Comprehensive Physiology Wiley Online Library

Evolutionary Physiology and Genomics in the Highly Adaptable Killifish (Fundulus heteroclitus)

Full Article on Wiley Online Library



Abstract

By investigating evolutionary adaptations that change physiological functions, we can enhance our understanding of how organisms work, the importance of physiological traits, and the genes that influence these traits. This approach of investigating the evolution of physiological adaptation has been used with the teleost fish Fundulus heteroclitus and has produced insights into (i) how protein polymorphisms enhance swimming and development; (ii) the role of equilibrium enzymes in modulating metabolic flux; (iii) how variation in DNA sequences and mRNA expression patterns mitigate changes in temperature, pollution, and salinity; and (iv) the importance of nuclear‐mitochondrial genome interactions for energy metabolism. Fundulus heteroclitus provides so many examples of adaptive evolution because their local population sizes are large, they have significant standing genetic variation, and they experience large ranges of environmental conditions that enhance the likelihood that adaptive evolution will occur. Thus, F. heteroclitus research takes advantage of evolutionary changes associated with exposure to diverse environments, both across the North American Atlantic coast and within local habitats, to contrast neutral versus adaptive divergence. Based on evolutionary analyses contrasting neutral and adaptive evolution in F. heteroclitus populations, we conclude that adaptive evolution can occur readily and rapidly, at least in part because it depends on large amounts of standing genetic variation among many genes that can alter physiological traits. These observations of polygenic adaptation enhance our understanding of how evolution and physiological adaptation progresses, thus informing both biological and medical scientists about genotype‐phenotype relationships. © 2020 American Physiological Society. Compr Physiol 10:637‐671, 2020.

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. Demographic patterns among F. heteroclitus populations. (A) Neighbor‐joining tree based on microsatellites 6,60,200. (B) Maximum parsimony tree for RFLP in mitochondria, LDH‐B coding region, and 309 bp of Cytochrome B 19. Dark circles and squares are northern populations (north of Hudson River), unfilled are southern populations (south of Hudson River). (C) PCA and population structure (k = 9) based on 354 SNP, 30 individuals per population and principal component analysis 203,204.
Figure 2. Figure 2. Allozymes in F. heteroclitus. (A) Allelic variations of protein enzymes (allozymes) for three different genes: LDH‐B*, Malate dehydrogenase A (M, MDH‐A), and Isocitrate dehydrogenase‐B (I, IDH‐B) (redrawn based on data in Refs 148,159). The frequencies of the northern type alleles are plotted versus latitude along the eastern seacoast of North America (degrees latitude N). (B) Enzyme kinetics for the three genotypes of LDH‐B [catalytic turn‐over (kcat) divided by the Km] measured at different temperatures at neutral pH ([OH] = [H]) (redrawn based on data in Ref. 144). (C) Enzyme kinetics (kcat/Km) for the three LDH‐B genotypes plotted against temperature and pH (redrawn based on data in Ref. 144). (D) Hatching time at 20°C for fish from single populations for the three LDH‐B genotypes (redrawn based on data in Ref. 51). Hatching times were defined among 20 randomly crossed pairs, and larvae were genotyped. Data represent larvae genotypes. Similar results were obtained by 4 replicate mass crosses with 40 male and 40 female heterozygotes in each cross (n > 1000/cross). (E) Critical swimming speed (maximum sustainable swimming speeds in body lengths per second) for the two LDH‐B homozygotes; all fish were from the same population (drawn from data in Ref. 52).
Figure 3. Figure 3. Phylogenetic analyses of enzyme amounts. (A) Fundulus phylogeny 141. There are 15 taxa: 2 populations from 7 species and a single population from the outgroup. Boxes represent taxa with similar environmental temperature variation: Blue–geographic variation in temperature with northern taxa being colder. Red–lack of geographic variation in temperature. (B) Two phylogenetic methods for correcting for species similarities in enzyme amount among 15 Fundulus taxa versus naturally occurring mean annual environmental temperatures. Only the three enzymes were significantly related to environmental temperature after correcting for phylogeny: glyceraldehyde‐3‐phosphate dehydrogenase (GAPDH), pyruvate kinase (PYK), and LDH‐B. (C) F. heteroclitus glucose‐dependent metabolism versus multiple factor equation using three phylogenetically important enzymes (GAPDH, PYK, and LDH‐B). r2 = 0.866 (p < 0.005) 146.
Figure 4. Figure 4. Adaptive evolution of LDH‐B proximal promoter. (A) LDH‐B mRNA versus protein for northern (Maine) and southern (Georgia) individuals. When fish are acclimated to 20°C, increasing amounts of mRNA are associated with larger amounts of LDH‐B protein (measured as maximal enzyme activity (r2 = 0.81, p < 0.01, data from Ref. 168). (B) LDH‐B transcriptional binding sites and sequence variation. Functional sites are DNA sequences that bind protein transcriptional factors or affect transcription. (C) Individual promoter activity (line extending above columns are standard errors) defined by linking LDH‐B proximal promoter to luciferase reporter gene and transfected into rainbow trout liver cell line 46. Promoter activities from northern individuals are significantly greater than promoter activities from southern individuals (p < 0.001). (D) Promoter activity with different proximal promoter elements. Binding site 6fp (but not intervening sequence), and SP1 reduce expression, and without SP1, the northern promoter activity is no longer greater than southern promoter activity. (E) Evolutionary relationship using nonfunctional sequences among Fundulus species and within F. heteroclitus. (F) Evolutionary relationship using functional DNA sequences (affect promoter activity). Northern F. heteroclitus functional sequences are derived and significantly different from southern and F. grandis promoter DNA sequences. (G) Sliding window of DNA sequence variation within and between northern and southern F. heteroclitus. Data derived from Refs 46,169.
Figure 5. Figure 5. Microarray: genome‐wide patterns of mRNA expression. (A) Heat map of adaptively significant mRNA expression. Red and green colors are the relative low or high expression. Notice northern individuals share similar expression patterns and are significantly different from southern F. heteroclitus and F. grandis. (B) Volcano plot: log2 expression relative to mean expression for each mRNA versus statistical significance as −log10 p‐values. Gray box highlights the most significant mRNA where northern mRNA (blue circle) is statistically larger than both southern F. heteroclitus (red square) and F. grandis (green circle). Redrawn from Ref. 133.
Figure 6. Figure 6. Clinal adaptive variation in mRNA expression. (A) Five sample sites and their phylogenetic relationship microsatellite‐derived neighbor‐joining tree with median annual temperatures (°C) averaged over 30 years. Branching pattern is a neighbor‐joining tree constructed from pairwise Cavalli‐Sforza and Edwards' chord distances 33 calculated from microsatellite allele frequencies. (B) Relationships between phylogenetic and ecological effects on variation in gene expression. For each gene, the explained variation (r2) for phylogeny [genetic distance based on Cavalli‐Sforza and Edwards' chord distances 33 calculated from microsatellite allele frequencies] versus the explained variation (r2) for habitat temperatures. Venn diagram is for the numbers of genes that have significant regression with habitat temperature (orange), phylogeny (green), or both temperature and phylogeny (blue). Colors of spots in the graph correspond to Venn diagram. Enlarged spots are the 13 genes that regress significantly with habitat temperature after correcting for phylogeny (red circle; Venn diagram) using the phylogenetic generalized least squares (PGLS) approach, and thus appear to be evolving by natural selection. (C) Variation in mRNA expression within or among populations. Plotted are the log of variation. Ratio of variation is indicative of evolutionary processes (directional, stabilizing, balancing, or neutral). Redrawn from Ref. 200.
Figure 7. Figure 7. mRNA expression and cardiac metabolism. (A) Relative levels of cardiac metabolism for 16 individuals (8 per Maine and Georgia population). Cardiac metabolism was measured using glucose, fatty acid, and LKA (lactate, ketones, and ethanol) as substrates 136. Red is at least 1.75‐fold greater and green is at least 1.75‐fold lower than the overall mean. (B) Significant mRNA expression differences between individuals within a population (negative log10 values, thus 2 is equal to a p‐value of 1%) versus the fold difference (log2 values, thus 1 = twofold difference). Fold differences are relative to the overall mean for each mRNA. Green background shadowing shows mRNA with 1.5‐fold or less differences. p‐Values are truncated at values more than 10−17. (C) Patterns of mRNA expression among all 16 individuals (green is relatively low, red is relatively high). A subset of mRNAs coding for metabolic genes that show shared expression within groups that is significantly different among groups. (D) Fatty acid metabolic rates are relative to the mRNA expression. mRNA expression summarized as one of three primary biochemical pathways (two principal components each for glycolysis, TCA cycle, and oxidative phosphorylation). Similar patterns occur for glucose and LKA supported cardiac metabolism 136.
Figure 8. Figure 8. Local osmotic adaptation. F. heteroclitus population variation along a salinity gradient in the Chesapeake Bay. (A) Map of salinity gradient in the Chesapeake Bay, where experiments contrasted physiology and genomics of marine‐native (M), brackish‐native (BW), and freshwater‐native (FW) populations. (B) Plot of genetic similarity of individuals collected from the three Chesapeake populations, where neighboring populations were equally genetically distant from each other. (C) Principal component analysis of genes that are differentially expressed between populations but not affected by salinity challenge. Genes where the pattern of population divergence matches the neutral expectation [e.g. as established by pattern of genetic relatedness shown in (B)] are included in the left panel and genes where the patterns of population divergences consistent with adaptation in the freshwater population (blue) are included in the right panel. Pie chart shows the proportion of genes within this set that show the neutral or adaptive pattern. (D) Principal component analysis of genes that are differentially expressed between populations and that are differentially expressed during salinity challenge. Genes where the pattern of population divergence matches the neutral expectation [e.g. as established by the pattern of genetic relatedness shown in (B)] are included in the left panel and genes where the patterns of population divergence consistent with adaptation in the freshwater population (blue) are included in the right panel. Pie chart shows the proportion of genes within this set that show the neutral or adaptive pattern. A greater proportion of genes that are transcriptionally responsive to salinity show the adaptive pattern than genes that are not responsive to salinity. Principal component analyses are redrawn from Ref. 202. Salinity gradient heatmap of the Chesapeake Bay was generated from the NOAA Chesapeake Bay Operational Forecast System (https://tidesandcurrents.noaa.gov/).
Figure 9. Figure 9. Local adaptation to warmer temperatures. Three populations (triads) were examined: a northern and southern reference population and a locally heated thermal effluent (TE) population. (A) Genetic structure among Oyster Creek TE site using all approximately 5400 SNPs. X and Y axes are the first and second principal components (linear equation maximizing the variation among populations). The first principal component separates all three sites, and the second separates the TE site (red) from both northern and southern reference sites. (B) Outlier SNPs with statistically large and unexpected FST values for paired comparisons between TE and references. SNPs evolving by natural selection are the 94 SNPs where TE differs from both reference populations but are not different between the pair of reference populations. (C) Structure plots using 94 outlier SNPs for 2, 3, or 4 groups of individuals (k = 2, 3, or 4). TE site is distinct in all comparisons. (D) Linkage disequilibrium as indicated by similar FST values relative to the DNA distance (base pair, bp). Dashed line is the mean, and shading is the 95% confidence bounds for the mean genome‐wide FST value estimate for both TE versus reference comparisons. Red is the decline in FST value for outlier SNPs, and blue is the mean FST value when TE and references are randomly permutated.
Figure 10. Figure 10. Rapid local adaptation to pollution. Analyses of polluted populations using changes in allele frequencies for 354 SNPs defined by mass spectrophotometry. (A) In each of three comparisons a polluted population (P) was compared to two clean reference populations (C). The Venn diagram for these three triads (C‐P‐C) was used to identify statistically significant SNPs based on an outlier test (Outlier, unexpectedly large FST), environmental association of SNPs (Assoc.), and changes in minor allele frequencies (MAF). The red number is the number of SNPs that are significant in all three tests. (B) Maximum parsimony tree of the 24 CYP1A promoter‐intron sequences use to test the effect of DNA sequence variation on gene expression. Blue highlights are sequences from polluted New Bedford populations. Red stars are for sequences with derived outlier SNP. (C) Induction of gene expression with exposure to persistent organic pollutants (POP) in cells in culture for CYP1A promoter from polluted and clean populations. (D) Average pollution‐induced gene expression from CYP1A promoter from the two clean (green and blue) and the polluted New Bedford population (red). Letters represent post‐hoc analysis indicating that the polluted New Bedford is significantly different from both clean populations, and there is no significant difference between the clean populations.
Figure 11. Figure 11. Genomics of adaptation to recent anthropogenic pollution. (A) Four pairs of populations were sampled: for each pair, one population inhabits highly polluted marine environments and individuals are tolerant to POPs (T), and the second population is in a clean, nonpolluted reference site and individuals are sensitive to POPs (S). (B) Pairs of mRNA expression for controls and POP exposure among tolerant (T) and sensitive (S) populations. Each population has mRNA expression for two sets of conditions: control and exposure to POP. In each row is the relative expression of an mRNA, with high expression as bright yellow. The lower panel highlights genes activated by ligand‐bound AHR protein. (C) Diagram of AHR signaling pathway including co‐regulators and transcriptional targets. Color boxes are color coded for location defined in (A). Filled boxes are genes identified as evolving by natural selection. (D) FST values and π (pi, nucleotide diversity) between tolerant (T) and sensitive (S) populations. Gray shading highlights DNA sequences with unusually large significant FST values, extreme pi values, or both. These regions of the genome also have significant Tajima's D (not shown).
Figure 12. Figure 12. Epistatic adaptive evolution. (A) Oxidative Phosphorylation pathway and the number of protein subunits encoded by mitochondrial and nuclear genomes 95. (B) Mitochondrial OxPhos dependent respiration (State 3) measured in Fundulus heteroclitus from a single New Jersey population. Individuals were acclimated to either 12 or 28°C and had either the northern or southern mitochondrial haplotype. Acclimation, acute (assay temperature), and mitochondrial effects were all significant. (C) Distribution of wFST values for 11,705 nuclear SNPs calculated between the two mitochondrial haplotypes within the single population. Plot contains wFST values and corresponding negative log10 p‐values (e.g. −log10(0.01) = 2). Blue values are significant with a p‐value less than 0.01, green values are significant with a 1% FDR correction, and purple values are significant with a Bonferroni correction. Histograms show wFST and p‐value distributions. (D) Mitochondrial OxPhos dependent respiration (State 3) as a function of the fraction of southern nuclear alleles. State 3 is the residual from a mixed model with body mass, acclimation, and assay temperatures. Individuals with greater than 75% northern nuclear alleles and the northern mitochondria are blue, individuals with less than 75% northern nuclear alleles, and the northern mitochondria are green. Individuals with less than 75% southern nuclear alleles and with southern mitochondria are orange. Individuals with greater than 75% southern alleles and the southern mitochondria are red.
Figure 13. Figure 13. Fine‐scale evolution among microhabitats. (A) Three New Jersey saltmarsh estuaries (Mantoloking, Rutgers University Marine Field Station, Stone Harbor) and an enlarged image of Rutgers University Marine Field Station with three microhabitats Basin (B), Creek (C), and Pond (P). The distance between microhabitats was never greater than 200 m and usually less than 50 m. (B) Evolutionary analyses among microhabitats for three populations, where each population has three analyses: (1) SNPs with significantly different FST values, (2) Lositan identified significant outlier SNPs, and (3) Arlequin identified significant outlier SNPs. Significant SNPs detected in all three analyses with joint FDR less than 1% were considered outlier SNPs. (C) Density of FST values within each of the three New Jersey saltmarsh populations. Plotted are large significant outlier SNPs (blue), 4352 nonoutlier SNPs (gold), and SNPs when population assignment is randomly permuted among microhabitats (red). (D) Density of outlier‐SNP FST values within and among populations. Significant outlier SNP‐specific FST values for within Rutgers University Marine Field Station (blue) and between Rutgers Marine Station and Stone Harbor (gold) or Mantoloking (red).


Figure 1. Demographic patterns among F. heteroclitus populations. (A) Neighbor‐joining tree based on microsatellites 6,60,200. (B) Maximum parsimony tree for RFLP in mitochondria, LDH‐B coding region, and 309 bp of Cytochrome B 19. Dark circles and squares are northern populations (north of Hudson River), unfilled are southern populations (south of Hudson River). (C) PCA and population structure (k = 9) based on 354 SNP, 30 individuals per population and principal component analysis 203,204.


Figure 2. Allozymes in F. heteroclitus. (A) Allelic variations of protein enzymes (allozymes) for three different genes: LDH‐B*, Malate dehydrogenase A (M, MDH‐A), and Isocitrate dehydrogenase‐B (I, IDH‐B) (redrawn based on data in Refs 148,159). The frequencies of the northern type alleles are plotted versus latitude along the eastern seacoast of North America (degrees latitude N). (B) Enzyme kinetics for the three genotypes of LDH‐B [catalytic turn‐over (kcat) divided by the Km] measured at different temperatures at neutral pH ([OH] = [H]) (redrawn based on data in Ref. 144). (C) Enzyme kinetics (kcat/Km) for the three LDH‐B genotypes plotted against temperature and pH (redrawn based on data in Ref. 144). (D) Hatching time at 20°C for fish from single populations for the three LDH‐B genotypes (redrawn based on data in Ref. 51). Hatching times were defined among 20 randomly crossed pairs, and larvae were genotyped. Data represent larvae genotypes. Similar results were obtained by 4 replicate mass crosses with 40 male and 40 female heterozygotes in each cross (n > 1000/cross). (E) Critical swimming speed (maximum sustainable swimming speeds in body lengths per second) for the two LDH‐B homozygotes; all fish were from the same population (drawn from data in Ref. 52).


Figure 3. Phylogenetic analyses of enzyme amounts. (A) Fundulus phylogeny 141. There are 15 taxa: 2 populations from 7 species and a single population from the outgroup. Boxes represent taxa with similar environmental temperature variation: Blue–geographic variation in temperature with northern taxa being colder. Red–lack of geographic variation in temperature. (B) Two phylogenetic methods for correcting for species similarities in enzyme amount among 15 Fundulus taxa versus naturally occurring mean annual environmental temperatures. Only the three enzymes were significantly related to environmental temperature after correcting for phylogeny: glyceraldehyde‐3‐phosphate dehydrogenase (GAPDH), pyruvate kinase (PYK), and LDH‐B. (C) F. heteroclitus glucose‐dependent metabolism versus multiple factor equation using three phylogenetically important enzymes (GAPDH, PYK, and LDH‐B). r2 = 0.866 (p < 0.005) 146.


Figure 4. Adaptive evolution of LDH‐B proximal promoter. (A) LDH‐B mRNA versus protein for northern (Maine) and southern (Georgia) individuals. When fish are acclimated to 20°C, increasing amounts of mRNA are associated with larger amounts of LDH‐B protein (measured as maximal enzyme activity (r2 = 0.81, p < 0.01, data from Ref. 168). (B) LDH‐B transcriptional binding sites and sequence variation. Functional sites are DNA sequences that bind protein transcriptional factors or affect transcription. (C) Individual promoter activity (line extending above columns are standard errors) defined by linking LDH‐B proximal promoter to luciferase reporter gene and transfected into rainbow trout liver cell line 46. Promoter activities from northern individuals are significantly greater than promoter activities from southern individuals (p < 0.001). (D) Promoter activity with different proximal promoter elements. Binding site 6fp (but not intervening sequence), and SP1 reduce expression, and without SP1, the northern promoter activity is no longer greater than southern promoter activity. (E) Evolutionary relationship using nonfunctional sequences among Fundulus species and within F. heteroclitus. (F) Evolutionary relationship using functional DNA sequences (affect promoter activity). Northern F. heteroclitus functional sequences are derived and significantly different from southern and F. grandis promoter DNA sequences. (G) Sliding window of DNA sequence variation within and between northern and southern F. heteroclitus. Data derived from Refs 46,169.


Figure 5. Microarray: genome‐wide patterns of mRNA expression. (A) Heat map of adaptively significant mRNA expression. Red and green colors are the relative low or high expression. Notice northern individuals share similar expression patterns and are significantly different from southern F. heteroclitus and F. grandis. (B) Volcano plot: log2 expression relative to mean expression for each mRNA versus statistical significance as −log10 p‐values. Gray box highlights the most significant mRNA where northern mRNA (blue circle) is statistically larger than both southern F. heteroclitus (red square) and F. grandis (green circle). Redrawn from Ref. 133.


Figure 6. Clinal adaptive variation in mRNA expression. (A) Five sample sites and their phylogenetic relationship microsatellite‐derived neighbor‐joining tree with median annual temperatures (°C) averaged over 30 years. Branching pattern is a neighbor‐joining tree constructed from pairwise Cavalli‐Sforza and Edwards' chord distances 33 calculated from microsatellite allele frequencies. (B) Relationships between phylogenetic and ecological effects on variation in gene expression. For each gene, the explained variation (r2) for phylogeny [genetic distance based on Cavalli‐Sforza and Edwards' chord distances 33 calculated from microsatellite allele frequencies] versus the explained variation (r2) for habitat temperatures. Venn diagram is for the numbers of genes that have significant regression with habitat temperature (orange), phylogeny (green), or both temperature and phylogeny (blue). Colors of spots in the graph correspond to Venn diagram. Enlarged spots are the 13 genes that regress significantly with habitat temperature after correcting for phylogeny (red circle; Venn diagram) using the phylogenetic generalized least squares (PGLS) approach, and thus appear to be evolving by natural selection. (C) Variation in mRNA expression within or among populations. Plotted are the log of variation. Ratio of variation is indicative of evolutionary processes (directional, stabilizing, balancing, or neutral). Redrawn from Ref. 200.


Figure 7. mRNA expression and cardiac metabolism. (A) Relative levels of cardiac metabolism for 16 individuals (8 per Maine and Georgia population). Cardiac metabolism was measured using glucose, fatty acid, and LKA (lactate, ketones, and ethanol) as substrates 136. Red is at least 1.75‐fold greater and green is at least 1.75‐fold lower than the overall mean. (B) Significant mRNA expression differences between individuals within a population (negative log10 values, thus 2 is equal to a p‐value of 1%) versus the fold difference (log2 values, thus 1 = twofold difference). Fold differences are relative to the overall mean for each mRNA. Green background shadowing shows mRNA with 1.5‐fold or less differences. p‐Values are truncated at values more than 10−17. (C) Patterns of mRNA expression among all 16 individuals (green is relatively low, red is relatively high). A subset of mRNAs coding for metabolic genes that show shared expression within groups that is significantly different among groups. (D) Fatty acid metabolic rates are relative to the mRNA expression. mRNA expression summarized as one of three primary biochemical pathways (two principal components each for glycolysis, TCA cycle, and oxidative phosphorylation). Similar patterns occur for glucose and LKA supported cardiac metabolism 136.


Figure 8. Local osmotic adaptation. F. heteroclitus population variation along a salinity gradient in the Chesapeake Bay. (A) Map of salinity gradient in the Chesapeake Bay, where experiments contrasted physiology and genomics of marine‐native (M), brackish‐native (BW), and freshwater‐native (FW) populations. (B) Plot of genetic similarity of individuals collected from the three Chesapeake populations, where neighboring populations were equally genetically distant from each other. (C) Principal component analysis of genes that are differentially expressed between populations but not affected by salinity challenge. Genes where the pattern of population divergence matches the neutral expectation [e.g. as established by pattern of genetic relatedness shown in (B)] are included in the left panel and genes where the patterns of population divergences consistent with adaptation in the freshwater population (blue) are included in the right panel. Pie chart shows the proportion of genes within this set that show the neutral or adaptive pattern. (D) Principal component analysis of genes that are differentially expressed between populations and that are differentially expressed during salinity challenge. Genes where the pattern of population divergence matches the neutral expectation [e.g. as established by the pattern of genetic relatedness shown in (B)] are included in the left panel and genes where the patterns of population divergence consistent with adaptation in the freshwater population (blue) are included in the right panel. Pie chart shows the proportion of genes within this set that show the neutral or adaptive pattern. A greater proportion of genes that are transcriptionally responsive to salinity show the adaptive pattern than genes that are not responsive to salinity. Principal component analyses are redrawn from Ref. 202. Salinity gradient heatmap of the Chesapeake Bay was generated from the NOAA Chesapeake Bay Operational Forecast System (https://tidesandcurrents.noaa.gov/).


Figure 9. Local adaptation to warmer temperatures. Three populations (triads) were examined: a northern and southern reference population and a locally heated thermal effluent (TE) population. (A) Genetic structure among Oyster Creek TE site using all approximately 5400 SNPs. X and Y axes are the first and second principal components (linear equation maximizing the variation among populations). The first principal component separates all three sites, and the second separates the TE site (red) from both northern and southern reference sites. (B) Outlier SNPs with statistically large and unexpected FST values for paired comparisons between TE and references. SNPs evolving by natural selection are the 94 SNPs where TE differs from both reference populations but are not different between the pair of reference populations. (C) Structure plots using 94 outlier SNPs for 2, 3, or 4 groups of individuals (k = 2, 3, or 4). TE site is distinct in all comparisons. (D) Linkage disequilibrium as indicated by similar FST values relative to the DNA distance (base pair, bp). Dashed line is the mean, and shading is the 95% confidence bounds for the mean genome‐wide FST value estimate for both TE versus reference comparisons. Red is the decline in FST value for outlier SNPs, and blue is the mean FST value when TE and references are randomly permutated.


Figure 10. Rapid local adaptation to pollution. Analyses of polluted populations using changes in allele frequencies for 354 SNPs defined by mass spectrophotometry. (A) In each of three comparisons a polluted population (P) was compared to two clean reference populations (C). The Venn diagram for these three triads (C‐P‐C) was used to identify statistically significant SNPs based on an outlier test (Outlier, unexpectedly large FST), environmental association of SNPs (Assoc.), and changes in minor allele frequencies (MAF). The red number is the number of SNPs that are significant in all three tests. (B) Maximum parsimony tree of the 24 CYP1A promoter‐intron sequences use to test the effect of DNA sequence variation on gene expression. Blue highlights are sequences from polluted New Bedford populations. Red stars are for sequences with derived outlier SNP. (C) Induction of gene expression with exposure to persistent organic pollutants (POP) in cells in culture for CYP1A promoter from polluted and clean populations. (D) Average pollution‐induced gene expression from CYP1A promoter from the two clean (green and blue) and the polluted New Bedford population (red). Letters represent post‐hoc analysis indicating that the polluted New Bedford is significantly different from both clean populations, and there is no significant difference between the clean populations.


Figure 11. Genomics of adaptation to recent anthropogenic pollution. (A) Four pairs of populations were sampled: for each pair, one population inhabits highly polluted marine environments and individuals are tolerant to POPs (T), and the second population is in a clean, nonpolluted reference site and individuals are sensitive to POPs (S). (B) Pairs of mRNA expression for controls and POP exposure among tolerant (T) and sensitive (S) populations. Each population has mRNA expression for two sets of conditions: control and exposure to POP. In each row is the relative expression of an mRNA, with high expression as bright yellow. The lower panel highlights genes activated by ligand‐bound AHR protein. (C) Diagram of AHR signaling pathway including co‐regulators and transcriptional targets. Color boxes are color coded for location defined in (A). Filled boxes are genes identified as evolving by natural selection. (D) FST values and π (pi, nucleotide diversity) between tolerant (T) and sensitive (S) populations. Gray shading highlights DNA sequences with unusually large significant FST values, extreme pi values, or both. These regions of the genome also have significant Tajima's D (not shown).


Figure 12. Epistatic adaptive evolution. (A) Oxidative Phosphorylation pathway and the number of protein subunits encoded by mitochondrial and nuclear genomes 95. (B) Mitochondrial OxPhos dependent respiration (State 3) measured in Fundulus heteroclitus from a single New Jersey population. Individuals were acclimated to either 12 or 28°C and had either the northern or southern mitochondrial haplotype. Acclimation, acute (assay temperature), and mitochondrial effects were all significant. (C) Distribution of wFST values for 11,705 nuclear SNPs calculated between the two mitochondrial haplotypes within the single population. Plot contains wFST values and corresponding negative log10 p‐values (e.g. −log10(0.01) = 2). Blue values are significant with a p‐value less than 0.01, green values are significant with a 1% FDR correction, and purple values are significant with a Bonferroni correction. Histograms show wFST and p‐value distributions. (D) Mitochondrial OxPhos dependent respiration (State 3) as a function of the fraction of southern nuclear alleles. State 3 is the residual from a mixed model with body mass, acclimation, and assay temperatures. Individuals with greater than 75% northern nuclear alleles and the northern mitochondria are blue, individuals with less than 75% northern nuclear alleles, and the northern mitochondria are green. Individuals with less than 75% southern nuclear alleles and with southern mitochondria are orange. Individuals with greater than 75% southern alleles and the southern mitochondria are red.


Figure 13. Fine‐scale evolution among microhabitats. (A) Three New Jersey saltmarsh estuaries (Mantoloking, Rutgers University Marine Field Station, Stone Harbor) and an enlarged image of Rutgers University Marine Field Station with three microhabitats Basin (B), Creek (C), and Pond (P). The distance between microhabitats was never greater than 200 m and usually less than 50 m. (B) Evolutionary analyses among microhabitats for three populations, where each population has three analyses: (1) SNPs with significantly different FST values, (2) Lositan identified significant outlier SNPs, and (3) Arlequin identified significant outlier SNPs. Significant SNPs detected in all three analyses with joint FDR less than 1% were considered outlier SNPs. (C) Density of FST values within each of the three New Jersey saltmarsh populations. Plotted are large significant outlier SNPs (blue), 4352 nonoutlier SNPs (gold), and SNPs when population assignment is randomly permuted among microhabitats (red). (D) Density of outlier‐SNP FST values within and among populations. Significant outlier SNP‐specific FST values for within Rutgers University Marine Field Station (blue) and between Rutgers Marine Station and Stone Harbor (gold) or Mantoloking (red).
References
 1.Able KW, Balletto JH, Hagan SM, Jivoff PR, Strait K. Linkages between salt marshes and other nekton habitats in Delaware Bay, USA. Rev Fish Sci 15: 1‐61, 2007.
 2.Able KW, Castagna M. Aspects of an undescribed reproductive behavior in Fundulus heteroclitus (Pisces: Cyprinodontae) from Virginia. Chesapeake Sci 16: 282, 1975.
 3.Able KW, Felley JD. Geographical variation in Fundulus‐heteroclitus – Tests for concordance between egg and adult morphologies. Am Zool 26: 145‐157, 1986.
 4.Able KW, Hagan SM, Brown SA. Habitat use, movement, and growth of young‐of‐the‐year Fundulus spp. in southern New Jersey salt marshes: Comparisons based on tag/recapture. J Exp Mar Biol Ecol 335: 177‐187, 2006.
 5.Able KW, Vivian DN, Petruzzelli G, Hagan SM. Connectivity among salt marsh subhabitats: Residency and movements of the mummichog (Fundulus heteroclitus). Estuaries Coast 35: 743‐753, 2012.
 6.Adams SM, Lindmeier JB, Duvernell DD. Microsatellite analysis of the phylogeography, Pleistocene history and secondary contact hypotheses for the killifish, Fundulus heteroclitus. Mol Ecol 15: 1109‐1123, 2006.
 7.Airaksinen R, Rantakokko P, Eriksson JG, Blomstedt P, Kajantie E, Kiviranta H. Association between type 2 diabetes and exposure to persistent organic pollutants. Diabetes Care 34: 1972‐1979, 2011.
 8.Anholt RRH, Mackay TFC. The road less traveled: From genotype to phenotype in flies and humans. Mamm Genome 29: 5‐23, 2018.
 9.Arnold SJ. Morphology, performance and fitness. Am Zool 23: 347‐361, 1983.
 10.Arnold SJ. Genetic correlation and the evolution of physiology. In: Feder ME, Bennett AE, Burggren WW, Huey RB, editors. New Directions in Ecological Physiology. Cambridge, New York, etc: Cambridge University Press, 1987, p. 189‐215.
 11.Arnot JA, Armitage JM, McCarty LS, Wania F, Cousins IT, Toose‐Reid L. Toward a consistent evaluative framework for POP risk characterization. Environ Sci Technol 45: 97‐103, 2011.
 12.Baris TZ, Blier PU, Pichaud N, Crawford DL, Oleksiak MF. Gene by environmental interactions affecting oxidative phosphorylation and thermal sensitivity. Am J Physiol Regul Integr Comp Physiol 311: R157‐R165, 2016.
 13.Baris TZ, Crawford DL, Oleksiak MF. Acclimation and acute temperature effects on population differences in oxidative phosphorylation. Am J Physiol Regul Integr Comp Physiol 310: R185‐R196, 2016.
 14.Baris TZ, Wagner DN, Dayan DI, Du X, Blier PU, Pichaud N, Oleksiak MF, Crawford DL. Evolved genetic and phenotypic differences due to mitochondrial‐nuclear interactions. PLoS Genet 13: e1006517, 2017.
 15.Bartholomew GA. Interspecific comparison as a tool for ecological physiologists. In: Feder ME, Bennett AF, Burggren WW, Huey RB, editors. New Directions in Ecological Physiology. Cambridge, New York, etc: Cambridge University Press, 1987, p. 11‐37.
 16.Bello SM, Franks DG, Stegeman JJ, Hahn ME. Acquired resistance to Ah receptor agonists in a population of Atlantic killifish (Fundulus heteroclitus) inhabiting a marine superfund site: In vivo and in vitro studies on the inducibility of xenobiotic metabolizing enzymes. Toxicol Sci 60: 77‐91, 2001.
 17.Bennett AF, Huey RB. Studying the evolution of physiological performance. Oxford Surv Evol Biol 7: 251‐284, 1990.
 18.Berg JJ, Coop G. A population genetic signal of polygenic adaptation. PLoS Genet 10: e1004412, 2014.
 19.Bernardi G, Sordino P, Powers DA. Concordant mitochondrial and nuclear DNA phylogenies for populations of the teleost fish Fundulus heteroclitus. Proc Natl Acad Sci U S A 90: 9271‐9274, 1993.
 20.Betancur RR, Orti G, Pyron RA. Fossil‐based comparative analyses reveal ancient marine ancestry erased by extinction in ray‐finned fishes. Ecol Lett 18: 441‐450, 2015.
 21.Bloom AL. Sea level and coastal changes. In: Wright WEJ, editor. The Holocene. Minneapolis: University of Minnesota, 1983, p. 26‐41.
 22.Bloom AL. Sea level and coastal morphology of the United States through the late Wisconsin glacial maximum. In: Porter SC, editor. The Late Pleistocene. Minneapolis: University of Minnesota, 1983, p. 215‐229.
 23.Boyle EA, Li YI, Pritchard JK. An expanded view of complex traits: From polygenic to omnigenic. Cell 169: 1177‐1186, 2017.
 24.Bozinovic G, Oleksiak MF. Genomic approaches with natural fish populations from polluted environments. Environ Toxicol Chem 30: 283‐289, 2011.
 25.Bozinovic G, Sit TL, Di Giulio R, Wills LF, Oleksiak MF. Genomic and physiological responses to strong selective pressure during late organogenesis: Few gene expression changes found despite striking morphological differences. BMC Genomics 14: 779, 2013.
 26.Bozinovic G, Sit TL, Hinton DE, Oleksiak MF. Gene expression throughout a vertebrate's embryogenesis. BMC Genomics 12: 132, 2011.
 27.Brennan RS, Healy TM, Bryant HJ, La MV, Schulte PM, Whitehead A. Genome‐wide selection scans integrated with association mapping reveal mechanisms of physiological adaptation across a salinity gradient in killifish. bioRxiv, 2018. DOI: 10.1101/254854.
 28.Brennan RS, Hwang R, Tse M, Fangue NA, Whitehead A. Local adaptation to osmotic environment in killifish, Fundulus heteroclitus, is supported by divergence in swimming performance but not by differences in excess post‐exercise oxygen consumption or aerobic scope. Comp Biochem Physiol A Mol Integr Physiol 196: 11‐19, 2016.
 29.Brieuc MSO, Waters CD, Drinan DP, Naish KA. A practical introduction to Random Forest for genetic association studies in ecology and evolution. Mol Ecol Resour 18: 755‐766, 2018.
 30.Brown PO, Botstein D. Exploring the new world of the genome with DNA microarrays. Nat Genet 21: 33‐37, 1999.
 31.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 genomics. Comp Biochem Physiol Part D Genomics Proteomics 2: 257‐286, 2007.
 32.Cashon RE, Van Beneden RJ, Powers DA. Biochemical genetics of Fundulus heteroclitus (L.). IV. Spatial variation in gene frequencies of Idh‐A, Idh‐B, 6‐Pgdh‐A, and Est‐S. Biochem Genet 19: 715‐728, 1981.
 33.Cavalli‐Sforza LL, Edwards AWF. Phylogenetic analysis: Models and estimation procedures. Evolution 21: 550‐570, 1967.
 34.Charlesworth B. Causes of natural variation in fitness: Evidence from studies of Drosophila populations. Proc Natl Acad Sci U S A 112: 1662‐1669, 2015.
 35.Chung DJ, Bryant HJ, Schulte PM. Thermal acclimation and subspecies‐specific effects on heart and brain mitochondrial performance in a eurythermal teleost (Fundulus heteroclitus). J Exp Biol 220: 1459‐1471, 2017.
 36.Chung DJ, Morrison PR, Bryant HJ, Jung E, Brauner CJ, Schulte PM. Intraspecific variation and plasticity in mitochondrial oxygen binding affinity as a response to environmental temperature. Sci Rep 7: 16238, 2017.
 37.Churchill GA. Fundamentals of experimental design for cDNA microarrays. Nat Genet 32 (Suppl): 490‐495, 2002.
 38.Clark BW, Bone AJ, Di Giulio RT. Resistance to teratogenesis by F1 and F2 embryos of PAH‐adapted Fundulus heteroclitus is strongly inherited despite reduced recalcitrance of the AHR pathway. Environ Sci Pollut Res Int 21: 13898‐13908, 2014.
 39.Cline RM, editor. Late Quaternary Paleoceanography and Paleoclimatology. Boulder: Geological Society of America, 1976.
 40.Crawford DL, Oleksiak MF. The biological importance of measuring individual variation. J Exp Biol 210: 1613‐1621, 2007.
 41.Crawford DL, Pierce VA, Segal JA. Evolutionary physiology of closely related taxa: Analyses of enzyme expression. Am Zool 39: 389‐400, 1999.
 42.Crawford DL, Place AR, Powers DA. Clinal variation in the specific activity of lactate dehydrogenase‐B. J Exp Zool 255: 110‐113, 1990.
 43.Crawford DL, Powers DA. Molecular‐basis of evolutionary adaptation at the lactate dehydrogenase‐B locus in the fish Fundulus‐heteroclitus. Proc Natl Acad Sci U S A 86: 9365‐9369, 1989.
 44.Crawford DL, Powers DA. Molecular adaptation to the thermal environment: Genetic and physiological mechanisms. In: Clegg MT, O'Brien SJ, editors. Molecular Evolution: Proceeding of a UCLA Colloquium 1989. New York: Wiley‐Liss, 1990, p. 213‐222.
 45.Crawford DL, Powers DA. Evolutionary adaptation to different thermal environments via transcriptional regulation. Mol Biol Evol 9: 806‐813, 1992.
 46.Crawford DL, Segal JA, Barnett JL. Evolutionary analysis of TATA‐less proximal promoter function. Mol Biol Evol 16: 194‐207, 1999.
 47.Crow JF. Some possibilities for measuring selection intensities in man. Am Anthropol 60: 1‐13, 1958.
 48.Dayan DI, Crawford DL, Oleksiak MF. Phenotypic plasticity in gene expression contributes to divergence of locally adapted populations of Fundulus heteroclitus. Mol Ecol 24: 3345‐3359, 2015.
 49.Dayan DI, Du X, Baris TZ, Wagner DN, Crawford DL, Oleksiak MF. Population genomics of rapid evolution in natural populations: Polygenic selection in response to power station thermal effluents. BMC Evol Biol 19: 61, 2019.
 50.DiMichele L, Paynter KT, Powers DA. Evidence of lactate dehydrogenase‐B allozyme effects in the teleost, Fundulus heteroclitus. Science 253: 898‐900, 1991.
 51.DiMichele L, Powers DA. LDH‐B genotype‐specific hatching times of Fundulus heteroclitus embryos. Nature 296: 563‐564, 1982.
 52.Dimichele L, Powers DA. Physiological‐basis for swimming endurance differences between Ldh‐B genotypes of Fundulus‐heteroclitus. Science 216: 1014‐1016, 1982.
 53.Dimichele L, Powers DA. Developmental and oxygen‐consumption rate differences between lactate dehydrogenase‐B genotypes of Fundulus‐heteroclitus and their effect on hatching time. Physiol Zool 57: 52‐56, 1984.
 54.Dimichele L, Powers DA. Allozyme variation, developmental rate, and differential mortality in the teleost Fundulus‐heteroclitus. Physiol Zool 64: 1426‐1443, 1991.
 55.DiMichele L, Westerman ME. Geographic variation in development rate between populations of the teleost Fundulus heteroclitus. Mar Biol 128: 1‐7, 1997.
 56.Donnelly JP, Driscoll NW, Uchupi E, Keigwin LD, Schwab WC, Thieler ER, Swift SA. Catastrophic meltwater discharge down the Hudson Valley: A potential trigger for the Intra‐Allerod cold period. Geology 33: 89‐92, 2005.
 57.Du X, Crawford DL, Nacci DE, Oleksiak MF. Heritable oxidative phosphorylation differences in a pollutant resistant Fundulus heteroclitus population. Aquat Toxicol 177: 44‐50, 2016.
 58.Du X, Crawford DL, Oleksiak MF. Effects of anthropogenic pollution on the oxidative phosphorylation pathway of hepatocytes from natural populations of Fundulus heteroclitus. Aquat Toxicol 165: 231‐240, 2015.
 59.Duforet‐Frebourg N, Luu K, Laval G, Bazin E, Blum MG. Detecting genomic signatures of natural selection with principal component analysis: Application to the 1000 genomes data. Mol Biol Evol 33: 1082‐1093, 2016.
 60.Duvernell DD, Lindmeier JB, Faust KE, Whitehead A. Relative influences of historical and contemporary forces shaping the distribution of genetic variation in the Atlantic killifish, Fundulus heteroclitus. Mol Ecol 17: 1344‐1360, 2008.
 61.Eisen MB, Brown PO. DNA arrays for analysis of gene expression. Methods Enzymol 303: 179‐205, 1999.
 62.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.
 63.Fader KA, Zacharewski TR. Beyond the aryl hydrocarbon receptor: Pathway interactions in the hepatotoxicity of 2,3,7,8‐tetrachlorodibenzo‐p‐dioxin and related compounds. Curr Opin Toxicol 2: 36‐41, 2017.
 64.Fangue NA, Hofmeister M, Schulte PM. Intraspecific variation in thermal tolerance and heat shock protein gene expression in common killifish, Fundulus heteroclitus. J Exp Biol 209: 2859‐2872, 2006.
 65.Fangue NA, Richards JG, Schulte PM. Do mitochondrial properties explain intraspecific variation in thermal tolerance? J Exp Biol 212: 514‐522, 2009.
 66.Feder ME, Bennett AF, Huey RB. Evolutionary physiology. Annu Rev Ecol Syst 31: 315‐341, 2000.
 67.Fisher B. Most unwanted. Environ Health Perspect 107: A18‐A23, 1999.
 68.Fraccalvieri D, Soshilov AA, Karchner SI, Franks DG, Pandini A, Bonati L, Hahn ME, Denison MS. Comparative analysis of homology models of the AH receptor ligand binding domain: Verification of structure‐function predictions by site‐directed mutagenesis of a nonfunctional receptor. Biochemistry 52: 714‐725, 2013.
 69.Gaggiotti OE, Bekkevold D, Jorgensen HB, Foll M, Carvalho GR, Andre C, Ruzzante DE. Disentangling the effects of evolutionary, demographic, and environmental factors influencing genetic structure of natural populations: Atlantic herring as a case study. Evolution 63: 2939‐2951, 2009.
 70.Garcia‐Ramos G, Kirkpatrick M. Genetic models of adaptation and gene flow in peripheral populations. Evolution 51: 21‐28, 1997.
 71.Garland T Jr, Adolph SC. Why not to do two‐species comparative studies: Limitations on inferring adaptation. Physiol Zool 67: 797‐828, 1994.
 72.Garland T Jr, Carter PA. Evolutionary physiology. Annu Rev Physiol 56: 579‐621, 1994.
 73.Garland T, Adolph SC. Physiological differentiation of vertebrate populations. Annu Rev Ecol Syst 22: 193‐228, 1991.
 74.Gonzalez‐Villasenor LI, Powers DA. Mitochondrial‐DNA restriction‐site polymorphisms in the teleost Fundulus Heteroclitus support secondary intergradation. Evolution 44: 27‐37, 1990.
 75.Griffith RW. Environment and salinity tolerance in the genus Fundulus. Copeia 1974: 319‐331, 1974.
 76.Haas HL, Freeman CJ, Logan JM, Deegan L, Gaines EF. Examining mummichog growth and movement: Are some individuals making intra‐season migrations to optimize growth? J Exp Mar Biol Ecol 369: 8‐16, 2009.
 77.Haldane JBS. The cost of natural selection. J Genet 55: 511‐524, 1957.
 78.Halpin PM. Habitat use patterns of the mummichog, Fundulus heteroclitus, in New England. I. Intramarsh variation. Estuaries 20: 618‐625, 1997.
 79.Haney RA, Dionne M, Puritz J, Rand DM. The comparative phylogeography of east coast estuarine fishes in formerly glaciated sites: Persistence versus recolonization in Cyprinodon variegatus ovinus and Fundulus heteroclitus macrolepidotus. J Hered 100: 284‐296, 2009.
 80.Hartl DL, Clark AG. Principles of Population Genetics. Sunderland, MA: Sinauer Associates, 2007.
 81.Hartl DL, Dykhuizen DE, Dean AM. Limits of adaptation: The evolution of selective neutrality. Genetics 111: 655‐674, 1985.
 82.Healy TM, Brennan RS, Whitehead A, Schulte PM. Tolerance traits related to climate change resilience are independent and polygenic. Glob Chang Biol 24: 5348‐5360, 2018.
 83.Healy TM, Bryant HJ, Schulte PM. Mitochondrial genotype and phenotypic plasticity of gene expression in response to cold acclimation in killifish. Mol Ecol 26: 814‐830, 2017.
 84.Healy TM, Schulte PM. Factors affecting plasticity in whole‐organism thermal tolerance in common killifish (Fundulus heteroclitus). J Comp Physiol B 182: 49‐62, 2012.
 85.Henn BM, Botigue LR, Bustamante CD, Clark AG, Gravel S. Estimating the mutation load in human genomes. Nat Rev Genet 16: 333‐343, 2015.
 86.Hermisson J, Pennings PS. Soft sweeps: Molecular population genetics of adaptation from standing genetic variation. Genetics 169: 2335‐2352, 2005.
 87.Hermisson J, Wagner GP. The population genetic theory of hidden variation and genetic robustness. Genetics 168: 2271‐2284, 2004.
 88.Hill GE. Mitonuclear mate choice: A missing component of sexual selection theory? Bioessays 40, 2018. DOI: 10.1002/bies.201700191.
 89.Hochachka PW, Somero GN. Biochemical Adaptation, Mechanism and Process in Physiological Evolution. New York, NY: Oxford University Press, 2002, p. 408.
 90.Huang W, Richards S, Carbone MA, Zhu D, Anholt RR, Ayroles JF, Duncan L, Jordan KW, Lawrence F, Magwire MM, Warner CB, Blankenburg K, Han Y, Javaid M, Jayaseelan J, Jhangiani SN, Muzny D, Ongeri F, Perales L, Wu YQ, Zhang Y, Zou X, Stone EA, Gibbs RA, Mackay TF. Epistasis dominates the genetic architecture of Drosophila quantitative traits. Proc Natl Acad Sci U S A 109: 15553‐15559, 2012.
 91.Hunter KL, Fox MG, Able KW. Habitat influences on reproductive allocation and growth of the mummichog (Fundulus heteroclitus) in a coastal salt marsh. Mar Biol 151: 617‐627, 2007.
 92.Imbrie J, McIntyre A, Moore TCJ. The ocean around North America at the last glacial maximum. In: Porter SC, editor. The Late Pleistocene. Minneapolis: University of Minnesota, 1983, p. 230‐238.
 93.Jha AR, Zhou D, Brown CD, Kreitman M, Haddad GG, White KP. Shared genetic signals of hypoxia adaptation in Drosophila and in high‐altitude human populations. Mol Biol Evol 33: 501‐517, 2016.
 94.Jombart T, Devillard S, Balloux F. Discriminant analysis of principal components: A new method for the analysis of genetically structured populations. BMC Genet 11: 94, 2010.
 95.Kanehisa M, Goto S. KEGG: Kyoto encyclopedia of genes and genomes. Nucleic Acids Res 28: 27‐30, 2000.
 96.Karami‐Mohajeri S, Abdollahi M. Toxic influence of organophosphate, carbamate, and organochlorine pesticides on cellular metabolism of lipids, proteins, and carbohydrates: A systematic review. Hum Exp Toxicol 30: 1119‐1140, 2011.
 97.Kerr MK, Churchill GA. Statistical design and the analysis of gene expression microarray data. Genet Res 77: 123‐128, 2001.
 98.Kimura M. Evolutionary rate at the molecular level. Nature 217: 624‐626, 1968.
 99.Kimura M. Recent development of the neutral theory viewed from the Wrightian tradition of theoretical population genetics. Proc Natl Acad Sci U S A 88: 5969‐5973, 1991.
 100.Kimura M, Crow JF. The number of alleles that can be maintained in a finite population. Genetics 49: 725‐738, 1964.
 101.Kneib RT. The role of Fundulus‐Heteroclitus in salt‐marsh trophic dynamics. Am Zool 26: 259‐269, 1986.
 102.Knox JC. Responses of river systems to holocene climates. In: Wright WEJ, editor. The Holocene. Minneapolis: University of Minnesota, 1983, p. 26‐41.
 103.Kong SW, Lee IH, Leshchiner I, Krier J, Kraft P, Rehm HL, Green RC, Kohane IS, MacRae CA, MedSeq P. Summarizing polygenic risks for complex diseases in a clinical whole‐genome report. Genet Med 17: 536‐544, 2015.
 104.Kreitman M. The neutral theory is dead. Long live the neutral theory. Bioessays 18: 678‐683; discussion 683, 1996.
 105.Laporte M, Pavey SA, Rougeux C, Pierron F, Lauzent M, Budzinski H, Labadie P, Geneste E, Couture P, Baudrimont M, Bernatchez L. RAD sequencing reveals within‐generation polygenic selection in response to anthropogenic organic and metal contamination in North Atlantic Eels. Mol Ecol 25: 219‐237, 2016.
 106.Lauerman T. The Functional Significance of the Amino Acid Differences between Allelic Isozymes of Heart Type Lactate Dehydrogenase in Fundulus heteroclitus. Ph.D. Thesis. In: Department of BiologyJohns Hopkins University, 1990.
 107.Lee DH, Lee IK, Song K, Steffes M, Toscano W, Baker B, Jacobs D. A strong dose‐response relation between serum concentrations of persistent organic pollutants and diabetes: Results from the National Health and Examination Survey 1999‐2002. Diabetes Care 29: 1638‐1644, 2006.
 108.Lee DH, Lind L, Jacobs DR Jr, Salihovic S, van Bavel B, Lind PM. Associations of persistent organic pollutants with abdominal obesity in the elderly: The Prospective Investigation of the Vasculature in Uppsala Seniors (PIVUS) study. Environ Int 40: 170‐178, 2012.
 109.Lee SH, Ra JS, Choi JW, Yim BJ, Jung MS, Kim SD. Human health risks associated with dietary exposure to persistent organic pollutants (POPs) in river water in Korea. Sci Total Environ 470‐471: 1362‐1369, 2014.
 110.Lewontin RC, Hubby JL. 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.
 111.Li QQ, Loganath A, Chong YS, Tan J, Obbard JP. Persistent organic pollutants and adverse health effects in humans. J Toxicol Environ Health A 69: 1987‐2005, 2006.
 112.Lim S, Cho YM, Park KS, Lee HK. Persistent organic pollutants, mitochondrial dysfunction, and metabolic syndrome. Ann N Y Acad Sci 1201: 166‐176, 2010.
 113.Lotrich VA. Summer home range and movements of Fundulus Heteroclitus (Pisces: Cyprinodontidae) in a tidal creek. Ecology 56: 191‐198, 1975.
 114.Mackay TF. Epistasis for quantitative traits in Drosophila. Methods Mol Biol 1253: 47‐70, 2015.
 115.Mackay TF, Moore JH. Why epistasis is important for tackling complex human disease genetics. Genome Med 6: 124, 2014.
 116.McBryan TL, Anttila K, Healy TM, Schulte PM. Responses to temperature and hypoxia as interacting stressors in fish: Implications for adaptation to environmental change. Integr Comp Biol 53: 648‐659, 2013.
 117.McBryan TL, Healy TM, Haakons KL, Schulte PM. Warm acclimation improves hypoxia tolerance in Fundulus heteroclitus. J Exp Biol 219: 474‐484, 2016.
 118.McDonald JH, Kreitman M. Adaptive protein evolution at the Adh locus in Drosophila. Nature 351: 652‐654, 1991.
 119.McIntyre A, Kipp NG, Be AWH, Crowley T, Kellogg T, Gardner JV, Prell W, Ruddiman WF. Glacial North America 18,000 years ago: A climate reconstruction. In: Cline RM, Hays JD, editors. Late Quaternary Paleoceanography and Paleoclimatology. Boulder: Geological Society of America, 1976, p. 43‐76.
 120.McKenzie JL, Dhillon RS, Schulte PM. Steep, coincident, and concordant clines in mitochondrial and nuclear‐encoded genes in a hybrid zone between subspecies of Atlantic killifish, Fundulus heteroclitus. Ecol Evol 6: 5771‐5787, 2016.
 121.Meyer J, Di Giulio R. Heritable adaptation and fitness costs in killifish (Fundulus heteroclitus) inhabiting a polluted estuary. Ecol Appl 13: 490‐503, 2003.
 122.Meyer JN, Wassenberg DM, Karchner SI, Hahn ME, Di Giulio RT. Expression and inducibility of aryl hydrocarbon receptor pathway genes in wild‐caught killifish (Fundulus heteroclitus) with different contaminant‐exposure histories. Environ Toxicol Chem 22: 2337‐2343, 2003.
 123.Mickelson DM, Clayton L, Fullerton DS, Borns HWJ. The late Wisconsin glacial record of laurentide ice sheet in the United States. In: Porter SC, editor. The Late Pleistocene. Minneapolis: University of Minnesota, 1983, p. 3‐37.
 124.Mitton JB, Koehn RK. Genetic organization and adaptive response of allozymes to ecological variables in Fundulus heteroclitus. Genetics 79: 97‐111, 1975.
 125.Morin RP, Able KW. Patterns of geographic‐variation in the egg morphology of the fundulid fish, Fundulus‐heteroclitus. Copeia 1983: 726‐740, 1983.
 126.Nacci D, Champlin D, Coiro L, McKinney R, Jayaraman S. Predicting the occurrence of genetic adaptation to dioxin like compounds in populations of the estuarine fish Fundulus heteroclitus. Environ Toxicol Chem 21: 1525‐1532, 2002.
 127.Nacci D, Coiro L, Champlin D, Jayaraman S, McKinney R, Gleason TR, Munns WR, Specker JL, Cooper KR. Adaptations of wild populations of the estuarine fish Fundulus heteroclitus to persistent environmental contaminants. Mar Biol 134: 9‐17, 1999.
 128.Nacci D, Huber M, Champlin D, Jayaraman S, Cohen S, Gauger E, Fong A, Gomez‐Chiarri M. Evolution of tolerance to PCBs and susceptibility to a bacterial pathogen (Vibrio harveyi) in Atlantic killifish (Fundulus heteroclitus) from New Bedford (MA, USA) harbor. Environ Pollut 157: 857‐864, 2009.
 129.Nacci DE, Champlin D, Jayaraman S. Adaptation of the estuarine fish Fundulus heteroclitus (Atlantic Killifish) to polychlorinated biphenyls (PCBs). Estuaries Coast 33: 853‐864, 2010.
 130.Nei M, Suzuki Y, Nozawa M. The neutral theory of molecular evolution in the genomic era. Annu Rev Genomics Hum Genet 11: 265‐289, 2010.
 131.Ohta T. The nearly neutral theory of molecular evolution. Annu Rev Ecol Syst 23: 263‐286, 1992.
 132.Oleksiak MF. Marine genomics: Insights and challenges. Brief Funct Genomics 15: 331‐332, 2016.
 133.Oleksiak MF, Churchill GA, Crawford DL. Variation in gene expression within and among natural populations. Nat Genet 32: 261‐266, 2002.
 134.Oleksiak MF, Crawford Douglas L. Functional genomics in fishes, insights into physiological complexity. In: Evan D, Claiborne J, editors. The Physiology of Fishes. Boca Raton: CRC Press, 2006, p. 523‐550.
 135.Oleksiak MF, Kolell KJ, Crawford DL. Utility of natural populations for microarray analyses: Isolation of genes necessary for functional genomic studies. Mar Biotechnol (NY) 3: S203‐S211, 2001.
 136.Oleksiak MF, Roach JL, Crawford DL. Natural variation in cardiac metabolism and gene expression in Fundulus heteroclitus. Nat Genet 37: 67‐72, 2005.
 137.Palumbi SR, Sidell BD, Vanbeneden R, Smith GD, Powers DA. Glucosephosphate isomerase (Gpi) of the teleost Fundulus‐heteroclitus (Linnaeus) – Isozymes, allozymes and their physiological roles. J Comp Physiol 138: 49‐57, 1980.
 138.Paynter KT, Dimichele L, Hand SC, Powers DA. Metabolic implications of Ldh‐B genotype during early development in Fundulus‐heteroclitus. J Exp Zool 257: 24‐33, 1991.
 139.Pierce VA, Crawford DL. Effect of acclimation on glycolytic enzymes in Atlantic and Gulf Coast Fundulus species. Am Zool 35: 99A, 1995.
 140.Pierce VA, Crawford DL. Variation in the glycolytic pathway: The role of evolutionary and physiological processes. Physiol Zool 69: 489‐508, 1996.
 141.Pierce VA, Crawford DL. Phylogenetic analysis of glycolytic enzyme expression. Science 276: 256‐259, 1997.
 142.Pierce VA, Crawford DL. Phylogenetic analysis of thermal acclimation of the glycolytic enzymes in the genus Fundulus. Physiol Zool 70: 597‐609, 1997.
 143.Place AR, Powers DA. Genetic bases for protein polymorphism in Fundulus heteroclitus (L.). I. Lactate dehydrogenase (Ldh‐B), malate dehydrogenase (Mdh‐A), glucosephosphate isomerase (Gpi‐B), and phosphoglucomutase (Pgm‐A). Biochem Genet 16: 577‐591, 1978.
 144.Place AR, Powers DA. Genetic variation and relative catalytic efficiencies: Lactate dehydrogenase B allozymes of Fundulus heteroclitus. Proc Natl Acad Sci U S A 76: 2354‐2358, 1979.
 145.Place AR, Powers DA. Kinetic characterization of the lactate dehydrogenase (LDH‐B4) allozymes of Fundulus heteroclitus. J Biol Chem 259: 1309‐1318, 1984.
 146.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.
 147.Powell WH, Bright R, Bello SM, Hahn ME. Developmental and tissue‐specific expression of AHR1, AHR2, and ARNT2 in dioxin‐sensitive and ‐resistant populations of the marine fish Fundulus heteroclitus. Toxicol Sci 57: 229‐239, 2000.
 148.Powers DA, Lauerman T, Crawford D, DiMichele L. Genetic mechanisms for adapting to a changing environment. Annu Rev Genet 25: 629‐659, 1991.
 149.Powers DA, Place AR. Biochemical genetics of Fundulus heterolitus (L.). I. Temporal and spatial variation in gene frequencies of Ldh‐B, Mdh‐A, Gpi‐B, and Pgm‐A. Biochem Genet 16: 593‐607, 1978.
 150.Powers DA, Ropson I, Brown DC, Vanbeneden R, Cashon R, Gonzalezvillasenor LI, Dimichele JA. Genetic‐variation in Fundulus‐heteroclitus – Geographic‐distribution. Am Zool 26: 131‐144, 1986.
 151.Powers DA, Schulte PM. Evolutionary adaptations of gene structure and expression in natural populations in relation to a changing environment: A multidisciplinary approach to address the million‐year saga of a small fish. J Exp Zool 282: 71‐94, 1998.
 152.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: Futuyma D, Antonovics J, editors. Oxford Surveys in Evolutionary Biology. New York, NY: Oxford University Press, 1993, p. 43‐108.
 153.Pritchard JK, Pickrell JK, Coop G. The genetics of human adaptation: Hard sweeps, soft sweeps, and polygenic adaptation. Curr Biol 20: R208‐R215, 2010.
 154.Pritchard JK, Stephens M, Donnelly P. Inference of population structure using multilocus genotype data. Genetics 155: 945‐959, 2000.
 155.Przeworski M, Coop G, Wall JD. The signature of positive selection on standing genetic variation. Evolution 59: 2312‐2323, 2005.
 156.Reid NM, Jackson CE, Gilbert D, Minx P, Montague MJ, Hampton TH, Helfrich LW, King BL, Nacci DE, Aluru N, Karchner SI, Colbourne JK, Hahn ME, Shaw JR, Oleksiak MF, Crawford DL, Warren WC, Whitehead A. The landscape of extreme genomic variation in the highly adaptable Atlantic killifish. Genome Biol Evol 9: 659‐676, 2017.
 157.Reid NM, Proestou DA, Clark BW, Warren WC, Colbourne JK, Shaw JR, Karchner SI, Hahn ME, Nacci D, Oleksiak MF, Crawford DL, Whitehead A. The genomic landscape of rapid repeated evolutionary adaptation to toxic pollution in wild fish. Science 354: 1305‐1308, 2016.
 158.Rockman MV. The QTN program and the alleles that matter for evolution: All that's gold does not glitter. Evolution 66: 1‐17, 2012.
 159.Ropson IJ, Brown EC, Powers DA. Biochemical genetics of Fundulus Heteroclitus (L.). Vi. Geographical variation in the gene frequencies of 15 loci. Evolution 44: 16‐26, 1990.
 160.Ruzzin J. Public health concern behind the exposure to persistent organic pollutants and the risk of metabolic diseases. BMC Public Health 12: 298, 2012.
 161.Schulte PM. What is environmental stress? Insights from fish living in a variable environment. J Exp Biol 217: 23‐34, 2014.
 162.Schulte PM. The effects of temperature on aerobic metabolism: Towards a mechanistic understanding of the responses of ectotherms to a changing environment. J Exp Biol 218: 1856‐1866, 2015.
 163.Schulte PM, Glemet HC, Fiebig AA, Powers DA. Adaptive variation in lactate dehydrogenase‐B gene expression: Role of a stress‐responsive regulatory element. Proc Natl Acad Sci U S A 97: 6597‐6602, 2000.
 164.Schulte PM, GomezChiarri M, Powers DA. Structural and functional differences in the promoter and 5' flanking region of Ldh‐B within and between populations of the teleost Fundulus heteroclitus. Genetics 145: 759‐769, 1997.
 165.Schultz ET, McCormick SD. Euryhalinity in an evolutionary context. Euryhaline Fishes 32: 477‐533, 2013.
 166.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.
 167.Segal JA, Barnett JL, Crawford DL. Functional analyses of natural variation in Sp1 binding sites of a TATA‐less promoter. J Mol Evol 49: 736‐749, 1999.
 168.Segal JA, Crawford DL. LDH‐B enzyme expression: The mechanisms of altered gene expression in acclimation and evolutionary adaptation. Am J Physiol 267: R1150‐R1153, 1994.
 169.Segal JA, Schulte PM, Powers DA, Crawford DL. Descriptive and functional characterization of variation in the Fundulus heteroclitus Ldh‐B proximal promoter. J Exp Zool 275: 355‐364, 1996.
 170.Sella G, Barton NH. Thinking about the evolution of complex traits in the era of genome‐wide association studies. Annu Rev Genomics Hum Genet 20: 461‐493, 2019.
 171.Shaw JR, Hampton TH, King BL, Whitehead A, Galvez F, Gross RH, Keith N, Notch E, Jung D, Glaholt SP, Chen CY, Colbourne JK, Stanton BA. Natural selection canalizes expression variation of environmentally induced plasticity‐enabling genes. Mol Biol Evol 31: 3002‐3015, 2014.
 172.Sidell BD, Johnston IA, Moerland TS, Goldspink G. The eurythermal myofibrillar protein complex of the mummichog (Fundulus heteroclitus): Adaptation to a fluctuating thermal enviroment. J Comp Physiol B 153: 167‐173, 1983.
 173.Skinner MA, Courtenay SC, Parker WR, Curry RA. Site fidelity of mummichogs (Fundulus heteroclitus) in an Atlantic Canadian Estuary. Water Qual Res J Can 40: 288‐298, 2005.
 174.Skinner MA, Courtenay SC, Parker WR, Curry RA. Stable isotopic assessment of site fidelity of mummichogs, Fundulus heteroclitus, exposed to multiple anthropogenic inputs. Environ Biol Fishes 94: 695‐706, 2012.
 175.Slatkin M. Gene flow and the geographic structure of natural populations. Science 236: 787‐792, 1987.
 176.Slatkin M. Gene flow and population structure. In: Real LA, editor. Ecological Genetics. Princeton, NJ: Princeton University Press, 1994, p. 3‐17.
 177.Smith KJ, Able KW. Dissolved oxygen dynamics in salt marsh pools and its potential impacts on fish assemblages. Mar Ecol Prog Ser 258: 223‐232, 2003.
 178.Smith MW, Chapman RW, Powers DA. Mitochondrial DNA analysis of Atlantic Coast, Chesapeake Bay, and Delaware Bay populations of the teleost Fundulus heteroclitus indicates temporally unstable distributions over geologic time. Mol Mar Biol Biotechnol 7: 79‐87, 1998.
 179.Smith MW, Glimcher MC, Powers DA. Genetic introgression of nuclear alleles between populations of the teleost Fundulus heteroclitus. Mol Mar Biol Biotechnol 1: 226‐238, 1992.
 180.Somero GN. Proteins and temperature. Annu Rev Physiol 57: 43‐68, 1995.
 181.Somero GN, Lockwood BL, Tomanek L. Biochemical Adaptation: Response to Environmental Challenges from Life's Origins to the Anthropocene. Oxford, UK: Oxford University Press, 2017, p. 108‐136.
 182.Srinivasan S, Bettella F, Mattingsdal M, Wang Y, Witoelar A, Schork AJ, Thompson WK, Zuber V, Schizophrenia Working Group of the Psychiatric Genomics Consortium TIHGC, Winsvold BS, Zwart JA, Collier DA, Desikan RS, Melle I, Werge T, Dale AM, Djurovic S, Andreassen OA. Genetic markers of human evolution are enriched in schizophrenia. Biol Psychiatry 80: 284‐292, 2016.
 183.Stockinger B, Di Meglio P, Gialitakis M, Duarte JH. The aryl hydrocarbon receptor: Multitasking in the immune system. Annu Rev Immunol 32: 403‐432, 2014.
 184.Storz JF, Bridgham JT, Kelly SA, Garland T Jr. Genetic approaches in comparative and evolutionary physiology. Am J Physiol Regul Integr Comp Physiol 309: R197‐R214, 2015.
 185.Strange RM, Burr BM. Intraspecific phylogeography of North American highland fishes: A test of the Pleistocene vicariance hypothesis. Evolution 51: 885‐897, 1997.
 186.Taylor MH, Michele LD, Leach GJ. Egg stranding in the life cycle of the Mummichog, Fundulus heteroclitus. Copeia 1977: 397, 1977.
 187.Teo SLH, Able KW. Growth and production of the mummichog (Fundulus heteroclitus) in a restored salt marsh. Estuaries 26: 51‐63, 2003.
 188.Teo SLH, Able KW. Habitat use and movement of the mummichog (Fundulus heteroclitus) in a restored salt marsh. Estuaries 26: 720‐730, 2003.
 189.Turchin MC, Chiang CWK, Palmer CD, Sankararaman S, Reich D, Hirschhorn JN. Evidence of widespread selection on standing variation in Europe at height‐associated SNPs. Nat Genet 44: 1015‐1019, 2012.
 190.Valiela I, Wright JE, Teal JM, Volkmann SB. Growth, production and energy transformations in the salt‐marsh killifish Fundulus heteroclitus. Mar Biol 40: 135‐144, 1977.
 191.Van Beneden RJ, Cashon RE, Powers DA. Biochemical genetics of Fundulus heteroclitus (L.). III. Inheritance of isocitrate dehydrogenase (Idh‐A and Idh‐B), 6‐phosphogluconate dehydrogenase (6‐Pgdh‐A), and serum esterase (Est‐S) polymorphisms. Biochem Genet 19: 701‐714, 1981.
 192.Vassy JL, Hivert MF, Porneala B, Dauriz M, Florez JC, Dupuis J, Siscovick DS, Fornage M, Rasmussen‐Torvik LJ, Bouchard C, Meigs JB. Polygenic type 2 diabetes prediction at the limit of common variant detection. Diabetes 63: 2172‐2182, 2014.
 193.Vega GC, Wiens JJ. Why are there so few fish in the sea? Proc Biol Sci 279: 2323‐2329, 2012.
 194.Visscher PM, Brown MA, McCarthy MI, Yang J. Five years of GWAS discovery. Am J Hum Genet 90: 7‐24, 2012.
 195.Wagner DN, Baris TZ, Dayan DI, Du X, Oleksiak MF, Crawford DL. Fine‐scale genetic structure due to adaptive divergence among microhabitats. Heredity (Edinb) 118: 594‐604, 2017.
 196.Wallace B. Fifty Years of Genetic Load: An Odyssey. Ithaca, NY: Cornell University Press, 1991, p. 174.
 197.Wayne ML, Simonsen KL. Statistical tests of neutrality in the age of weak selection. Trends Ecol Evol 13: 236‐240, 1998.
 198.Whitehead A. Comparative mitochondrial genomics within and among species of killifish. BMC Evol Biol 9: 11, 2009.
 199.Whitehead A. The evolutionary radiation of diverse osmotolerant physiologies in killifish (Fundulus sp.). Evolution 64: 2070‐2085, 2010.
 200.Whitehead A, Crawford DL. Neutral and adaptive variation in gene expression. Proc Natl Acad Sci U S A 103: 5425‐5430, 2006.
 201.Whitehead A, Crawford DL. Variation within and among species in gene expression: Raw material for evolution. Mol Ecol 15: 1197‐1211, 2006.
 202.Whitehead A, Roach JL, Zhang S, Galvez F. Genomic mechanisms of evolved physiological plasticity in killifish distributed along an environmental salinity gradient. Proc Natl Acad Sci U S A 108: 6193‐6198, 2011.
 203.Williams LM, Ma X, Boyko AR, Bustamante CD, Oleksiak MF. SNP identification, verification, and utility for population genetics in a non‐model genus. BMC Genet 11: 32, 2010.
 204.Williams LM, Oleksiak MF. Signatures of selection in natural populations adapted to chronic pollution. BMC Evol Biol 8: 282, 2008.
 205.Williams LM, Oleksiak MF. Ecologically and evolutionarily important SNPs identified in natural populations. Mol Biol Evol 28: 1817‐1826, 2011.
 206.Williams LM, Oleksiak MF. Evolutionary and functional analyses of cytochrome P4501A promoter polymorphisms in natural populations. Mol Ecol 20: 5236‐5247, 2011.
 207.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.
 208.Yang J, Lee T, Kim J, Cho MC, Han BG, Lee JY, Lee HJ, Cho S, Kim H. Ubiquitous polygenicity of human complex traits: Genome‐wide analysis of 49 traits in Koreans. PLoS Genet 9: e1003355, 2013.
 209.Yeaman S. Local adaptation by alleles of small effect. Am Nat 186 (Suppl 1): S74‐S89, 2015.
 210.Yu HY, Guo Y, Zeng EY. Dietary intake of persistent organic pollutants and potential health risks via consumption of global aquatic products. Environ Toxicol Chem 29: 2135‐2142, 2010.

Teaching Material

Douglas L. Crawford, Patricia M. Schulte, Andrew Whitehead, and Marjorie F. Oleksiak. Evolutionary Physiology and Genomics in the Highly Adaptable Killifish (Fundulus heteroclitus). Compr Physiol 10 : 2020, 637-671.

Didactic Synopsis

Major Teaching Points:

Evolutionary studies in the small teleost fish Fundulus heteroclitus have revealed how genetic variation influences biochemical and physiological function.

o These findings include:

o Genetic variation between LDH-B alleles that alters swimming abilities and development.

o DNA sequence variation that changes LDH-B enzyme expression.

o Quantitative differences in glycolytic enzyme expression related to thermal environment and how they modify cardiac metabolism.

o mRNA expression changes that are influenced by physiological acclimation and alter metabolism.

o Population-specific variation in DNA sequences and mRNA expression that mitigates variation in temperature, pollution, and osmotic environments.

o DNA sequence variation within and between nuclear and mitochondrial genomes and their interactions that alter mitochondrial metabolism.

o Based on research contrasting neutral and adaptive hypotheses, we suggest adaptive evolution in Fundulus is common because local populations have significant standing genetic variation and adaptive evolution typically proceeds by polygenic selection.

o If evolutionary adaptation often proceeds by polygenic selection involving small changes in allele frequencies in many genes, as is the case in Fundulus, then results from this system can enhance our general understanding of biochemical, physiological and evolutionary processes.

Didactic Legends

The following legends to the figures that appear throughout the article are written to be useful for teaching.

Figure 1. Teaching Points: Populations of F. heteroclitus along the eastern North Atlantic seacoast are more closely related on either side of the Hudson River. This pattern is seen in 1A with microsatellites (neutral markers) where Maine (ME) and Connecticut (CT) cluster separately from New Jersey (NJ); even though NJ and CT are geographically closer. Other molecular markers provide similar data. These include LDH-B sequences, mitochondrial DNA and 354 variable sites in the nuclear genome.

Figure 2. Teaching Points: A: Along the eastern seacoast of North America, F. heteroclitus has a clinal variation in the genetic variants (alleles) for many different enzymes. Three of these patterns are shown here: MDH-A with a steep change of allele frequencies at the Hudson River, IDH-B, with a small change in allele frequencies and LDH-B with a gradual change in allele frequencies. B &C: LDH-B enzyme reaction rates (kinetic constants) with different genotypes (homozygotes for the northern allele type, heterozygotes and homozygotes for southern allele type). These different LDH-B genotypes affect both hatching times (D) and swimming speeds (E). For the enzyme kinetics, hatching and swimming speeds, all fish are from the same population, ensuring a random genetic background, thus differences can be attributed to LDH-B genotypes.

Figure 3. Teaching Points: A. The evolutionary relationships among 15 populations and species of Fundulus. Blue boxes have large differences in natural environmental temperature between populations. B. Phylogenetically corrected enzyme amounts for the three enzymes that are related to environmental temperatures. Phylogenetic correction is analogous to correcting for body mass and provides measures of enzymes as a phylogenetically independent value for each population or species. C. Metabolic rates for northern and southern F. heteroclitus populations are a function of the amount of three enzymes GAPDH- glyceraldehyde-3-phosphate dehydrogenase; PYK pyruvate kinase and LDH-B.

Figure 4. Teaching Points: A. LDH-B mRNA expression defines the amount of LDH-B enzyme, and the variation in mRNA expression is related to the LDH-B proximal promoter. (B) DNA sequence variation where there is surprisingly more variation in functional sites that bind transcription factors than in non-functional sites. C. LDH-B proximal promoters from northern and southern individuals demonstrate that the DNA sequence variation between F. heteroclitus populations effect a difference in mRNA expression. D. Deleting functional DNA but not non-functional sites affects promoter activity and mRNA expression. E and F. Evolutionary relationships within and between species demonstrating that non-functional DNA sequence are similar to neutral expectation, but the functional sites demonstrate much greater difference between populations than between species. This pattern is indicative of evolution by natural selection. G. Analysis of DNA sequence variation indicating non-neutral patterns for functional DNA sequences.

Figure 5. Teaching Points: Genome wide patterns of mRNA expression measured with microarrays. A. Relative expression of adaptively important genes showing that northern individuals have different expression than both southern and F. grandis individuals. B. Plot of relative expression level (as log2 values, thus 1.0 = 2 fold difference) and probability of being significantly different (as negative log10 values, thus 2 = 0.01 for northern versus both southern and F. grandis. Gray box highlights p-value ~0.0001, with less than 1.5 fold difference.

Figure 6. Teaching Points:. mRNA expression was measured among five F. heteroclitus populations distributed along the eastern seacoast of North America (A). Because demography and thus neutral evolutionary processes can create difference among these five populations, the effect of demography was compared (Y-axis) to the effect of habitat temperature (B). The enlarged spots are adaptively important because there is a significant relationship between mRNA expression and habitat after removing demographic effects. (C) Patterns of variation within and between populations are indicative of different evolutionary effects.

Figure 7. Teaching Points: A. mRNA expression predicts heart metabolic rates when supported by glucose, fatty-acid or mixture of secondary metabolites (lactate, ketones and ethyl alcohol). Individuals differ in their overall metabolism and which substrate supports the highest metabolic rates. B. Significant difference in mRNA expression among individuals within each population (Maine and Georgia). Most mRNAs have a significant difference between individuals (p < 0.01 to 10-17) even though the magnitude of the difference (fold-change) is relatively small (few genes have 2-fold or more differences among individuals). C. Three groups of individuals that share similar patterns of expression such that individuals in one group are up (red in group 3) and individuals in another group are down (green group 2). D. mRNA expression was summarized as a linear combination of genes in three primary biochemical pathways: glycolysis, TCA-cycle and oxidative phosphorylation. The mRNA from different biochemical pathway predict fatty-acid metabolic rates for different groups of individuals. For example, TCA mRNAs explain 64% of the variation in fatty-acid dependent cardiac metabolism for group 1, while oxidative phosphorylation mRNAs or glycolytic mRNAs explain group 2 and 3 cardiac metabolism (respectively).

Figure 8. Teaching Points:.F. heteroclitus populations live in marine environments with different salinities that are a few hundred kilometers apart (A). These populations differ genetically due to both neutral and potentially adaptive genetic variation (B). The change in mRNA expression in response to different salinities depends on whether the genes evolve by neutral or adaptive processes.

Figure 9. Teaching Points: With a triads (3 populations: one hot TE and two cool reference populations north and south of the TE), (A) we can distinguish between populations, and more importantly, TE is different from both references. We can identify the SNPs that have the greatest effect on genetic distance (FST) as "outlier SNPs": SNPs with statistically unexpectedly large FST values. For this study (B), the 94 "adaptive SNPs" or SNPs evolving by natural selection have statistically large FST values between the TE and both references but are not different between references. These 94 SNPs reveal population structure where the TE population is unique and different from both references (C). Surprisingly, there is little linkage among SNPs (D). That is, the FST values for outlier SNPs declines to random expectation after 25 bp of DNA.

Figure 10. Teaching Points: A. illustrates the three locations where each location has a polluted and two clean reference populations (triad). Each of these locations was used to define changes in DNA sequences for 354 separate SNPs. The Venn diagram illustrated the number of SNPs with DNA changes shared for three different statistical tests. The red number is the number of SNPs that are statistically significant for all three tests. B and C, show how DNA promoter sequences effect a change in gene expression when exposed to pollution. Each promoter was linked to a reporter gene and cultured cells were transformed with these DNA sequences. Thus, in the same genetic cell line the promoter sequences from polluted populations induce higher gene expression than promoter sequences from clean reference populations. In C, populations with different letters are statistically different.

Figure 11. Teaching Points: A. Shows the pairs of populations used in this study where one population inhabiting polluted water was compared with a population inhabiting clean water. One of the comparisons (B) was mRNA expression patterns revealing that sensitive populations, but not tolerant populations, responded to pollution exposure, and this was particularly relevant to AHR regulated genes (lower panel in B). The diagram (C) illustrates AHR regulatory pathway with filled boxes indicating genes evolving by natural selection. These change in mRNA expression, especially associated with the AHR pathway, have patterns of DNA sequence variation indicative of evolution by natural selection (D).

Figure 12. Teaching Points: A. Mitochondrial metabolism produces most of the cells ATP via the Oxidative Phosphorylation pathway. This pathway has five enzyme complexes with a total 91 proteins – 13 from the mitochondrial genome and 78 from the nuclear genome. B. In a single population the mitochondrial respiration is affected by acclimation, assay temperature (acute effects), and an individual's mitochondria. C. Within this population are 349 nuclear DNA sequence variants (SNPs) that have large and statistically unlikely differences in allele frequencies between mitochondrial haplotypes. This difference in nuclear allele frequencies is denoted as wFST value—FST values within a population between mitochondrial types. D. The difference in nuclear SNPs affect mitochondrial respiration: individuals with more "southern" nuclear alleles, regardless of which mitochondria have higher metabolic rates.

 

Figure 13. Teaching Points: Three saltmarsh estuaries were examined along the New Jersey shore. In each saltmarsh there are permanent ponds and intertidal creeks that drain into basins (A). To determine if there was significant, potentially adaptive divergence, three statistical tests were applied to SNP allele frequencies. SNPs that were significant for all three tests were considered outlier SNPs that are most likely evolving by natural selection (B). The FST values for outlier, non-outlier and randomized data sets is shown for each population (C) and compared between Rutgers NJ and either Stone Harbor, NJ or Mantoloking, NJ.


Related Articles:

Teaching Material

Contact Editor

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

* Required Field

How to Cite

Douglas L. Crawford, Patricia M. Schulte, Andrew Whitehead, Marjorie F. Oleksiak. Evolutionary Physiology and Genomics in the Highly Adaptable Killifish (Fundulus heteroclitus). Compr Physiol 2020, 10: 637-671. doi: 10.1002/cphy.c190004