes JS, Garlant T Jr. Phenotypic and evolutionary plasticity in organ massed in response to voluntary exercise in house mice. Integr Comp Biol 45: 426‐437, 2005.
 171. Swallow JG, Wroblewska AK, Waters RP, Renner KJ, Britton SL, Koch LG. Phenotypic plasticity of body composition in rats selectively bred for high endurance capacity. J Appl Physiol 109: 778‐785, 2010.
 172. Sword GA, Simpson SJ, El Hadi OTM, Wilps H. Density‐dependent aposematism in the desert locust. Proc R Soc Lond B 267: 63‐68, 2000.
 173. Swynghedauw B. Phenotypic plasticity of adult myocardium: Molecular mechanisms. J Exp Biol 209: 2320‐2327, 2006.
 174. Timmerman CM, Chapman LJ. Behavioral and physiological compensation for chronic hypoxia in the sailfin molly (Poecilia latipinna). Physiol Biochem Zool 77: 601‐10, 2004.
 175. Tomy S, Ching Y, Chen Y, Cao J, Wang T, Chang C. Salinity effects on the expression of osmoregulatory genes in the euryhaline black porgy Acanthopagrus schlegeli. Gen Comp Endocrin 161: 123‐132, 2009.
 176. US Department of Health and Human Services. Physical Activity and Health: A Report of the Surgeon General. Atlanta, GA: U.S. Department of Health and Human Services, Centers for Disease Control and Prevention, National Center for Chronic Disease Prevention and Health Promotion, 1996.
 177. Van Veld PA, Nacci DE. Toxicity resistance. In: Di Giulio RT, Hinton DE, editors. The Toxicology of Fishes. Boca Raton, FL: Taylor and Francis, 2008.
 178.Via S, Gomulkiewicz R, de Jong G, Scheiner S, Schlichting CD, van Tienderen PH. Adaptive phenotypic plasticity: Consensus and controversy. Trends Ecol Evol 10: 212‐217, 1995.
 179. Wang X, Zhu H, Snieder H, Su S, Munn D, Harshfield G, Maria BL, Dong Y, Trieber F, Gutin B, Shi H. Obesity related methylation changes in DNA of peripheral blood leukocytes. BMC Medicine 8: 87‐94, 2010.
 180. Weismann A. Studies in the Theory of Descent, Vol 1. London: Sampon, Low, Marston, Searle and Rivington, 1882.
 181. West–Eberhard, MJ. Developmental Plasticity and Evolution, New York: Oxford University Press, 2002.
 182. Westerterp KR. Physical activity, food intake, and body weight regulation: Insights from doubly labeled water studies. Nutr Rev 68: 148‐154, 2010.
 183. Whitehead A. The evolutionary radiation of diverse osmotolerant physiologies in killifish (Fundulus sp). Evolution 64: 2070‐2085, 2010.
 184. Whitehead A, Galvez F, Zhang S, Williams LM, Oleksiak MF. Functional genomics of physiological plasticity and local adaptation in killifish. J Hered 102: 499‐511, 2011.
 185. Williams GC. Adaptation and Natural Selection. Princeton NJ: Princeton University Press, 1966.
 186. Wilson K, Thomas MB, Blanford S, Doggett M, Simpson SJ, Moore SL. Coping with crowds: Density‐dependent disease resistance in desert locusts. Proc Natl Acad Sci U S A 99: 5471‐5475, 2002.
 187. Wilson RS, Franklin CE. Testing the beneficial acclimation hypothesis. Trends Ecol Evol 17: 66‐70, 2002.
 188. Woltereck, R. Weitere experimentelle untersuchungen über artveränderung, speziellüber das wesen quantitativer artunterschiede bei daphnien. Verh Dtsch Zool Ges 19: 110–173, 1909.
 189. Wood CM, Marshall WS. Ion balance, acid‐base regulation, and chloride cell‐function in the common killifish, Fundulus heteroclitus – a euryhaline estuarine teleost. Estuaries 17: 34‐52, 1994.
 190. Woods HA. Evolution of homeostatic physiological systems In: Whitman DW, Ananthakrishnan TN, editors. Phenotypic Plasticity of Insects: Mechanisms and Consequences. Enfield, New Hampshire, USA: Science Publishers, 2009.
 191. World Health Organization. Obesity: Preventing and managing the global epidemic. World Health Organ Tech Rep Ser 1‐253, 2000.
 192. Zuk M, Stoehr AM. Immune defense and host life‐history. Am Nat 160: S9‐S22, 2002.

Related Articles:

Metabolic Flexibility: Hibernation, Torpor, and Estivation
Alterations in Myofibril Size and Structure During Growth, Exercise, and Changes in Environmental Temperature

Contact Editor

Submit a note to the editor about this article by filling in the form below.

* Required Field

How to Cite

Scott A. Kelly, Tami M. Panhuis, Andrew M. Stoehr. Phenotypic Plasticity: Molecular Mechanisms and Adaptive Significance. Compr Physiol 2012, 2: 1417-1439. doi: 10.1002/cphy.c110008