Skip to main content
  • Research article
  • Open access
  • Published:

Parallel selection on ecologically relevant gene functions in the transcriptomes of highly diversifying salmonids

Abstract

Background

Salmonid fishes are characterised by a very high level of variation in trophic, ecological, physiological, and life history adaptations. Some salmonid taxa show exceptional potential for fast, within-lake diversification into morphologically and ecologically distinct variants, often in parallel; these are the lake-resident charr and whitefish (several species in the genera Salvelinus and Coregonus). To identify selection on genes and gene categories associated with such predictable diversifications, we analysed 2702 orthogroups (4.82 Mbp total; average 4.77 genes/orthogroup; average 1783 bp/orthogroup). We did so in two charr and two whitefish species and compared to five other salmonid lineages, which do not evolve in such ecologically predictable ways, and one non-salmonid outgroup.

Results

All selection analyses are based on Coregonus and Salvelinus compared to non-diversifying taxa. We found more orthogroups were affected by relaxed selection than intensified selection. Of those, 122 were under significant relaxed selection, with trends of an overrepresentation of serine family amino acid metabolism and transcriptional regulation, and significant enrichment of behaviour-associated gene functions. Seventy-eight orthogroups were under significant intensified selection and were enriched for signalling process and transcriptional regulation gene ontology terms and actin filament and lipid metabolism gene sets. Ninety-two orthogroups were under diversifying/positive selection. These were enriched for signal transduction, transmembrane transport, and pyruvate metabolism gene ontology terms and often contained genes involved in transcriptional regulation and development. Several orthogroups showed signs of multiple types of selection. For example, orthogroups under relaxed and diversifying selection contained genes such as ap1m2, involved in immunity and development, and slc6a8, playing an important role in muscle and brain creatine uptake. Orthogroups under intensified and diversifying selection were also found, such as genes syn3, with a role in neural processes, and ctsk, involved in bone remodelling.

Conclusions

Our approach pinpointed relevant genomic targets by distinguishing among different kinds of selection. We found that relaxed, intensified, and diversifying selection affect orthogroups and gene functions of ecological relevance in salmonids. Because they were found consistently and robustly across charr and whitefish and not other salmonid lineages, we propose these genes have a potential role in the replicated ecological diversifications.

Background

Identifying the molecular mechanisms underlying adaptive phenotypic divergence is a central challenge for evolutionary biology; a key first step is to detect genes under selection rather than reflecting background neutral evolutionary processes. Parallel or convergent evolution at the molecular level may, or may not, be associated with phenotypic parallelisms across species, but the idea remains compelling [1,2,3] and has been an important analytical framework to advance research in non-model systems [4,5,6]. Molecular parallelism or convergence can be inferred either from nucleotide site-specific changes [5, 7,8,9] or at a higher level, in the sense of similar genes being targeted by similar selective forces [10,11,12,13].

Fishes have proven a powerful ecological and evolutionary group for comparisons of genes under selection and that are associated with ecological and evolutionary novelty. Sticklebacks, for example, have become a model group of repeated ecological adaptation of Holarctic marine and freshwater distribution [14,15,16]. In cichlid fishes, adaptive potential and highly malleable phenotypes are spread throughout the family. In some cases, it has been shown that relaxed selection [17,18,19,20] or positive selection [21, 22] correlate with phenotypic diversification. However, ecological opportunity differs dramatically among cichlid lineages [23,24,25,26,27], which makes it difficult to pinpoint taxa in which adaptive potential is elevated due to a shared genetic toolset [3, 6]. In contrast, freshwater lake-resident salmonids of different species and genera have similar ecological opportunity and commonly sympatric distributions across the northern hemisphere [28,29,30]. Furthermore, the freshwater habitats of northern fishes were all colonised on a similar postglacial timescale [30, 31], unlike the dramatically different and complex colonisation histories of cichlids [4, 32,33,34].

Salmonid fishes are increasingly used as model organisms in evolutionary research, because of their ecological diversity, economic value, and replicated evolution of distinct ecomorphs and traits in some taxa, such as depth specialisation and alternative migratory tactics [35,36,37,38,39,40,41,42]. Two salmonid genera in particular, the whitefishes (Coregonus) and the charrs (Salvelinus), are not sister taxa but exhibit parallel (or convergent) adaptive tendencies in freshwaters across the northern hemisphere. They have repeatedly diverged into various within-lake ecomorphs along the depth axis over short evolutionary time spans that are unmatched in any other salmonid species [2, 43,44,45,46,47]. The evolutionary and molecular basis for why Coregonus and Salvelinus show such a high degree of ecomorphological adaptability while other salmonid species do not is, however, unknown [39, 46, 47].

Determining how single and combined effects of selection act at the molecular level is facilitated by new analytical tools [48,49,50]. These selective processes are associated with adaptive evolution in different ways and are most powerful when linked with known lineage-specific phenotype changes or phenotypic diversification [51, 52]. Two selective processes – relaxed and intensified selection – are on opposite ends of the spectrum. Relaxed selection decreases the selective constraints of a gene and can lead to the accumulation of nonsynonymous substitutions and consequently changes in the amino acid sequence of a protein. By releasing a gene of selective constraints, relaxed selection can potentially foster phenotypic novelty, plasticity, and evolutionary innovation [4, 19, 49, 53]. In contrast, intensified selection increases selective constraints but can also manifest as positive intensified selection leading to more differences at some sites of a gene [54, 55]. Intensified selection implies changes to a gene have strong fitness consequences [49]. Additionally, lineage-specific episodic diversifying selection, or positive selection, will leave other signatures at the sequence level, such as more nonsynonymous changes than expected under neutrality at a subset of positions in a gene on some branches in the phylogenetic tree (i.e., branch-site model) [50, 56]. While relaxed and intensified selection are antithetical, in either case diversifying (positive) selection can simultaneously act at a proportion of sites in a gene [57, 58].

It has long been proposed that the propensity for ecological speciation in some salmonid lineages is associated with shared patterns of relaxed or diversifying selection on ecologically relevant genes and gene function terms [2, 3, 19, 39, 59, 60]. We focus on the well-characterised and richly diversifying genera Coregonus and Salvelinus, which show repeated within-lake divergences into distinct ecomorphs across the northern hemisphere [46, 47]. The Coregonus species assessed are lake resident at least since postglacial times and have high rates of within-lake adaptive divergence [45, 46, 61]. Salvelinus species are mostly freshwater residents and have undergone frequent adaptive divergence into ecomorphs along the depth axis, predominantly within lakes [38, 47, 62, 63]. Representatives of all other major salmonid lineages, Oncorhynchus, Salmo, and Thymallus, were also included in the dataset; these are generally riverine or anadromous genera that do not extensively diversify within lakes [64, 65]. By assessing consistency across two non-sister lineages, Coregonus and Salvelinus, our approach mitigates against false positives. The focus on orthogroups within and across species, rather than single genes, alleviates the problem of differing relaxation of selective constraint in duplicated compared to non-duplicated or rediploidised genes [66], which is particularly important in salmonids due to the whole-genome duplication (WGD) that their common ancestor experienced 80–103 Mya [67,68,69].

Here, we use a genome-wide comparative approach to test for shared evidence of selection at particular categories of genes, gene functions, and gene ontology terms in the two highly diversifying lineages, Coregonus and Salvelinus, relative to all other major salmonid lineages. We test a comprehensive suite of 2702 orthologous protein-coding gene sets (orthogroups) for signals of parallel relaxed, intensified, and diversifying/positive selection in Coregonus and Salvelinus (average of 4.77 genes per orthogroup; 4.82 Mbp in total; average of 1783 aligned bp per orthogroup). By distinguishing among different kinds of selection in replicate across two independent lineages, our approach can pinpoint the action of selective pressure more accurately. We find that different types of selection target different gene sets and functions in salmonids, with novel and established ecological relevance for repeated, parallel diversification potential.

Results

Selection parameter distribution and number of orthogroups under selection

Shared molecular response to selection in two whitefish and two charr species was inferred relative to six background species (five salmonids and one pike, Fig. 1). The selection parameter k in whitefish and charr, ranging from 0 (very relaxed) to 50 (very intensified), had a median value of 0.992 across orthogroups and was significantly different from the neutral expectation of 1 (Wilcoxon signed-rank test: V = 1,995,900, p = 8.859E-06). Visually, there was an excess of orthogroups with k close to 0, indicating a high number of orthogroups under pronounced relaxed selection (Fig. 2). The number of orthogroups with k < 1 (1387; relaxed selection prevailing) was slightly higher than the number of orthogroups with k > 1 (1308; intensified selection prevailing), but not significantly so (Fisher’s Exact Test, p = 0.288).

Fig. 1
figure 1

Maximum-likelihood phylogenetic tree of nine salmonid species (foreground taxa Coregonus and Salvelinus in red) and outgroup northern pike (Esox lucius; in blue). Node support values are bootstrap values from 1000 bootstrap replications. Branch lengths correspond to the number of substitutions per site. All pictures used here are under public domain. Coregonus and Salvelinus are two genera with exceptional ability for repeated, rapid diversification into ecomorphs within lakes that is unmatched in other salmonid taxa [2, 43,44,45,46,47]

Fig. 2
figure 2

Histogram of the distribution of selection parameter k values (exponent of dN/dS ratio, i.e., k in ωk) from RELAX analysis in the 2702 orthogroups. The values shown are for Coregonus and Salvelinus compared to the other five salmonid species and the outgroup (pike). The neutral expectation of k = 1 is shown as a vertical line. k values above 2 are omitted for visibility purposes in the above plot (416 orthogroups have k values above 2; the maximum possible k value is 50; see Additional file 1 for total set of k values)

On 2702 orthogroups in the final dataset we conducted analyses of relaxed and intensified selection (in RELAX) and diversifying/positive selection (in aBSREL with branch-site model). We inferred 138 orthogroups to be under relaxed selection (k < 1, false discovery rate (FDR) < 0.10) in either Coregonus or Salvelinus, of which 122 were found in both Coregonus and Salvelinus. On the other hand, 105 orthogroups showed signals of intensified selection (k > 1, FDR < 0.10), of which 78 included both Coregonus and Salvelinus. The number of relaxed orthogroups in Coregonus and Salvelinus was significantly higher compared to the number of intensified orthogroups (one-sided Fisher’s Exact Test, p = 0.035). Of the 2702 orthogroups, branch-site selection analyses inferred 111 orthogroups as being under significant diversifying/positive selection (FDR < 0.10), of which 92 included both Coregonus and Salvelinus. Thus, these orthogroups harbour a proportion of sites with significantly elevated dN/dS (= ω) values in at least one of the foreground branches leading to Coregonus or Salvelinus taxa.

After averaging selection parameter values for each gene ontology (GO) term, 13 of 1478 GO terms showed significant deviations from the null expectation of k = 1 (Wilcoxon signed-rank tests: p < 0.05; Fig. 3). Eight of these had significantly elevated k values, indicating intensified selection. The other five had significantly lowered k values, which is evidence for relaxed selection. The GO term enrichment results agreed with the general shift of selection (distribution of k) in all orthogroups. The GO terms ‘carbohydrate metabolic process’ and ‘obsolete acyl-carrier-protein biosynthetic process’, for example, were also present in the orthogroups under intensified selection. The ‘ATPase activity’ and ‘proton transmembrane transport’ GO terms were also found among orthogroups under relaxed selection. Other deviating GO terms were ‘DNA repair’ and ‘protein deubiquitination’, with evidence for intensified selection, and ‘exocytosis’ and ‘protein dephosphorylation’, with evidence for relaxed selection (Fig. 3).

Fig. 3
figure 3

Selection parameter k distributions for the gene ontology (GO) terms that deviated from the null expectation in the total set of 2702 orthogroups. k values are selection parameter values (exponent of dN/dS ratio, i.e., k in ωk) in Coregonus and Salvelinus compared to the five other salmonid species and the outgroup (pike). The red horizontal bars indicate the medians of k

Gene functions under relaxed selection

Blast2GO annotation and UniProt/Swiss-Prot literature research on the orthogroups under relaxed selection identified gene functions with potential relevance for the diversification process in charr and whitefish. Such functions included visual perception (e.g., ‘peripherin-2-like’), gene and gene product regulation (e.g., ‘E3 ubiquitin-protein ligase RNF128-like’), lipid metabolism (e.g., ‘calcium-independent phospholipase A2-gamma-like’), muscle and heart growth (e.g., ‘dual specificity protein phosphatase 6’), locomotion (e.g., ‘serine/threonine-protein phosphatase PP1-beta catalytic subunit’), and immunity (e.g., ‘adaptor-related protein complex 1’, ‘natterin-3-like’), but also genes with a role for various nervous system processes (e.g., ‘POU domain, class 4, transcription factor 3-like’; results of relaxed and intensified selection analyses: Additional file 1).

We observed compelling trends of GO term enrichment (one-tailed Fisher’s Exact Tests, uncorrected p < 0.05 but FDR > 0.10) in the orthogroups under relaxed, intensified, and diversifying selection that largely agree with the research literature on the genes contained in those orthogroups (Fig. 4, Table 1). We found the 122 orthogroups under significant relaxed selection in Coregonus and Salvelinus were enriched for a total of 11 GO terms associated with transcriptional regulation, serine family amino acid metabolism, lipid metabolism, and oxidoreductase activity, amongst others (Table 1). The REVIGO redundancy analysis results showed transcriptional regulation, serine family amino acid metabolism, lipid metabolism, and acrosome reaction to be among the few non-redundant GO terms (frequency and significance plot of non-redundant GO terms: Fig. 4a, includes clustering by semantic similarity). Transcriptional regulation and serine family amino acid metabolism were the most frequent non-redundant GO terms. In total, six of 11 GO terms were found to be non-redundant.

Fig. 4
figure 4

Gene Ontology (GO) terms with trends of overrepresentation (p < 0.05) in a the orthogroups under relaxed selection (FDR < 0.10), b the orthogroups under intensified selection (FDR < 0.10), and c the orthogroups under diversifying selection (FDR < 0.10). The orthogroups used are present in both Coregonus and Salvelinus. Bubble colour in the indicated colour spectrum corresponds to the log10 p-value of overrepresentation (blue = higher significance, red = lower significance). Bubble size corresponds to the frequency of a GO term in the orthogroups under selection. Highly similar GO terms are linked by edges in the graph, where the line width indicates the degree of similarity. The bubble placement corresponds to the position along two semantic space axes based on SimRel clustering as described in Material & methods [70]. Asterisks indicate overlap with enriched GO terms in orthogroups under other types of selection

Table 1 Over- and underrepresented terms from a GO enrichment analysis. Cor = Coregonus, Salv = Salvelinus, OGs = orthogroups

Among the top ten enriched functions in relaxed orthogroups in both Coregonus and Salvelinus, behaviour and many neural function GO terms and KEGG pathways were found in gene set enrichment analyses (Fig. 5). This is in agreement with the neural process orthogroups and serine family amino acid metabolism GO terms obtained in the GO term enrichment analysis above. The behaviour gene set was the only gene set that was significantly enriched after FDR correction (Fig. 5). Other overrepresented functions included, for example, negative regulation of signalling, urogenital system development, the peroxisome pathway, vascular smooth muscle contraction, and the AGE-RAGE signalling pathway, which plays a major role in inflammation and infection.

Fig. 5
figure 5

Top ten enriched a biological process GO terms and b KEGG pathways based on a gene set enrichment analysis (GSEA) of all orthogroups in Coregonus and Salvelinus relative to non-diversifying taxa. Terms with negative scores (yellow) were inferred to be under relaxed selection and with positive scores (blue) under intensified selection. The normalised enrichment score is calculated from transformed k values of each gene in a gene set. ** indicates significant gene set enrichment after multiple correction (FDR < 0.2); * indicates significant gene set enrichment without multiple correction ( 0.05)

Gene functions under intensified selection

The gene functions of the 78 orthogroups under intensified selection in both Coregonus and Salvelinus (results of relaxed and intensified selection analyses: Additional file 1) were found from literature search to be frequently involved in functions relevant for lipid and carbohydrate metabolism (e.g., ‘acetyl-CoA carboxylase beta’ and ‘endoplasmic reticulum mannosyl-oligosaccharide 1,2-alpha-mannosidase-like’, respectively) as well as neurological and bone development (e.g., neurological development: ‘synapsin-3’, bone development: ‘cathepsin K precursor’ and ‘paired like homeodomain 1’).

The orthogroups under intensified selection in Coregonus and Salvelinus were enriched for transcriptional regulation GO terms, but also for those associated with ubiquitine-related processes and steroid hormone receptor activity, amongst others. A total of 18 GO terms were overrepresented (Table 1). Transcriptional regulation and several signalling processes were the only high-frequency GO terms among the few non-redundant GO terms in the REVIGO analysis (frequency and significance plot of non-redundant GO terms: Fig. 4b, includes clustering by semantic similarity). In total, nine of 18 GO terms remained after the REVIGO redundancy analysis.

In the gene set enrichment analysis of all intensified selection orthogroups present in Coregonus and Salvelinus, the ‘actin filament-based process’ GO term, the ‘spliceosome’ and several signalling KEGG pathways were among the top enriched functions (Fig. 5). Other functions included ‘cellular protein-containing complex assembly’, ‘fatty acid elongation’, ‘progesterone-mediated oocyte maturation’, and ‘steroid biosynthesis’. Overall, the gene enrichment results (Fig. 5) mostly agree with the GO term overrepresentations (Fig. 4b).

Gene functions under diversifying selection

A large number of the 92 orthogroups under diversifying selection in Coregonus and Salvelinus were found in literature search to contain genes involved in regulation of gene expression, signal transduction and transmembrane transporter genes, but also immunity-related genes and a gene of the FOX set of genes, ‘FOX I1-ema’, a tissue-specific splicing factor important in otic placode formation and jaw development in zebrafish [71] (orthogroups under diversifying selection: Additional file 2, includes associated GO terms).

Orthogroups under diversifying selection were enriched for GO terms associated with transmembrane transport, phospholipid metabolic processes, acetyl-CoA carboxylase activity, various lipid metabolic processes, regulation of Wnt signalling pathway, and RNA splicing, amongst others (Fig. 4c, Table 1). A total of 47 GO terms were overrepresented, of which 23 remained after the REVIGO redundancy analysis. Pyruvate metabolism, several signal transduction processes, lipid metabolism, and transmembrane transport processes were shown to be amongst the non-redundant GO terms in the REVIGO analysis (frequency and significance plot of non-redundant GO terms: Fig. 4c, includes clustering by semantic similarity). Compared to the orthogroups under relaxed or intensified selection (Fig. 4a,b), the orthogroups under diversifying selection included a higher number of rather dissimilar low-frequency GO terms, apart from a cluster of similar metabolic GO terms (Fig. 4c). Only one GO term, ‘DNA binding’, was underrepresented (p < 0.05); with zero occurrences in the orthogroups under diversifying selection in Coregonus and Salvelinus but 77 occurrences in all other orthogroups.

Overlap between selection types

We identified nine orthogroups that showed both signals of relaxed selection (RELAX) and diversifying selection (aBSREL) and 12 orthogroups that showed both signals of intensified selection (RELAX) and diversifying selection (aBSREL) (Fig. 6a, Table 2). The overlap between orthogroups under relaxed and diversifying selection was higher than expected by chance, but not significantly so (hypergeometric expectation: 3.6 vs observed 9; one-tailed Fisher’s Exact Test: p = 0.126). The overlap between orthogroups under intensified and diversifying selection was significantly higher than expected by chance (hypergeometric expectation: 2.1 vs observed 12; one-tailed Fisher’s Exact Test: p = 0.004).

Fig. 6
figure 6

Overlapping orthogroups under selection and biological process GO terms. a Overlap between orthogroups under relaxed and diversifying or intensified and diversifying selection in both Coregonus and Salvelinus. b Biological process GO terms of orthogroups under both relaxed and diversifying selection. c Biological process GO terms of orthogroups under both intensified and diversifying selection. In b and c, rectangles are clustered into larger, coloured rectangles based on relationships among GO terms [70]

Table 2 Orthogroups (OGs) with signals of relaxed and diversifying selection or intensified and diversifying selection. k values are selection parameter values (exponent of dN/dS ratio, i.e., k in ωk). FDR values are from Benjamini-Hochberg FDR correction. Orthogroups shown are present in both Coregonus and Salvelinus. For aBSREL, p-values after correction for multiple testing (FDR < 0.10) are shown

Based on UniProt/Swiss-Prot gene information, the orthogroups with both signals of relaxed and diversifying selection are associated with functions such as immunity (5 of 9 orthogroups, e.g., ‘adaptor-related protein complex 1, mu 2 subunit (ap1m2)’), the nervous system (4 of 9 orthogroups, e.g., ‘protein kinase C epsilon type-like (prkce)’), muscle function (2 of 9 orthogroups, e.g., ‘solute carrier family 6 (neurotransmitter transporter), member 8 (slc6a8)’), blood pressure (1 of 9 orthogroups, ‘endoplasmic reticulum aminopeptidase 1-like (LOC106570844)’), and transcriptional regulation (1 of 9 orthogroups, ‘probable histone deacetylase 1-B (hdac1-b)’) (Table 2). This is in agreement with the more general GO term functions inferred using Blast2GO and associated tools (Fig. 6b – biological process GO terms, Table 2), such as neurotransmitter transport, calcium-mediated signalling, antigen presentation, regulation of blood pressure, and serine family amino acid metabolism.

Based on UniProt/Swiss-Prot gene information, orthogroups with both signals of intensified and diversifying selection are associated with transcriptional regulation (4 of 12 orthogroups, e.g., ‘paired amphipathic helix protein Sin3a-like (sin3a)’), lipid metabolism (3 of 12 orthogroups, e.g., ‘acetyl-CoA carboxylase beta (acacb)’), nervous system function (3 of 12 orthogroups, e.g., ‘synapsin-3 (syn3)’), carbohydrate metabolism (2 of 12 orthogroups, e.g., ‘alpha-2,8-sialyltransferase 8F-like (st8sia6)’), organelle function (2 of 12 orthogroups, e.g., ‘sterile alpha motif domain-containing protein 9-like (samd9l)’), bone growth (1 of 12 orthogroups, ‘cathepsin K precursor (ctsk)’), and immunity (1 of 12 orthogroups, ‘furin-1-like (fur1)’), amongst others (Table 2). Again, this agrees with the more general GO term functions inferred using Blast2GO and associated tools (Fig. 6c – biological process GO terms, Table 2), such as neurotransmitter secretion, negative regulation of transcription, serine family amino acid metabolism, fatty acid biosynthesis, and carbohydrate metabolism.

Discussion

Our analyses of shared selection in the highly diversifying taxa Coregonus and Salvelinus identified genes and gene functions with deviating signatures of selection compared to five relatively less diversifying salmonid taxa and one non-salmonid species used as background. We identified more orthogroups under relaxed selection (k value < 1) than under intensified selection (k value > 1) (Figs. 2 and 6, Additional file 1). Further, we identified a set of 92 orthogroups that reflect signals of diversifying selection, nine and 12 of which additionally experienced either relaxed or intensified selection, respectively (Fig. 6, Table 2). We explored the underlying biological relevance and associated gene functions of these orthogroups under selection and observed trends of overrepresented GO terms in all three sets of orthogroups under selection (Fig. 4, Table 1). What is more, we identified 13 GO terms with k distributions deviating from the null expectation (Fig. 3).

In our functional analyses of the above selected orthogroups, genes in orthogroups under relaxed selection were found to be involved in processes such as transcriptional regulation, nervous system function, muscle and heart growth, lipid metabolism, and immunity, while orthogroups under intensified selection were found to be more often involved in signalling processes, actin filament-based processes, RNA splicing, carbohydrate metabolism, transcriptional regulation as well as nervous system and bone development (Additional file 1). Orthogroups under relaxed selection were enriched for ‘serine family amino acid metabolism’ while those under intensified selection tended to be enriched more strongly for various signalling pathways (Fig. 4, Table 1). ‘Regulation of transcription’ was overrepresented in orthogroups under both relaxed and intensified selection. Orthogroups with signals of diversifying selection showed a clear trend towards enrichment for signal transduction, transmembrane transport, proteolysis, lipid metabolism, and pyruvate metabolism GO terms (Fig. 4, Table 1) and often contained genes involved in transcriptional regulation, development, lipid metabolism, and immunity (orthogroups under diversifying selection, Additional file 2, including associated GO terms). Therefore, our results imply different intensities of selection on different gene functions and processes.

Selection parameter distribution – shift to relaxation?

Our results agree with the repeated finding of a relaxation of selective constraint in rapidly diverging taxa [18, 19, 72, 73]. One possible cause behind this observation could be an increase of ecological opportunity due to the emergence of new habitat and a decrease in interspecies competition, for example in caves [72] or within postglacial lakes [74,75,76,77,78,79]. The adaptation to life in lakes, before or in the course of (repeated) postglacial colonisations, such as in charr and whitefish, could be such an ecological opportunity, entailing a release of selective pressures on a substantial number of genes.

From the above results we can also conclude that relaxation of selective constraints does not affect the whole set of protein-coding genes to the same degree. Purifying selection and selective constraint still affect a large number of genes with functions that ensure an organism’s integrity, survival, and reproduction (see orthogroups under intensified selection) [80]. However, in charr and whitefish we observed a larger number of orthogroups under relaxed selection that therefore escaped the effects of purifying selection.

The observation of largely different GO terms in orthogroups under relaxed selection or intensified selection agrees with a large number of studies that demonstrated different selective pressures on different gene functions [81, 82]. Various functional and structural constraints govern the substitution patterns within genes. For example, amino acid alterations in some groups of proteins can easily make them dysfunctional [81], while other kinds of proteins are under less constraint to vary [83] or are even beneficial to fitness when being different from the majority of the population (i.e., negative frequency-dependent or balanced selection) [84,85,86]. The observation of intensified selection on DNA repair in our dataset and the overrepresented GO terms related to oxidative stress and apoptotic signalling, for example, could highlight the importance of an effective response to environmental stress in rapidly diversifying taxa. Also, carbohydrate and ribosome-related processes seem to be under strong selective constraint. On the other hand, relaxation of selective constraint on ATPase activity and proton transmembrane transport as well as various lipid metabolism genes, amongst others, could have led to the physiological variety observed in extant highly diversifying salmonids.

Gene functions under relaxed selection in Coregonus and Salvelinus

Many of the orthogroups under parallel relaxed selection, i.e. those that experienced a release of selective constraint in Coregonus and Salvelinus, have been linked to the divergence of these species along the benthic-limnetic axis within lakes found in earlier research [74, 87,88,89,90,91]. Examples include the ‘peripherin-2-like’ gene, which has been implicated in visual adaptation in S. namaycush in Lake Superior [92] and the ‘natterin-3-like’ gene, which plays a putative role as an immunopeptide and experienced pronounced expression divergence in Icelandic S. alpinus ecomorphs [93, 94]. Based on our analyses, we propose these are linked to the evolution of novel phenotypes and the variation in feeding and swimming behaviour, energy storage, and the release of competition in postglacial lakes (behaviour, muscle and heart growth, locomotion, lipid metabolism, nervous system function) as well as a release or change of parasite and pathogen burden (immunity genes) in postglacial lakes [95]. This implies changes in selective pressure on these gene functions as a consequence of shifts in the ecology of highly diversifying salmonids.

Gene functions under intensified selection in Coregonus and Salvelinus

Genes under intensified selection are expected to have a strong role for survival and reproduction (i.e., fitness) of organisms [49]. We found the orthogroups under intensified selection showed a high number of processes associated with key metabolic pathways (mainly of carbohydrates and lipids), signal transduction, and regulation of gene expression and the immune system, indicating strong selective pressures on these processes and functions in Coregonus and Salvelinus. The occurrence of immunity genes under intensified selection in our results would rather imply changing parasite and pathogen regimes rather than a complete release of pathogen burden [96]. Interestingly, our results also indicated strong selective pressures on actin filament-based processes and RNA splicing as well as protein modification, which are functions with a crucial role in development [96,97,98,99]. Among these genes, some were found to be directly linked to differences in bone development [e.g., ‘paired like homeodomain 1’ and ‘cathepsin K precursor’) [100,101,102]. We speculate that these developmental process genes may be relevant in the extreme morphological diversity that can be found among closely related Coregonus species and particularly Salvelinus species [90, 100,101,102,103,104].

Gene functions under diversifying selection in Coregonus and Salvelinus

Diversifying (i.e., positive) selection on genes, which in this context means that selective pressure favours amino acid polymorphism in a particular gene, affected 92 orthogroups in Coregonus and Salvelinus relative to the non-diversifying taxa. The enrichment of lipid, glycerolipid, and pyruvate metabolism genes is a strong indication that these metabolic processes experienced diversification/positive selection in Coregonus and Salvelinus, which could imply a role in ecological diversification and habitat shifts across the depth axis in lakes. Lipid metabolism has been commonly found to be under diversifying selection in freshwater fishes, such as sticklebacks [105] and salmonids [89, 106]. Also, diversifying/positive selection on transmembrane transport, for example, has been shown repeatedly in teleost fish adapting to the freshwater environment, such as sculpin and alewife, putatively because of its role in osmoregulation [107, 108].

Overlap between selection types

A combination of approaches, such as aBSREL and RELAX as we apply here, has been shown to reliably infer genes under selection and is also able to appropriately distinguish between positive selection and relaxed selection, which are hard to disentangle using single conventional methods [109,110,111]. This is particularly important since it has been known for some time that the number of false positives in tests for positive or diversifying selection can be substantial [112,113,114,115]. Overlapping sets of orthogroups under selection inferred with different analytical approaches, as we resolve here, can therefore implicate consistent, strong, and specific signals of selection. For example, orthogroups with signals of relaxed and diversifying selection likely experienced a release of purifying selection, with the potential for evolutionary diversification [19, 49, 116].

One of the orthogroups with the most extreme signal of relaxed and diversifying selection was ‘adaptor-related protein complex 1, mu 2 subunit (ap1m2)’ (Table 2). Apart from a role in endothelial and intestinal immunity [117], ap1m2 plays a central role in the development of endoderm-derived organs (e.g., stomach and intestine) in zebrafish [118]. Given the differing parasite pressure [119, 120] and the diversity of trophic niches in lake systems [121, 122], genes under reduced selective constraint could play a crucial role in phenotypic diversity in the often lake-dwelling Salvelinus and Coregonus. Another gene showing a strong signal of relaxed and diversifying selection was ‘solute carrier family 6 (neurotransmitter transporter), member 8 (slc6a8)’, which regulates creatine uptake in tissues with high energy demands such as muscle and brain tissue [122,123,124]. Recent research on C. mykiss and C. maraena found two copies of this gene, which were either mainly expressed in muscle or kidney tissue, with the strongest expression overall in muscle [125]. Changes of selection regimes on these functions are expected, given that Salvelinus and Coregonus experienced a diversification in trophic and swimming behaviour, ecology, and life history (e.g., Salvelinus and Coregonus can be free swimming, lake bottom dwelling, or restricted to the littoral zone of lakes). Our analyses independently corroborate an important functional role of this gene in highly diversifying salmonids.

Orthogroups showing both intensified and diversifying selection likely constitute targets of “true” positive selection rather than relaxed selection misinterpreted as positive selection. Genes in these orthogroups are thus under stronger selective pressure and are potentially of higher importance to survival and overall fitness [49] in our analyses of Coregonus and Salvelinus. Two of the most strongly significant orthogroups under intensified and diversifying selection involved the genes ‘synapsin-3 (syn3)’ and ‘cathepsin K precursor (ctsk)’. syn3 is involved in neurotransmitter secretion [126] and plays a crucial role in early neuronal differentiation and in neuronal progenitor cell development in fish [127] as well as in mammals [128]. ctsk, encoding a collagenase, has an important function in bone modelling and remodelling as well as bone homeostasis in vertebrates [129, 130] and is differentially expressed in S. alpinus [102]. Both syn3 and ctsk are among the genes experiencing the highest selective pressure in Coregonus and Salvelinus as compared to other salmonid taxa, which is an indication of their importance for survival in these two genera and warrants further research.

Not only specific genes but also the gene functions associated with orthogroups under selection were found to have biological relevance. Changes in heart and muscle function, immunity genes, lipid metabolism, and transcriptional regulation have previously been implicated in diversification in salmonids [81, 131,132,133,134,135,136], cichlids [33], ants [83], and tonguefishes [84]. Therefore, our findings from molecular evolution agree with evidence from ecomorphological divergence of Coregonus and Salvelinus of immune response, feeding, metabolism, and locomotion [46, 137, 138].

Whole-genome duplication and potential effects on relaxed selection

Multiple lines of evidence suggest that whole-genome duplication (WGD) events give rise to vast amounts of genetic material that can subsequently experience elevated substitution rates and relaxation of selective constraints [139,140,141]. The lineage ancestral to salmonids experienced a WGD around 80–103 Mya [67,68,69]. It has been speculated that this may have contributed to new phenotypic innovations and to the ecological success of salmonids [67, 142,143,144,145], although a contribution to diversification rates is contentious [67, 146]. While conventional approaches for the detection of positive and relaxed selection based on orthologues in salmonids would potentially be biased in the case of duplicated and non-rediploidised genes, approaches based on orthogroups, as we applied here, alleviate this issue by explicitly comparing selection shared in Coregonus and Salvelinus relative to all other salmonid lineages [58]. This approach circumvents the difficulty of inferring orthologues in taxa that experienced multiple duplications and unequal rediploidisation.

While genes duplicated during the salmonid WGD might be more susceptible to relaxed selection [49, 147], charr and whitefish are in different subfamilies and not sister genera (Fig. 1) [41, 148]. Therefore, based on evolutionary relationships, they should have rediploidisation patterns more similar to their sister genera than to each other and there is no reason to suspect unequal rediploidisation among genera to be the cause of the relaxed selection signals we identified in Coregonus and Salvelinus. It is more likely that these two genera experienced similar selective pressures due to common diversification patterns, environmental conditions, or other common ecological or life history traits. Our results from molecular evolution analyses reflect these shared ecological and evolutionary patterns. However, recent results on gene expression in duplicated genes of salmonids have shown that relaxed selection often occurs in downregulated gene duplicates [149], which might be associated with dosage balance effects [150, 151]. Therefore, future research should ideally analyse selection in concert with gene expression and duplication.

Conclusions

Our findings suggest that there is some evidence for a parallel relaxation of selective constraint in the repeatedly diversifying salmonid lineages Coregonus and Salvelinus compared to all other salmonid lineages, which are known to be less diversifying. Genes in orthogroups under relaxed selection are involved in functions with a potential role in the rapid diversification and ecological adaptation that can be observed in wild populations of Coregonus and Salvelinus and experienced a release of selective pressure, including genes involved in behaviour, muscle function, and infection. On the other hand, orthogroups under intensified selection, such as various signal transduction and regulatory genes, are under stronger selective pressure and are consequently expected to have higher fitness effects in charr and whitefish compared to other salmonids. An independent analysis of orthogroups under diversifying selection showed an enrichment for signal transduction genes and GO terms of various metabolic processes, while genes involved in DNA binding were underrepresented. Orthogroups with both signals of relaxed and diversifying or intensified and diversifying selection can give further hints as to what selective processes govern the evolution of these genes and are also important candidates for gene sets under particularly weak (relaxed) selective pressure (while still experiencing molecular diversification), or strong (intensified) selective pressure. The discovered orthogroups under selection are a valuable resource for further investigations into the importance of certain genes for rapid diversification in salmonids and beyond.

Material & methods

Taxonomic dataset

All analyses were performed on RNAseq transcriptome datasets drawn from a search in NCBI and a literature search in Google Scholar (keywords: each salmonid genus and species, in combination with “transcriptome” and/or “RNAseq” or “RNA-seq”) to identify all available studies. The final dataset represents one transcriptome per species of transcriptomes of all available taxa as of October 2016, with additional in-house data for Salmo salar, Salmo trutta, Salvelinus alpinus, and Coregonus lavaretus that was published at a later point [88]. We aimed to retain transcriptomes with a maximal number of overlapping transcripts across salmonids, which consisted of nine species representing all five major genera (Coregonus, Oncorhynchus, Salmo, Salvelinus, and Thymallus), and the closest relative as outgroup (Esox lucius) (Fig. 1, Table 3). The selected transcriptome assemblies were all derived from several tissues so as to obtain almost complete sets of transcripts (see references in Table 3). The topology for foreground branch definitions is based on the phylogenetic tree shown in Fig. 1, which was obtained using maximum-likelihood tree estimation based on a preliminary set of 1564 orthologues (derived from the same set of transcriptomes; orthologue alignments available upon request) and in agreement with the current understanding of salmonid phylogenetic relationships [41, 148]. From the full dataset, ranging from 59.0 to 209.2 Mbp per transcriptome (Table 3), we obtained 2702 sequences of orthogroups (mean number of orthologues per orthogroup: 4.77; mean length 1783 bp; range 607–14,743 bp; total 4.82 Mbp). Using a Blast2GO [156] analysis (for details see below), we confirmed that the dataset was not enriched for any particular GO categories (Fisher’s Exact Test: all p > 0.10) and therefore not a biased representation of transcripts.

Table 3 Sources of transcriptome assemblies used in this study

Orthogroup and OG tree inference

The longest open reading frames (ORFs) in the transcriptome sequences of the ten species (excluding mitochondrial genes due to different genetic codes that could bias the selection analyses) and, additionally, protein-coding genes of the salmon genome (NCBI RefSeq assembly accession: GCF_000233375.1) [68] were determined using TransDecoder v2.1 (https://github.com/TransDecoder/TransDecoder). Next, the standard mode of OrthoFinder v2.2.3 [157], which uses a five-step algorithm that mitigates gene length bias and other biases introduced by other methods, was used to infer orthogroups from the TransDecoder ORFs. Based on the OrthoFinder orthogroup assignments, gene sequences were then extracted from the whole transcriptomes with custom Python scripts (Python v3.6.5) if they were also present in the protein-coding sequences of the Atlantic salmon genome v2 (NCBI RefSeq assembly accession: GCF_000233375.1) [68] to exclude spurious transcripts. To further minimise the number of spurious transcripts retained, only orthogroups with a higher or equal number (but not larger than four) of orthologous genes in the focal species compared to the salmon genome were included to avoid inclusion of collapsed orthologues (i.e., multiple genes that were merged into one during assembly, which seems to be a common issue in the salmonid transcriptomes published so far according to in-house results and A. Yurchenko and M. Carruthers, personal communication). Orthologues were then combined into orthogroups using custom Python scripts if they matched the same gene sets in the salmon reference genome. Duplicated orthogroups, inferred as multiple occurrences of the same orthologue sets, orthogroups with a number of sequences exceeding 80, which could include an inflated number of duplicates, or lower than seven, to avoid orthogroups unrepresentative of the whole dataset, and those with fewer than seven taxa were removed with custom Python scripts. The longest ORFs with a minimum length of 200 bp were then extracted from this orthogroup set with get_orfs_or_cdss.py (https://github.com/peterjc/pico_galaxy/blob/master/tools/get_orfs_or_cdss/get_orfs_or_cdss.py). A codon-based alignment was then produced using PRANK v.140603 [158]. The aligned orthogroups were then trimmed and filtered with trimAl (−resoverlap 1.0, −seqoverlap 0.38 -noallgaps) [159], optimising both alignment quality and the number of retained alignments. Using the final set of orthogroups as input for FastME [160], we then constructed phylogenies of each orthogroup separately for use in the selection analyses below. The number of retained genes per orthogroup was calculated using custom BASH (BASH v3.2) and Python scripts.

Selection analyses

Foreground branches for subsequent selection analyses were labelled using a custom R script (R v3.5.1) and the R package ape [161,162,163]. All selection analyses were then performed in HYPHY v2.3.13.20180521beta(MP) [48]. Orthogroup alignment FASTA file headers were shortened using custom Python scripts and FASTA and tree files were combined with custom BASH scripts. Using the above foreground definitions, RELAX [49], included in the HYPHY package, was run to infer orthogroups with significant signals of relaxed or intensified selection in the foreground branches relative to the background branches and to infer the selection strength parameter k (as defined in [49]). The software aBSREL (adaptive Branch-Site Random-Effects Likelihood inference) [50], also contained in the HYPHY package, was then used to obtain orthogroups with a proportion of sites under significant diversifying or positive selection in the foreground branches relative to the background branches. Relevant information from both the RELAX and aBSREL output was extracted with custom BASH scripts. Multiple testing was accounted for with false-discovery rate (FDR) correction [164] or Bonferroni correction [165, 166] in R. We report on the FDR level of significance throughout the manuscript but also report the orthogroups under selection after Bonferroni correction in Additional file 1. Significant differences in the number of orthogroups under relaxed or intensified selection were quantified using Fisher’s Exact Tests in R. A two-sided one-sample Wilcoxon signed-rank test in R was used to infer whether the overall selection parameter (k) distribution differed significantly from the null expectation of 1. Selection parameter distributions were visualised using the R package ggplot2 [167]. Hypergeometric expectations of orthogroups under multiple types of selection were calculated in R.

Gene ontology analyses

GO terms were inferred using a complementary set of methods in Blast2GO v5.2.5 [156]. We used BLASTn (megablast) [168, 169] using QBlast (NCBI Blast Server) and standard web blast against non-redundant nucleotide sequences and in a second step against protein sequences. A BLAST expectation value (e-value) threshold of 1.0E-3 was used. Only the top blast hit was retained per orthogroup BLAST search. BLAST descriptors were extracted for later annotations. BLASTn was run with a word size of 28, a low complexity filter, a high-scoring segment pair (HSP) length cutoff of 33, and an HSP-hit coverage of 0. GO mapping was performed with the Goa version 2018.12 database [170]. Orthogroups were remapped with InterProScan [171,172,173] using EMBL-EBI InterPro to improve mapping success. Annotations were then created using an annotation cut-off of 55, GO weight of 5, an e-value filter of 1.0E-6, an HSP-hit coverage cut-off of 0, and a hit filter of 500. Further annotations were obtained using merging of InterProScan GOs with existing annotations. Annotations were then augmented by ANNEX [174]. Remaining unannotated sequences were annotated using blast descriptions, applying a minimum similarity of 85 and validation of annotations. Gene functions were inferred based on Blast2GO and Uniprot/Swiss-Prot gene information for fish, mice, and humans (last accessed on 27/01/2019).

GO term enrichment analyses were performed using Fisher’s Exact Tests in Blast2GO on the whole dataset and all sets of orthogroups under selection (from RELAX and aBSREL with Bonferroni multiple correction, FDR correction, or no correction) with FDR multiple correction (FDR threshold of 0.10). GO terms were extracted from the Blast2GO output using custom R scripts. Overlap of sets of orthogroups under selection was determined with custom R scripts. Additionally, GO term enrichment was tested using custom R scripts to see trends in over- and underrepresentation (Blast2GO only outputs significantly over- and underrepresented GO terms after FDR correction). Violin plots of selection parameter k distributions, compared across GO terms, were plotted with the ggplot2 R package. The significance of deviations from the null expectation of k = 1 was tested for each GO term using two-sided one-sample Wilcoxon signed-rank tests in R.

To summarise and visualise the obtained GO terms in the different sets of orthogroups under selection, we used the REVIGO online analysis tool with an allowed similarity score of GO terms of 0.9, the zebrafish reference GO term database available in REVIGO, and SimRel as semantic similarity measure (last access: 01.10.2019) [70], which makes use of a semantic clustering algorithm and the p-values from the GO term enrichment analysis in R above. In this context, semantic similarity refers to the degree of relatedness between two entities/GO terms by the similarity in meaning of their annotations [175]. REVIGO clusters GO terms of sets of genes based on this algorithm in a two-dimensional semantic similarity space. Additionally, REVIGO ranks the GO terms of genes according to their redundancy.

Gene set enrichment analysis

Gene set enrichment analysis (GSEA) was performed on gene symbols of zebrafish homologues of human genes derived from Blast2GO. Zebrafish homologues were obtained from http://www.informatics.jax.org/downloads/reports/HOM_AllOrganism.rpt using a custom R script. The selection parameter (k) values from the above RELAX analysis were transformed into -1 to 1 scores in R, with negative values representing relaxed selection orthogroups and positive values representing intensified selection orthogroups. We then used WebGestalt [176,177,178,179] on the zebrafish gene symbols and associated transformed k scores to derive a) the top ten enriched biological process GO terms and b) the top ten enriched KEGG pathways in both the relaxed and intensified orthogroups. We performed 10,000 permutations for each GSEA run. All other parameters were kept at default settings (last use: 02.10.2019).

Availability of data and materials

The scripts used for orthogroup inference, selection, and gene ontology analyses, the scripts for intermediate steps and running software, and the orthogroup alignments are made publically available in the University of Glasgow Enlighten repository, DOI: https://doi.org/10.5525/gla.researchdata.913. The transcriptomes and genomes used are from the National Center for Biotechnology Information (NCBI; https://www.ncbi.nlm.nih.gov/) (GenBank accession: GATF00000000.1, Gene Expression Omnibus accession: GSE59756, RefSeq assembly accession: GCF_000233375.1), the Oregon State University’s Salmon Transcriptome website (salmon.cgrb.oregonstate.edu; Oncorhynchus mykiss), the PhyloFish project website (phylofish.sigenae.org; Coregonus clupeaformis, Salvelinus fontinalis, Thymallus thymallus, last access: 05.10.2016), and in-house data (earlier versions of NCBI GenBank accessions: GGAO00000000.1, GGAP00000000.1, GGAQ00000000.1, GGAR00000000.1; available upon request).

Abbreviations

aBSREL:

Adaptive branch-site random effects likelihood

AGE:

Advanced glycation end products

BASH:

Bourne-again shell

BLAST:

Basic local alignment search tool

BLASTn:

Nucleotide basic local alignment Search Tool

bp:

Base pair

dN/dS:

Ratio of the number of nonsynonymous substitutions per nonsynonymous site to the number of synonymous substitutions per synonymous site

DNA:

Deoxyribonucleic acid

e-value:

Expectation value in BLAST search

FASTA:

Text-based format for representing nucleotide and amino acid sequences

FDR:

False discovery rate

FOX:

Forkhead box

GO:

Gene ontology

GSEA:

Gene set enrichment analysis

HSP:

High-scoring segment pair

KEGG:

Kyoto Encyclopedia of Genes and Genomes

Mbp:

Megabase pair

Mya:

Million years ago

NCBI:

National Center for Biotechnology Information

ORF:

Open reading frame

RAGE:

Receptor for advanced glycation end products

RELAX:

Hypothesis testing framework for detection of relaxed selection

RNAseq:

Ribonucleic acid sequencing

Swiss-Prot:

An annotated protein sequence database

UniProt:

Universal protein resource

WGD:

Whole-genome duplication

ω:

Synonymous to dN/dS

References

  1. Manceau M, Domingues VS, Linnen CR, Rosenblum EB, Hoekstra HE. Convergence in pigmentation at multiple levels: mutations, genes and function. Philos Trans R Soc B. 2010;365:2439–50.

    Article  CAS  Google Scholar 

  2. Elmer KR, Meyer A. Adaptation in the age of ecological genomics: insights from parallelism and convergence. Trends Ecol Evol. 2011;26:298–306.

    Article  PubMed  Google Scholar 

  3. Berner D, Salzburger W. The genomics of organismal diversification illuminated by adaptive radiations. Trends Genet. 2015;31:491–9.

    Article  CAS  PubMed  Google Scholar 

  4. Brawand D, Wagner CE, Li YI, Malinsky M, Keller I, Fan S, et al. The genomic substrate for adaptive radiation in African cichlid fish. Nature. 2014;513:375–81.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  5. Foote AD, Liu Y, Thomas GW, Vinař T, Alföldi J, Deng J, et al. Convergent evolution of the genomes of marine mammals. Nat Genet. 2015;47:272–5.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  6. Torres-Dowdall J, Henning F, Elmer KR, Meyer A. Ecological and lineage specific factors drive the molecular evolution of rhodopsin in cichlid fishes. Mol Biol Evol. 2015;32:2876–82.

    Article  CAS  PubMed  Google Scholar 

  7. Castoe TA, de Koning AJ, Kim HM, Gu W, Noonan BP, Naylor G, et al. Evidence for an ancient adaptive episode of convergent molecular evolution. Proc Natl Acad Sci. 2009;106:8986–91.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  8. Dobler S, Dalla S, Wagschal V, Agrawal AA. Community-wide convergent evolution in insect adaptation to toxic cardenolides by substitutions in the Na, K-ATPase. Proc Natl Acad Sci. 2012;109:13040–5.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  9. Bailey SF, Rodrigue N, Kassen R. The effect of selection environment on the probability of parallel evolution. Mol Biol Evol. 2015;32:1436–48.

    Article  CAS  PubMed  Google Scholar 

  10. Farhat MR, Shapiro BJ, Kieser KJ, Sultana R, Jacobson KR, Victor TC, et al. Genomic analysis identifies targets of convergent positive selection in drug-resistant Mycobacterium tuberculosis. Nat Genet. 2013;45:1183–9.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  11. Laayouni H, Oosting M, Luisi P, Ioana M, Alonso S, Ricaño-Ponce I, et al. Convergent evolution in European and Rroma populations reveals pressure exerted by plague on Toll-like receptors. Proc Natl Acad Sci. 2014;111:2668–73.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  12. Chikina M, Robinson JD, Clark NL. Hundreds of genes experienced convergent shifts in selective pressure in marine mammals. Mol Biol Evol. 2016;33:2182–92.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  13. Castiglione GM, Schott RK, Hauser FE, Chang BS. Convergent selection pressures drive the evolution of rhodopsin kinetics at high altitudes via nonparallel mechanisms. Evolution. 2018;72:170–86.

    Article  CAS  PubMed  Google Scholar 

  14. Hatfield T, Schluter D. Ecological speciation in sticklebacks: environment-dependent hybrid fitness. Evolution. 1999;53:866–73.

    Article  PubMed  Google Scholar 

  15. Berner D, Grandchamp AC, Hendry AP. Variable progress toward ecological speciation in parapatry: stickleback across eight lake-stream transitions. Evolution. 2009;63:1740–53.

    Article  PubMed  Google Scholar 

  16. Schluter D, Conte GL. Genetics and ecological speciation. Proc Natl Acad Sci. 2009;106:9955–62.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  17. Houle D, Hughes KA, Hoffmaster DK, Ihara J, Assimacopoulos S, Charlesworth B. The effects of spontaneous mutation on quantitative traits. I Variances and covariances of life history traits. Genetics. 1994;138:773–85.

    CAS  PubMed  PubMed Central  Google Scholar 

  18. Yoder JB, Clancey E, Des Roches S, Eastman JM, Gentry L, Godsoe W, et al. Ecological opportunity and the origin of adaptive radiations. J Evol Biol. 2010;23:1581–96.

    Article  CAS  PubMed  Google Scholar 

  19. Hunt BG, Ometto L, Wurm Y, Shoemaker D, Soojin VY, Keller L, Goodisman MA. Relaxed selection is a precursor to the evolution of phenotypic plasticity. Proc Natl Acad Sci. 2011;108:15936–41.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  20. Velotta JP, McCormick SD, O’Neill RJ, Schultz ET. Relaxed selection causes microevolution of seawater osmoregulation and gene expression in landlocked alewives. Oecologia. 2014;175:1081–92.

    Article  PubMed  Google Scholar 

  21. Elmer KR, Fan S, Gunter HM, Jones JC, Boekhoff S, Kuraku S, Meyer A. Rapid evolution and selection inferred from the transcriptomes of sympatric crater lake cichlid fishes. Mol Ecol. 2010;19:197–211.

    Article  CAS  PubMed  Google Scholar 

  22. Fan S, Elmer KR, Meyer A. Positive Darwinian selection drives the evolution of the morphology-related gene, EPCAM, in particularly species-rich lineages of African cichlid fishes. J Mol Evol. 2011;73:1–9.

    Article  CAS  PubMed  Google Scholar 

  23. Wagner CE, Harmon LJ, Seehausen O. Ecological opportunity and sexual selection together predict adaptive radiation. Nature. 2012;487:366–9.

    Article  CAS  PubMed  Google Scholar 

  24. Recknagel H, Elmer KR, Meyer A. Crater lake habitat predicts morphological diversity in adaptive radiations of cichlid fishes. Evolution. 2014;68:2145–55.

    Article  PubMed  Google Scholar 

  25. Koblmüller S, Odhiambo EA, Sinyinza D, Sturmbauer C, Sefc KM. Big fish, little divergence: phylogeography of Lake Tanganyika’s giant cichlid, Boulengerochromis microlepis. Hydrobiologia. 2015;748:29–38.

    Article  PubMed  PubMed Central  CAS  Google Scholar 

  26. Arbour JH, López-Fernández H. Continental cichlid radiations: functional diversity reveals the role of changing ecological opportunity in the Neotropics. Proc R Soc B. 2016;283:20160556.

    Article  PubMed  PubMed Central  Google Scholar 

  27. Stroud JT, Losos JB. Ecological opportunity and adaptive radiation. Annu Rev Ecol Evol Syst. 2016;47:507–32.

    Article  Google Scholar 

  28. Matthews WJ. Patterns in freshwater fish ecology. New York: The University of Chicago Press; 1998.

    Book  Google Scholar 

  29. Stearns SC, Hendry AP. The salmonid contribution to key issues in evolution. Evolution illuminated: Salmon and their relatives. 2004;3–19.

  30. Griffiths D. The direct contribution of fish to lake phosphorus cycles. Ecol Freshw Fish. 2006;15:86–95.

    Article  Google Scholar 

  31. Bernatchez L, Wilson CC. Comparative phylogeography of Nearctic and Palearctic fishes. Mol Ecol. 1998;7:431–52.

    Article  Google Scholar 

  32. Elmer KR, Reggio C, Wirth T, Verheyen E, Salzburger W, Meyer A. Pleistocene desiccation in East Africa bottlenecked but did not extirpate the adaptive radiation of Lake Victoria haplochromine cichlid fishes. Proc Natl Acad Sci U S A. 2009;106:13404–9.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  33. López-Fernández H, Arbour JH, Winemiller K, Honeycutt RL. Testing for ancient adaptive radiations in Neotropical cichlid fishes. Evolution. 2013;67:1321–37.

    PubMed  Google Scholar 

  34. Matschiner M, Musilová Z, Barth JM, Starostová Z, Salzburger W, Steel M, Bouckaert R. Bayesian phylogenetic estimation of clade ages supports trans-Atlantic dispersal of cichlid fishes. Syst Biol. 2017;66:3–22.

    Article  PubMed  Google Scholar 

  35. Primmer CR. Genetics of local adaptation in salmonid fishes. Heredity. 2011;106:401.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  36. Dodson JJ, Aubin-Horth N, Thériault V, Páez DJ. The evolutionary ecology of alternative migratory tactics in salmonid fishes. Biol Rev. 2013;88:602–25.

    Article  PubMed  Google Scholar 

  37. Filteau M, Pavey SA, St-Cyr J, Bernatchez L. Gene coexpression networks reveal key drivers of phenotypic divergence in lake whitefish. Mol Biol Evol. 2013;30:1384–96.

    Article  CAS  PubMed  Google Scholar 

  38. Chavarie L, Muir AM, Zimmerman MS, Baillie SM, Hansen MJ, Nate NA, et al. Challenge to the model of lake charr evolution: shallow-and deep-water morphs exist within a small postglacial lake. Biol J Linn Soc. 2016;120:578–603.

    Google Scholar 

  39. Elmer KR. Genomic tools for new insights to variation, adaptation, and evolution in the salmonid fishes: a perspective for charr. Hydrobiologia. 2016;783:191–208.

    Article  Google Scholar 

  40. Laporte M, Dalziel AC, Martin N, Bernatchez L. Adaptation and acclimation of traits associated with swimming capacity in Lake whitefish (Coregonus clupeaformis) ecotypes. BMC Evol Biol. 2016;16:160.

    Article  PubMed  PubMed Central  CAS  Google Scholar 

  41. Macqueen DJ, Primmer CR, Houston RD, et al. Functional annotation of all salmonid genomes (FAASG): an international initiative supporting future salmonid research, conservation and aquaculture. BMC Genomics. 2017;18:484.

    Article  PubMed  PubMed Central  Google Scholar 

  42. Moore JS, Harris LN, Le Luyer J, Sutherland BJ, Rougemont Q, Tallman RF, et al. Genomics and telemetry suggest a role for migration harshness in determining overwintering habitat choice, but not gene flow, in anadromous Arctic char. Mol Ecol. 2017;26:6784–800.

    Article  CAS  PubMed  Google Scholar 

  43. Jonsson B, Jonsson N. Polymorphism and speciation in Arctic charr. J Fish Biol. 2001;58:605–38.

    Article  Google Scholar 

  44. Kottelat M, Freyhof J. Handbook of European freshwater fishes. Cornol and Freyhof, Berlin: Publications Kottelat; 2007.

    Google Scholar 

  45. Vonlanthen P, Roy D, Hudson AG, Largiadèr CR, Bittner D, Seehausen O. Divergence along a steep ecological gradient in lake whitefish (Coregonus sp.). J Evol Biol. 2009;22:498–514.

    Article  CAS  PubMed  Google Scholar 

  46. Bernatchez L, Renaut S, Whiteley AR, Derome N, Jeukens J, Landry L, Lu G, Nolte AW, Østbye K, Rogers SM, St-Cyr J. On the origin of species: insights from the ecological genomics of the lake whitefish. Philos Trans R Soc Lond B. 2010;365:1783–800.

    Article  CAS  Google Scholar 

  47. Muir AM, Hansen MJ, Bronte CR, Krueger CC. If Arctic charr Salvelinus alpinus is ‘the most diverse vertebrate’, what is the lake charr Salvelinus namaycush? Fish Fish. 2016;17:1194–207.

    Article  Google Scholar 

  48. Pond SLK, Muse SV. HyPhy: hypothesis testing using phylogenies. In: Statistical Methods in Molecular Evolution. New York, NY: Springer; 2005. p. 125–81.

    Chapter  Google Scholar 

  49. Wertheim JO, Murrell B, Smith MD, Kosakovsky Pond SL, Scheffler K. RELAX: detecting relaxed selection in a phylogenetic framework. Mol Biol Evol. 2014;32:820–32.

    Article  PubMed  PubMed Central  CAS  Google Scholar 

  50. Smith MD, Wertheim JO, Weaver S, Murrell B, Scheffler K, Kosakovsky Pond SL. Less is more: an adaptive branch-site random effects model for efficient detection of episodic diversifying selection. Mol Biol Evol. 2015;32:1342–53.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  51. Lamichhaney S, Fan G, Widemo F, Gunnarsson U, Thalmann DS, Hoeppner MP, et al. Structural genomic changes underlie alternative reproductive strategies in the ruff (Philomachus pugnax). Nat Genet. 2016;48:84.

    Article  CAS  PubMed  Google Scholar 

  52. Morandin C, Tin MM, Abril S, Gómez C, Pontieri L, Schiøtt M, et al. Comparative transcriptomics reveals the conserved building blocks involved in parallel evolution of diverse phenotypic traits in ants. Genome Biol. 2016;17:43.

    Article  PubMed  PubMed Central  CAS  Google Scholar 

  53. Sharma V, Hecker N, Roscito JG, Foerster L, Langer BE, Hiller M. A genomics approach reveals insights into the importance of gene losses for mammalian adaptations. Nat Commun. 2018;9:1215.

    Article  PubMed  PubMed Central  CAS  Google Scholar 

  54. Finseth FR, Bondra E, Harrison RG. Selective constraint dominates the evolution of genes expressed in a novel reproductive gland. Mol Biol Evol. 2014;31:3266–81.

    Article  CAS  PubMed  Google Scholar 

  55. Afanasyeva A, Bockwoldt M, Cooney CR, Heiland I, Gossmann TI. Human long intrinsically disordered protein regions are frequent targets of positive selection. Genome Res. 2018;28:975–82.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  56. Yang Z, Nielsen R. Codon-substitution models for detecting molecular adaptation at individual sites along specific lineages. Mol Biol Evol. 2002;19:908–17.

    Article  CAS  PubMed  Google Scholar 

  57. Bielawski JP, Yang Z. Positive and negative selection in the DAZ gene family. Mol Biol Evol. 2001;18:523–9.

    Article  CAS  PubMed  Google Scholar 

  58. Murrell B, Weaver S, Smith MD, Wertheim JO, Murrell S, Aylward A, et al. Gene-wide identification of episodic selection. Mol Biol Evol. 2015;32:1365–71.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  59. Orr HA, Masly JP, Presgraves DC. Speciation genes. Curr Opin Genet Dev. 2004;14:675–9.

    Article  CAS  PubMed  Google Scholar 

  60. Benderoth M, Textor S, Windsor AJ, Mitchell-Olds T, Gershenzon J, Kroymann J. Positive selection driving diversification in plant secondary metabolism. Proc Natl Acad Sci. 2006;103:9118–23.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  61. Hudson AG, Lundsgaard-Hansen B, Lucek K, Vonlanthen P, Seehausen O. Managing cryptic biodiversity: fine-scale intralacustrine speciation along a benthic gradient in Alpine whitefish (Coregonus spp.). Evol Appl. 2017;10:251–66.

    Article  PubMed  Google Scholar 

  62. Adams CE, Fraser D, Huntingford FA, Greer RB, Askew CM, Walker AF. Trophic polymorphism amongst Arctic charr from loch Rannoch, Scotland. J Fish Biol. 1998;52:1259–71.

    Article  Google Scholar 

  63. Chavarie L, Howland K, Harris L, Tonn W. Polymorphism in lake trout in Great Bear Lake: intra-lake morphological diversification at two spatial scales. Biol J Linn Soc. 2015;114:109–25.

    Article  Google Scholar 

  64. Klemetsen A, Amundsen PA, Dempson JB, Jonsson B, Jonsson N, O'Connell MF, Mortensen E. Atlantic salmon Salmo salar L., brown trout Salmo trutta L. and Arctic charr Salvelinus alpinus (L.): a review of aspects of their life histories. Ecol Freshw Fish. 2003;12:1–59.

    Article  Google Scholar 

  65. McPhee MV, Utter F, Stanford JA, Kuzishchin KV, Savvaitova KA, Pavlov DS, Allendorf FW. Population structure and partial anadromy in Oncorhynchus mykiss from Kamchatka: relevance for conservation strategies around the Pacific rim. Ecol Freshw Fish. 2007;16:539–47.

    Article  Google Scholar 

  66. Hodgins KA, Bock DG, Hahn MA, Heredia SM, Turner KG, Rieseberg LH. Comparative genomics in the Asteraceae reveals little evidence for parallel evolutionary change in invasive taxa. Mol Ecol. 2015;24:2226–40.

    Article  CAS  PubMed  Google Scholar 

  67. Macqueen DJ, Johnston IA. A well-constrained estimate for the timing of the salmonid whole genome duplication reveals major decoupling from species diversification. Proc R Soc B. 2014;281:20132881.

    Article  PubMed  PubMed Central  Google Scholar 

  68. Lien S, Koop BF, Sandve SR, Miller JR, Kent MP, Nome T, et al. The Atlantic salmon genome provides insights into rediploidization. Nature. 2016;533:200–5.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  69. Robertson FM, Gundappa MK, Grammes F, Hvidsten TR, Redmond AK, Lien S, Martin SAM, Holland PWH, Sandve SR, Macqueen DJ. Lineage-specific rediploidization is a mechanism to explain time-lags between genome duplication and evolutionary diversification. Genome Biol. 2017;18:111.

    Article  PubMed  PubMed Central  CAS  Google Scholar 

  70. Supek F, Bošnjak M, Škunca N, Šmuc T. REVIGO summarizes and visualizes long lists of gene ontology terms. PLoS One. 2011;6:e21800.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  71. Solomon KS, Kudoh T, Dawid IB, Fritz A. Zebrafish foxi1 mediates otic placode formation and jaw development. Development. 2003;130:929–40.

    Article  CAS  PubMed  Google Scholar 

  72. Calderoni L, Rota-Stabelli O, Frigato E, Panziera A, Kirchner S, Foulkes NS, et al. Relaxed selective constraints drove functional modifications in peripheral photoreception of the cavefish P. andruzzii and provide insight into the time of cave colonization. Heredity. 2016;117:383.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  73. Škaloud P, Škaloudová M, Doskočilová P, Kim JI, Shin W, Dvořák P. Speciation in protists: spatial and ecological divergence processes cause rapid species diversification in a freshwater chrysophyte. Mol Ecol. 2019;28(5):1084–95. https://doi.org/10.1111/mec.15011.

    Article  PubMed  Google Scholar 

  74. Doenz CJ, Bittner D, Vonlanthen P, Wagner CE, Seehausen O. Rapid buildup of sympatric species diversity in Alpine whitefish. Ecol Evol. 2018;8:9398–412.

    Article  PubMed  PubMed Central  Google Scholar 

  75. Siwertsson A, Knudsen R, Kahilainen K, Præbel K, Primicerio R, Amundsen PA. Sympatric diversification as influenced by ecological opportunity and historical contingency in a young species lineage of whitefish. Evol Ecol Res. 2010;12:929–47.

    Google Scholar 

  76. Wellborn GA, Langerhans RB. Ecological opportunity and the adaptive diversification of lineages. Ecol Evol. 2015;5:176–95.

    Article  PubMed  Google Scholar 

  77. Schluter D. Speciation, ecological opportunity, and latitude: (American Society of Naturalists Address). Am Nat. 2016;187:1–18.

    Article  PubMed  Google Scholar 

  78. Recknagel H, Hooker OE, Adams CE, Elmer KR. Ecosystem size predicts eco-morphological variability in a postglacial diversification. Ecol Evol. 2017;7:5560–70.

    Article  PubMed  PubMed Central  Google Scholar 

  79. Miller SE, Roesti M, Schluter D. A single interacting species leads to widespread parallel evolution of the stickleback genome. Curr Biol. 2019;29:530–7 e6.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  80. Shakhnovich BE, Koonin EV. Origins and impact of constraints in evolution of gene families. Genome Res. 2006;16:1529–36.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  81. Roux J, Liu J, Robinson-Rechavi M. Selective constraints on coding sequences of nervous system genes are a major determinant of duplicate gene retention in vertebrates. Mol Biol Evol. 2017;34:2773–91.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  82. Mitterboeck TF, Liu S, Adamowicz SJ, Fu J, Zhang R, Song W, et al. Positive and relaxed selection associated with flight evolution and loss in insect transcriptomes. GigaScience. 2017;6:1–14.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  83. Roux J, Privman E, Moretti S, Daub JT, Robinson-Rechavi M, Keller L. Patterns of positive selection in seven ant genomes. Mol Biol Evol. 2014;31:1661–85.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  84. Du M, Chen SL, Liu YH, Liu Y, Yang JF. MHC polymorphism and disease resistance to Vibrio anguillarum in 8 families of half-smooth tongue sole (Cynoglossus semilaevis). BMC Genet. 2011;12:78.

    Article  PubMed  PubMed Central  Google Scholar 

  85. Gagnaire PA, Normandeau E, Côté C, Hansen MM, Bernatchez L. The genetic consequences of spatially varying selection in the panmictic American eel (Anguilla rostrata). Genetics. 2012;190:725–36.

    Article  PubMed  PubMed Central  Google Scholar 

  86. Yan J, Zhang Y, Cheng S, Kang B, Peng J, Zhang X, et al. Common genetic heterogeneity of human interleukin-37 leads to functional variance. Cell Mol Immunol. 2017;14:783.

    Article  CAS  PubMed  Google Scholar 

  87. Bernatchez S, Laporte M, Perrier C, Sirois P, Bernatchez L. Investigating genomic and phenotypic parallelism between piscivorous and planktivorous lake trout (Salvelinus namaycush) ecotypes by means of RAD seq and morphometrics analyses. Mol Ecol. 2016;25:4773–92.

    Article  CAS  PubMed  Google Scholar 

  88. Carruthers M, Yurchenko A, Augley JJ, Adams CE, Herzyk P, Elmer KR. De novo transcriptome assembly, annotation and comparison of four ecological and evolutionary model salmonid fish species. BMC Genomics. 2018;19:32.

    Article  PubMed  PubMed Central  CAS  Google Scholar 

  89. Varadharajan S, Sandve SR, Gillard GB, Tørresen OK, Mulugeta TD, Hvidsten TR, et al. The Grayling genome reveals selection on gene expression regulation after whole-genome duplication. Gen Biol Evol. 2018;10:2785–800.

    Article  CAS  Google Scholar 

  90. Jacobs A, Carruthers M, Yurchenko A, Gordeeva N, Alekseyev S, Hooker O, et al. Convergence in form and function overcomes non-parallel evolutionary histories in a Holarctic fish. bioRxiv. 2019;1:265272v2.

    Google Scholar 

  91. Jacobs A, Carruthers M, Eckmann R, Yohannes E, Adams CE, Behrmann-Godel J, Elmer KR. Rapid niche expansion by selection on functional genomic variation after ecosystem recovery. Nature Ecol Evol. 2019;3:77.

    Article  Google Scholar 

  92. Perreault-Payette A, Muir AM, Goetz F, Perrier C, Normandeau E, Sirois P, Bernatchez L. Investigating the extent of parallelism in morphological and genomic divergence among lake trout ecotypes in Lake Superior. Mol Ecol. 2017;26:1477–97.

    Article  CAS  PubMed  Google Scholar 

  93. Steinhäuser SS. Characterization of natterin-like genes in Arctic charr (Salvelinus alpinus). (doctoral dissertation). Reykjavíc, Iceland: University of Iceland; 2013.

    Google Scholar 

  94. Gudbrandsson J, Ahi EP, Franzdottir SR, Kapralova KH, Kristjánsson BK, Steinhäuser SS, et al. The developmental transcriptome of contrasting Arctic charr (Salvelinus alpinus) morphs. F1000Research. 2015;4:136.

    Article  PubMed  Google Scholar 

  95. Poulin R. Greater diversification of freshwater than marine parasites of fish. Int J Parasitol. 2016;46:275–9.

    Article  PubMed  Google Scholar 

  96. Xing Y, Lee C. Alternative splicing and RNA selection pressure—evolutionary consequences for eukaryotic genomes. Nat Rev Genet. 2006;7:499.

    Article  CAS  PubMed  Google Scholar 

  97. Rottner K, Stradal TE. Actin dynamics and turnover in cell motility. Curr Opin Cell Biol. 2011;23:569–78.

    Article  CAS  PubMed  Google Scholar 

  98. Tan K, An L, Wang SM, Wang XD, Zhang ZN, Miao K, et al. Actin disorganization plays a vital role in impaired embryonic development of in vitro-produced mouse preimplantation embryos. PLoS One. 2015;10:e0130382.

    Article  PubMed  PubMed Central  CAS  Google Scholar 

  99. Reimand J, Wagih O, Bader GD. Evolutionary constraint and disease associations of post-translational modification sites in human genomes. PLoS Genet. 2015;11:e1004919.

    Article  PubMed  PubMed Central  CAS  Google Scholar 

  100. Ahi EP, Kapralova KH, Pálsson A, Maier VH, Gudbrandsson J, Snorrason SS, et al. Transcriptional dynamics of a conserved gene expression network associated with craniofacial divergence in Arctic charr. EvoDevo. 2014;5:40.

    Article  PubMed  PubMed Central  CAS  Google Scholar 

  101. Ahi EP. Studies of craniofacial gene expression during embryonic development in divergent Arctic charr (Salvelinus alpinus) morphs. (doctoral dissertation). Reykjavíc, Iceland: University of Iceland; 2016.

    Google Scholar 

  102. Beck SV, Räsänen K, Ahi EP, Kristjánsson BK, Skúlason S, Jónsson ZO, Leblanc CA. Gene expression in the phenotypically plastic Arctic charr (Salvelinus alpinus): a focus on growth and ossification at early stages of development. Evol Dev. 2019;21:16.

    Article  CAS  PubMed  Google Scholar 

  103. Rougeux C, Gagnaire PA, Praebel K, Seehausen O, Bernatchez L. Polygenic selection drives the evolution of convergent transcriptomic landscapes across continents within a Nearctic sister species complex. Mol Ecol. 2019;28:4388.

    Article  PubMed  Google Scholar 

  104. Striberny A, Jørgensen EH, Klopp C, Magnanou E. Arctic charr brain transcriptome strongly affected by summer seasonal growth but only subtly by feed deprivation. BMC Genomics. 2019;20:529.

    Article  PubMed  PubMed Central  CAS  Google Scholar 

  105. Bruneaux M, Johnston SE, Herczeg G, Merilä J, Primmer CR, Vasemägi A. Molecular evolutionary and population genomic analysis of the nine-spined stickleback using a modified restriction-site-associated DNA tag approach. Mol Ecol. 2013;22:565–82.

    Article  CAS  PubMed  Google Scholar 

  106. Chen X, Wang J, Yue W, Lei S, Dobjay S, Li Z, Wang C. Integrated transcriptome provides resources and insights into the adaptive evolution of colonized brown trout (Salmo trutta fario) in the Tibetan plateau. J World Aquacult Soc. 2019. https://doi.org/10.1111/jwas.12621.

  107. Dennenmoser S, Vamosi SM, Nolte AW, Rogers SM. Adaptive genomic divergence under high gene flow between freshwater and brackish-water ecotypes of prickly sculpin (Cottus asper) revealed by Pool-Seq. Mol Ecol. 2017;26:25–42.

    Article  CAS  PubMed  Google Scholar 

  108. Velotta JP, Wegrzyn JL, Ginzburg S, Kang L, Czesny S, O'Neill RJ, et al. Transcriptomic imprints of adaptation to fresh water: parallel evolution of osmoregulatory gene expression in the alewife. Mol Ecol. 2017;26:831–48.

    Article  CAS  PubMed  Google Scholar 

  109. Suzuki Y, Nei M. False-positive selection identified by ML-based methods: examples from the Sig1 gene of the diatom Thalassiosira weissflogii and the tax gene of a human T-cell lymphotropic virus. Mol Biol Evol. 2004;21:914–21.

    Article  CAS  PubMed  Google Scholar 

  110. Nozawa M, Suzuki Y, Nei M. Reliabilities of identifying positive selection by the branch-site and the site-prediction methods. Proc Natl Acad Sci. 2009;106:6700–5.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  111. Jordan G, Goldman N. The effects of alignment error and alignment filtering on the sitewise detection of positive selection. Mol Biol Evol. 2011;29:1125–39.

    Article  PubMed  CAS  Google Scholar 

  112. Markova-Raina P, Petrov D. High sensitivity to aligner and high rate of false positives in the estimates of positive selection in the 12 Drosophila genomes. Genome Res. 2011;21:863–74.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  113. Zhang J. Frequent false detection of positive selection by the likelihood method with branch-site models. Mol Biol Evol. 2004;21:1332–9.

    Article  CAS  PubMed  Google Scholar 

  114. Zhang J, Nielsen R, Yang Z. Evaluation of an improved branch-site likelihood method for detecting positive selection at the molecular level. Mol Biol Evol. 2005;22:2472–9.

    Article  CAS  PubMed  Google Scholar 

  115. Yang Z. PAML 4: phylogenetic analysis by maximum likelihood. Mol Biol Evol. 2007;24:1586–91.

    Article  CAS  PubMed  Google Scholar 

  116. Shinde SS, Teekas L, Sharma S, Vijay N. Signatures of relaxed selection in the CYP8B1 gene of birds and mammals. J Mol Evol. 2019;87:209–20.

    Article  CAS  PubMed  Google Scholar 

  117. Takahashi D, Hase K, Kimura S, Nakatsu F, Ohmae M, Mandai Y, et al. The epithelia-specific membrane trafficking factor AP-1B controls gut immune homeostasis in mice. Gastroenterology. 2011;141:621–32.

    Article  CAS  PubMed  Google Scholar 

  118. Zizioli D, Forlanelli E, Guarienti M, Nicoli S, Fanzani A, Bresciani R, et al. Characterization of the AP-1 μ1A and μ1B adaptins in zebrafish (Danio rerio). Dev Dyn. 2010;239:2404–12.

    Article  PubMed  CAS  Google Scholar 

  119. Siwertsson A, Refsnes B, Frainer A, Amundsen PA, Knudsen R. Divergence and parallelism of parasite infections in Arctic charr morphs from deep and shallow lake habitats. Hydrobiologia. 2016;783:131–43.

    Article  CAS  Google Scholar 

  120. Brunner FS, Anaya-Rojas JM, Matthews B, Eizaguirre C. Experimental evidence that parasites drive eco-evolutionary feedbacks. Proc Natl Acad Sci. 2017;114:3678–83.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  121. Lu G, Bernatchez L. Correlated trophic specialization and genetic divergence in sympatric lake whitefish ecotypes (Coregonus clupeaformis): support for the ecological speciation hypothesis. Evolution. 1999;53:1491–505.

    PubMed  Google Scholar 

  122. Hooker OE, Barry J, Van Leeuwen TE, Lyle A, Newton J, Cunningham P, Adams CE. Morphological, ecological and behavioural differentiation of sympatric profundal and pelagic Arctic charr (Salvelinus alpinus) in loch Dughaill Scotland. Hydrobiologia. 2016;783:209–21.

    Article  CAS  Google Scholar 

  123. Braissant O, Bachmann C, Henry H. Expression and function of AGAT, GAMT and CT1 in the mammalian brain. In: Creatine and creatine kinase in health and disease. Dordrecht: Springer; 2007. p. 67–81.

    Chapter  Google Scholar 

  124. Borchel A, Verleih M, Rebl A, Kühn C, Goldammer T. Creatine metabolism differs between mammals and rainbow trout (Oncorhynchus mykiss). SpringerPlus. 2014;3:510.

    Article  PubMed  PubMed Central  CAS  Google Scholar 

  125. Borchel A, Verleih M, Kühn C, Rebl A, Goldammer T. Evolutionary expression differences of creatine synthesis-related genes: implications for skeletal muscle metabolism in fish. Sci Rep. 2019;9:5429.

    Article  PubMed  PubMed Central  CAS  Google Scholar 

  126. Feng J, Chi P, Blanpied TA, Xu Y, Magarinos AM, Ferreira A, et al. Regulation of neurotransmitter release by synapsin III. J Neurosci. 2002;22:4372–80.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  127. Garbarino G, Costa S, Pestarino M, Candiani S. Differential expression of synapsin genes during early zebrafish development. Neuroscience. 2014;280:351–67.

    Article  CAS  PubMed  Google Scholar 

  128. Kao HT, Li P, Chao HM, Janoschka S, Pham K, Feng J, et al. Early involvement of synapsin III in neural progenitor cell development in the adult hippocampus. J Comp Neurol. 2008;507:1860–70.

    Article  CAS  PubMed  Google Scholar 

  129. Saftig P, Hunziker E, Wehmeyer O, Jones S, Boyde A, Rommerskirch W, et al. Impaired osteoclastic bone resorption leads to osteopetrosis in cathepsin-K-deficient mice. Proc Natl Acad Sci. 1998;95:13453–8.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  130. To TT, Witten PE, Huysseune A, Winkler C. An adult osteopetrosis model in medaka reveals the importance of osteoclast function for bone remodeling in teleost fish. Comp Biochem Physiol C Toxicol Pharmacol. 2015;178:68–75.

    Article  PubMed  CAS  Google Scholar 

  131. Kapralova KH, Gudbrandsson J, Reynisdottir S, Santos CB, Baltanás VC, Maier VH, et al. Differentiation at the MHCIIα and Cath2 loci in sympatric Salvelinus alpinus resource morphs in Lake Thingvallavatn. PLoS One. 2013;8:e69402.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  132. Conejeros P, Phan A, Power M, O'Connell M, Alekseyev S, Salinas I, Dixon B. Differentiation of sympatric Arctic char morphotypes using major histocompatibility class II genes. Trans Am Fish Soc. 2014;143:586–94.

    Article  CAS  Google Scholar 

  133. Johnston SE, Orell P, Pritchard VL, Kent MP, Lien S, Niemelä E, et al. Genome-wide SNP analysis reveals a genetic basis for sea-age variation in a wild population of Atlantic salmon (Salmo salar). Mol Ecol. 2014;23:3452–68.

    Article  CAS  PubMed  Google Scholar 

  134. Gillard G, Harvey TN, Gjuvsland A, Jin Y, Thomassen M, Lien S, et al. Life-stage-associated remodelling of lipid metabolism regulation in Atlantic salmon. Mol Ecol. 2018;27:1200–13.

    Article  CAS  PubMed  Google Scholar 

  135. Jacobs A, Womack R, Chen M, Gharbi K, Elmer KR. Significant synteny and colocalization of ecologically relevant quantitative trait loci within and across species of salmonid fishes. Genetics. 2017;207:741–54.

    CAS  PubMed  PubMed Central  Google Scholar 

  136. Larson WA, Dann TH, Limborg MT, McKinney GJ, Seeb JE, Seeb LW. Parallel signatures of selection at genomic islands of divergence and the MHC in ecotypes of sockeye salmon across Alaska. Mol Ecol. 2019. https://doi.org/10.1111/mec.15082.

  137. Jeukens J, Renaut S, St-Cyr J, Nolte AW, Bernatchez L. The transcriptomics of sympatric dwarf and normal lake whitefish (Coregonus clupeaformis spp., Salmonidae) divergence as revealed by next-generation sequencing. Mol Ecol. 2010;19:5389–403.

    Article  PubMed  Google Scholar 

  138. Bernatchez L, Renaut S, Whiteley AR, Derome N, Jeukens J, Landry L, et al. On the origin of species: insights from the ecological genomics of lake whitefish. Philos Trans R Soc B Biol Sci. 2010;365:1783–800.

    Article  CAS  Google Scholar 

  139. Lynch M, Conery JS. The evolutionary fate and consequences of duplicate genes. Science. 2000;290:1151–5.

    Article  CAS  PubMed  Google Scholar 

  140. Kondrashov FA, Rogozin IB, Wolf YI, Koonin EV. Selection in the evolution of gene duplications. Genome Biol. 2002;3:8.1–9.

    Article  Google Scholar 

  141. Brunet FG, Crollius HR, Paris M, Aury JM, Gibert P, Jaillon O, et al. Gene loss and evolutionary rates following whole-genome duplication in teleost fishes. Mol Biol Evol. 2006;23:1808–16.

    Article  CAS  PubMed  Google Scholar 

  142. Taylor JS, Van de Peer Y, Meyer A. Genome duplication, divergent resolution and speciation. Trends Genet. 2001;17:299–301.

    Article  CAS  PubMed  Google Scholar 

  143. Hoegg S, Brinkmann H, Taylor JS, Meyer A. Phylogenetic timing of the fish-specific genome duplication correlates with the diversification of teleost fish. J Mol Evol. 2004;59:190–203.

    Article  CAS  PubMed  Google Scholar 

  144. Phillips RB, Keatley KA, Morasch MR, Ventura AB, Lubieniecki KP, Koop BF, et al. Assignment of Atlantic salmon (Salmo salar) linkage groups to specific chromosomes: conservation of large syntenic blocks corresponding to whole chromosome arms in rainbow trout (Oncorhynchus mykiss). BMC Genet. 2009;10:46.

    Article  PubMed  PubMed Central  CAS  Google Scholar 

  145. Crête-Lafrenière A, Weir LK, Bernatchez L. Framing the Salmonidae family phylogenetic portrait: a more complete picture from increased taxon sampling. PLoS One. 2012;7:e46662.

    Article  PubMed  PubMed Central  CAS  Google Scholar 

  146. Santini F, Harmon LJ, Carnevale G, Alfaro ME. Did genome duplication drive the origin of teleosts? A comparative study of diversification in ray-finned fishes. BMC Evol Biol. 2009;9:194.

    Article  PubMed  PubMed Central  CAS  Google Scholar 

  147. Innan H, Kondrashov F. The evolution of gene duplications: classifying and distinguishing between models. Nat Rev Genet. 2010;11:97.

    Article  CAS  PubMed  Google Scholar 

  148. Lecaudey LA, Schliewen UK, Osinov AG, Taylor EB, Bernatchez L, Weiss SJ. Inferring phylogenetic structure, hybridization and divergence times within Salmoninae (Teleostei: Salmonidae) using RAD-sequencing. Mol Phylogenet Evol. 2018;124:82–99.

    Article  CAS  PubMed  Google Scholar 

  149. Sandve SR, Rohlfs RV, Hvidsten TR. Subfunctionalization versus neofunctionalization after whole-genome duplication. Nat Genet. 2018;50:908–9.

    Article  CAS  PubMed  Google Scholar 

  150. Rosado A, Raikhel NV. Application of the gene dosage balance hypothesis to auxin-related ribosomal mutants in Arabidopsis. Plant Signal Behav. 2010;5:450–2.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  151. Birchler JA, Veitia RA. Gene balance hypothesis: connecting issues of dosage sensitivity across biological disciplines. Proc Natl Acad Sci. 2012;109:14746–53.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  152. Pasquier J, Cabau C, Nguyen T, Jouanno E, Severac D, Braasch I, et al. Gene evolution and gene expression after whole genome duplication in fish: the PhyloFish database. BMC Genomics. 2016;17:368.

    Article  PubMed  PubMed Central  CAS  Google Scholar 

  153. Rondeau EB, Minkley DR, Leong JS, Messmer AM, Jantzen JR, von Schalburg KR, et al. The genome and linkage map of the northern pike (Esox lucius): conserved synteny revealed between the salmonid sister group and the Neoteleostei. PLoS One. 2014;9:e102089.

    Article  PubMed  PubMed Central  CAS  Google Scholar 

  154. Fox SE, Christie MR, Marine M, Priest HD, Mockler TC, Blouin MS. Sequencing and characterization of the anadromous steelhead (Oncorhynchus mykiss) transcriptome. Mar Genomics. 2014;15:13–5.

    Article  PubMed  Google Scholar 

  155. Tomalty KM, Meek MH, Stephens MR, Rincón G, Fangue NA, May BP, Baerwald MR. Transcriptional response to acute thermal exposure in juvenile Chinook salmon determined by RNAseq. G3: Genes. Genomes, Genetics. 2015;5:1335–49.

    CAS  Google Scholar 

  156. Conesa A, Götz S, García-Gómez JM, Terol J, Talón M, Robles M. Blast2GO: a universal tool for annotation, visualization and analysis in functional genomics research. Bioinformatics. 2005;21:3674–6.

    Article  CAS  PubMed  Google Scholar 

  157. Emms DM, Kelly S. OrthoFinder: solving fundamental biases in whole genome comparisons dramatically improves orthogroup inference accuracy. Genome Biol. 2015;16:157.

    Article  PubMed  PubMed Central  CAS  Google Scholar 

  158. Löytynoja A. Phylogeny-aware alignment with PRANK. In Multiple Sequence Alignment Methods. Totowa, NJ: Humana Press; 2014. p. 155–70.

    Book  Google Scholar 

  159. Capella-Gutiérrez S, Silla-Martínez JM, Gabaldón T. trimAl: a tool for automated alignment trimming in large-scale phylogenetic analyses. Bioinformatics. 2009;25:1972–3.

    Article  PubMed  PubMed Central  CAS  Google Scholar 

  160. Lefort V, Desper R, Gascuel O. FastME 2.0: a comprehensive, accurate, and fast distance-based phylogeny inference program. Mol Biol Evol. 2015;32:2798–800.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  161. R Core Team. R: a language and environment for statistical computing. Vienna, Austria: R Foundation for Statistical Computing; 2016. https://www.R-project.org/

    Google Scholar 

  162. Paradis E, Claude J, Strimmer K. APE: analyses of phylogenetics and evolution in R language. Bioinformatics. 2004;20:289–90.

    Article  CAS  PubMed  Google Scholar 

  163. Popescu AA, Huber KT, Paradis E. Ape 3.0: new tools for distance-based phylogenetics and evolutionary analysis in R. Bioinformatics. 2012;28:1536–7.

    Article  CAS  PubMed  Google Scholar 

  164. Benjamini Y, Hochberg Y. Controlling the false discovery rate: a practical and powerful approach to multiple testing. J R Stat Soc Ser B Methodol. 1995;57:289–300.

    Google Scholar 

  165. Bonferroni C. Teoria statistica delle classi e calcolo delle probabilita. Pubblicazioni del R Istituto Superiore di Scienze Economiche e Commericiali di Firenze. 1936;8:3–62.

    Google Scholar 

  166. Dunn OJ. Multiple comparisons among means. J Am Stat Assoc. 1961;56:52–64.

    Article  Google Scholar 

  167. Wickham H. ggplot2. Wiley Interdisciplinary Reviews: Computational Statistics. 2011;3:180–5.

    Article  Google Scholar 

  168. Altschul SF, Gish W, Miller W, Myers EW, Lipman DJ. Basic local alignment search tool. J Mol Biol. 1990;215:403–10.

    Article  CAS  PubMed  Google Scholar 

  169. Morgulis A, Coulouris G, Raytselis Y, Madden TL, Agarwala R, Schäffer AA. Database indexing for production MegaBLAST searches. Bioinformatics. 2008;24:1757–64.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  170. Camon E, Magrane M, Barrell D, Lee V, Dimmer E, Maslen J, et al. The gene ontology annotation (Goa) database: sharing knowledge in uniprot with gene ontology. Nucleic Acids Res. 2004;32:D262–6.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  171. Zdobnov EM, Apweiler R. InterProScan – an integration platform for the signature-recognition methods in InterPro. Bioinformatics. 2001;17:847–8.

    Article  CAS  PubMed  Google Scholar 

  172. Quevillon E, Silventoinen V, Pillai S, Harte N, Mulder N, Apweiler R, Lopez R. InterProScan: protein domains identifier. Nucleic Acids Res. 2005;33:W116–20.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  173. Jones P, Binns D, Chang HY, Fraser M, Li W, McAnulla C, et al. InterProScan 5: genome-scale protein function classification. Bioinformatics. 2014;30:1236–40.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  174. Myhre S, Tveit H, Mollestad T, Lægreid A. Additional gene ontology structure for improved biological reasoning. Bioinformatics. 2006;22:2020–7.

    Article  CAS  PubMed  Google Scholar 

  175. Pesquita C, Faria D, Falcao AO, Lord P, Couto FM. Semantic similarity in biomedical ontologies. PLoS Comput Biol. 2009;5:e1000443.

    Article  PubMed  PubMed Central  CAS  Google Scholar 

  176. Zhang B, Kirov SA, Snoddy JR. WebGestalt: an integrated system for exploring gene sets in various biological contexts. Nucleic Acids Res. 2005;33(web server issue):W741–8.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  177. Wang J, Duncan D, Shi Z, Zhang B. WEB-based GEne SeT AnaLysis toolkit (WebGestalt): update 2013. Nucleic Acids Res. 2013;41(web server issue):W77–83.

    Article  PubMed  PubMed Central  Google Scholar 

  178. Wang J, Vasaikar S, Shi Z, Greer M, Zhang B. WebGestalt 2017: a more comprehensive, powerful, flexible and interactive gene set enrichment analysis toolkit. Nucleic Acids Res. 2017;45:W130–7.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  179. Liao Y, Wang J, Jaehnig E, Shi Z, Zhang B. WebGestalt 2019: gene set analysis toolkit with revamped UIs and APIs. Nucleic Acids Res. 2019;47:W199–205.

    Article  PubMed  PubMed Central  Google Scholar 

Download references

Acknowledgements

We thank lab members A. Yurchenko, M. Carruthers, H. Recknagel, and A. Jacobs for discussions and support during the project and M. Carruthers for contributing unpublished data.

Funding

We thank the ERASMUS+ Internship programme and the Fisheries Society of the British Isles PhD Studentship programme for funding awarded to KS and Wellcome Trust-Glasgow Polyomics ISSF Catalyst funding to KRE (Wellcome Trust [097821/Z/11/Z]). The funding bodies played no role in the design of the study and collection, analysis, and interpretation of data and in writing the manuscript.

Author information

Authors and Affiliations

Authors

Contributions

KS and KRE designed the study with input from CEA. KS analysed the data and created the figures and tables. KS wrote the manuscript, with contributions from KRE and comments from CEA. All authors read and approved the final manuscript.

Corresponding author

Correspondence to Kathryn R. Elmer.

Ethics declarations

Ethics approval and consent to participate

Not applicable.

Consent for publication

Not applicable.

Competing interests

The authors declare that they have no competing interests.

Additional information

Publisher’s Note

Springer Nature remains neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Supplementary information

Additional file 1.

Total set of orthogroups with associated GO terms and results of relaxed and intensified selection analyses.

Additional file 2.

Orthogroups under diversifying selection (FDR < 0.10) in both Coregonus and Salvelinus.

Rights and permissions

Open Access This article is distributed under the terms of the Creative Commons Attribution 4.0 International License (http://creativecommons.org/licenses/by/4.0/), which permits unrestricted use, distribution, and reproduction in any medium, provided you give appropriate credit to the original author(s) and the source, provide a link to the Creative Commons license, and indicate if changes were made. The Creative Commons Public Domain Dedication waiver (http://creativecommons.org/publicdomain/zero/1.0/) applies to the data made available in this article, unless otherwise stated.

Reprints and permissions

About this article

Check for updates. Verify currency and authenticity via CrossMark

Cite this article

Schneider, K., Adams, C.E. & Elmer, K.R. Parallel selection on ecologically relevant gene functions in the transcriptomes of highly diversifying salmonids. BMC Genomics 20, 1010 (2019). https://doi.org/10.1186/s12864-019-6361-2

Download citation

  • Received:

  • Accepted:

  • Published:

  • DOI: https://doi.org/10.1186/s12864-019-6361-2

Keywords