Next Article in Journal
In Vivo Hematopoietic Stem Cell Genome Editing: Perspectives and Limitations
Next Article in Special Issue
Heterogeneous Genomic Divergence Landscape in Two Commercially Important European Scallop Species
Previous Article in Journal
Interplay of Ecological Opportunities and Functional Traits Drives the Evolution and Diversification of Millettiod Legumes (Fabaceae)
Previous Article in Special Issue
Complete Mitochondrial Genome of Two Ectoparasitic Capsalids (Platyhelminthes: Monogenea: Monopisthocotylea): Gene Content, Composition, and Rearrangement
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Review

Salmonidae Genome: Features, Evolutionary and Phylogenetic Characteristics

by
Artem P. Dysin
1,*,
Yuri S. Shcherbakov
1,
Olga A. Nikolaeva
1,
Valerii P. Terletskii
2,
Valentina I. Tyshchenko
1 and
Natalia V. Dementieva
1
1
Russian Research Institute of Farm Animal Genetics and Breeding-Branch of the L.K. Ernst Federal Research Center for Animal Husbandry, Pushkin, 196601 St. Petersburg, Russia
2
All-Russian Research Veterinary Institute of Poultry Science-Branch of the Federal Scientific Center, All-Russian Research and Technological Poultry Institute (ARRVIPS), Lomonosov, 198412 St. Petersburg, Russia
*
Author to whom correspondence should be addressed.
Genes 2022, 13(12), 2221; https://doi.org/10.3390/genes13122221
Submission received: 12 September 2022 / Revised: 19 October 2022 / Accepted: 24 November 2022 / Published: 27 November 2022
(This article belongs to the Special Issue Genomics in Aquaculture and Fisheries)

Abstract

:
The salmon family is one of the most iconic and economically important fish families, primarily possessing meat of excellent taste as well as irreplaceable nutritional and biological value. One of the most common and, therefore, highly significant members of this family, the Atlantic salmon (Salmo salar L.), was not without reason one of the first fish species for which a high-quality reference genome assembly was produced and published. Genomic advancements are becoming increasingly essential in both the genetic enhancement of farmed salmon and the conservation of wild salmon stocks. The salmon genome has also played a significant role in influencing our comprehension of the evolutionary and functional ramifications of the ancestral whole-genome duplication event shared by all Salmonidae species. Here we provide an overview of the current state of research on the genomics and phylogeny of the various most studied subfamilies, genera, and individual salmonid species, focusing on those studies that aim to advance our understanding of salmonid ecology, physiology, and evolution, particularly for the purpose of improving aquaculture production. This review should make potential researchers pay attention to the current state of research on the salmonid genome, which should potentially attract interest in this important problem, and hence the application of new technologies (such as genome editing) in uncovering the genetic and evolutionary features of salmoniforms that underlie functional variation in traits of commercial and scientific importance.

1. Introduction

Due to their great diversity, morphology, ecology, genetics, and genomics, as well as their higher economic value, biology, and culture, fish are a very interesting group of animals. In this regard, more and more intensive efforts are now being devoted to the development of aquaculture. Efficient fish farming makes it possible to produce large volumes of high-quality products, which contribute to the conservation of natural populations of valuable fish species [1,2]. One of these interesting species of fish is the Salmonidae family, consisting of salmon, trout, char, grayling, whitefish, and their relatives, which has a scientific, social, and economic value that is unique among fish [3]. They are mainly known for performing some specific key functions in ecology, such as bringing marine nutrients to freshwater ecosystems [4].
Combined with their tetraploid origin, life history diversity, and rate of diversification, they have generated significant interest from the research community. Equally important, salmon is considered an excellent source of nutrients such as amino acids, lipids, vitamins, and minerals [5,6,7,8,9,10,11,12,13]. The problem remains that many salmon populations are declining, and great efforts are directed toward their conservation, especially in relation to anthropogenic factors [14]. Salmonids include 70 to 200 species with a wide variety of adaptations and life cycle strategies [15]. Salmon aquaculture and fisheries (mainly Atlantic salmon Salmo salar L. and Oncorhynchus spp.) play an important role in the economic and/or food security of several countries, accounting for 7.2/16.6% of all fish traded in terms of weight ratio/price [16].
Fortunately, genomics has given us new tools for addressing fundamental fishery management concerns, including stock recognition, population structure, and adaptive responses to ecological change [17,18,19]. The application of comparative genomics, molecular cell biology, and in situ hybridization of sequences and chromosomes using model and non-model fish species considerably helps our understanding of gene function, evolution, speciation, selection and adaptation, and species diversity [17]. The identifying of SNP markers by NGS has expanded the capacity to trace fishing resources or products back to the original site, allowing regulations for some industrially useful fish species to be enforced [20]. The Japanese pufferfish Fugu rubripes was the first to have its whole genome sequenced [21]. Since the first main commercial species, the Atlantic cod, was sequenced in 2011, aquaculture and fisheries genomics have made substantial progress [22]. Over 200 fish genomes have been sequenced and made accessible in public repositories since the invention and enhancement of massively parallel sequencing technology began in about 2005 [23]. Population genomic analysis using RAD sequencing or genome resequencing has been implemented for a few commercially important species, such as Asian sea bass [24], European sea bass [25], and Atlantic cod [22,26], revealing the genetic basis of fisheries-driven evolution and the possible repercussions of changes in the environment. High-quality genomes for numerous fish species have recently been released, providing insights into the molecular basis of evolution and biology [27,28,29,30]. A large amount of knowledge has been gained by sequencing the teleost fish genome. For example, the spotted gar (Lepisosteus oculatus) genome has shown how immunity evolution, mineralization, and development can be mediated by Hox, ParaHox, and microRNA genes [24]. The genetic foundation of polyploidy and parallelism between basal and fish-derived lineages to acquire duplications has been shown by genome sequencing of ancient fish lineages such as the starfish (Acipenser ruthenus) [31]. The Genome 10 k project, for instance, was established in 2009 by a partnership of biologists and genomic researchers to sequence the full genomes of 10,000 vertebrate species, including around 4000 fish species, in order to comprehend vertebrates evolution and conserve many endangered species [32,33,34].
Returning to the specific object of our discussion, the fish of the salmon family, we flesh out this discussion regarding them. The NCBI database currently provides data on 41 genomic assemblages of salmoniforms, which may indicate a dynamic trend in their study [35]. The high phenotypic diversity among salmonids has provided an excellent research framework for understanding adaptive divergence and ecological speciation [36,37] and potentially contributed to the WGD in their common ancestor, an event known as Ss4R [38], about 95 Ma ago [39,40]. Thus, one of the most remarkable features of the evolutionary history of salmonids is their autopolyploid origin [41,42]. This makes salmon an ideal organism for studying the consequences of the genome and gene duplications [43], which are thought to have played a key role in creating the gene diversity and functional specialization found in modern vertebrates [44,45]. In this regard, the genome of the Atlantic salmon (Salmo salar) was chosen as the reference sequence for all salmonids due to its importance in the aquaculture industry and because so much research on this species had previously been undertaken at the genomic level [46]. Significant progress in understanding salmon biology has been made by sequencing two salmon genomes, as well as the genome of the northern pike Esox lucius, a sister line that has not undergone salmon-specific whole genome duplication [47]. By taking into account these discoveries, it can be concluded that genome reconstruction and comparative genomic analysis of various fish species have become a successful approach to understanding their evolution. In this regard, in our work, we present an updated and comprehensive review of recent achievements in the genomics of salmonids and highlight the key results of various comparative genomic approaches for their study.

2. Characterization of the Salmon Genome

2.1. WGD

As noted above, one of the most notable features of the evolutionary history of salmonids is their autopoliploid origin [48]. It has been suggested that whole genome duplication (WGD), which characterizes all inherited salmonids [48], occurred, according to various estimates, from 58–63 to 88–103 million years ago [40], resulting in a single tetraploid ancestor in salmonids, whose autotetraploidization has been termed an ‘autopolyploidization’ (Ss4R) event [38,49]. Rediploidization occurs after all WGD occurrences, resulting in additional genomic diversity [50].
After WGD in the same species (autopolyploidization), rediploidization entails a change during meiosis from polyvalent (tetraploid inheritance) to bivalent chromosomal mating (diploid inheritance) [38,51]. As a result, recombination between the four alleles stops, causing sequence divergence between duplicated genes on distinct chromosomes (ohnologues) [50,51]. This, in turn, creates new pathways of functional evolution, unlike before WGD [50,52,53]. However, genetic studies of organisms with relatively recent WGD can be hampered by the inability to distinguish alleles and sequences from the same chromosome from alleles and sequences from a duplicated chromosome [42]. Fortunately, approaches using gamete manipulation, whole-genome sequencing, and long-read sequencing have improved the ability of researchers to characterize duplicated regions. Linkage mapping to haploid and double haploid has facilitated the analysis of duplicated regions in salmonids [54,55]. Furthermore, long-read sequencing methods have enabled the assembly of large genomes with complicated duplication histories [56]. These technical developments have improved our understanding of the genomic architecture of species that have dynamically extended to multiple species after ancestral WGD in lineages such as salmonids [38,57,58].
Mutations that promote preferred bivalent mating during meiosis, such as structural rearrangements and mobile element insertions, are required for rediploidization [38,59]. The same rediploidization process does not occur in allopolyploids, where WGD emerges after hybridization between different species, proving immediate preferential bivalent mating of subgenomes originating from each parental species [60,61], but it can possibly occur whenever sequence similarity is sufficient to generate polyvalent pairs, as in segmental allopolyploids [62]. It is widely believed that WGD promotes evolutionary diversification through mechanisms that are still not fully understood [63]. Some authors have suggested that WGD, which is salmonid-specific, is followed by radiation exposure [64,65]. Autotetraploidization involves the spontaneous doubling of all chromosomes distinct from the other major WGD classes, allotetraploidization, which involves hybridization between different species.
It has been observed that interspecific hybridization in salmonids is a frequent phenomenon. Some publications describe hybrids between Arctic char S. alpinus and S. malma [66], Dolly Varden and brook trout S. confluentus [67], S. confluentus and S. fontinalis [68], Arctic char and river trout [69], mottled char S. leucomaenis and S. levanidovi [70], mallow and white-spotted char [71]. As a rule, hybrids might be identified accordingly to their external morphology [72,73,74]. Some researchers consider interspecific hybridization in loaches as a consequence of the formation and existence of so-called contact zones between species during post-glacial colonization [69]. Other authors attribute cases of interspecific hybridization to the anthropogenic transformation of fish habitats or the invasion of alien species [75,76,77,78,79]. Here, there is no reason to say that the system is practically in a pristine state [80]. More recently, Gruzdeva et al. (2018) used genetic marker analysis to find hybrids formed by female white-spotted char and male northern Dolly Varden [81].
Although the impact of WGDs on long-term processes such as organism diversification is unknown [82,83], findings demonstrate that WGDs cause numerous problems during mitosis and meiosis [84]. These concerns are reflected in the fact that most neopolyploid species revert to diploidy rapidly following duplication [85]. Campbell et al. (2019) reported that gene conservation is a prominent result of WGD in salmonids. Although approximately 10% of the rainbow trout genome is inherited tetrasomally, ~28% of conserved ohnologue pairs are found in tetrasomal regions, suggesting that tetrasomal inheritance inhibits neofunctionalization and molecular adaptation. Furthermore, both the function and expression of tetrasomal ohnologues are preserved [57].
Early research has revealed that the most common outcome of a duplicated gene is for it to become dysfunctional and lost [86]. Although most ohnologists’ ultimate fate is gene loss following WGD, some ohnologists remain [87,88]. Genes predisposed to haploinsufficiency may be more likely to be conserved as functional duplicates to counteract the potential negative consequences of mutations in a single copy, i.e., the dangerous duplicate hypothesis [89,90]. However, it has been suggested that the selection for duplicate conservation and prevention of haploinsufficiency is weak, limiting the potential of this explanation for duplicate conservation [91,92]. Nevertheless, duplicates involved in immune function, cell cycle regulation, transcriptional control, and cell signaling pathways are conserved during vertebrate evolution [93,94,95]. However, the enriched GO term analysis presented by Campbell et al. (2019) found limited evidence for genes associated with such functions in tetrasomal regions of the genome [57].
The underlying determinants of delayed rediploidization remain a mystery. Rediploidization may be severely inhibited in places with delayed rediploidization due to negative implications for specific genes or favorable benefits of keeping genes as tetraploid. Gene set enrichment and gene expression analysis, for example, may not resolve such possibilities but can give hints for comprehending what appears to be a complicated process. Consequently, selection for conservation and loss of duplicated genes is probably multifaceted when delayed rediploidization is involved [96]. For Ss4R, it is assumed that selection has maintained tetraploidy in SSR regions for tens of millions of years [97]. It’s also worth wondering why a second round of rediploidization was permitted. Because it is inextricably linked to the evolutionary origin of salmonid subfamilies, it is possible that the events leading to speciation coincided with a reduction in effective population size, reducing the efficiency of selection while also introducing the adverse effects of rediploidization [38,39,64]. Perhaps novel selected forces linked with early species diversification (e.g., related to the earliest formation of anadromy) affected selection during rediploidization and anthropogenic divergence in other ways owing to impacts on individual genes [98]. Long-read assemblages containing all salmonid species will provide a mechanistic understanding of the relationship between mobile element evolution, speciation, and lineage-specific rediploidization [99].

2.2. LORe and AORe

Previous research on salmonid fish has revealed that rediploidization happened at various stages of development for distinct parts of the genome following Ss4R, which dates back 88–103 million years [61]. After previous occurrences at the root of vertebrate [31] and bony fish evolution, Ss4R is the fourth WGD in salmonid evolutionary history [31,39,58,59,61]. Variable degrees of sequence divergence among large synthetic ohnologue blocks reflect the different times of rediploidization for duplicated regions left from Ss4R [38]. Although rediploidization occurred in the ancestors of all modern salmonids in large genomic regions, including several complete chromosomes, speciation occurred before rediploidization was completed in several large genomic segments [70]. Consequently, some duplicated chromosomal arms have very high sequence similarity (>95%) in all salmonid species [38,100,101,102] and experience anthropological divergence independently in three subfamilies of salmonids that diverged about 50 million years ago [70]. This process has been termed “lineage-specific ohnological resolution” (“LORe”) and is thought to be possible whenever an evolutionary transition from multivalent to bivalent mating [61,70].
Late rediploidization and LORe are crucial for phylogenetic reconstruction [70,100]. If global WGD expectations are not taken into account, LORe can easily be mistaken for lineage-specific (e.g., tandem) gene duplication during phylogenetic analyses [70]. In terms of the evolutionary significance of WGD events, it has been hypothesized that LORe promotes lineage-specific adaptation [70] and provides a viable explanation for the commonly observed temporal delays between a WGD event and subsequent species or phenotypic diversification regimes [103,104,105]. Many authors have noted the potential relevance of late rediploidization and LORe in divergent species after the discovery of these events in salmonids [40,64,104,105,106]. Despite the fact that salmonids are a well-established research system, many researchers’ understanding of rediploidization discoveries in this group of fishes remains fragmented as a result of whole genome sequence data and/or insufficient phylogenetic resolution in prior reconstructions.
Delayed rediploidization and LORe have yet to be unequivocally demonstrated beyond salmonids but have been proposed to track WGD events specific to bony fish (Ts3R) and WGD events at the base of vertebrates [70,71,106,107]. Several writers have shown that delayed and nested models of species diversity in the aftermath of WGD accord with LORe [64,104,105,107,108]. Given the large and growing number of high-quality genomic sequences in various eukaryotic lineages with a history of WGD, Gundappa et al. (2022) adapted their phylogenomic approach to rediploidization results for WGD events of comparable or earlier age for Ss4R, using synteny/collinearity to distinguish ohnologues from other duplicate genes. Because of the large evolutionary distances involved, genome alignment approaches would be inappropriate for studying earlier WGD events like Ts3R. The complex LORe patterns characteristic of large sections of the salmonid genome strongly violate the usual assumption for such methods that ohnologue divergence begins at a particular branch of the species tree [39].
To enhance the ability to reconstruct rediploidization dynamics during salmonid evolution, Gundappa et al. set out to reconstruct the rediploidization process after Ss4R and its results with significantly increased genomic and phylogenetic resolution compared with earlier studies, for which they created a high-quality genome sequence of Danube salmon (Hucho hucho) and developed a genome alignment approach to capture Ss4R-matched regions in several species. They sequenced the genome of a salmonid species with a highly interesting phylogenetic position and devised a whole-genome alignment method to catch ethnographic areas in genome assemblies previously released for many salmonid species. Using phylogenetic approaches, they were able to recreate whole-genome rediploidization dynamics, capturing two major waves of rediploidization as well as intricate histories of lineage-specific ohnologue divergence, the complexity of which rises with speciation history. The authors created multispecies alignments for a priori defined blocks of syntenic ohnologues left over from Ss4R [38] in two regions of the Hucho hucho genome where rediploidization was either ancestral to all salmonids (AORe regions with ancestral resolution [70]) or occurred after the separation of Salmoninae and Thymallinae. The alignments they obtained included species from Salmoninae and Thymallinae as well as a representative of Esociformes, a sister lineage of salmonids that diverged to Ss4R [38,40,70]. Their predecessor studies showed that about 25% of the Salmoninae genome underwent rediploidization after separation from Thymallinae [70,108].
Lien et al. (2016) showed that both tissue expression bias [38] and rediploidization time are important factors in ohnologue regulatory divergence in Ss4R ohnologues, with duplicates in SSR regions showing a higher correlation in tissue expression than AORe regions [70]. The results obtained by Gundappa et al. support these findings and show that expression level is an additional factor to be considered in the evolution of ohnologue regulation, thereby capturing differences in functional enrichment between ohnologues and singletons, which is consistent with past studies [109,110,111,112] but, in contrast to previous work [57,70], indicate general deviations in functional enrichment between Ss4R ohnologues with different rediploidization ages and suggest enrichment of a small number of unique functions in ohnologues from regions with different rediploidization ages [39].
In an earlier Ss4R study, all rainbow trout ohnologues classified as belonging to SSR regions were considered tetrasomic [113,114]. Gundappa et al. (2022) emphasize that their results require caution when confusing delayed rediploidization with no rediploidization [39]. As a result, tetrasomal heredity across many millions of years in rainbow trout areas designated as tetrasomal seems implausible [57]. Sequence similarity analysis using reference genomes by Blumstein et al. (2020) showed that all species possessed traits of residual tetrasomal heritability in seven homeologous pairs. By using linkage maps, the same seven homeologous pairs were found to be tetrasomal in Oncorhynchus [115,116], Salvelinus [117,118], Salmo [70], and probably Coregonus [119], which strongly suggests that tetravalent meiosis can and do form between these homeologs in all species examined to date. For several karyotypes, the evidence for residual tetrasomy differed between the linkage map and genome methods in Blumstein et al. (2020), with one categorized as tetrasomal in linkage mapping studies but not in genome analysis, and another categorized as tetrasomal in genome analysis but not in linkage maps [101]. This finding suggests that the frequency of diploidization in Salvelinus differs compared to other salmonids for this homeologous pair. A study of Arctic char found numerous duplicate markers in this karyotype [97]. The discovery that the karyotype is not tetrasomic, on the other hand, is most likely owing to differences in determining the level of residual tetraploidy between linkage mapping and genome assembly techniques. Linkage mapping in Oncorhynchus and Salvelinus consistently reveals the presence of tetrasomal inheritance [118,120], but genome analysis by Blumstein et al. (2020) and by Campbell et al. (2019) for rainbow trout found that this karyotype exhibits similarity to intermediate sequences consistent with disomic homeologs [59,101].
The last two genomes in the cell are usually different enough to split into two sets of bivalents during meiosis, which removes the incompatibility in mating between the hybridizing species before WGD [121]. Conversely, autotetraploidization results in four sets of chromosomes that initially diverge randomly during meiosis after WGD; bivalent mating must be restored before duplicate genes created by WGD can go beyond the allelic state [38,70,122]. In lake whitefish, aneuploidy has been documented across populations and historical circumstances [123,124]. Results from Blumstein et al. (2020) show that ambiguity in homologous relationships persists for at least five chromosome arms in all three coregonins [101]. This degree of ambiguity was much higher than that documented by Sutherland et al. (2017) in Salmo, Oncorhynchus, and Salvelinus, where there were only two ambiguities in these groups [117]. Whitefish appears to have many relatively small acrocentric chromosomes [125], some of which contain large numbers of duplicated loci, making it difficult to construct linkage maps compared with other salmonids [102,126].
Sutherland et al. (2017) investigated the history of fusions in various Oncorhynchus species and discovered that most species have numerous species-specific fusions (e.g., 17 species-specific fusions in pink salmon). Sutherland et al. (2017), on the other hand, researched only one species from the genera Coregonus and Salvelinus since that was all that was available at the time of publishing, and none from Thymallus [117]. The physiological consequence of proposed fusion histories is yet unknown, although it is a significant problem in distinguishing Thymallus, Salvelinus, and Coregonus spp. from many other salmonids. When high-quality genomes for a growing amount of salmon species become available, the accuracy with which it is possible to uncover changes in genome content and structure after whole genome duplication increases significantly. Additional data from genome sequencing studies, such as the European whitefish genome of De-Kayne et al. (2020), should allow for crucial additional research assessing genomic events and organization in species with distinct merging histories [127]. Using the northern pike’s chromosome numbering, these authors identified pairs of homeologous whitefish scaffolds and their associated ancestral chromosome. This should make future comparisons across the salmon family easier (as Blumstein et al. (2020) did) to examine the separate process of rediploidization in distinct salmon lineages. They also discovered a number of whitefish scaffolds with no known homeologs and demonstrated that some of these areas formed through the fusion of sequentially identical regions [128]. Diversity studies in salmoniforms are of essential scientific importance in order to comprehend the processes that produce and support such diversification, as well as to assist the conservation of this varied group. It will especially help to understand the genetic basis of their adaptability, including determining the degree of parallelism. In this regard, Table 1 reflects key data on the previously studied salmonid genomes (Table 1).
Many salmonid species have documented instances of residual tetrasomy and enhanced sequence similarity between homeologous chromosomes, implying that certain rediploidization might remain [38]. Although some species-specific differences in residual tetrasomal regions have been reported, tetraploidy appears to persist among salmonid species in the 7–8 homeologous pairs of chromosomes [117]. Although the evolution of the European grayling karyotype is otherwise comparable to that of most salmonids, such evolution is also observed in the European grayling genome assembly based on common linkage maps in the case of residual tetrasomal regions [129]. The evolutionary importance of persistent residual tetrasomy is unknown, but its existence in the European grayling ancestral genome demonstrates that tetrasomy is autonomous of the chromosomal fusions seen in other salmonids [38,125] rather than favors another cause. Therefore, we have elucidated the key characteristics of salmonid karyotypic evolution.

2.3. Mobile Elements in the Genome

Mobile elements could also play a key role in genome evolutionary processes [128]. Furthermore, these components may be significant in the processes of rediploidization, causing sequence divergence that would split homologs. Mechanistically, rediploidization involves the proliferation of mobile elements in the genome, which cause rearrangements that terminate polyvalent meiotic pairs, limiting anthropologic divergence [130]. Because bursts of mobile element activity are known throughout salmonid evolution [38], lineage-specific mobile element proliferation may be causally related to delayed lineage-specific rediploidization.
Unfortunately, the current generation of salmonid genomes does not allow such ideas to be tested because of their poor representation of mobile elements and genomic regions showing very recent rediploidization. In particular, it may be of interest to compare classes of mobile elements between organisms with very different genomic rearrangements, for example, between European grayling and Atlantic salmon. For example, Sävilammi et al. (2019) found that retrotransposons (class I mobile elements) are more abundant compared to DNA transposons (class II mobile elements) with a higher abundance of 1.7 and 1.3 times in the genomes of the European grayling and Atlantic salmon, respectively [129]. Differential accumulation of mobile elements between lineages may play a significant role in genome evolutionary processes, but the complexity of the underlying reasons for such differences must be explored to determine it. Elements specific to Atlantic salmon include DNA transposons, which together encompass 3.57% of the Atlantic salmon genome but are completely absent in European grayling. These mobile elements are one of the most prevalent types of mobile elements in salmonids, and they are thought to play a significant role in Atlantic salmon rediploidization [38]. The Copia-12 retrotransposon, which belongs to the Copia retrotransposon superfamily and has recently been proposed to have a role in chromosomal diversity and speciation in other bony fish, is one example of an Atlantic salmon-specific element [128].
Tc1-Mariner transposons, for example, DNA hAT transposons, are cut-and-insert elements with transposition processes that may actively drive genomic alternations as well as indirect pathways to make homologous recombination copies of elements [131]. The concentration of a specific mobile element in one of the two species could be interpreted as evidence of lineage-specific mobile element functioning. Understanding of mobile elements in the genome can give unique information to inspire future studies into the molecular mechanisms of these various genome evolutionary changes. More sensitive techniques, including such recurrent sequence grouping, may be employed to detect them since they may allow for a more accurate assessment of repeating element content [132]. Moreover, they may allow for a better understanding of repeat community structure and identification of key elements using network approaches [133], allowing for a more detailed investigation of repeat distribution dynamics among salmonids. It has been suggested that periods of impaired-purifying selection occurring in bottlenose populations is necessary to perpetuate the deleterious effects of chromosomal rearrangements [134]. Although chromosomal evolution may be initially chaotic, the impact of chromosome evolution on mutation and recombination frequencies might result in directed evolution [135] and phenotypic changes. Additionally, lineage-specific loss of duplicated gene copies after gene duplication [136] or possibly evolutionary divergence of expression, as observed in European grayling and Atlantic salmon [137], may contribute to the formation of new species.
The activity of mobile elements with lineage-specific changes, such as those observed between Atlantic salmon and European grayling, is a primary driver of genome evolution [138] and could have also been engaged in numerous genome evolutionary changes. Furthermore, it has been claimed that chromosomal inversions, such as those often seen in the European grayling genome, have a significant impact on the processes of adaptation and speciation [139]. For instance, they enhance genome sequence divergence between saltwater and freshwater clades of the stickleback Pungitius pungitius [140], as well as non-migratory and migratory Atlantic cod ecotypes (Gadus morhua) [141,142]. Computer models supported these findings and demonstrated how chromosomal inversions could hasten speciation, particularly if adaptability includes several genes with tiny individual adaption effects [143]. A study on domestic mice (Mus musculus domesticus) revealed the possibility of fast differentiation driven by Robbertson fusions [144].

2.4. Local Genome Features of Different Salmonids

The literature found makes it clear that the genome of many salmonids has been studied in fish inhabiting specific local water bodies and rivers, therefore possessing their own uniqueness. Typically, one study will focus on a single population inhabiting a particular water body. However, this tends to reveal both the distinctive characteristics of individual salmonid populations and their relatedness. This section, therefore, focuses on examples from studies of some of the local groups that inhabit local watercourses and rivers.
Kanjuh et al. (2020) set out to map coho populations in poorly studied and unexplored rivers in the Danube basin in Croatia and to study their molecular diversity, gaining insight into phylogeographic haplotype patterns using different molecular markers [145]. Kalayci et al. (2018) suggested that all brown trout populations belonging to the Danube lineage should have nominate species status [146]. This is consistent with the estimate that diversification (i.e., genetic differentiation) between lineages should be high enough for absolute reproductive isolation between them [147,148,149]. Conservative morphology and low levels of diversity between trout taxa from different phylogeographic lineages on continuous morphological traits also confirm their close evolutionary relationships [150]. Initially, low genetic variability was observed within the Danube lineage [151], but thanks to numerous studies covering many populations, records now show higher values of genetic diversity [152,153,154]. Simonovic et al. (2017) found that Da1 is the most widely distributed local haplotype of coho salmon [155]. However, an anthropogenic factor is also considered important for their spread beyond their natural range, primarily because of the long tradition of the stocking with factory fish of AT origin that exists in many countries [156]. Similarly, mainly due to the activities of fishermen, wild trout of the AT line has been introduced in suitable conditions in Croatia [157]. Thus, the stocking of AT line trout outside the territory of the Republic of Croatia has been recognised as the main reason for the loss of native genetic diversity of the genus Salmo spp. [150]. A disjunctive distribution of haplotype Da22 allowed Simonovic et al. (2017) to hypothesize that the coho salmon populations still possessing this haplotype are remnants of a once widely distributed population, probably of Late Miocene and Early Pliocene Ponto-Messinian age [155]. The same authors believe that the recent/modern molecular diversity of coho salmon in the area is the result of their dynamic evolution. In the middle section of the Una River, the Da22 haplotype was accompanied by the Da2 and At1 haplotypes, both of which were considered non-native and were probably introduced into the population by the stocking of factory-grown coho salmon [158].
The mixing of these home-run trout with local populations has caused introgressive hybridization and reduced the genetic integrity of local trout so that in some geographical areas, the original populations have been almost completely mixed or replaced by non-native trout [159,160]. This context of phenotypic plasticity and ecological adaptation has contributed to a confusing nomenclatural pattern of Italian native trout with many morphs described as species (or subspecies) whose validity is questionable [161]. Rossi et al. (2019) previously carried out work to identify and assess the genetic composition of residual local populations of Mediterranean trout from different sampling locations in the Italian Lazio region, for which they applied two molecular markers for the diagnostic origin and geographic origin of trout. The main similarity of Italian trout is that they are part of the Salmo trutta complex, defined on a molecular basis [156]. The results obtained by Rossi et al. confirm the harmful effects of the mass introduction of domestic trout into Italian rivers already observed in other geographical regions of Italy [159,160,161] or Latium [162], with some exceptions specific to some island territories [163,164]. The most frequent AT haplotype (At18) corresponds to the haplotype originally identified in samples from Norway, Denmark, and Spain [165,166]. These data show that the practice of stocking domestic trout has some effect, spreading among allochthonous genotypes [160]. Atlantic trout have almost completely displaced native Mediterranean trout at some sites or at least mixed with them [167].
Interesting results were also found in some regions of Russia. Gordeeva et al. (2018) analyzed sequence variability of the mtDNA control region (537–547 bp) in 25 populations of European Russia and Siberia [168]. Currently, the Siberian part of the C. Albula–C. sardinella range is relatively isolated from the European part; however, none of the studied genetic markers (allozymes and different mtDNA sequences) show interspecific differences between grouse and grouse populations. The introgression of northern Dolly Varden haplotypes (Bering phylogenetic group) in the Arctic char genome has been found in both North America [169,170] and Asia [171,172,173,174]. Furthermore, analysis of the nuclear sequences of rag1 did not reveal any substitutions strictly specific to C. albula or C. sardinella. Moreover, populations in which some fish have morphological and genetic features of C. vandesius, and others have features of cisco occur in a wide area covering water bodies of the Pechora River basin and adjacent regions of Eastern Europe [175]. The authors also discovered the haplotype BER10 in the lower Anabar River; this finding shifts the boundary of its distribution along the Siberian Arctic coast further westward. It should be noted that many other species native to northern ecosystems with a wide range were originally described in different parts of the range as separate taxonomic units [176,177]. However, a thorough examination of these animals’ morphological, ecological, and genetic characteristics, particularly those occupying overlapping subspecies/species ranges, frequently finds species with intermediary morphological and genetic characteristics, raising questions about their authenticity [178,179,180,181]. The presence of the SIB25 haplotype in Lake Siysk from the same area indicates that the ranges of Eurasian and Beringian phylogenetic groups overlap there, as in the lower Lena and Olenek. The introgressive hybridization with Dolly Varden could have occurred in the Pacific basin during one of the glacial maxima; from there, the Arctic loach with the Bering group haplotype dispersed along the Arctic Ocean coast during the subsequent climate warming. One can also acknowledge the former westward expansion of the chrysalis itself and its hybridization with the Arctic char in the Arctic regions of Siberia, with its subsequent disappearance. Either way, the evidence suggests that the Beringian cutthroat participated in the colonization of the eastern Arctic coast of Siberia, and the presence of only one haplotype over a vast area from Anabar to the Indigirka delta indicates that their ancestors passed through a narrow neck [168,182].
Thus, we have reviewed the main questions and points that have been asked in salmonid genome research to date (Table 1).
Table 1. Data on salmon genomes ever obtained. As we can assess, the great majority of studies focus on the mitochondrial genome.
Table 1. Data on salmon genomes ever obtained. As we can assess, the great majority of studies focus on the mitochondrial genome.
ViewGenomeGenome Data Acquisition ApproachCharacteristics and Features of the GenomeReference
S. alpinus erythrinusMitochondrialSanger sequencing and annotation by comparison with other sequences using Geneious R11
Genome structure determined using MEGA X
Size: 16,652 bp; two ribosomal RNA (rRNA) genes, 13 protein-coding genes, 22 tRNA genes; the overall base composition was 28.0% A, 26.4% T, 17.0% G, and 28.6% C, and the mitogenome GC content was 45.6%[183]
S. alpinus alpinusMitochondrialPrimers were designed using the mitoPrimer_V1 program; sequences
annotated by comparison with published charr mitogenome sequences using Geneious R11
Size: 16,655 and 16,657 bp; GC content: 45.6%; 24 SNPs. The highest was the variability of NADH dehydrogenase subunit genes (42.6%Table
of all variable sites).
[184]
Salvelinus taranetziMitochondrialThe sequenced fragments were de novo assembled into a complete mitochondrial genome and annotated by comparing with published genome sequences of charr using Geneious R11Size: 16,654 b.p.
The overall base composition was 28.0% A, 26.4% T, 28.6% C, and 17.0% G, with a slight A + T bias (54.5%).
Two substitutions were found in the control region and 12S rRNA. Other single nucleotide substitutions were found in common protein-coding sequences
The total sequence discrepancy (D xy) was 0.0011 ± 0.0002
[185]
Salvelinus sp. Mitochondrial Size 16,654 b.p.
The organization of the genome was identical to that of typical salmon genomes, including 2 rRNA genes, 13 protein-coding genes, 22 tRNA genes, a light chain origin of replication (OL), and a control region (CR). The overall base composition was 28.0% A, 26.4% T, 17.0% G, and 28.6% C, and the GC content was 45.6%.
[186]
Salvelinus boganidae and Salvelinus elgyticusMitochondrialLibraries were prepared using an Ion Plus Fragment Library Kit and unique adapters (Ion Xpress), and preliminary fragmentation of PCR products was performed on a Covaris M220 ultrasonicator. Libraries were sequenced on the Ion S5 platform (Thermo Fisher Scientific). Clean reads were assembled into contig with the Bowtie2 algorithm in Geneious R11
Mitogenome of S. taranezi was used as a reference for correct contig position and orientation
16,654 b.p. for s. elgyticus and 16,655 b.p. for s. boganidae. The gene arrangement, composition, and size are similar to s. Taranezi
The genomic organization is identical to typical salmon mitogenomes, including two rRNA genes, 13 protein-coding genes, 22 tRNA genes, a light chain origin of replication (OL), and a control region (CR).
[187]
Salvelinus levanidoviMitochondrialUsing Geneious R11, the sequenced fragments were combined into a full mitogenome and described by reference to previous char mitogenomes.The complete mitogenome of native S. levanidovi was 16,624 bp long.
the overall base composition was 28.1% A, 26.4% T, 28.6% C, and 16.9% G, with a slight A + T bias (54.5%)
[188]
T. thymallusMitochondrialThe consensus sequence was controlled and annotated using MitoAnnotator.Mitogenome has a total length of 16,660 bp and includes 13 protein-coding genes, 22 transfer RNA genes including 2 tRNA-Leu and 2 tRNA-Ser, two ribosomal RNA genes, and a control region following the standard vertebrate order.
Intergenic spaces and overlapping sequences were found.
Six coding genes have an incomplete codon stop: NADH2, COII, ATP6, NADH3, NADH4, and cytb. The base composition of the entire genome was 27.5% for A, 27.9% for T, 17.5% for G and 27.1% for C.
[189]
Sl. SvetovidovMitochondrialGenomes were matched using the MAFFT algorithm in Geneious, and maximum likelihood (ML) analysis was performed based on Tamura-Nei (TrN93) plus a nucleotide substitution gamma model.
The machine learning tree was built using the MEGA X software and validated by bootstrap analysis.
Genome length 16,655 bp. The overall base composition was 28.0% A, 26.4% T, 28.7% C, and 16.9% G with a slight A + T bias (54.5%) 18 single nucleotide differences and no difference in length between sequences MK695627, MK695628 and MK695629; only 13 substitutions were found in all sequences encoding the protein and five were found in the control region.[190]
O. masouMitochondrial Genome length 16,648 bp, with 13 protein-coding genes, two rRNA genes, 22 tRNA genes, a putative control region (CR), and one light chain origin of replication.
The overall base composition is 28.6% A, 26.8% T, 28.1% C, and 16.4% G, respectively, with a slight AT bias (55.4%).
[191]
S. fontinalis × S. malma sp. hybridMitochondrial Genome length 16,623 bp.
The overall base composition was 28.3% A, 26.5% T, 28.4% C and 16.8% G, with a slight A + T bias (54.8%)
two non-coding regions, the L-chain origin of replication was located between tRNA-Asn and tRNA-Cys, and the control region was located within tRNA-Pro and tRNA-Phe. Eight tRNAs and the ND6 gene were encoded on the L-chain, the rest were encoded on the H-chain
The complete mitogenome sequence contained 16s RNA and 12s RNA, which were located between tRNA-Phe and tRNA-Leu and were separated by the tRNA-Val gene.
[192]
S. malmaMitochondrial Genome length 16,652 bp.
The overall base composition was 28.1% A, 26.4% T, 28.5% C, and 17.0% G.
The 13 protein-coding genes code for 3808 amino acids in their entirety.
[193]
S. truttaGeneralSequencing with Illumina HiSeqX;
Assembly of the genome using a software package;
Gene annotation with Ensembl
Size: 2.37 Gbp-p; most of the assembly consists of 40 chromosomal pseudomolecules; 43,935 protein-coding and 4441 non-protein-coding genes[194]

3. Phylogenetic Characteristics of Salmon

The number of teleost species far outnumbers any other group of fish or any other vertebrate, and this has been associated with a whole genome duplication prior to their emission during the Cretaceous [27,195,196,197,198]. Salmonids present a unique opportunity to explore a range of evolutionary and ecological concepts, including the mechanisms of speciation [199,200], the evolution of complex life cycles [197,201,202,203,204], patterns of chromosomal evolution [101], and genome duplication [40], as well as the role of hybridization in evolution [205]. The WGD mentioned above has played an important role in the long-term evolutionary success of salmonids by providing lineage-specific physiological adaptations such as anadromy, therefore potentially promoting evolutionary diversification and speciation [70]. According to the same authors, most of the salmon genome has returned to a state of diploid inheritance before subfamily divergence [70]. In this regard, salmonids occupy a unique phylogenetic position compared to fish species whose genomes have been or are in the process of being sequenced since they belong to the Protacanthopterygii, the most primitive group of teleosts [206]. Thus, salmonids provide a key phylogenetic link between the evolution of teleost fish and the evolution of non-bony fish, as well as other vertebrates. Salmonid evolutionary connections have been the focus of considerable systematic and phylogenetic investigations for several decades [207,208,209,210,211,212,213,214,215,216,217,218,219,220,221,222,223]. The derivation of a reliable phylogeny for this group is important for comparative analysis of salmon adaptation [224,225,226], comparative genomics [227,228], and for assessing conservation priorities [38,47,215,229]. Population genomics is now commonly applied to salmonids without genome sequencing by using conserved synteny with rainbow trout or Atlantic salmon, for example [38,47,48,229,230,231].
Salmonidae is a phylogenetically distinct family that arose, according to the latest data, from 50.8 to 64 million years ago [40,232,233,234,235], and this is confirmed by morphological comparison data of 11 genera belonging to three subfamilies: Coregoninae (ciscoes, whitefish, and inconnu), Thymallinae (grayling) and Salmoninae (huchen, lenok, trout, char, and salmon) [211,236,237,238]. The most numerous of these, Salmoninae, includes seven genera distributed throughout the world: Brachymystax (flax), Hucho (huchen and taimen), Oncorhynchus (Pacific trout and salmon), Salmo (Atlantic salmon and trout), Parahucho, Salvethymus and Salvelinus [43,140,217]. There is strong evidence that each is a monophyletic clade, a natural group that includes all the descendants of its last common ancestor [75,239,240]. Coregoninae is divided into Coregonus, Prosopium, and Stenodus. The Thymallinae has one genus, Thymallus. Three of these genera (Parahucho, Salvethymus, and Stenodus) are monotypic, and their precise location in their respective subfamilies is being debated. Originally, Salmoninae phylogenetic rebuilding has permitted Parahucho perryi to also be found in a variety of positions within the group, including sibling Salvelinus [218,241], Salmo [158,223], and the rest of Salmoninae [232,242,243]. Based on these data, it has been suggested that Coregoninae is a sister group to the rest of the Salmonidae [217,244]. Recent genetic investigations of the subfamily Thymallinae utilizing genome-wide mitochondrial genome sequencing confirmed the relatedness of the three subfamilies Coregoninae, Thymallinae, and Salmoninae but left confusion regarding the evolutionary connections between the long-recognized subfamilies [244,245]. Other monotypic genera, such as Salvethymus in the Salmoninae and Stenodus in the Coregoninae, have distinct morphologies and karyotypes that set them apart from other genera in their respective subfamilies, although certain molecular data show that they possess not justify genus classification [185,218,236,241,243,246,247,248]. Comparison of currently obtained mitogenomes with 27 related-group mitogenomes available from GenBank, including the genera Salvelinus, Parahucho, and Salmo, indicates a close relationship of S. alpinus erythrinus to related species, S. taranetzi (and closely related taxa). The study of chromosome patterns served to decipher the pathways of evolution [249,250] and genome duplication [251,252]. A theory of tetraplodization of salmon as a result of exposure to radiation exposure between 25 and 100 million years ago was proposed; this period is a distinctive feature of the family [48,253].
Unclear issues concerning Salmonidae interactions could be clarified by expanding the sample size of both taxa and characteristics. So, Crête-Lafrenière et al. (2012) produced one of the most significant contributions to the study of salmon phylogeny, choosing 63 species, more than twice the amount in use by Stearley and Smith [217], and completing a complete morphological investigation to present. They hypothesized that increasing the sample of taxa could separate long branches and that this would allow for a more precise establishment of phylogenetic relationships [232]. Also, in their opinion, an increase in the sample of taxa could be useful in estimating parameters for models of molecular evolution [254,255] and various types of phylogenetic tests, including rooting analysis [256] estimating the time of divergence [257]. According to their results, Parahucho perryi is a sister group (Salvelinus, Oncorhynchus), which was confirmed later [100], and Stenodus and Salvethymus cannot be highlighted as independent genera [246,248,258]. Crête-Lafrenière et al. (2012) employed a variety of genes to infer salmon phylogeny, in addition to expanding the number of species studied. Single gene phylogenies are intrinsically restricted in their capacity to reliably establish taxonomic relationships and are prone to random mistakes. As a result, they integrated gene sequences into a “supermatrix” to boost phylogenetic signal and node support. However, simultaneous analysis of linked gene sequences should be treated with caution, as biases can interfere with phylogenetic inferences due to strongly supportive clades erroneously grouped based on multiple substitution artifacts (e.g., heterogeneity of nucleotide composition [259,260], change in speed at different sites [261,262] and an increase in the sample of taxa [254,255,263,264].

3.1. Subfamily Salmoninae

Salmoninae became a sister subfamily of the Coregonidae, some 35.6 mya. Hucho and Brachymystax were found to be sister genera with a common evolutionary lineage that diverged earlier than other members of the subfamily [265]. Another example is the positioning of the monotypic genus Parahucho as a sister genus of Salmo with an average divergence time of 21.9 Ma. This placement also appeared in Crespi and Fulton (2004) [223] as well as Alexandrou et al. (2013) [266] but in contrast to several other studies that have either grouped Parahucho with Salvelinus or simply as a sister group to the Oncorhynchus/Salvelinus clade or Oncorhynchus/Salmo clade [230,232,233,267]. Parahucho and Salvelinus share an evolutionary lineage with a common ancestor with Oncorhynchus, with the genus Salmo placed between these latter three genera (Oncorhynchus, Salvelinus, Parahucho) and others (Hucho, Brachymistax). In Gong et al. (2017), within each genus Salmoninae, the phylogenetic relationships between species differed slightly from those previously published [191,265]. Within the genus Salvelinus, for example, the phylogenetic relationships of Salvelinus fontinalis and S. leucomaensis appeared similar in Ma et al. (2016) [268], where both belong to different evolutionary lineages, but various in Sahoo et al. (2016) [269] and Balakirev et al. (2016) [270], where they shared an evolutionary lineage. To further verify the newly-defined sequences and establish their taxonomic status in Salmoninae, phylogenetic trees, including all Salmoninae species available in databases such as GenBank, were constructed based on maximum likelihood analysis [271].
Identifying the evolutionary relationships of monotypic genera of the subfamily Salmoninae is rather complicated and identifying the relationship between Salvelinus, Oncorhynchus and Salmo has been a source of debate. Monotypic genera are simply one source of ambiguity in determining evolutionary connections inside the Salmonidae subfamily. As a result of genetic investigations, the lengthy classification of Oncorhynchus and Salmo as sister species has indeed been substituted by the group Oncorhynchus and Salvelinus [220,223,272]. The greatest issues with phylogenetic reconstructing inside the genera are centered on three more species-specific taxa, Oncorhynchus, Salvelinus, and Coregonus. In molecular investigations of Oncorhynchus, uncertain and contentious [223,273,274], and many issues emerge concerning the connection between Pacific trout, which is disguised by repeated hybridization [238]. Likewise, Salvelinus species frequently hybridize, resulting in discrepancies in phylogenetic analyses [217,275,276].
The long-standing definition of Oncorhynchus and Salmo as related species was divided into the Oncorhynchus and Salvelinus groups based on genetic studies [220,223,238,272,277]. Ninua et al. (2018) studied the morphology and mitochondrial phylogeny of five nominate trout species from the Western Caucasus. According to their data, trout from the Black Sea and Caspian Sea basins represent a monophyletic evolutionary lineage (matrilineal clade) distinct from trout from other parts of western Eurasia, including those from the Atlantic, Mediterranean, and Indian Ocean basins [278]. Earlier, Mari et al. (2014) analyzed grayling from the Kama and Ural basins and discovered previously documented haplotypes that form a sister clade with Scandinavian haplotypes when compared to accessible published studies [279,280,281,282]. Analysis of the data they presented implies that the Caspian and Scandinavian clades separated 0.33–0.92 million years ago, based on a replacement rate of 0.5% every million years. Despite the short number of examined populations and the small sample size, the Caspian haplotype polymorphism is equivalent to that of European groups [280,283]. Individuals from the Bugurla and Barangulovka sample sites had considerable genetic similarities, indicating extensive gene flow between these two proximal streams, giving rise to a single population [282]. The basal diversification within Salmo that separated S. salar from the bull-trout lineage occurred, according to different authors, from 9.6 to 15.4 Ma [232,234,266,267,284]. The separation between the widespread European S. trutta and the recently described species S. ohridanus and S. obtusirostris occurred between 3 and 9 Ma, in the Late Miocene or Pliocene [40,232,267,285].

Salvelinus 

All populations of Arctic char, including Taranian char, form a monophyletic group. Thus, until a recent study, several forms of char were described as separate species [222,286,287]. It was established that three species of loaches of the genus Salvelinus inhabit Lake Elgygytgyn, two of which are considered endemics of this water body. The first species is the long-finned loach S. svetovidovi, originally described in a separate genus Salvethymus; moreover, it was assumed that its age is comparable to that of the lake itself [233,258,287,288,289]. Another endemic species of this lake is the smallmouth char S. elgyticus. A third species of charr, S. boganidae charr, was thought to be identical to the form of the same name from Taimyr [289]. Although there is very little data on nuclear gene comparisons between Taimyr and Chukotka boganidae, mtDNA data suggest a polyphyletic origin [290,291,292].
Gonen et al. (2015) presented the first phylogenetic relationships of salmonids based on a RAD sequencing dataset, including five salmonid species and 3050 loci in the analysis, and found that the number of identified orthologous Sbf I RAD loci decreased as the evolutionary distance between species increased: several thousand loci were preserved in five salmonid species (~50 million year divergence), and several hundred were preserved in more distantly related bony species (~100–100 million year divergence). Most (>70%) of the loci identified between more distantly related species were of genetic origin [293]. Later, Lecaudey et al. (2018) presented the first phylogeny of salmonid fishes based on a large RAD sequencing dataset with an extensive sample of family taxa [233].
The use of molecular genetic approaches enabled us to transcend the restrictions placed by S’s parallel diversity in the morphological features of S. alpinus [294]. Current ideas about the phylogenetic and phylogeographic structure of Arctic char are based on studies by Brunner et al. (2001) of the diversity of the mtDNA regulatory region in different populations across its range, as well as in Dolly Varden (S. malma), a similarly-related species from the North Pacific. Five phylogenetic haplotype groups were revealed, one (Beringian) associated to the northern form of Dolly Varden and four (Arctic, Atlantic, Siberian, and Acadian) associated with Arctic char. Extensive research into charr mtDNA has substantially extended our understanding of their genetic diversity and evolutionary linkages in the Pacific basin [70,184,246,294,295,296,297,298,299,300,301,302,303,304,305,306,307,308], in Eastern Siberia [171,173], and in the North American Arctic [170]. A revision of the systematics of the genus has been proposed, involving consideration of monophyletic groups identified by mtDNA analysis as separate species, in particular, S. taranetzi [300], which includes Asian and North American loach with Arctic group haplotypes [288,291,300].
Brunner et al. (2007) recognized loach phylogenetic haplotype groups as clusters with strong initial support in their phylogenetic haplotype trees [291]. Loaches from North America’s Arctic regions, the Canadian Arctic Archipelago, Greenland, and the Asian Taranian loach were included in the Arctic group; loaches from Quebec and Maine were included in the Acadian group; loaches from Siberia, Finland, and Spitsbergen were included in the Siberian group; and loaches from Northern Europe, the Alps, the British Isles, Iceland, Greenland, Newfoundland, and Labrador were included in the Atlantic. The latter two groups were discovered as being the most closely linked, and they were joined with the Acadian group to form the Atlantic-Siberian-Acadian supergroup. Haplotypes from the fifth (Beringian) group were discovered in Arctic char after being discovered in the northern Varden river, which is thought to be the consequence of the previous hybridization between the two species [169,170,171,173,294,296]. As new data accumulated, the differences between the Atlantic and Siberian groups became increasingly blurred. The insertion of additional Siberian haplotypes caused the latter to disintegrate, and bootstrap analysis revealed no strong support for Siberian haplotypes or their big combinations [171]. Moore et al. (2015) achieved a similar finding for both the Siberian and Atlantic groups [170]. Also, Yamamoto et al.’s (2014) tree, which includes just Siberian haplotypes released by Brunner et al. (2001) [291,301], is divided into two clusters. One cluster combines with the Atlantic group cluster, although with poor bootstrapping support. The discreteness of the Atlantic and Siberian clusters seen by Brunner et al. (2001) might be attributable to both a paucity of Siberian material and sequencing errors [170,287].
Some of the forms that originated in this region are reproductively isolated [287,309,310]. Boganidae (S. boganidae) and Dryagina’s loach (S. drjagini) are described as different species [222]. Thanks to this study, boganids from lakes Elgygygytgyn and Lama are attributed to different phylogenetic groups of the Arctic char lineage and are not considered as a single species, S. boganidae [70,287,288]. These forms are now strictly separated, just like the Boganidae S. boganidae and the smallmouth S. elgyticus [286]. The results of the RAG1 gene analysis also unambiguously point to the belonging of Lake Cherechenskii char to the Arctic char lineage. However, the two species of loaches did not form a supported clade on the phyloo graph; S. boganidae grouped with S. taranetzi from Chukotka, as previously suggested by some authors [287]. Oleinik et al. (2020) tested two alternative hypotheses about the origin of loaches in Lake Elgygygytgyn. The data presented are most consistent with the conclusion that two different in-time colonization events by ancestral lineages of Taranec shiners occurred in Lake Elgygytgyn during the postglacial periods. Despite the stability of the topology, contradictory signals or alternative phylogenetic histories were found within the Arctic phylogenetic group. Taking into account the previous genetic work of these species based on microsatellites [287], they suggested that the reticulations detected by the Neighbor-Net revealed several signals of hybridization events. The phylogenetic network showed that past hybridization between S. boganidae and S. elgyticus is possible, although no hybrids were observed between them [287]. S. boganidae also shows potential hybridization in the past with S. taranetzi. Among loaches of the genus Salvelinus, phylogenetic relationships and taxonomy are most problematic within the group that includes the Arctic loach S. alpinus, the trout S. malma, and closely related forms and species of loaches [284,288,311,312]. These two groups of char are often considered within two species complexes (S. alpinus complex and S. malma complex) [312] or one supercomplex S. alpinus-S. malma complex [243].
Gordeeva et al. (2018) found Holts with Atlantic subgroup haplotypes in Karelia, the Kola Peninsula, Novaya Zemlya, Polar Urals, and Taimyr. Their finds indicate a wide distribution of several Atlantic haplotypes. The Siberian subgroup borders the Arctic subgroup to the east of the Indigirka river basin; however, the location of this boundary and the extent to which their ranges overlap is uncertain. The Siberian subgroup is almost entirely represented by sedentary continental populations; anadromous and insular populations have not been noted so far, except for one anadromous population from the Novaya River, Khatanga Bay basin. Different sets of haplotypes are detected in different isolated regions of the Siberian group range. However, these locations are not used to categorize the haplotypes in the phylogenetic trees. Earlier, common haplotypes SIB8 and SIB10 were found in Transbaikalia and the Yana basin [171]. These findings support the shared origin and tight evolutionary relationships of the Siberian subgroup’s Arctic loaches from various parts of Siberia, as well as putative migrations between these locations during the Pleistocene glacial maximum, but not as extensive as in the Atlantic subgroup [168].
As can be seen, current ideas about the phylogeny of loaches are based on studying the variability of the mtDNA control region [171,291,301]. Oleinik et al. (2015) identified several phylogenetic groups uniting closely related species of loaches (Arctic, Atlantic, Siberian, Acadian, Beringian, Western Pacific, and Eastern Pacific). Their results show that specimens of S. boganidae and S. elgyticus belong to the aforementioned Arctic group, S. taranetzi. Originally described as a separate species, it later became synonymous with the Arctic char S. alpinus taranetzi [312]. There is still a lack of molecular data to support this view [313]. Most previous studies of S. taranetzi, along with other loaches, are limited to analyzing only short fragments of a few mitochondrial and nuclear genes [173]. Molecular data based on ten microsatellite loci and sequences of the mtDNA control region and cytochrome b gene [287], cytochrome b and cytochrome c oxidase I genes [232], and RAD sequencing support the assignment of Salvethymus to the genus Salvelinus [233]. In a phylogeny based on the mitogenomes of the loaches, Sl. svetovidov represents the latest branch that diverged after the main group of species (S. fontinalis, S. leucomaenis, S. levanidovi, S. namaycush). The authors in this study sequenced and described two complete mitochondrial genomes of S. taranetzi for further study and more precise phylogenetic analysis. S. levanidovi was phylogenetically placed alongside other loaches but showed substantial divergence from them. S. fontinalis, S. levanidovi, S. leucomaensis, and S. namaycus represent a basal group of species in terms of mitogenome diversity, each corresponding to a separate evolutionary lineage. As a result, Oleinik et al. (2020) revealed that S. levanidovi is highly correlated to the single origin of the genus Salvelinus [184].
To determine the taxonomic status of hybrid salmon, Zhang et al. (2019) reconstructed the phylogeny of this salmon stock with other natural salmon populations based on complete mitogenome sequences. The phylogenetic tree showed that hybrid salmon are related to Salvelinus and are more closely related to S. fontinalis, distinct from S. malma sp. The complete mitochondrial genome sequence of the S. fontinalis × S. malma hybrid provided an important data set for a better understanding of mitogenomic diversity and evolution of salmonid fishes, as well as new genetic markers for studying population genetics and species identification [192].
Analysis of the nucleotide sequences of the control region and the cytb mtDNA gene indicates that the “boganids” and smallmouth characters from Elgygygytgyn are young species whose origin is most likely related to the events of the last ice age [287]. Both species have Arctic haplogroup mtDNA haplotypes and are sister species that may have formed in the lake during the postglacial time. The results of the phylogenetic analysis of the nucleotide data of the mtDNA site ATPase6-NADH4L indicate an allopatric origin of the smallmouth and “richnid” loaches of Lake Elgygytgyn. This contradicts the results of the analysis of the CR and cytb genes, according to which their sympatric origin is more likely [287]. Despite the inconsistency in the available mtDNA data, data on the mtDNA control region [170,258,287,313] suggest that the level of nucleotide and haplotype diversity in the Taranec shin population in North America and Greenland is much lower than previously assumed [280] and slightly lower than in some Asian populations. Previously, based on an analysis of their own and published mtDNA data, the difference between the two studies is that the authors included smallmouth and "richnid" char from Elgygytgyn in the first subgroup [258], and in contrast, the ATPase6-NADH4L mtDNA site data divided the two species into different subgroups [289].
Introgressive hybridization across distinct lineages or taxa has been widely established for Salvelinus species as a result of previous secondary interactions [71,233,314,315], such as between S. malma and S. alpinus, which recently diverged from each other, about 1.5 million years ago, and have overlapping current distributions [315]. The presence of parallel edges in the network, on the other hand, does not guarantee hybridization, simply the probability of hybridization [316]. The results of the phylogenetic analysis by Oleinik et al. (2021) are consistent with earlier studies based on mtDNA fragments [232,246,287,301]. In comparison to earlier studies’ phylogenetic trees, their phylogenetic tree based on the mitogenomes of the char species grew more strong and more trustworthy. Longer DNA sequences have previously been proven to give an acceptable resolution of relatively high connections in fish [317]. A study by Oleinik et al. (2021) additionally showed that mitogenomes could make up a reliable phylogenetic tree and resolve relationships between closely related species of loaches [318]. On the basis of morphological similarity, the boganids of Chukotka and Taimyr were attributed to the same species [286,306]. However, the results of mtDNA analysis indicate a probable polyphyletic origin of allopatric populations of Boganidae [287,308]. Boganida charr’s morphological resemblance in this example might be a symptom of parallelism, which is defined as the autonomous development of shared traits among closely linked taxa during evolution [287]. Data on the RAG1 gene confirm the monophyletic origin of Arctic charr and Dolly Varden complexes [242]. Results from Lecaudey et al. (2018) specifically show hybridization signals between S. malma from the Bering clade and S. alpinus. An earlier study showed S. alpinus individuals with introgressed haplotypes from the Bering clade of S. malma along the eastern coast of Siberia to be parapatric [171]. In a very recent study, mtDNA introgression from S. malma malma to S. taranetzi near the Sea of Okhotsk, due to which representatives of the first species did not remain in this location completely [318]. In addition, postglacial hybridization between different glacial lineages of S. alpinus has been demonstrated to survive in separate refugia [170].
The mitochondrial gene relationship between S. confluentus and the rest of the S. alpinus group has not been revealed by nuclear genome data, confirming lots of evidence referring to introgression of the Arctic char’s mitochondrial genome in this species that might conceal the sister taxon’s affection with S. leucomaenis [319]. With the emphasis on the subfamily Salmoninae and the extensive coverage of taxa of the genus Salvelinus, the topology recovered from >28,000 loci is well-defined and well-supported by all methods used, providing clear answers to several phylogenetic uncertainties revealed by conflicting results of previous studies.

3.2. The Subfamily Coregoninae

Coregonus 

In whitefish, species identification in phylogenetic studies is complicated by parallel evolution [237,320,321,322], phenotypic plasticity [323], trophic population variability [324,325], and hybridization [326,327]. The greatest difficulty in identifying stems from the variability of the two phenotypes that distinguish whitefish and grayling, each of which was formerly thought to be a monophyletic subgenus [238]. Thus, within the subfamily Coregonidae about 33.1 million years ago, the genus Prosopium showed an evolutionary lineage that had previously deviated from the Stenodus/Coregonus lineage. Coregonus nasus and C. chadary had a common ancestor in the same evolutionary lineage that previously diverged from the other species [328,329]. After them, C. clupeaformes and C. autumnalis also showed a common and divergent evolutionary lineage [269,330]. Phylogenetic analysis using a maximum likelihood tree among 42 complete mitogenomes of the Salmonidae family and one Danio rerio sequence showed 99% sequence similarity to the S. trutta genome [269]. Nevertheless, Baikal omul, C. migratorius, possesses morphological characteristics that suggest it is whitefish, although it is more strongly linked to whitefish based on molecular similarities [331,332]. S. trutta fario L. has been phylogenetically placed with other S. trutta species, with obvious divergence [330]. Hence, the whitefish could be classed as a paraphyletic clade based on morphological categorization.
Graylings are also a paraphyletic group, according to molecular evidence [333], with the smallest grayling, C. sardinella, becoming more strongly linked to whitefish than other graylings. Previous research has demonstrated that C. huntsmani has a unique evolutionary lineage [334,335], although it does not keep this status in the genus. Various research studies have found parallels between Stenodus and Coregonus [232,336,337]. Crête-Lafrenière et al. (2012) found slightly different and reversing results. First, C. huntsmani was recognized as a sister species to the other Coregonus species; second, Stenodus is probably a separate genus rather than a subgenus within Coregonus. Species such as Coregonus artedi, Coregonus hoyi, Coregonus kiyi, Coregonus nigripinnis, and Coregonus zenethicus, according to their data, are not related as previously thought [338,339,340,341], while Coregonus pollan and Coregonus fallalis, by contrast, tend to be more conspecific [232].
The only species of another distinct lineage that have not left the lake after separating from the aforementioned two are Baikal indigenous whitefish, lake omul (Coregonus migratorius), and bottom lake whitefish (Coregonus baicalensis). In contrast to Europe and America, where morphologically similar species pairs are the consequence of postglacial secondary contacts between glacial isolates, sympatric limnetic-benthic divergence was repeated here several times within the same body of water over a long geological period due to Pleistocene fluctuations. According to cytb mtDNA, phylogenetic analyses indicated the interspecific linkages described for Baikal whitefish [342]. Comparative genetic ranges across the three lines of responsibility: Baikal endemics, North American lake whitefish, and European whitefish, imply comparable ages [343].
Another interesting representative of whitefish is muksun. The revealed low level of genetic differentiation of muksun and whitefish, along with the available literature data, indicates that they belong to the same species, C. lavaretus. Borovikova and Budin’s (2020) analysis of the genetic variability of muksun from the Khatanga River points to its polyphyletic origin. The presence of common haplotypes, on the one hand, indicates their divergence from a common ancestor within the region under study; on the other hand, each form includes representatives of phylogenetic lines originating outside it. However, analysis of the haplotype network suggests the presence of several large mtDNA phylogenetic lineages. The molecular genetic analysis clarifies the phylogenetic relationships of high-tip and low-tip muksun of the Khatanga River, as well as with other Coregonus species. It also allowed us to clarify the origin of these forms in the basin of this river. It turned out that the degree of genetic differentiation between the forms is low and does not exceed the intraspecific level. It is important that the intrapopulation differentiation of muksun is comparable to the level of differences between it and whitefish: 0.2–0.3 and 0.2–0.5%, respectively. It should be noted that the differentiation level of 0.5% is also typical for different whitefish populations. Besides, it should be noted that the polyphyletic origin of low stamen and high stamen density forms of muksun does not allow us to refer them to different taxa and raise their status even to the subspecies level. Obviously, muksun is characterized by independent morphogenesis in different water systems, as in the case of European whitefish [344,345]. The phylogenetic kinship of whitefish and muksun is evidenced by the presence of the same common sequence variants of all three genetic polymorphism markers used in this study (ND1, COI, and ITS1) [346].
The species of roach (C. peled) is genetically similar to the aforementioned vendace. Peled differs from C. albula and C. sardinella, first of all, by the number of chromosomes. Differences in the number and chromosomal arrangement of the nucleus organizer regions between the vendace and the peled confirm that they are separate species [347]. C. peled’s emergence as a species may be an instance of recent chromosomal differentiation that enabled reproductive isolation from C. albula and C. sardinella. The most compelling evidence that C. peled is a distinct species is the qualitative difference between its nuclear genome and that of C. albula/C. sardinella [348]. Nevertheless, for the sea bass (genus Sebastes), a similar scenario exists: S. mentella Travin, 1951 and S. marinus (Linnaeus, 1758) (S. norvegicus (Ascanius, 1772)) might be separated from each other [349]. As previously stated, the specimen discovered in Lake Sobachye was classified as C. sardinella morphologically but carried the C. peled haplotype N42 [350].
Individual phylogenetic lineages might well be readily identified using median connective networks that indicate the taxon’s evolutionary history, although cladograms (dendrograms) are frequently useless. These species, like C. peled, C. albula, and C. sardinella, were not distinguishable by barcoding; nonetheless, haplotype analysis of another mtDNA sequence, D-loop, revealed that they were separate species [351]. The nd1 sequence distinguishes C. peled from C. albula and C. sardinella, but its nd1 haplotypes form a single unique branch, unlike the nd1 haplotypes of both cisco and lesser cisco. Furthermore, we cannot rule out the possibility that this species is a descendent of hybridization between C. sardinella and C. peled, which might have taken place in the past because hybridization is prevalent in whitefish even in natural reservoirs [352,353]. Therefore, Borovikova and Artamonova (2021) observe that in the instance of mtDNA sequences that do not generally recombine, the most significant criteria is their monophyletic or polyphyletic origin rather than the difference between haplotype sets [180].

3.3. The Subfamily Tymallinae

The subfamily Tymallinae originated about 29.5 Ma [265]. The T. burejensis / T. tugarinae clade was the first to diverge, next by T. yaluensis, another that included T. thymallus, and the last one that included all other species of the genus. Despite these topological variations, T. arcticus and were sister species, as recently proposed by Liu et al. (2016) and Balakirev, Romanov, and Ayala (2017) [354,355]. Ma et al. (2016) showed that mitochondrial genomes could be a powerful marker for determining phylogeny within Thymallinae. Their study confirmed that Yalu grayling should be synonymous with Amur grayling (Thymallus grubii) at the whole mitogenome level [268]. Other studies addressing both the nuclear mitogenome [61] and the salmonid nuclear genome [232,268] proposed that Thymallinae and Coregoninae are sister subfamilies, while others studying mitochondrial [245,267] and mitonuclear DNA [230] have proposed Thymallinae and Salmoninae as sister subfamilies. Horreo’s (2017) study with the mitogenome of 46 different salmonid species contributed much to the aspect of phylogeny construction by mitogenome, asserts Coregoninae and Salmoninae as sister groups in the family Salmonidae, and provides new insight into the phylogenetic relationships between genera and species in this family, including the molecular nodes [265]. The Thymallus sequences are divided into two monophyletic groups. The first (including the newly sequenced mitogen) includes four nominal species: T. thymallus, T. brevirostris, T. Arcticus, and T. baicalolenensis. The second includes three nominal species T. tugarinae, T. grubii, and T. yaluensis [189].
Despite a significant amount of effort acknowledging various elements of salmonid phylogeny, several of their phylogenetic connections and evolutionary history remain unknown. The degree of resolution of these concerns varies at different levels of biological organization, from the proper position of the salmonid tree root to the significance of introgression in species or subspecies classifications [218] and systematic constraints (which include completely inadequate taxon or gene sampling) [223,224,225]. In future studies, the focus of this study on the genus Salvelinus should also be on the genera Salmo and Thymallus. In the case of Salmo, there is still considerable uncertainty regarding the evolutionary history of several known taxa, such as Salmo marmoratus (marbled trout) and S. obtusirostris (soft-snout trout), as well as S. carpio (carpione) [356] and other larger phenotypes over the entire range of the S. trutta species complex, all of which may have been involved in important hybridization events. The genus Thymallus requires a comprehensive molecular study in both East and Central Asia. This is due to the comparatively high species diversity in East Asia and the somewhat exceptional link between contemporary taxonomy and phenotypic diversity in Central Asia. More thorough full-genome investigations of individual taxa like Salvelinus, Salmo, and Coregonus could give highly helpful information on both evolutionary radiation processes and the distinguishing features of individual taxa for all salmonids.
Thus, observing trends in salmonid genetics and genomics research, we can notice that molecular genetic and phylogenetic studies reveal an understanding of the history of salmonids as a family. This allows us to identify salmonid ancestors and establish intervals of genetic variation in Salmonidae. The complexity of the duplicated salmonid genome causes difficulties in its study but, on the other hand, increases the duplicability and genetic diversity of this family. The use of genomic techniques in commercial fisheries is mainly viable and cost-effective; nevertheless, the conversion of genomic data into management strategies has halted [17,357].

4. Conclusions

Thus, by observing trends in salmonid genetics and genomics research, we can conclude that molecular genetic and phylogenetic studies reveal an understanding of the evolutionary history of salmonids as a family. This allows us to identify salmonid ancestors and establish intervals of genetic variation in Salmonidae. The complex duplicated genome of salmonids poses a challenge to study, but on the other hand, it increases the plasticity and genetic diversity of this family. Applying genomic approaches to fisheries management is mainly feasible and cost-effective [17]. Salmonid species provide exciting possibilities to research speciation and adaptation mechanisms within an ecological and evolutionary context. In particular, it provides an opportunity to study the effect of hybridization and genome duplication on the evolution of species. Thus, Thompson et al. (2020) discovered that the complicated migratory phenotype is the product of a single gene area, which will aid in the conservation and recovery of Chinook salmon [358]. In formulating fisheries management rules and recommendations, genomic technologies and their capacities for recognizing species, establishing management units, and monitoring natural resources should be thoroughly explored [359,360,361].

Funding

The work was financially supported by the Ministry of Science and Higher Education of the Russian Federation, grant no. 075-15-2021-1037 (internal no. 15.BRK.21.0001).

Institutional Review Board Statement

Not applicable.

Informed Consent Statement

Not applicable.

Conflicts of Interest

The authors declare no conflict of interest. The funders had no role in the design of the study; in the collection, analyses, or interpretation of data; in the writing of the manuscript; or in the decision to publish the results.

References

  1. Ahrens, R.N.M.; Allen, M.S.; Walters, C.; Arlinghaus, R. Saving Large Fish through Harvest Slots Outperforms the Classical Minimum-length Limit When the Aim Is to Achieve Multiple Harvest and Catch-related Fisheries Objectives. Fish Fish. 2020, 21, 483–510. [Google Scholar] [CrossRef] [Green Version]
  2. Lawson, R.M. Economics of Fisheries Development; Bloomsbury Academic: London, UK, 2013; pp. 121–255. [Google Scholar]
  3. Davidson, W.S.; Koop, B.F.; Jones, S.J.; Iturra, P.; Vidal, R.; Maass, A.; Jonassen, I.; Lien, S.; Omholt, S.W. Sequencing the Genome of the Atlantic Salmon (Salmo salar). Genome Biol. 2010, 11, 403. [Google Scholar] [CrossRef] [Green Version]
  4. Cooke, S.J. "The Behavior and Ecology of Pacific Salmon and Trout" by Thomas P. Quinn [Book Review]. Can. Field Nat. 2005, 119, 147. [Google Scholar] [CrossRef] [Green Version]
  5. Erdem, Ö.A.; Alkan, B.; Dinçer, M.T. Comparison on Nutritional Properties of Wild and Cultured Brown Trout and Atlantic Salmon. Ege J. Fish. Aquat. Sci. 2020, 37, 37–41. [Google Scholar]
  6. Colombo, S.M.; Mazal, X. Investigation of the Nutritional Composition of Different Types of Salmon Available to Canadian Consumers. J. Agric. Food Res. 2020, 2, 100056. [Google Scholar] [CrossRef]
  7. Yoon, H.D.; Karaulova, E.P.; Shulgina, L.V.; Yakush, E.V.; Mok, J.S.; Lee, S.S.; Xie, C.; Kim, J.G. Nutritional Value and Bioactive Properties of Enzymatic Hydrolysates Prepared from the Livers of Oncorhynchus keta and Oncorhynchus gorbuscha (Pacific Salmon). Fish. Aquat. Sci. 2015, 18, 13–20. [Google Scholar] [CrossRef]
  8. Abdollahi, M.; Undeland, I. Structural, Functional, and Sensorial Properties of Protein Isolate Produced from Salmon, Cod, and Herring by-Products. Food Bioproc. Technol. 2018, 11, 1733–1749. [Google Scholar] [CrossRef] [Green Version]
  9. Esaiassen, M.; Jensen, T.K.; Edvinsen, G.K.; Otnæs, C.H.A.; Ageeva, T.N.; Mæhre, H.K. Nutritional Value and Storage Stability in Commercially Produced Organically and Conventionally Farmed Atlantic Salmon (Salmo salar L.) in Norway. Appl. Food Res. 2022, 2, 100033. [Google Scholar] [CrossRef]
  10. Khalili Tilami, S.; Sampels, S. Nutritional Value of Fish: Lipids, Proteins, Vitamins, and Minerals. Rev. Fish. Sci. Aquac. 2018, 26, 243–253. [Google Scholar] [CrossRef]
  11. Opheim, M.; Šližytė, R.; Sterten, H.; Provan, F.; Larssen, E.; Kjos, N.P. Hydrolysis of Atlantic Salmon (Salmo salar) Rest Raw Materials—Effect of Raw Material and Processing on Composition, Nutritional Value, and Potential Bioactive Peptides in the Hydrolysates. Process Biochem. 2015, 50, 1247–1257. [Google Scholar] [CrossRef]
  12. Tocher, D.R. Omega-3 Long-Chain Polyunsaturated Fatty Acids and Aquaculture in Perspective. Aquaculture 2015, 449, 94–107. [Google Scholar] [CrossRef]
  13. Calder, P.C. Marine Omega-3 Fatty Acids and Inflammatory Processes: Effects, Mechanisms and Clinical Relevance. Biochim. Biophys. Acta (BBA)-Mol. Cell Biol. Lipids 2015, 1851, 469–484. [Google Scholar] [CrossRef] [PubMed]
  14. Hindar, K.; de Leniz, C.G.; Koljonen, M.-L.; Tufto, J.; Youngson, A.F. Fisheries Exploitation. In The Atlantic Salmon; Blackwell Publishing Ltd.: Oxford, UK, 2007; pp. 299–324. [Google Scholar]
  15. Samy, J.K.A.; Mulugeta, T.D.; Nome, T.; Sandve, S.R.; Grammes, F.; Kent, M.P.; Lien, S.; Våge, D.I. SalmoBase: An Integrated Molecular Data Resource for Salmonid Species. BMC Genom. 2017, 18, 482. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  16. Pomeroy, R.; Ureta, B.E.B.; Solis, D.; Johnston, R.J. Bioeconomic Modelling and Salmon Aquaculture: An Overview of the Literature. Int. J. Environ. Pollut. 2008, 33, 485. [Google Scholar] [CrossRef] [Green Version]
  17. Bernatchez, L.; Wellenreuther, M.; Araneda, C.; Ashton, D.T.; Barth, J.M.I.; Beacham, T.D.; Maes, G.E.; Martinsohn, J.T.; Miller, K.M.; Naish, K.A.; et al. Harnessing the Power of Genomics to Secure the Future of Seafood. Trends Ecol. Evol. 2017, 32, 665–680. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  18. Dudgeon, C.L.; Blower, D.C.; Broderick, D.; Giles, J.L.; Holmes, B.J.; Kashiwagi, T.; Krück, N.C.; Morgan, J.A.T.; Tillett, B.J.; Ovenden, J.R. A Review of the Application of Molecular Genetics for Fisheries Management and Conservation of Sharks and Rays. J. Fish Biol. 2012, 80, 1789–1843. [Google Scholar] [CrossRef]
  19. Li, Y.-H.; Wang, H.-P. Advances of Genotyping-by-Sequencing in Fisheries and Aquaculture. Rev. Fish Biol. Fish. 2017, 27, 535–559. [Google Scholar] [CrossRef]
  20. Valenzuela-Quiñonez, F. How Fisheries Management Can Benefit from Genomics? Brief. Funct. Genom. 2016, 15, 352–357. [Google Scholar] [CrossRef] [Green Version]
  21. Aparicio, S.; Chapman, J.; Stupka, E.; Putnam, N.; Chia, J.; Dehal, P.; Christoffels, A.; Rash, S.; Hoon, S.; Smit, A.; et al. Whole-Genome Shotgun Assembly and Analysis of the Genome of Fugu rubripes. Science 2002, 297, 1301–1310. [Google Scholar] [CrossRef] [Green Version]
  22. Heather, J.M.; Chain, B. The Sequence of Sequencers: The History of Sequencing DNA. Genomics 2016, 107, 1–8. [Google Scholar] [CrossRef] [Green Version]
  23. Vij, S.; Kuhl, H.; Kuznetsova, I.S.; Komissarov, A.; Yurchenko, A.A.; van Heusden, P.; Singh, S.; Thevasagayam, N.M.; Prakki, S.R.S.; Purushothaman, K.; et al. Chromosomal-Level Assembly of the Asian Seabass Genome Using Long Sequence Reads and Multi-Layered Scaffolding. PLoS Genet. 2016, 12, e1005954. [Google Scholar]
  24. Tine, M.; Kuhl, H.; Gagnaire, P.-A.; Louro, B.; Desmarais, E.; Martins, R.S.T.; Hecht, J.; Knaust, F.; Belkhir, K.; Klages, S.; et al. European Sea Bass Genome and Its Variation Provide Insights into Adaptation to Euryhalinity and Speciation. Nat. Commun. 2014, 5, 5770. [Google Scholar] [CrossRef] [PubMed]
  25. Bradbury, I.R.; Hubert, S.; Higgins, B.; Borza, T.; Bowman, S.; Paterson, I.G.; Snelgrove, P.V.R.; Morris, C.J.; Gregory, R.S.; Hardie, D.C.; et al. Parallel Adaptive Evolution of Atlantic Cod on Both Sides of the Atlantic Ocean in Response to Temperature. Proc. R. Soc. B Biol. Sci. 2010, 277, 3725–3734. [Google Scholar] [CrossRef] [Green Version]
  26. Lu, L.; Zhao, J.; Li, C. High-Quality Genome Assembly and Annotation of the Big-Eye Mandarin Fish (Siniperca knerii). G3 Genes Genomes Genet. 2020, 10, 877–880. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  27. Yuan, D.; Chen, X.; Gu, H.; Zou, M.; Zou, Y.; Fang, J.; Tao, W.; Dai, X.; Xiao, S.; Wang, Z. Chromosomal Genome of Triplophysa bleekeri Provides Insights into Its Evolution and Environmental Adaptation. Gigascience 2020, 9, giaa132. [Google Scholar] [CrossRef] [PubMed]
  28. Wang, K.; Wang, J.; Zhu, C.; Yang, L.; Ren, Y.; Ruan, J.; Fan, G.; Hu, J.; Xu, W.; Bi, X.; et al. African Lungfish Genome Sheds Light on the Vertebrate Water-to-Land Transition. Cell 2021, 184, 1362–1376.e18. [Google Scholar] [CrossRef]
  29. Warren, W.C.; Boggs, T.E.; Borowsky, R.; Carlson, B.M.; Ferrufino, E.; Gross, J.B.; Hillier, L.; Hu, Z.; Keene, A.C.; Kenzior, A.; et al. A Chromosome-Level Genome of Astyanax mexicanus Surface Fish for Comparing Population-Specific Genetic Differences Contributing to Trait Evolution. Nat. Commun. 2021, 12, 1447. [Google Scholar] [CrossRef]
  30. Du, K.; Stöck, M.; Kneitz, S.; Klopp, C.; Woltering, J.M.; Adolfi, M.C.; Feron, R.; Prokopov, D.; Makunin, A.; Kichigin, I.; et al. The Sterlet Sturgeon Genome Sequence and the Mechanisms of Segmental Rediploidization. Nat. Ecol. Evol. 2020, 4, 841–852. [Google Scholar] [CrossRef] [Green Version]
  31. Bernardi, G.; Wiley, E.O.; Mansour, H.; Miller, M.R.; Orti, G.; Haussler, D.; O’Brien, S.J.; Ryder, O.A.; Venkatesh, B. The Fishes of Genome 10 K. Mar. Genom. 2012, 7, 3–6. [Google Scholar] [CrossRef]
  32. Koepfli, K.-P.; Paten, B.; O’Brien, S.J. The Genome 10K Project: A Way Forward. Annu. Rev. Anim. Biosci. 2015, 3, 57–111. [Google Scholar] [CrossRef] [Green Version]
  33. NCBI—Genome. Available online: https://www.ncbi.nlm.nih.gov/data-hub/genome/?taxon=8015 (accessed on 18 October 2022).
  34. Utter, F. Population Genetics, Conservation and Evolution in Salmonids and Other Widely Cultured Fishes: Some Perspectives over Six Decades. J. Fish Biol. 2004, 65, 323–324. [Google Scholar] [CrossRef]
  35. Bernatchez, L.; Renaut, S.; Whiteley, A.R.; Derome, N.; Jeukens, J.; Landry, L.; Lu, G.; Nolte, A.W.; Østbye, K.; Rogers, S.M.; et al. On the Origin of Species: Insights from the Ecological Genomics of Lake Whitefish. Philos. Trans. R. Soc. B Biol. Sci. 2010, 365, 1783–1800. [Google Scholar] [CrossRef] [PubMed]
  36. Lien, S.; Koop, B.F.; Sandve, S.R.; Miller, J.R.; Kent, M.P.; Nome, T.; Hvidsten, T.R.; Leong, J.S.; Minkley, D.R.; Zimin, A.; et al. The Atlantic Salmon Genome Provides Insights into Rediploidization. Nature 2016, 533, 200–205. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  37. Macqueen, D.J.; Johnston, I.A. A Well-Constrained Estimate for the Timing of the Salmonid Whole Genome Duplication Reveals Major Decoupling from Species Diversification. Proc. R. Soc. B Biol. Sci. 2014, 281, 20132881. [Google Scholar] [CrossRef] [Green Version]
  38. Vasil’ev, V.P. Mechanisms of Polyploid Evolution in Fish: Polyploidy in Sturgeons. In Biology, Conservation and Sustainable Development of Sturgeons; Springer: Dordrecht, The Netherlands, 2009; pp. 97–117. [Google Scholar]
  39. Limborg, M.T.; Seeb, L.W.; Seeb, J.E. Sorting Duplicated Loci Disentangles Complexities of Polyploid Genomes Masked by Genotyping by Sequencing. Mol. Ecol. 2016, 25, 2117–2129. [Google Scholar] [CrossRef] [Green Version]
  40. Waples, R.S.; Hendry, A.P. Special Issue: Evolutionary Perspectives on Salmonid Conservation and Management. Evol. Appl. 2008, 1, 183–188. [Google Scholar] [CrossRef]
  41. Formenti, G.; Rhie, A.; Balacco, J.; Haase, B.; Mountcastle, J.; Fedrigo, O.; Brown, S.; Capodiferro, M.R.; Al-Ajli, F.O.; Ambrosini, R.; et al. Complete Vertebrate Mitogenomes Reveal Widespread Repeats and Gene Duplications. Genome Biol. 2021, 22, 120. [Google Scholar] [CrossRef]
  42. Robinson-Rechavi, M.; Boussau, B.; Laudet, V. Phylogenetic Dating and Characterization of Gene Duplications in Vertebrates: The Cartilaginous Fish Reference. Mol. Biol. Evol. 2004, 21, 580–586. [Google Scholar] [CrossRef]
  43. Cañestro, C.; Albalat, R.; Irimia, M.; Garcia-Fernàndez, J. Impact of Gene Gains, Losses and Duplication Modes on the Origin and Diversification of Vertebrates. Semin. Cell Dev. Biol. 2013, 24, 83–94. [Google Scholar] [CrossRef]
  44. Rondeau, E.B.; Minkley, D.R.; Leong, J.S.; Messmer, A.M.; Jantzen, J.R.; von Schalburg, K.R.; Lemon, C.; Bird, N.H.; Koop, B.F. 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. [Google Scholar] [CrossRef] [Green Version]
  45. Allendorf, F.W.; Thorgaard, G.H. Tetraploidy and the Evolution of Salmonid Fishes. In Evolutionary Genetics of Fishes; Springer: Boston, MA, USA, 1984; pp. 1–53. [Google Scholar]
  46. Glasauer, S.M.K.; Neuhauss, S.C.F. Whole-Genome Duplication in Teleost Fishes and Its Evolutionary Consequences. Mol. Genet. Genom. 2014, 289, 1045–1060. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  47. Conant, G.C.; Wolfe, K.H. Turning a Hobby into a Job: How Duplicated Genes Find New Functions. Nat. Rev. Genet. 2008, 9, 938–950. [Google Scholar] [CrossRef]
  48. Furlong, R.F.; Holland, P.W.H. Were Vertebrates Octoploid? Philos. Trans. R Soc. Lond. B Biol. Sci. 2002, 357, 531–544. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  49. Innan, H.; Kondrashov, F. The Evolution of Gene Duplications: Classifying and Distinguishing between Models. Nat. Rev. Genet. 2010, 11, 97–108. [Google Scholar] [CrossRef] [PubMed]
  50. Kassahn, K.S.; Dang, V.T.; Wilkins, S.J.; Perkins, A.C.; Ragan, M.A. Evolution of Gene Function and Regulatory Control after Whole-Genome Duplication: Comparative Analyses in Vertebrates. Genome Res. 2009, 19, 1404–1418. [Google Scholar] [CrossRef] [Green Version]
  51. Brieuc, M.S.O.; Waters, C.D.; Seeb, J.E.; Naish, K.A. A Dense Linkage Map for Chinook Salmon (Oncorhynchus tshawytscha) Reveals Variable Chromosomal Divergence after an Ancestral Whole Genome Duplication Event. G3 Genes Genomes Genet. 2014, 4, 447–460. [Google Scholar] [CrossRef] [Green Version]
  52. Waples, R.K.; Seeb, L.W.; Seeb, J.E. Linkage Mapping with Paralogs Exposes Regions of Residual Tetrasomic Inheritance in Chum Salmon (Oncorhynchus keta). Mol. Ecol. Resour. 2016, 16, 17–28. [Google Scholar] [CrossRef]
  53. Kyriakidou, M.; Tai, H.H.; Anglin, N.L.; Ellis, D.; Strömvik, M.V. Current Strategies of Polyploid Plant Genome Sequence Assembly. Front. Plant. Sci. 2018, 9, 1660. [Google Scholar] [CrossRef]
  54. Campbell, M.A.; Hale, M.C.; McKinney, G.J.; Nichols, K.M.; Pearse, D.E. Long-Term Conservation of Ohnologues Through Partial Tetrasomy Following Whole-Genome Duplication in Salmonidae. G3 Genes Genomes Genet. 2019, 9, 2017–2028. [Google Scholar]
  55. Lappin, F.M.; Shaw, R.L.; Macqueen, D.J. Targeted Sequencing for High-Resolution Evolutionary Analyses Following Genome Duplication in Salmonid Fish: Proof of Concept for Key Components of the Insulin-like Growth Factor Axis. Mar. Genom. 2016, 30, 15–26. [Google Scholar] [CrossRef] [Green Version]
  56. Luo, J.; Chai, J.; Wen, Y.; Tao, M.; Lin, G.; Liu, X.; Ren, L.; Chen, Z.; Wu, S.; Li, S.; et al. From Asymmetrical to Balanced Genomic Diversification during Rediploidization: Subgenomic Evolution in Allotetraploid Fish. Sci. Adv. 2020, 6, eaaz7677. [Google Scholar] [CrossRef] [PubMed]
  57. Mason, A.S.; Wendel, J.F. Homoeologous Exchanges, Segmental Allopolyploidy, and Polyploid Genome Evolution. Front. Genet. 2020, 11, 1014. [Google Scholar] [CrossRef] [PubMed]
  58. Parey, E.; Louis, A.; Montfort, J.; Guiguen, Y.; Roest Crollius, H.; Berthelot, C. An Atlas of Fish Genome Evolution Reveals Delayed Rediploidization Following the Teleost Whole-Genome Duplication. Genome Res. 2022, 32, 1685–1697. [Google Scholar] [CrossRef] [PubMed]
  59. Kuhl, H.; Du, K.; Schartl, M.; Kalous, L.; Stöck, M.; Lamatsch, D.K. Equilibrated Evolution of the Mixed Auto-/Allopolyploid Haplotype-Resolved Genome of the Invasive Hexaploid Prussian Carp. Nat. Commun. 2022, 13, 4092. [Google Scholar] [CrossRef] [PubMed]
  60. Soltis, P.S.; Marchant, D.B.; van de Peer, Y.; Soltis, D.E. Polyploidy and Genome Evolution in Plants. Curr. Opin. Genet. Dev. 2015, 35, 119–125. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  61. van de Peer, Y.; Mizrachi, E.; Marchal, K. The Evolutionary Significance of Polyploidy. Nat. Rev. Genet. 2017, 18, 411–424. [Google Scholar] [CrossRef]
  62. Soltis, P.S.; Soltis, D.E. Plant Genomes: Markers of Evolutionary History and Drivers of Evolutionary Change. Plants People Planet 2021, 3, 74–82. [Google Scholar] [CrossRef]
  63. Singh, P.P.; Isambert, H. OHNOLOGUES v2: A Comprehensive Resource for the Genes Retained from Whole Genome Duplication in Vertebrates. Nucleic Acids Res. 2019, 48, D724–D730. [Google Scholar]
  64. Ahmad, S.F.; Jehangir, M.; Srikulnath, K.; Martins, C. Fish Genomics and Its Impact on Fundamental and Applied Research of Vertebrate Biology. Rev. Fish Biol. Fish. 2022, 32, 357–385. [Google Scholar] [CrossRef]
  65. Robertson, F.M.; Gundappa, M.K.; Grammes, F.; Hvidsten, T.R.; Redmond, A.K.; Lien, S.; Martin, S.A.M.; Holland, P.W.H.; Sandve, S.R.; Macqueen, D.J. Lineage-Specific Rediploidization Is a Mechanism to Explain Time-Lags between Genome Duplication and Evolutionary Diversification. Genome Biol. 2017, 18, 111. [Google Scholar] [CrossRef] [Green Version]
  66. Berthelot, C.; Brunet, F.; Chalopin, D.; Juanchich, A.; Bernard, M.; Noël, B.; Bento, P.; da Silva, C.; Labadie, K.; Alberti, A.; et al. The Rainbow Trout Genome Provides Novel Insights into Evolution after Whole-Genome Duplication in Vertebrates. Nat. Commun. 2014, 5, 3657. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  67. May-McNally, S.L.; Quinn, T.P.; Taylor, E.B. Low Levels of Hybridization between Sympatric Arctic Char (Salvelinus alpinus) and Dolly Varden Char (Salvelinus malma) Highlights Their Genetic Distinctiveness and Ecological Segregation. Ecol. Evol. 2015, 5, 3031–3045. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  68. Baxter, J.S.; Taylor, E.B.; Devlin, R.H.; Hagen, J.; McPhail, J.D. Evidence for Natural Hybridization between Dolly Varden (Salvelinus malma) and Bull Trout (Salvelinus confluentus) in a Northcentral British Columbia Watershed. Can. J. Fish. Aquat. Sci. 1997, 54, 421–429. [Google Scholar] [CrossRef]
  69. DeHaan, P.W.; Schwabe, L.T.; Ardren, W.R. Spatial Patterns of Hybridization between Bull Trout, Salvelinus confluentus, and Brook Trout, Salvelinus fontinalis in an Oregon Stream Network. Conserv. Genet. 2010, 11, 935–949. [Google Scholar] [CrossRef]
  70. Radchenko, V.I.; Mathisen, O.A. Distribution, Growth, and Feeding of Sockeye Salmon in the Western Bering Sea. Trans. Am. Fish. Soc. 2004, 133, 606–621. [Google Scholar] [CrossRef]
  71. Yamamoto, S.; Kitano, S.; Maekawa, K.; Koizumi, I.; Morita, K. Introgressive Hybridization between Dolly Varden Salvelinus Malma and White-Spotted Charr Salvelinus leucomaenis on Hokkaido Island, Japan. J. Fish Biol. 2006, 68, 68–85. [Google Scholar] [CrossRef]
  72. Araujo, H.A.; Duguid, W.D.P.; Withler, R.; Supernault, J.; Schulze, A.D.; Mckenzie, J.L.; Pellett, K.; Beacham, T.D.; Jonsen, K.; Gummer, A. Chinook and Coho Salmon Hybrids Linked to Habitat and Climatic Changes on Vancouver Island, British Columbia. Ecol. Evol. 2021, 11, 16874–16889. [Google Scholar] [CrossRef]
  73. Fraser, T.W.K.; Hansen, T.J.; Sambraus, F.; Fjelldal, P.G. Vertebral Deformities in Interspecific Diploid and Triploid Salmonid Hybrids. J. Fish Biol. 2021, 98, 1059–1070. [Google Scholar] [CrossRef]
  74. Fukui, S.; May-McNally, S.L.; Taylor, E.B.; Koizumi, I. Maladaptive Secondary Sexual Characteristics Reduce the Reproductive Success of Hybrids between Native and Non-Native Salmonids. Ecol. Evol. 2018, 8, 12173–12182. [Google Scholar] [CrossRef] [Green Version]
  75. Redenbach, Z.; Taylor, E.B. Evidence for historical introgression along a contact zone between two species of char (Pisces: Salmonidae) in northwestern North America. Evolution 2002, 56, 1021–1035. [Google Scholar] [CrossRef]
  76. Jerry, D.R.; Raadik, T.A.; Cairns, S.C.; Baverstock, P.R. Evidence for Natural Interspecific Hybridization between the Australian Bass (Macquaria novemaculeata) and Estuary Perch (M. colonorum). Mar. Freshw. Res. 1999, 50, 661. [Google Scholar] [CrossRef]
  77. Montanari, S.R.; Hobbs, J.-P.A.; Pratchett, M.S.; van Herwerden, L. The Importance of Ecological and Behavioural Data in Studies of Hybridisation among Marine Fishes. Rev. Fish Biol. Fish. 2016, 26, 181–198. [Google Scholar] [CrossRef]
  78. Chafin, T.K.; Douglas, M.R.; Martin, B.T.; Douglas, M.E. Hybridization Drives Genetic Erosion in Sympatric Desert Fishes of Western North America. Heredity 2019, 123, 759–773. [Google Scholar] [CrossRef] [PubMed]
  79. Rahman, M.A.; Lee, S.-G.; Yusoff, F.M.; Rafiquzzaman, S.M. Hybridization and Its Application in Aquaculture. In Sex Control in Aquaculture; John Wiley & Sons, Ltd.: Chichester, UK, 2018; pp. 163–178. [Google Scholar]
  80. Fraser, D.J. How Well Can Captive Breeding Programs Conserve Biodiversity? A Review of Salmonids. Evol. Appl. 2008, 1, 535–586. [Google Scholar] [CrossRef]
  81. Gruzdeva, M.A.; Kuzishchin, K.V.; Semenova, A.V.; Ponomareva, E.V.; Volkov, A.A.; Pavlov, D.S. A Rare Case of Permanent Introgressive Hybridization in Char of the Genus Salvelinus (Salmonidae: Salmoniformes) in the Utkholok River, Western Kamchatka. Russ. J. Mar. Biol. 2018, 44, 442–451. [Google Scholar] [CrossRef]
  82. Mayrose, I.; Zhan, S.H.; Rothfels, C.J.; Magnuson-Ford, K.; Barker, M.S.; Rieseberg, L.H.; Otto, S.P. Recently Formed Polyploid Plants Diversify at Lower Rates. Science 2011, 333, 1257. [Google Scholar] [CrossRef] [Green Version]
  83. Tank, D.C.; Eastman, J.M.; Pennell, M.W.; Soltis, P.S.; Soltis, D.E.; Hinchliff, C.E.; Brown, J.W.; Sessa, E.B.; Harmon, L.J. Nested Radiations and the Pulse of Angiosperm Diversification: Increased Diversification Rates Often Follow Whole Genome Duplications. New Phytol. 2015, 207, 454–467. [Google Scholar] [CrossRef] [Green Version]
  84. Landis, J.B.; Soltis, D.E.; Li, Z.; Marx, H.E.; Barker, M.S.; Tank, D.C.; Soltis, P.S. Impact of Whole-genome Duplication Events on Diversification Rates in Angiosperms. Am. J. Bot. 2018, 105, 348–363. [Google Scholar] [CrossRef] [Green Version]
  85. Gerstein, A.C.; Otto, S.P. Ploidy and the Causes of Genomic Evolution. J. Hered. 2009, 100, 571–581. [Google Scholar] [CrossRef] [Green Version]
  86. Ortíz-Barrientos, D.; Counterman, B.A.; Noor, M.A.F. Gene Expression Divergence and the Origin of Hybrid Dysfunctions. Genetica 2006, 129, 71–81. [Google Scholar] [CrossRef]
  87. Schneider, H.; Liu, H.-M.; Chang, Y.-F.; Ohlsen, D.; Perrie, L.R.; Shepherd, L.; Kessler, M.; Karger, D.N.; Hennequin, S.; Marquardt, J.; et al. Neo- and Paleopolyploidy Contribute to the Species Diversity of Asplenium -the Most Species-Rich Genus of Ferns. J. Syst Evol. 2017, 55, 353–364. [Google Scholar] [CrossRef] [Green Version]
  88. Mandáková, T.; Joly, S.; Krzywinski, M.; Mummenhoff, K.; Lysak, M.A. Fast Diploidization in Close Mesopolyploid Relatives of Arabidopsis. Plant. Cell 2010, 22, 2277–2290. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  89. Singh, R.D.; Banerjee, J.; Arora, A. Prebiotic Potential of Oligosaccharides: A Focus on Xylan Derived Oligosaccharides. Bioact. Carbohydr. Diet. Fibre 2015, 5, 19–30. [Google Scholar] [CrossRef]
  90. Casanova-Sáez, R.; Candela, H.; Micol, J.L. Combined Haploinsufficiency and Purifying Selection Drive Retention of RPL36a Paralogs in Arabidopsis. Sci. Rep. 2015, 4, 4122. [Google Scholar] [CrossRef]
  91. Pires, J.C.; Conant, G.C. Robust Yet Fragile: Expression Noise, Protein Misfolding, and Gene Dosage in the Evolution of Genomes. Annu. Rev. Genet. 2016, 50, 113–131. [Google Scholar] [CrossRef]
  92. Picolo, F.; Grandchamp, A.; Piégu, B.; Rolland, A.D.; Veitia, R.A.; Monget, P. Genes Encoding Teleost Orthologs of Human Haploinsufficient and Monoallelically Expressed Genes Remain in Duplicate More Frequently Than the Whole Genome. Int. J. Genom. 2021, 2021, 9028667. [Google Scholar] [CrossRef]
  93. Pires-daSilva, A.; Sommer, R.J. The Evolution of Signalling Pathways in Animal Development. Nat. Rev. Genet. 2003, 4, 39–49. [Google Scholar] [CrossRef]
  94. Berckmans, B.; de Veylder, L. Transcriptional Contro.ol of the Cell Cycle. Curr. Opin. Plant. Biol. 2009, 12, 599–605. [Google Scholar] [CrossRef]
  95. Ning, X.; Wang, R.; Li, X.; Wang, S.; Zhang, M.; Xing, Q.; Sun, Y.; Wang, S.; Zhang, L.; Hu, X.; et al. Genome-Wide Identification and Characterization of Five MyD88 Duplication Genes in Yesso Scallop (Patinopecten Yessoensis) and Expression Changes in Response to Bacterial Challenge. Fish. Shellfish Immunol 2015, 46, 181–191. [Google Scholar] [CrossRef]
  96. de Smet, R.; Adams, K.L.; Vandepoele, K.; van Montagu, M.C.E.; Maere, S.; van de Peer, Y. Convergent Gene Loss Following Gene and Genome Duplications Creates Single-Copy Families in Flowering Plants. Proc. Natl. Acad. Sci. USA 2013, 110, 2898–2903. [Google Scholar] [CrossRef] [Green Version]
  97. Smith, S.R.; Normandeau, E.; Djambazian, H.; Nawarathna, P.M.; Berube, P.; Muir, A.M.; Ragoussis, J.; Penney, C.M.; Scribner, K.T.; Luikart, G.; et al. A Chromosome-anchored Genome Assembly for Lake Trout (Salvelinus namaycush). Mol. Ecol. Resour 2022, 22, 679–694. [Google Scholar] [CrossRef] [PubMed]
  98. Nugent, C.M.; Easton, A.A.; Norman, J.D.; Ferguson, M.M.; Danzmann, R.G. A SNP Based Linkage Map of the Arctic Charr (Salvelinus alpinus) Genome Provides Insights into the Diploidization Process After Whole Genome Duplication. G3 Genes Genomes Genet. 2017, 7, 543–556. [Google Scholar] [CrossRef] [Green Version]
  99. Colgan, T.J.; Moran, P.A.; Archer, L.C.; Wynne, R.; Hutton, S.A.; McGinnity, P.; Reed, T.E. Evolution and Expression of the Immune System of a Facultatively Anadromous Salmonid. Front. Immunol. 2021, 12, 568729. [Google Scholar] [CrossRef] [PubMed]
  100. Warren, I.A.; Naville, M.; Chalopin, D.; Levin, P.; Berger, C.S.; Galiana, D.; Volff, J.-N. Evolutionary Impact of Transposable Elements on Genomic Diversity and Lineage-Specific Innovation in Vertebrates. Chromosome Res. 2015, 23, 505–531. [Google Scholar] [CrossRef] [PubMed]
  101. Campbell, M.A.; Buser, T.J.; Alfaro, M.E.; López, J.A. Addressing Incomplete Lineage Sorting and Paralogy in the Inference of Uncertain Salmonid Phylogenetic Relationships. PeerJ 2020, 8, e9389. [Google Scholar] [CrossRef]
  102. Blumstein, D.M.; Campbell, M.A.; Hale, M.C.; Sutherland, B.J.G.; McKinney, G.J.; Stott, W.; Larson, W.A. Comparative Genomic Analyses and a Novel Linkage Map for Cisco (Coregonus artedi) Provide Insights into Chromosomal Evolution and Rediploidization Across Salmonids. G3 Genes Genomes Genet. 2020, 10, 2863–2878. [Google Scholar] [CrossRef]
  103. De-Kayne, R.; Feulner, P.G.D. A European Whitefish Linkage Map and Its Implications for Understanding Genome-Wide Synteny Between Salmonids Following Whole Genome Duplication. G3 Genes Genomes Genet. 2018, 8, 3745–3755. [Google Scholar] [CrossRef] [Green Version]
  104. Clark, J.W.; Puttick, M.N.; Donoghue, P.C.J. Origin of Horsetails and the Role of Whole-Genome Duplication in Plant Macroevolution. Proc. R. Soc. B Biol. Sci. 2019, 286, 20191662. [Google Scholar] [CrossRef]
  105. Carretero-Paulet, L.; van de Peer, Y. The Evolutionary Conundrum of Whole-genome Duplication. Am. J. Bot. 2020, 107, 1101–1105. [Google Scholar] [CrossRef]
  106. Clark, J.W.; Donoghue, P.C.J. Whole-Genome Duplication and Plant Macroevolution. Trends Plant Sci. 2018, 23, 933–945. [Google Scholar] [CrossRef] [Green Version]
  107. Rozenfeld, C.; Blanca, J.; Gallego, V.; García-Carpintero, V.; Herranz-Jusdado, J.G.; Pérez, L.; Asturiano, J.F.; Cañizares, J.; Peñaranda, D.S. De Novo European Eel Transcriptome.e Provides Insights into the Evolutionary History of Duplicated Genes in Teleost Lineages. PLoS ONE 2019, 14, e0218085. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  108. van de Peer, Y.; Ashman, T.-L.; Soltis, P.S.; Soltis, D.E. Polyploidy: An Evolutionary and Ecological Force in Stressful Times. Plant. Cell 2021, 33, 11–26. [Google Scholar] [CrossRef] [PubMed]
  109. Li, X.; Guo, B. Substantially Adaptive Potential in Polyploid Cyprinid Fishes: Evidence from Biogeographic, Phylogenetic and Genomic Studies. Proc. R. Soc. B Biol. Sci. 2020, 287, 20193008. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  110. Devlin, R.H.; Biagi, C.A.; Sakhrani, D.; Fujimoto, T.; Leggatt, R.A.; Smith, J.L.; Yesaki, T.Y. An Assessment of Hybridization Potential between Atlantic and Pacific Salmon. Can. J. Fish. Aquat. Sci. 2022, 79, 670–676. [Google Scholar] [CrossRef]
  111. Parey, E.; Louis, A.; Cabau, C.; Guiguen, Y.; Roest Crollius, H.; Berthelot, C. Synteny-Guided Resolution of Gene Trees Clarifies the Functional Impact of Whole-Genome Duplications. Mol. Biol. Evol. 2020, 37, 3324–3337. [Google Scholar]
  112. Inoue, J.; Sato, Y.; Sinclair, R.; Tsukamoto, K.; Nishida, M. Rapid Genome Reshaping by Multiple-Gene Loss after Whole-Genome Duplication in Teleost Fish Suggested by Mathematical Modeling. Proc. Natl. Acad. Sci. USA 2015, 112, 14918–14923. [Google Scholar] [CrossRef] [Green Version]
  113. Pasquier, J.; Braasch, I.; Batzel, P.; Cabau, C.; Montfort, J.; Nguyen, T.; Jouanno, E.; Berthelot, C.; Klopp, C.; Journot, L.; et al. Evolution of Gene Expression after Whole-Genome Duplication: New Insights from the Spotted Gar Genome. J. Exp. Zool. B Mol. Dev. Evol. 2017, 328, 709–721. [Google Scholar] [CrossRef] [Green Version]
  114. Gillard, G.B.; Grønvold, L.; Røsæg, L.L.; Holen, M.M.; Monsen, Ø.; Koop, B.F.; Rondeau, E.B.; Gundappa, M.K.; Mendoza, J.; Macqueen, D.J.; et al. Comparative Regulomics Supports Pervasive Selection on Gene Dosage Following Whole Genome Duplication. Genome Biol. 2021, 22, 103. [Google Scholar] [CrossRef]
  115. Allendorf, F.W.; Bassham, S.; Cresko, W.A.; Limborg, M.T.; Seeb, L.W.; Seeb, J.E. Effects of Crossovers Between Homeologs on Inheritance and Population Genomics in Polyploid-Derived Salmonid Fishes. J. Hered. 2015, 106, 217–227. [Google Scholar] [CrossRef] [Green Version]
  116. May, B.; Delany, M.E. Meiotic Models to Explain Classical Linkage, Pseudolinkage, and Chromosomal Pairing in Tetraploid Derivative Salmonid Genomes: II. Wright Is Still Right. J. Hered. 2015, 106, 762–766. [Google Scholar] [CrossRef] [Green Version]
  117. Kodama, M.; Brieuc, M.S.O.; Devlin, R.H.; Hard, J.J.; Naish, K.A. Comparative Mapping Between Coho Salmon (Oncorhynchus kisutch) and Three Other Salmonids Suggests a Role for Chromosomal Rearrangements in the Retention of Duplicated Regions Following a Whole Genome Duplication Event. G3 Genes Genomes Genet. 2014, 4, 1717–1730. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  118. McKinney, G.J.; Pascal, C.E.; Templin, W.D.; Gilk-Baumer, S.E.; Dann, T.H.; Seeb, L.W.; Seeb, J.E. Dense SNP Panels Resolve Closely Related Chinook Salmon Populations. Can. J. Fish. Aquat. Sci. 2020, 77, 451–461. [Google Scholar] [CrossRef]
  119. Sutherland, B.J.G.; Rico, C.; Audet, C.; Bernatchez, L. Sex Chromosome Evolution, Heterochiasmy, and Physiological QTL in the Salmonid Brook Charr Salvelinus fontinalis. G3 Genes Genomes Genet. 2017, 7, 2749–2762. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  120. Mérot, C.; Stenløkk, K.S.R.; Venney, C.; Laporte, M.; Moser, M.; Normandeau, E.; Árnyasi, M.; Kent, M.; Rougeux, C.; Flynn, J.M.; et al. Genome Assembly, Structural Variants, and Genetic Differentiation between Lake Whitefish Young Species Pairs (Coregonus sp.) with Long and Short Reads. Mol. Ecol. 2022, 1–20. [Google Scholar] [CrossRef]
  121. Larson, W.A.; Limborg, M.T.; McKinney, G.J.; Schindler, D.E.; Seeb, J.E.; Seeb, L.W. Genomic Islands of Divergence Linked to Ecotypic Variation in Sockeye Salmon. Mol. Ecol. 2017, 26, 554–570. [Google Scholar] [CrossRef]
  122. Blasio, F.; Prieto, P.; Pradillo, M.; Naranjo, T. Genomic and Meiotic Changes Accompanying Polyploidization. Plants 2022, 11, 125. [Google Scholar] [CrossRef]
  123. Morgan, C.; Zhang, H.; Henry, C.E.; Franklin, F.C.H.; Bomblies, K. Derived Alleles of Two Axis Proteins Affect Meiotic Traits in Autotetraploid Arabidopsis Arenosa. Proc. Natl. Acad. Sci. USA 2020, 117, 8980–8988. [Google Scholar] [CrossRef] [Green Version]
  124. Dion-Côté, A.-M.; Symonová, R.; Ráb, P.; Bernatchez, L. Reproductive Isolation in a Nascent Species Pair Is Associated with Aneuploidy in Hybrid Offspring. Proc. R. Soc. B Biol. Sci. 2015, 282, 20142862. [Google Scholar] [CrossRef]
  125. Dion-Côté, A.-M.; Symonová, R.; Lamaze, F.C.; Pelikánová, Š.; Ráb, P.; Bernatchez, L. Standing Chromosomal Variation in Lake Whitefish Species Pairs: The Role of Historical Contingency and Relevance for Speciation. Mol. Ecol. 2017, 26, 178–192. [Google Scholar] [CrossRef]
  126. Gagnaire, P.-A.; Pavey, S.A.; Normandeau, E.; Bernatchez, L. The genetic architecture of reproductive isolation during speciation-with-gene-flow in lake whitefish species pairs assessed by RAD sequencing. Evolution 2013, 67, 2483–2497. [Google Scholar] [CrossRef]
  127. De-Kayne, R.; Zoller, S.; Feulner, P.G.D. A de Novo Chromosome-level Genome Assembly of Coregonus sp. “Balchen”: One Representative of the Swiss Alpine Whitefish Radiation. Mol. Ecol. Resour. 2020, 20, 1093–1109. [Google Scholar] [CrossRef] [PubMed]
  128. Sävilammi, T.; Primmer, C.R.; Varadharajan, S.; Guyomard, R.; Guiguen, Y.; Sandve, S.R.; Vøllestad, L.A.; Papakostas, S.; Lien, S. The Chromosome-Level Genome Assembly of European Grayling Reveals Aspects of a Unique Genome Evolution Process Within Salmonids. G3 Genes Genomes Genet. 2019, 9, 1283–1294. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  129. Auvinet, J.; Graça, P.; Belkadi, L.; Petit, L.; Bonnivard, E.; Dettaï, A.; Detrich, W.H.; Ozouf-Costaz, C.; Higuet, D. Mobilization of Retrotransposons as a Cause of Chromosomal Diversification and Rapid Speciation: The Case for the Antarctic Teleost Genus Trematomus. BMC Genom. 2018, 19, 339. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  130. Wang, S.; Diaby, M.; Puzakov, M.; Ullah, N.; Wang, Y.; Danley, P.; Chen, C.; Wang, X.; Gao, B.; Song, C. Divergent Evolution Profiles of DD37D and DD39D Families of Tc1/Mariner Transposons in Eukaryotes. Mol. Phylogenet. Evol. 2021, 161, 107143. [Google Scholar] [CrossRef] [PubMed]
  131. Flynn, J.M.; Hubley, R.; Goubert, C.; Rosen, J.; Clark, A.G.; Feschotte, C.; Smit, A.F. RepeatModeler2 for Automated Genomic Discovery of Transposable Element Families. Proc. Natl. Acad. Sci. USA 2020, 117, 9451–9457. [Google Scholar] [CrossRef] [PubMed]
  132. Flaxman, S.M.; Wacholder, A.C.; Feder, J.L.; Nosil, P. Theoretical Models of the Influence of Genomic Architecture on the Dynamics of Speciation. Mol. Ecol. 2014, 23, 4074–4088. [Google Scholar] [CrossRef] [PubMed]
  133. Dapper, A.L.; Payseur, B.A. Connecting Theory and Data to Understand Recombination Rate Evolution. Philos. Trans. R. Soc. B Biol. Sci. 2017, 372, 20160469. [Google Scholar] [CrossRef] [Green Version]
  134. Ortiz-Barrientos, D.; Engelstädter, J.; Rieseberg, L.H. Recombination Rate Evolution and the Origin of Species. Trends Ecol. Evol. 2016, 31, 226–236. [Google Scholar] [CrossRef] [Green Version]
  135. Fortna, A.; Kim, Y.; MacLaren, E.; Marshall, K.; Hahn, G.; Meltesen, L.; Brenton, M.; Hink, R.; Burgers, S.; Hernandez-Boussard, T.; et al. Lineage-Specific Gene Duplication and Loss in Human and Great Ape Evolution. PLoS Biol. 2004, 2, e207. [Google Scholar] [CrossRef] [Green Version]
  136. Varadharajan, S.; Sandve, S.R.; Gillard, G.B.; Tørresen, O.K.; Mulugeta, T.D.; Hvidsten, T.R.; Lien, S.; Asbjørn Vøllestad, L.; Jentoft, S.; Nederbragt, A.J.; et al. The Grayling Genome Reveals Selection on Gene Expression Regulation after Whole-Genome Duplication. Genome Biol. Evol. 2018, 10, 2785–2800. [Google Scholar] [CrossRef] [Green Version]
  137. Collens, A.B.; Katz, L.A. Opinion: Genetic Conflict With Mobile Elements Drives Eukaryotic Genome Evolution, and Perhaps Also Eukaryogenesis. J. Hered. 2021, 112, 140–144. [Google Scholar] [CrossRef] [PubMed]
  138. Wellenreuther, M.; Bernatchez, L. Eco-Evolutionary Genomics of Chromosomal Inversions. Trends Ecol. Evol. 2018, 33, 427–440. [Google Scholar] [CrossRef] [PubMed]
  139. Nelson, T.C.; Cresko, W.A. Ancient Genomic Variation Underlies Repeated Ecological Adaptation in Young Stickleback Populations. Evol. Lett 2018, 2, 9–21. [Google Scholar] [CrossRef] [PubMed]
  140. Kirubakaran, T.G.; Grove, H.; Kent, M.P.; Sandve, S.R.; Baranski, M.; Nome, T.; de Rosa, M.C.; Righino, B.; Johansen, T.; Otterå, H.; et al. Two Adjacent Inversions Maintain Genomic Differentiation between Migratory and Stationary Ecotypes of Atlantic Cod. Mol. Ecol. 2016, 25, 2130–2143. [Google Scholar] [CrossRef] [PubMed]
  141. Breistein, B.; Dahle, G.; Johansen, T.; Besnier, F.; Quintela, M.; Jorde, P.E.; Knutsen, H.; Westgaard, J.; Nedreaas, K.; Farestveit, E.; et al. Geographic Variation in Gene Flow from a Genetically Distinct Migratory Ecotype Drives Population Genetic Structure of Coastal Atlantic Cod (Gadus morhua L.). Evol. Appl. 2022, 15, 1162–1176. [Google Scholar] [CrossRef]
  142. Feder, J.L.; Nosil, P.; Flaxman, S.M. Assessing When Chromosomal Rearrangements Affect the Dynamics of Speciation: Implications from Computer Simulations. Front. Genet. 2014, 5, 295. [Google Scholar] [CrossRef] [Green Version]
  143. Garagna, S.; Page, J.; Fernandez-Donoso, R.; Zuccotti, M.; Searle, J.B. The Robertsonian Phenomenon in the House Mouse: Mutation, Meiosis and Speciation. Chromosoma 2014, 123, 529–544. [Google Scholar] [CrossRef]
  144. Kanjuh, T.; Marić, A.; Piria, M.; Špelić, I.; Maguire, I.; Simonović, P. Diversity of Brown Trout, Salmo Trutta (Actinopterygii: Salmoniformes: Salmonidae), in the Danube River Basin of Croatia Revealed by Mitochondrial DNA. Acta Ichthyol. Piscat. 2020, 50, 291–300. [Google Scholar] [CrossRef]
  145. Kalayci, G.; Ozturk, R.C.; Capkin, E.; Altinok, I. Genetic and Molecular Evidence That Brown Trout Salmo trutta Belonging to the Danubian Lineage Are a Single Biological Species. J. Fish Biol. 2018, 93, 792–804. [Google Scholar] [CrossRef]
  146. Ellingson, R.A.; Krug, P.J. Reduced Genetic Diversity and Increased Reproductive Isolation Follow Population-Level Loss of Larval Dispersal in a Marine Gastropod. Evolution 2016, 70, 18–37. [Google Scholar] [CrossRef]
  147. van Doren, B.M.; Campagna, L.; Helm, B.; Illera, J.C.; Lovette, I.J.; Liedvogel, M. Correlated Patterns of Genetic Diversity and Differentiation across an Avian Family. Mol. Ecol. 2017, 26, 3982–3997. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  148. Vendramin, N.; Cuenca, A.; Sørensen, J.; Alencar, A.L.F.; Christiansen, D.H.; Jacobsen, J.A.; Axen, C.; Lieffrig, F.; Ruane, N.M.; Martin, P.; et al. Presence and Genetic Variability of Piscine orthoreovirus Genotype 1 (PRV-1) in Wild Salmonids in Northern Europe and North Atlantic Ocean. J. Fish Dis. 2019, 42, 1107–1118. [Google Scholar] [CrossRef]
  149. Simonović, P. Risks to Stocks of Native Trout of the Genus Salmo (Actinopterygii: Salmoniformes: Salmonidae) of Serbia and Management for Their Recovery. Acta Ichthyol. Piscat. 2015, 45, 161–173. [Google Scholar] [CrossRef]
  150. Lehoczky, I.; Magyary, I.; Hancz, C.; Weiss, S. Preliminary Studies on the Genetic Variability of Six Hungarian Common Carp Strains Using Microsatellite DNA Markers. Hydrobiologia 2005, 533, 223–228. [Google Scholar] [CrossRef]
  151. Kohout, J.; Šedivá, A.; Apostolou, A.; Stefanov, T.; Marić, S.; Gaffaroğlu, M.; Šlechta, V. Genetic Diversity and Phylogenetic Origin of Brown Trout Salmo Trutta Populations in Eastern Balkans. Biologia 2013, 68, 1229–1237. [Google Scholar] [CrossRef]
  152. Tosic, A.; Skraba, D.; Nikolic, V.; Mrdak, D.; Simonovic, P. New Mitochondrial DNA Haplotype of Brown Trout Salmo Trutta L. from Crni Timok Drainage Area in Serbia. Turk. J. Fish. Aquat Sci. 2013, 14, 37–42. [Google Scholar] [CrossRef]
  153. Siah, A.; Breyta, R.B.; Warheit, K.I.; Gagne, N.; Purcell, M.K.; Morrison, D.; Powell, J.F.F.; Johnson, S.C. Genomes Reveal Genetic Diversity of Piscine Orthoreovirus in Farmed and Free-Ranging Salmonids from Canada and USA. Virus Evol. 2020, 6, veaa054. [Google Scholar] [CrossRef] [PubMed]
  154. Simonović, P.; Tošić, A.; Škraba Jurlina, D.; Nikolić, V.; Piria, M.; Tomljanović, T.; Šprem, N.; Mrdak, D.; Milošević, D.; Bećiraj, A.; et al. Diversity of Brown Trout Salmo Cf. Trutta in the River Danube Basin of Western Balkans as Assessed from the Structure of Their Mitochondrial Control Region Haplotypes. J. Ichthyol. 2017, 57, 603–616. [Google Scholar] [CrossRef]
  155. Sanz, N. Phylogeographic History of Brown Trout. In Brown Trout; John Wiley & Sons, Ltd.: Chichester, UK, 2017; pp. 15–63. [Google Scholar]
  156. Piria, M.; Špelić, I.; Rezić, A.; Šprem, N. Morfološke Značajke i Kondicija Potočne Pastrve Salmo Trutta Žumberačkih i Samoborskih Potoka. J. Cent. Eur. Agric. 2020, 21, 231–245. [Google Scholar] [CrossRef]
  157. Simonović, P.; Marić, S.; Nikolić, V. Trout Salmo Spp. Complex in Serbia and Adjacent Regions of the Western Balkans: Reconstruction of Evolutionary History from External Morphology. J. Fish Biol. 2007, 70, 359–380. [Google Scholar] [CrossRef]
  158. Splendiani, A.; Ruggeri, P.; Giovannotti, M.; Pesaresi, S.; Occhipinti, G.; Fioravanti, T.; Lorenzoni, M.; Nisi Cerioni, P.; Caputo Barucchi, V. Alien Brown Trout Invasion of the Italian Peninsula: The Role of Geological, Climate and Anthropogenic Factors. Biol. Invasions 2016, 18, 2029–2044. [Google Scholar] [CrossRef]
  159. Splendiani, A.; Giovannotti, M.; Righi, T.; Fioravanti, T.; Cerioni, P.N.; Lorenzoni, M.; Carosi, A.; la Porta, G.; Barucchi, V.C. Introgression despite Protection: The Case of Native Brown Trout in Natura 2000 Network in Italy. Conserv. Genet. 2019, 20, 343–356. [Google Scholar] [CrossRef]
  160. Tougard, C.; Justy, F.; Guinand, B.; Douzery, E.J.P.; Berrebi, P. Salmo Macrostigma (Teleostei, Salmonidae): Nothing More than a Brown Trout (S. trutta) Lineage? J. Fish Biol. 2018, 93, 302–310. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  161. Caputo, V.; Giovannotti, M.; Nisi Cerioni, P.; Caniglia, M.L.; Splendiani, A. Genetic Diversity of Brown Trout in Central Italy. J. Fish Biol. 2004, 65, 403–418. [Google Scholar] [CrossRef]
  162. Fabiani, A.; Gratton, P.; Zappes, I.A.; Seminara, M.; D’Orsi, A.; Sbordoni, V.; Allegrucci, G. Investigating the Genetic Structure of Trout from the Garden of Ninfa (Central Italy): Suggestions for Conservation and Management. Fish. Manag. Ecol. 2018, 25, 1–11. [Google Scholar] [CrossRef]
  163. Zaccara, S.; Trasforini, S.; Antognazza, C.M.; Puzzi, C.; Britton, J.R.; Crosa, G. Morphological and Genetic Characterization of Sardinian Trout Salmo Cettii Rafinesque, 1810 and Their Conservation Implications. Hydrobiologia 2015, 760, 205–223. [Google Scholar] [CrossRef]
  164. Berrebi, P.; Caputo Barucchi, V.; Splendiani, A.; Muracciole, S.; Sabatini, A.; Palmas, F.; Tougard, C.; Arculeo, M.; Marić, S. Brown Trout (Salmo trutta L.) High Genetic Diversity around the Tyrrhenian Sea as Revealed by Nuclear and Mitochondrial Markers. Hydrobiologia 2019, 826, 209–231. [Google Scholar] [CrossRef]
  165. Mo, T.A.; Gahr, A.; Hansen, H.; Hoel, E.; Oaland, Ø.; Poppe, T.T. Presence of Anisakis Simplex (Rudolphi, 1809 Det. Krabbe, 1878) and Hysterothylacium Aduncum (Rudolphi, 1802) (Nematoda; Anisakidae) in Runts of Farmed Atlantic Salmon, Salmo salar L. J. Fish. Dis. 2014, 37, 135–140. [Google Scholar] [CrossRef]
  166. Cortey, M.; Garcia-Marin, J.-L. Evidence for Phylogeographically Informative Sequence Variation in the Mitochondrial Control Region of Atlantic Brown Trout. J. Fish Biol. 2002, 60, 1058–1063. [Google Scholar] [CrossRef]
  167. Rossi, A.R.; Petrosino, G.; Milana, V.; Martinoli, M.; Rakaj, A.; Tancioni, L. Genetic Identification of Native Populations of Mediterranean Brown Trout Salmo trutta L. Complex (Osteichthyes: Salmonidae) in Central Italy. Eur. Zool. J. 2019, 86, 424–431. [Google Scholar] [CrossRef] [Green Version]
  168. Gordeeva, N.V.; Alekseyev, S.S.; Kirillov, A.F.; Vokin, A.I.; Samusenok, I.V. Distribution, Composition and Relationships of Phylogenetic Groups of Arctic Charr Salvelinus alpinus (L.) (Salmoniformes, Salmonidae) in the European Part of Russia and in Siberia as Revealed by the Analysis of Nucleotide Sequences of Mitochondrial DNA. J. Ichthyol. 2018, 58, 808–818. [Google Scholar] [CrossRef]
  169. Taylor, E.B.; Lowery, E.; Lilliestråle, A.; Elz, A.; Quinn, T.P. Genetic Analysis of Sympatric Char Populations in Western Alaska: Arctic Char (Salvelinus alpinus) and Dolly Varden (Salvelinus malma) Are Not Two Sides of the Same Coin. J. Evol. Biol. 2008, 21, 1609–1625. [Google Scholar] [CrossRef]
  170. Moore, J.-S.; Bajno, R.; Reist, J.D.; Taylor, E.B. Post-Glacial Recolonization of the North American Arctic by Arctic Char (Salvelinus alpinus): Genetic Evidence of Multiple Northern Refugia and Hybridization between Glacial Lineages. J. Biogeogr. 2015, 42, 2089–2100. [Google Scholar] [CrossRef]
  171. Alekseev, V.R.; Abramson, N.I.; Sukhikh, N.M. Introduction of Sibling Species to the Ecosystem of the Baltic Sea. Dokl. Biol. Sci. 2009, 429, 544–547. [Google Scholar] [CrossRef]
  172. Senchukova, A.L.; Pavlov, S.D.; Esin, E.V.; Markevich, G.N.; Mugue, N.S. Charrs of the Genus Salvelinus from Nachikinskoe Lake (Kamchatka Peninsula) and Their Position in the Phylogenetic System of the S. alpinus-S. malma Complex. J. Ichthyol 2015, 55, 97–104. [Google Scholar] [CrossRef]
  173. Osinov, A.G.; Volkov, A.A.; Alekseyev, S.S.; Sergeev, A.A.; Oficerov, M.V.; Kirillov, A.F. On the Origin and Phylogenetic Position of Arctic Charr (Salvelinus Alpinus Complex, Salmonidae) from Lake Cherechen’ (Middle Kolyma River Basin): Controversial Genetic Data. Polar Biol. 2017, 40, 777–786. [Google Scholar] [CrossRef]
  174. Esin, E.V.; Bocharova, E.S.; Mugue, N.S.; Markevich, G.N. Occurrence of Sympatric Charr Groups, Salvelinus, Salmonidae, in the Lakes of Kamchatka: A Legacy of the Last Glaciations. J. Fish Biol. 2017, 91, 628–644. [Google Scholar] [CrossRef] [PubMed]
  175. Borovikova, E.A.; Alekseeva, Y.I.; Schreider, M.J.; Artamonova, V.S.; Makhrov, A.A. Morphology and Genetics of the Ciscoes (Actinopterygii: Salmoniformes: Salmonidae: Coregoninae: Coregonus) from the Solovetsky Archipelago (White Sea) as a Key to Determination of the Taxonomic Position of Ciscoes in Northeastern Europe. Acta Ichthyol. Piscat. 2013, 43, 183–194. [Google Scholar] [CrossRef] [Green Version]
  176. Naseka, A.M.; Bogutskaya, N.G. Contribution to Taxonomy and Nomenclature of Freshwater Fishes of the Amur Drainage Area and the Far East (Pisces, Osteichthyes). Zoosyst. Ross. 2004, 12, 279–290. [Google Scholar] [CrossRef]
  177. Kawamura, K. "Handbook of European Freshwater Fishes" by M. Kottelat and J. Freyhof (2007). Ichthyol. Res. 2008, 55, 99. [Google Scholar] [CrossRef]
  178. Etheridge, E.C.; Adams, C.E.; Bean, C.W.; Durie, N.C.; Gowans, A.R.D.; Harrod, C.; Lyle, A.A.; Maitland, P.S.; Winfield, I.J. Are Phenotypic Traits Useful for Differentiating among a Priori Coregonus Taxa? J. Fish Biol. 2012, 80, 387–407. [Google Scholar] [CrossRef] [PubMed]
  179. Bolotov, I.N.; Makhrov, A.A.; Bespalaya, Y.V.; Vikhrev, I.V.; Aksenova, O.V.; Aspholm, P.E.; Gofarov, M.Y.; Ostrovskii, A.N.; Popov, I.Y.; Pal’tser, I.S.; et al. Results of Testing the Comparatory Method: The Curvature of the Shell Valve Frontal Section Is Inappropriate as a Systematic Character for the Freshwater Pearl Mussel of the Genus Margaritifera. Biol. Bull. 2013, 40, 221–231. [Google Scholar] [CrossRef]
  180. Artamonova, V.S.; Kucheryavyy, A.V.; Makhrov, A.A. Nucleotide Sequence Diversity of the Mitochondrial Cytochrome Oxidase Subunit I (COI) Gene of the Arctic Lamprey (Lethenteron camtschaticum) in the Eurasian Part of the Range. Hydrobiologia 2015, 757, 197–208. [Google Scholar] [CrossRef]
  181. Levin, B.A.; Simonov, E.P.; Ermakov, O.A.; Levina, M.A.; Interesova, E.A.; Kovalchuk, O.M.; Malinina, Y.A.; Mamilov, N.S.; Mustafayev, N.J.; Pilin, D.V.; et al. Phylogeny and Phylogeography of the Roaches, Genus Rutilus (Cyprinidae), at the Eastern Part of Its Range as Inferred from MtDNA Analysis. Hydrobiologia 2017, 788, 33–46. [Google Scholar] [CrossRef]
  182. Borovikova, E.A.; Artamonova, V.S. Vendace (Coregonus albula) and Least Cisco (Coregonus sardinella) Are a Single Species: Evidence from Revised Data on Mitochondrial and Nuclear DNA Polymorphism. Hydrobiologia 2021, 848, 4241–4262. [Google Scholar] [CrossRef]
  183. Oleinik, A.G.; Kukhlevsky, A.D.; Skurikhina, L.A. Complete Mitochondrial Genomes of the Charr Salvelinus alpinus erythrinus (Salmoniformes: Salmonidae) from Arctic Canada. Mitochondrial DNA Part B 2022, 7, 231–233. [Google Scholar] [CrossRef]
  184. Oleinik, A.G.; Skurikhina, L.A.; Kukhlevsky, A.D.; Semenchenko, A.A. Complete Mitochondrial Genomes of the Arctic Charr Salvelinus alpinus alpinus Linnaeus (Salmoniformes: Salmonidae). Mitochondrial DNA Part B 2020, 5, 2895–2897. [Google Scholar] [CrossRef]
  185. Oleinik, A.G.; Skurikhina, L.A.; Kukhlevsky, A.D.; Semenchenko, A.A. Complete Mitochondrial Genome and Phylogenetic Position of the Taranetz Charr Salvelinus taranetzi Kaganovsky, 1955 (Salmoniformes: Salmonidae). Mitochondrial DNA Part B 2019, 4, 2491–2492. [Google Scholar] [CrossRef] [Green Version]
  186. Oleinik, A.G.; Kukhlevsky, A.D.; Skurikhina, L.A. Complete Mitochondrial Genomes of Endemic Charr of the Genus Salvelinus from Lake Nachikinskoe (Kamchatka, Russia). Mitochondrial DNA Part B 2022, 7, 110–111. [Google Scholar] [CrossRef]
  187. Oleinik, A.G.; Skurikhina, L.A.; Kukhlevsky, A.D.; Semenchenko, A.A. The First Complete Mitochondrial Genomes of Two Species of Charr, Salvelinus Boganidae and Salvelinus Elgyticus, from Lake El’gygytgyn (Chukotka). Polar Biol. 2021, 44, 1209–1217. [Google Scholar] [CrossRef]
  188. Oleinik, A.G.; Skurikhina, L.A.; Kukhlevsky, A.D.; Semenchenko, A.A. Complete Mitochondrial Genome and Phylogenetic Position of the Levanidov’s Charr Salvelinus levanidovi Chereshnev, Skopetz et Gudkov, 1989 (Salmoniformes, Salmonidae). Mitochondrial DNA Part B 2020, 5, 2514–2515. [Google Scholar] [CrossRef] [PubMed]
  189. Denys, G.P.J.; Secci-Petretto, G.; Gomes-Dos-Santos, A. The Complete Mitochondrial Genome of Thymallus thymallus (Linnaeus, 1758) (Actinopterygii, Salmonidae) Obtained by Long Range PCRs and Double Multiplexing. Cybium 2020, 44, 91–94. [Google Scholar]
  190. Oleinik, A.G.; Skurikhina, L.A.; Kukhlevsky, A.D.; Semenchenko, A.A. First Report of Three Complete Mitochondrial Genomes of the Long-Finned Charr Salvethymus svetovidovi Chereshnev et Skopetz, 1990 (Salmoniformes: Salmonidae) with Phylogenetic Consideration. Mitochondrial DNA Part B 2019, 4, 2464–2466. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  191. Gong, L.; Liu, L.-Q.; Guo, B.-Y.; Ye, Y.-Y.; Lü, Z.-M. The Complete Mitochondrial Genome of Oncorhynchus masou formosanus (Salmoniformes: Salmonidae) and Phylogenetic Studies of Salmoninae. Conserv Genet. Resour. 2017, 9, 281–284. [Google Scholar] [CrossRef]
  192. Zhang, Y.; Meng, F.; Yang, L.; Wang, R.; Shi, G. The Complete Mitochondrial Genome of the Hybrid Fish Salvelinus fontinalis (♀) × Salvelinus malma sp. (♂). Mitochondrial DNA Part B 2019, 4, 1306–1307. [Google Scholar] [CrossRef] [Green Version]
  193. Yang, L.; Meng, F.; Wang, R.; Shi, G. Complete Mitochondrial Genome of the Salvelinus malma sp. (Salmoniformes, Salmonidae) with Phylogenetic Consideration. Mitochondrial DNA Part B 2017, 2, 889–890. [Google Scholar] [CrossRef]
  194. Hansen, T.; Fjelldal, P.G.; Lien, S.; Smith, M.; Corton, C.; Oliver, K.; Skelton, J.; Betteridge, E.; Doulcan, J.; Fedrigo, O.; et al. The Genome Sequence of the Brown Trout, Salmo Trutta Linnaeus 1758. Wellcome Open Res. 2021, 6, 108. [Google Scholar] [CrossRef]
  195. Bower, S.; Lennox, R.; Cooke, S. Is There a Role for Freshwater Protected Areas in the Conservation of Migratory Fish? Inland Waters 2015, 5, 1–6. [Google Scholar] [CrossRef]
  196. Meyer, A. Evolutionary Celebrities. Nature 2001, 410, 17–18. [Google Scholar] [CrossRef]
  197. Christoffels, A.; Koh, E.G.L.; Chia, J.; Brenner, S.; Aparicio, S.; Venkatesh, B. Fugu Genome Analysis Provides Evidence for a Whole-Genome Duplication Early During the Evolution of Ray-Finned Fishes. Mol. Biol. Evol. 2004, 21, 1146–1151. [Google Scholar] [CrossRef]
  198. Meyer, A.; van de Peer, Y. From 2R to 3R: Evidence for a Fish-Specific Genome Duplication (FSGD). BioEssays 2005, 27, 937–945. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  199. Santini, F.; Harmon, L.J.; Carnevale, G.; Alfaro, M.E. Did Genome Duplication Drive the Origin of Teleosts? A Comparative Study of Diversification in Ray-Finned Fishes. BMC Evol. Biol. 2009, 9, 194. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  200. Vandepoele, K.; de Vos, W.; Taylor, J.S.; Meyer, A.; van de Peer, Y. Major Events in the Genome Evolution of Vertebrates: Paranome Age and Size Differ Considerably between Ray-Finned Fishes and Land Vertebrates. Proc. Natl. Acad. Sci. USA 2004, 101, 1638–1643. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  201. de Boer, J.G.; Yazawa, R.; Davidson, W.S.; Koop, B.F. Bursts and Horizontal Evolution of DNA Transposons in the Speciation of Pseudotetraploid Salmonids. BMC Genom. 2007, 8, 422. [Google Scholar] [CrossRef] [Green Version]
  202. Salisbury, S.J.; Ruzzante, D.E. Genetic Causes and Consequences of Sympatric Morph Divergence in Salmonidae: A Search for Mechanisms. Annu. Rev. Anim Biosci. 2022, 10, 81–106. [Google Scholar] [CrossRef]
  203. Dodson, J.J.; Aubin-Horth, N.; Thériault, V.; Páez, D.J. The Evolutionary Ecology of Alternative Migratory Tactics in Salmonid Fishes. Biol. Rev. 2013, 88, 602–625. [Google Scholar] [CrossRef]
  204. Okamura, B.; Hartikainen, H.; Schmidt-Posthaus, H.; Wahli, T. Life Cycle Complexity, Environmental Change and the Emerging Status of Salmonid Proliferative Kidney Disease. Freshw. Biol. 2011, 56, 735–753. [Google Scholar] [CrossRef]
  205. Morrissey, M.B.; Ferguson, M.M. Individual Variation in Movement throughout the Life Cycle of a Stream-Dwelling Salmonid Fish. Mol. Ecol. 2011, 20, 235–248. [Google Scholar] [CrossRef]
  206. Thorpe, J.E. Salmonid Life-History Evolution as a Constraint on Marine Stock Enhancement. Bull. Mar. Sci. 1998, 62, 465–475. [Google Scholar]
  207. Castillo, A.G.F.; Ayllon, F.; Moran, P.; Izquierdo, J.I.; Martinez, J.L.; Beall, E.; Garcia-Vazquez, E. Interspecific Hybridization and Introgression Are Associated with Stock Transfers in Salmonids. Aquaculture 2008, 278, 31–36. [Google Scholar] [CrossRef]
  208. Steinke, D.; Salzburger, W.; Meyer, A. Novel Relationships Among Ten Fish Model Species Revealed Based on a Phylogenomic Analysis Using ESTs. J. Mol. Evol. 2006, 62, 772–784. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  209. Pleyte, K.A.; Duncan, S.D.; Phillips, R.B. Evolutionary Relationships of the Salmonid Fish Genus Salvelinus Inferred from DNA Sequences of the First Internal Transcribed Spacer (ITS 1) of Ribosomal DNA. Mol. Phylogenet. Evol. 1992, 1, 223–230. [Google Scholar] [CrossRef] [PubMed]
  210. Levin, D.B.; Dann, S.G.; Ted Allison, W.; Taylor, J.S.; Hawryshyn, C.W. Salmonid Opsin Sequences Undergo Positive Selection and Indicate an Alternate Evolutionary Relationship in Oncorhynchus. J. Mol. Evol. 2004, 58, 400–412. [Google Scholar] [CrossRef] [PubMed]
  211. McDowall, R.M. The Evolution of Diadromy in Fishes (Revisited) and Its Place in Phylogenetic Analysis. Rev. Fish Biol. Fish. 1997, 7, 443–462. [Google Scholar] [CrossRef]
  212. Behnke, R.J. The Application of Cytogenetic and Biochemical Systematics to Phylogenetic Problems in the Family Salmonidae. Trans. Am. Fish. Soc. 1970, 99, 237–248. [Google Scholar] [CrossRef]
  213. Norden, C.R. Comparative Osteology of Representative Salmonid Fishes, with Particular Reference to the Grayling (Thymallus arcticus) and Its Phylogeny. J. Fish. Res. Board Can. 1961, 18, 679–791. [Google Scholar] [CrossRef]
  214. Behnke, R.J. The Systematics of Salmonid Fishes of Recently Glaciated Lakes. J. Fish. Res. Board Can. 1972, 29, 639–671. [Google Scholar] [CrossRef]
  215. McCormick, S.D. Ontogeny and Evolution of Salinity Tolerance in Anadromous Salmonids: Hormones and Heterochrony. Estuaries 1994, 17, 26. [Google Scholar] [CrossRef]
  216. Koskinen, M.T.; Haugen, T.O.; Primmer, C.R. Contemporary Fisherian Life-History Evolution in Small Salmonid Populations. Nature 2002, 419, 826–830. [Google Scholar] [CrossRef]
  217. Crespi, B.J.; Teo, R. Comparative phylogenetic analysis of the evolution of semelparity and life history in salmonid fishes. Evolution 2002, 56, 1008–1020. [Google Scholar] [CrossRef]
  218. McPhail, J.D. The Origin and Speciation of Oncorhynchus Revisited. In Pacific Salmon & Their Ecosystems; Springer: Boston, MA, USA, 1997; pp. 29–38. [Google Scholar]
  219. Stearley, R.F.; Smith, G.R. Phylogeny of the Pacific Trouts and Salmons (Oncorhynchus) and Genera of the Family Salmonidae. Trans. Am. Fish. Soc. 1993, 122, 1–33. [Google Scholar] [CrossRef]
  220. Phillips, R.B.; Oakley, T.H. Phylogenetic Relationships among the Salmoninae Based on Nuclear and Mitochondrial DNA Sequences. In Molecular Systematics of Fishes; Academic Press: London, UK, 1997; pp. 145–162. [Google Scholar]
  221. Domanico, M.J.; Phillips, R.B. Phylogenetic Analysis of Pacific Salmon (Genus Oncorhynchus) Based on Mitochondrial DNA Sequence Data. Mol. Phylogenet. Evol. 1995, 4, 366–371. [Google Scholar] [CrossRef] [PubMed]
  222. Oakley, T.H.; Phillips, R.B. Phylogeny of Salmonine Fishes Based on Growth Hormone Introns: Atlantic (Salmo) and Pacific (Oncorhynchus) Salmon Are Not Sister Taxa. Mol. Phylogenet. Evol. 1999, 11, 381–393. [Google Scholar] [CrossRef]
  223. Domanico, M.J.; Phillips, R.B.; Oakley, T.H. Phylogenetic Analysis of Pacific Salmon (Genus Oncorhynchus) Using Nuclear and Mitochondrial DNA Sequences. Can. J. Fish. Aquat. Sci. 1997, 54, 1865–1872. [Google Scholar] [CrossRef]
  224. McDowall, R.M. The Origin of Salmonid Fishes: Marine, Freshwater... or Neither? Rev. Fish Biol. Fish. 2001, 11, 171–179. [Google Scholar] [CrossRef]
  225. Crespi, B.J.; Fulton, M.J. Molecular Systematics of Salmonidae: Combined Nuclear Data Yields a Robust Phylogeny. Mol. Phylogenet. Evol. 2004, 31, 658–679. [Google Scholar] [CrossRef]
  226. Rasmussen, J.A.; Villumsen, K.R.; Duchêne, D.A.; Puetz, L.C.; Delmont, T.O.; Sveier, H.; von Gersdorff Jørgensen, L.; Præbel, K.; Martin, M.D.; Bojesen, A.M.; et al. Genome-Resolved Metagenomics Suggests a Mutualistic Relationship between Mycoplasma and Salmonid Hosts. Commun. Biol. 2021, 4, 579. [Google Scholar] [CrossRef]
  227. Miya, M.; Friedman, M.; Satoh, T.P.; Takeshima, H.; Sado, T.; Iwasaki, W.; Yamanoue, Y.; Nakatani, M.; Mabuchi, K.; Inoue, J.G.; et al. Evolutionary Origin of the Scombridae (Tunas and Mackerels): Members of a Paleogene Adaptive Radiation with 14 Other Pelagic Fish Families. PLoS ONE 2013, 8, e73535. [Google Scholar]
  228. Wang, Y.; Xiong, F.; Song, Z. Molecular Phylogeny and Adaptive Mitochondrial DNA Evolution of Salmonids (Pisces: Salmonidae). Front. Genet. 2022, 13, 903240. [Google Scholar] [CrossRef]
  229. Gharbi, K.; Gautier, A.; Danzmann, R.G.; Gharbi, S.; Sakamoto, T.; Høyheim, B.; Taggart, J.B.; Cairney, M.; Powell, R.; Krieg, F.; et al. A Linkage Map for Brown Trout (Salmo trutta): Chromosome Homeologies and Comparative Genome Organization with Other Salmonid Fish. Genetics 2006, 172, 2405–2419. [Google Scholar] [CrossRef] [Green Version]
  230. Woram, R.A.; Gharbi, K.; Sakamoto, T.; Hoyheim, B.; Holm, L.-E.; Naish, K.; McGowan, C.; Ferguson, M.M.; Phillips, R.B.; Stein, J.; et al. Comparative Genome Analysis of the Primary Sex-Determining Locus in Salmonid Fishes. Genome Res. 2003, 13, 272–280. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  231. Presa, P.; Guyomard, R. Conservation of Microsatellites in Three Species of Salmonids. J. Fish Biol. 1996, 49, 1326–1329. [Google Scholar] [CrossRef]
  232. Yano, A.; Nicol, B.; Jouanno, E.; Quillet, E.; Fostier, A.; Guyomard, R.; Guiguen, Y. The Sexually Dimorphic on the Y-chromosome Gene (SdY) Is a Conserved Male-specific Y-chromosome Sequence in Many Salmonids. Evol. Appl. 2013, 6, 486–496. [Google Scholar] [CrossRef]
  233. Palti, Y.; Genet, C.; Gao, G.; Hu, Y.; You, F.M.; Boussaha, M.; Rexroad, C.E.; Luo, M.-C. A Second Generation Integrated Map of the Rainbow Trout (Oncorhynchus mykiss) Genome: Analysis of Conserved Synteny with Model Fish Genomes. Mar. Biotechnol. 2012, 14, 343–357. [Google Scholar] [CrossRef]
  234. Crête-Lafrenière, A.; Weir, L.K.; Bernatchez, L. Framing the Salmonidae Family Phylogenetic Portrait: A More Complete Picture from Increased Taxon Sampling. PLoS ONE 2012, 7, e46662. [Google Scholar] [CrossRef] [Green Version]
  235. Lecaudey, L.A.; Schliewen, U.K.; Osinov, A.G.; Taylor, E.B.; Bernatchez, L.; Weiss, S.J. Inferring Phylogenetic Structure, Hybridization and Divergence Times within Salmoninae (Teleostei: Salmonidae) Using RAD-Sequencing. Mol. Phylogenet. Evol. 2018, 124, 82–99. [Google Scholar] [CrossRef]
  236. Campbell, M.A.; López, J.A.; Sado, T.; Miya, M. Pike and Salmon as Sister Taxa: Detailed Intraclade Resolution and Divergence Time Estimation of Esociformes+Salmoniformes Based on Whole Mitochondrial Genome Sequences. Gene 2013, 530, 57–65. [Google Scholar] [CrossRef]
  237. Ma, B.; Jiang, H.; Sun, P.; Chen, J. Complete Mitochondrial Genome of Thymallus grubii (Salmonidae: Thymallinae). Mitochondrial DNA 2015, 26, 815–816. [Google Scholar] [CrossRef]
  238. Borovikova, E.A.; Artamonova, V.S. Morphological Specificities of Vendace (Salmoniformes: Salmonidae: Coregoninae: Coregonus Albula) Population in Lake Pleshcheyevo (the Volga River Basin): Relationships of Two Phylogenetic Lineages in a New Zone of Secondary Contact. Org. Divers. Evol. 2018, 18, 355–366. [Google Scholar] [CrossRef]
  239. Pigeon, D.; Chouinard, A.; Bernatchez, L. Multiple modes of speciation involved in the parallel evolution of sympatric morphotypes of lake whitefish (Coregonus clupeaformis, Salmonidae). Evolution 1997, 51, 196–205. [Google Scholar]
  240. Murata, S.; Takasaki, N.; Saitoh, M.; Okada, N. Determination of the Phylogenetic Relationships among Pacific Salmonids by Using Short Interspersed Elements (SINEs) as Temporal Landmarks of Evolution. Proc. Natl. Acad. Sci. USA 1993, 90, 6995–6999. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  241. Ramsden, S.D.; Brinkmann, H.; Hawryshyn, C.W.; Taylor, J.S. Mitogenomics and the Sister of Salmonidae. Trends Ecol. Evol. 2003, 18, 607–610. [Google Scholar] [CrossRef]
  242. Phillips, R.B.; Oakley, T.H.; Davis, E.L. Evidence Supporting the Paraphyly of Hucho (Salmonidae) Based on Ribosomal DNA Restriction Maps. J. Fish Biol. 1995, 47, 956–961. [Google Scholar] [CrossRef]
  243. Shedko, S.V.; Miroshnichenko, I.L.; Nemkova, G.A. Phylogeny of Salmonids (Salmoniformes: Salmonidae) and Its Molecular Dating: Analysis of Nuclear RAG1 Gene. Russ. J. Genet. 2012, 48, 575–579. [Google Scholar] [CrossRef]
  244. Osinov, A.G.; Lebedev, V.S. Genetc Divergence and Phylogeny of the Salmoninae Based on Allozyme Data. J. Fish Biol. 2000, 57, 354–381. [Google Scholar] [CrossRef]
  245. Koop, B.F.; von Schalburg, K.R.; Leong, J.; Walker, N.; Lieph, R.; Cooper, G.A.; Robb, A.; Beetz-Sargent, M.; Holt, R.A.; Moore, R.; et al. A Salmonid EST Genomic Study: Genes, Duplications, Phylogeny and Microarrays. BMC Genom. 2008, 9, 545. [Google Scholar] [CrossRef] [Green Version]
  246. Yasuike, M.; Jantzen, S.; Cooper, G.A.; Leder, E.; Davidson, W.S.; Koop, B.F. Grayling (Thymallinae) Phylogeny within Salmonids: Complete Mitochondrial DNA Sequences of Thymallus Arcticus and Thymallus Thymallus. J. Fish Biol. 2010, 76, 395–400. [Google Scholar] [CrossRef]
  247. Oleinik, A.G.; Skurikhina, L.A.; Brykov, V.A. Phylogeny of Charrs of the Genus Salvelinus Based on Mitochondrial DNA Data. Russ. J. Genet. 2015, 51, 55–68. [Google Scholar] [CrossRef]
  248. Makhrov, A. A Narrowing of the Phenotypic Diversity Range after Large Rearrangements of the Karyotype in Salmonidae: The Relationship between Saltational Genome Rearrangements and Gradual Adaptive Evolution. Genes 2017, 8, 297. [Google Scholar] [CrossRef]
  249. Brunner, P.C.; Douglas, M.R.; Osinov, A.; Wilson, C.C.; Bernatchez, L. Holarctic Phylogeography of Arctic Charr (Salvelinus alpinus L.) Inferred from Mitochondrial DNA Sequences. Evolution 2001, 55, 573–586. [Google Scholar] [CrossRef]
  250. Phillips, R.; Rab, P. Chromosome Evolution in the Salmonidae (Pisces): An Update. Biol. Rev. Camb Philos Soc. 2001, 76, S1464793100005613. [Google Scholar]
  251. Phillips, R.B. Evolution of the Sex Chromosomes in Salmonid Fishes. Cytogenet. Genome Res. 2013, 141, 177–185. [Google Scholar] [CrossRef] [PubMed]
  252. Guyomard, R.; Boussaha, M.; Krieg, F.; Hervet, C.; Quillet, E. A Synthetic Rainbow Trout Linkage Map Provides New Insights into the Salmonid Whole Genome Duplication and the Conservation of Synteny among Teleosts. BMC Genet. 2012, 13, 15. [Google Scholar] [CrossRef] [PubMed]
  253. Sutherland, B.J.G.; Gosselin, T.; Normandeau, E.; Lamothe, M.; Isabel, N.; Audet, C.; Bernatchez, L. Salmonid Chromosome Evolution as Revealed by a Novel Method for Comparing RADseq Linkage Maps. Genome Biol. Evol. 2016, 8, 3600–3617. [Google Scholar] [CrossRef]
  254. Walsh, B. Population-Genetic Models of the Fates of Duplicate Genes. Genetica 2003, 18, 279–294. [Google Scholar] [CrossRef]
  255. Townsend, J.P.; Leuenberger, C. Taxon Sampling and the Optimal Rates of Evolution for Phylogenetic Inference. Syst Biol. 2011, 60, 358–365. [Google Scholar] [CrossRef] [Green Version]
  256. Nabhan, A.R.; Sarkar, I.N. The Impact of Taxon Sampling on Phylogenetic Inference: A Review of Two Decades of Controversy. Brief. Bioinform. 2012, 13, 122–134. [Google Scholar] [CrossRef] [Green Version]
  257. Cohen, B.L. Rerooting the RDNA Gene Tree Reveals Phoronids to Be ‘Brachiopods without Shells’; Dangers of Wide Taxon Samples in Metazoan Phylogenetics (Phoronida; Brachiopoda). Zool. J. Linn. Soc. 2013, 167, 82–92. [Google Scholar] [CrossRef] [Green Version]
  258. Cao, Y.; Dietrich, C.H.; Zahniser, J.N.; Dmitriev, D.A. Dense Sampling of Taxa and Characters Improves Phylogenetic Resolution among Deltocephaline Leafhoppers (Hemiptera: Cicadellidae: Deltocephalinae). Syst. Entomol. 2022, 47, 430–444. [Google Scholar] [CrossRef]
  259. Yang, M.; Li, J.; Su, S.; Zhang, H.; Wang, Z.; Ding, W.; Li, L. The Mitochondrial Genomes of Tortricidae: Nucleotide Composition, Gene Variation and Phylogenetic Performance. BMC Genom. 2021, 22, 755. [Google Scholar] [CrossRef]
  260. Rosenberg, M.S. Heterogeneity of Nucleotide Frequencies Among Evolutionary Lineages and Phylogenetic Inference. Mol. Biol. Evol. 2003, 20, 610–621. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  261. Brocchieri, L. Phylogenetic Inferences from Molecular Sequences: Review and Critique. Theor. Popul. Biol. 2001, 59, 27–40. [Google Scholar] [CrossRef]
  262. Suzuki, Y.; Glazko, G.V.; Nei, M. Overcredibility of Molecular Phylogenies Obtained by Bayesian Phylogenetics. Proc. Natl. Acad. Sci. USA 2002, 99, 16138–16143. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  263. Rokas, A.; Carroll, S.B. More Genes or More Taxa? The Relative Contribution of Gene Number and Taxon Number to Phylogenetic Accuracy. Mol. Biol. Evol. 2005, 22, 1337–1344. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  264. Pollock, D.D.; Zwickl, D.J.; McGuire, J.A.; Hillis, D.M. Increased Taxon Sampling Is Advantageous for Phylogenetic Inference. Syst. Biol. 2002, 51, 664–671. [Google Scholar] [CrossRef] [Green Version]
  265. Horreo, J.L. Revisiting the Mitogenomic Phylogeny of Salmoninae: New Insights Thanks to Recent Sequencing Advances. PeerJ 2017, 5, e3828. [Google Scholar] [CrossRef] [Green Version]
  266. Alexandrou, M.A.; Swartz, B.A.; Matzke, N.J.; Oakley, T.H. Genome Duplication and Multiple Evolutionary Origins of Complex Migratory Behavior in Salmonidae. Mol. Phylogenet. Evol. 2013, 69, 514–523. [Google Scholar] [CrossRef]
  267. Shedko, S.V.; Miroshnichenko, I.L.; Nemkova, G.A. Phylogeny of Salmonids (Salmoniformes: Salmonidae) and Its Molecular Dating: Analysis of MtDNA Data. Russ. J. Genet. 2013, 49, 623–637. [Google Scholar] [CrossRef]
  268. Ma, B.; Jiang, H.; Sun, P.; Chen, J.; Li, L.; Zhang, X.; Yuan, L. Phylogeny and Dating of Divergences within the Genus Thymallus (Salmonidae: Thymallinae) Using Complete Mitochondrial Genomes. Mitochondrial DNA Part A 2016, 27, 3602–3611. [Google Scholar] [CrossRef]
  269. Sahoo, P.K.; Singh, L.; Sharma, L.; Kumar, R.; Singh, V.K.; Ali, S.; Singh, A.K.; Barat, A. The Complete Mitogenome of Brown Trout (Salmo trutta fario) and Its Phylogeny. Mitochondrial DNA Part A 2016, 27, 4563–4565. [Google Scholar] [CrossRef]
  270. Balakirev, E.S.; Parensky, V.A.; Kovalev, M.Y.; Ayala, F.J. Complete Mitochondrial Genome of the Stone Char Salvelinus kuznetzovi (Salmoniformes, Salmonidae). Mitochondrial DNA Part B 2016, 1, 287–288. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  271. Angers, B.; Bernatchez, L. Complex Evolution of a Salmonid Microsatellite Locus and Its Consequences in Inferring Allelic Divergence from Size Information. Mol. Biol. Evol. 1997, 14, 230–238. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  272. Shed’ko, S.V.; Ginatulina, L.K.; Parpura, I.Z.; Ermolenko, A.V. Evolutionary and Taxonomic Relationships among Far-Eastern Salmonid Fishes Inferred from Mitochondrial DNA Divergence. J. Fish Biol. 1996, 49, 815–829. [Google Scholar] [CrossRef]
  273. Murata, S.; Takasaki, N.; Saitoh, M.; Tachida, H.; Okada, N. Details of Retropositional Genome Dynamics That Provide a Rationale for a Generic Division: The Distinct Branching of All the Pacific Salmon and Trout (Oncorhynchus) From the Atlantic Salmon and Trout (Salmo). Genetics 1996, 142, 915–926. [Google Scholar] [CrossRef]
  274. Bernatchez, L.; Glémet, H.; Wilson, C.C.; Danzmann, R.G. Introgression and Fixation of Arctic Char (Salvelinus alpinus) Mitochondrial Genome in an Allopatric Population of Brook Trout (Salvelinus fontinalis). Can. J. Fish. Aquat. Sci. 1995, 52, 179–185. [Google Scholar] [CrossRef]
  275. Wilson, C.C.; Bernatchez, L. The Ghost of Hybrids Past: Fixation of Arctic Charr (Salvelinus alpinus) Mitochondrial DNA in an Introgressed Population of Lake Trout (S. namaycush). Mol. Ecol. 1998, 7, 127–132. [Google Scholar] [CrossRef] [Green Version]
  276. Taylor, E.B. Evolution in Mixed Company Evolutionary Inferences from Studies of Natural Hybridization in Salmonidae. In Evolution Illuminated: Salmon and Their Relatives; Hendry, A.P., Stearns, S.C., Eds.; Oxford University Press: New York, NY, USA, 2004; pp. 232–263. [Google Scholar]
  277. Ninua, L.; Tarkhnishvili, D.; Gvazava, E. Phylogeography and Taxonomic Status of Trout and Salmon from the Ponto-Caspian Drainages, with Inferences on European Brown Trout Evolution and Taxonomy. Ecol. Evol. 2018, 8, 2645–2658. [Google Scholar] [CrossRef] [Green Version]
  278. Froufe, E.; Alekseyev, S.; Knizhin, I.; Alexandrino, P.; Weiss, S. Comparative Phylogeography of Salmonid Fishes (Salmonidae) Reveals Late to Post-Pleistocene Exchange between Three Now-Disjunct River Basins in Siberia. Divers. Distrib. 2003, 9, 269–282. [Google Scholar] [CrossRef]
  279. Uiblein, F.; Jagsch, A.; Honsig-Erlenburg, W.; Weiss, S. Status, Habitat Use, and Vulnerability of the European Ggrayling in Austrian Waters. J. Fish Biol. 2001, 59, 223–247. [Google Scholar] [CrossRef]
  280. Koskinen, M.T.; Ranta, E.; Piironen, J.; Veselov, A.; Titov, S.; Haugen, T.O.; Nilsson, J.; Carlstein, M.; Primmer, C.R. Genetic Lineages and Postglacial Colonization of Grayling (Thymallus thymallus, Salmonidae) in Europe, as Revealed by Mitochondrial DNA Analyses. Mol. Ecol. 2000, 9, 1609–1624. [Google Scholar] [CrossRef]
  281. Marić, S.; Askeyev, I.V.; Askeyev, O.V.; Monakhov, S.P.; Bravničar, J.; Snoj, A. Phylogenetic and Population Genetic Analysis of Thymallus thymallus (Actinopterygii, Salmonidae) from the Middle Volga and Upper Ural Drainages. Hydrobiologia 2014, 740, 167–176. [Google Scholar] [CrossRef]
  282. Marić, S.; Kalamujić, B.; Snoj, A.; Razpet, A.; Lukić-Bilela, L.; Pojskić, N.; Bajec, S.S. Genetic Variation of European Grayling (Thymallus thymallus) Populations in the Western Balkans. Hydrobiologia 2012, 691, 225–237. [Google Scholar] [CrossRef]
  283. Dunham, J.; Baxter, C.; Fausch, K.; Fredenberg, W.; Kitano, S.; Koizumi, I.; Morita, K.; Nakamura, T.; Rieman, B.; Savvaitova, K.; et al. Evolution, Ecology, and Conservation of Dolly Varden, White Spotted Char, and Bull Trout. Fisheries 2008, 33, 537–550. [Google Scholar] [CrossRef]
  284. Hedges, S.B.; Marin, J.; Suleski, M.; Paymer, M.; Kumar, S. Tree of Life Reveals Clock-Like Speciation and Diversification. Mol. Biol. Evol. 2015, 32, 835–845. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  285. Viktorovskii, R.M. Review of the Book Salmonids of Northeast Russia by I.A. Chereshnev, V.V. Volobuev, A.V. Shestakov, and S.V. Frolov (Vladivostok: Dal’nauka, 2002, 496 pp.). Russ. J. Mar. Biol. 2003, 29, 403–406. [Google Scholar] [CrossRef]
  286. Osinov, A.G.; Senchukova, A.L.; Mugue, N.S.; Pavlov, S.D.; Chereshnev, I.A. Speciation and Genetic Divergence of Three Species of Charr from Ancient Lake El’gygytgyn (Chukotka) and Their Phylogenetic Relationships with Other Representatives of the Genus Salvelinus. Biol. J. Linn. Soc. 2015, 116, 63–85. [Google Scholar] [CrossRef] [Green Version]
  287. Osinov, A.G.; Lebedev, V.S. Salmonid Fishes (Salmonidae, Salmoniformes): The Systematic Position in the Superorder Protacanthopterygii, the Main Stages of Evolution, and Molecular Dating. J. Ichthyol. 2004, 44, 690–715. [Google Scholar]
  288. Osinov, A.G.; Volkov, A.A. On the Origin of Two Species of Charrs (Salvelinus, Salmonidae) in Lake Elgygytgyn: Data on the ATPase6–NADH4L Region of Mitochondrial DNA. J. Ichthyol. 2020, 60, 618–625. [Google Scholar] [CrossRef]
  289. Salmenkova, E.A.; Omel’chenko, V.T.; Radchenko, O.A.; Gordeeva, N.V.; Rubtsova, G.A.; Romanov, N.S. Genetic Divergence of Chars of the Genus Salvelinus from Kronotsky Lake (Kamchatka Peninsula). Russ. J. Genet. 2005, 41, 897–906. [Google Scholar] [CrossRef]
  290. Osinov, A.G.; Volkov, A.A.; Mugue, N.S. Charrs of the Genus Salvelinus (Salmonidae): Hybridization, Phylogeny and Evolution. Hydrobiologia 2021, 848, 705–726. [Google Scholar] [CrossRef]
  291. Gonen, S.; Bishop, S.C.; Houston, R.D. Exploring the Utility of Cross-Laboratory RAD-Sequencing Datasets for Phylogenetic Analysis. BMC Res. Notes 2015, 8, 299. [Google Scholar] [CrossRef] [PubMed]
  292. Esin, E.V.; Markevich, G.N.; Pichugin, M.Y. Juvenile Divergence in Adaptive Traits among Seven Sympatric Fish Ecomorphs Arises before Moving to Different Lacustrine Habitats. J. Evol. Biol. 2018, 31, 1018–1034. [Google Scholar] [CrossRef] [PubMed]
  293. Gilbey, J.; Soshnina, V.A.; Volkov, A.A.; Zelenina, D.A. Comparative Genetic Variability of Pink Salmon from Different Parts of Their Range: Native Pacific, Artificially Introduced White Sea and Naturally Invasive Atlantic Scottish Rivers. J. Fish Biol. 2022, 100, 549–560. [Google Scholar] [CrossRef]
  294. Pavlov, S.D.; Kuzishchin, K.V.; Gruzdeva, M.A.; Senchukova, A.L.; Pivovarov, E.A. Phenetic Diversity and Spatial Structure of Chars (Salvelinus) of the Kronotskaya Riverine-Lacustrine System (Eastern Kamchatka). J. Ichthyol. 2013, 53, 662–686. [Google Scholar] [CrossRef]
  295. Shedko, S.V.; Ginatulina, L.K.; Miroshnichenko, I.L.; Nemkova, G.A. Phylogeography of Mitochondrial DNA in South Asian Dolly Varden Char Salvelinus Curilus Pallas, 1814 (Salmoniformes, Salmonidae): Mediated Gene Introgression? Russ. J. Genet. 2007, 43, 165–176. [Google Scholar] [CrossRef]
  296. Balakirev, E.S.; Romanov, N.S.; Ayala, F.J. Complete Mitochondrial Genomes of the Northern (Salvelinus Malma) and Southern (Salvelinus curilus) Dolly Varden Chars (Salmoniformes, Salmonidae). Mitochondrial DNA 2016, 27, 1016–1017. [Google Scholar] [CrossRef] [Green Version]
  297. Skurikhina, L.A.; Oleinik, A.G.; Kukhlevsky, A.D.; Malyar, V.V. Intraspecific Polymorphism of MtDNA in Sakhalin Taimen Parahucho Perryi. Russ. J. Genet. 2013, 49, 924–936. [Google Scholar] [CrossRef]
  298. Oleinik, A.G.; Skurikhina, L.A.; Bondar, E.I.; Brykov, V.A. Phylogeography of Northern Dolly Varden Salvelinus Malma (Salmoniformes: Salmonidae) from Asia and North America: An Analysis Based on the Mitochondrial DNA Genealogy. J. Ichthyol. 2013, 53, 820–832. [Google Scholar] [CrossRef]
  299. Yamamoto, S.; Maekawa, K.; Morita, K.; Crane, P.A.; Oleinik, A.G. Phylogeography of the Salmonid Fish, Dolly Varden Salvelinus Malma: Multiple Glacial Refugia in the North Pacific Rim. Zool. Sci. 2014, 31, 660–670. [Google Scholar] [CrossRef]
  300. Oleinik, A.G.; Skurikhina, L.A.; Brykov, V.A. Population Genetic Structure of Northern Dolly Varden Char Salvelinus Malma Malma in Asia and North America. Russ. J. Genet. 2011, 47, 1456–1467. [Google Scholar] [CrossRef]
  301. Oleinik, A.G.; Skurikhina, L.A. Mitochondrial DNA Diversity and Relationships of Endemic Charrs of the Genus Salvelinus from Lake Kronotskoye (Kamchatka Peninsula). Hydrobiologia 2010, 650, 145–159. [Google Scholar] [CrossRef]
  302. Oleinik, A.G.; Skurikhina, L.A.; Brykov, V.A. Divergence of Salvelinus Species from Northeastern Asia Based on Mitochondrial DNA. Ecol. Freshw Fish. 2007, 16, 87–98. [Google Scholar] [CrossRef]
  303. Oleinik, A.G.; Skurikhina, L.A.; Brykov, V.A.; Crane, P.A.; Wenburg, J.K. Differentiation of Dolly Varden Char Salvelinus Malma from Asia and North America Inferred from PCR-RFLP Analysis of Mitochondrial DNA. Russ. J. Genet. 2005, 41, 501–508. [Google Scholar] [CrossRef]
  304. Oleinik, A.G.; Skurikhina, L.A.; Frolov, S.V.; Brykov, V.A.; Chereshnev, I.A. Differences Between Two Subspecies of Dolly Varden, Salvelinus Malma, Revealed by RFLP—PCR Analysis of Mitochondrial DNA. Environ. Biol. Fishes 2004, 69, 449–459. [Google Scholar] [CrossRef]
  305. Oleinik, A.G.; Skurikhina, L.A.; Brykov, V.A. Genetic Differentiation of Three Sympatric Charr Species from the Genus SalvelinusInferred from PCR–RFLP Analysis of Mitochondrial DNA. Russ. J. Genet. 2003, 39, 924–929. [Google Scholar] [CrossRef]
  306. Radchenko, O.A. Comparative Restriction Enzyme Analysis of Mitochondrial DNA in Several Populations of Lake Resident Chars of the Genus Salvelinus. Russ. J. Genet. 2003, 39, 87–90. [Google Scholar] [CrossRef]
  307. Osinov, V.A.; Gudehus, G. Plane Shear Waves and Loss of Stability in a Saturated Granular Body. Mech. Cohesive-Frict. Mater. 1996, 1, 25–44. [Google Scholar] [CrossRef]
  308. Pichugin, M.Y. Development of Skeleton Elements in Juvenile Dwarf and Small Sympatric Forms of Salvelinus Alpinus Complex from Lake Davatchan (Transbaikalia). J. Ichthyol. 2009, 49, 741–758. [Google Scholar] [CrossRef]
  309. Nevoux, M.; Finstad, B.; Davidsen, J.G.; Finlay, R.; Josset, Q.; Poole, R.; Höjesjö, J.; Aarestrup, K.; Persson, L.; Tolvanen, O.; et al. Environmental Influences on Life History Strategies in Partially Anadromous Brown Trout (Salmo trutta, Salmonidae). Fish Fish. 2019, 20, 1051–1082. [Google Scholar] [CrossRef]
  310. Phillips, R.B.; Pleyte, K.A.; Hartley, S.E. Stock-specific differences in the number and chromosome positions of the nucleolar organizer regions in arctic char (Salvelinus alpinus). Cytogenet. Genome Res. 1988, 48, 9–12. [Google Scholar] [CrossRef]
  311. Wilson, A.J.; Hutchings, J.A.; Ferguson, M.M. Dispersal in a Stream Dwelling Salmonid: Inferences from Tagging and Microsatellite Studies. Conserv. Genet. 2004, 5, 25–37. [Google Scholar] [CrossRef]
  312. Alekseyev, S.S.; Samusenok, V.P.; Matveev, A.N.; Pichugin, M.Y. Diversification, Sympatric Speciation, and Trophic Polymorphism of Arctic Charr, Salvelinus alpinus Complex, in Transbaikalia. In Ecology, Behaviour and Conservation of the Charrs, Genus Salvelinus; Springer: Dordrecht, The Netherlands, 2002; pp. 97–114. [Google Scholar]
  313. Oleinik, A.G.; Skurikhina, L.A.; Kukhlevsky, A.D. The Lactate Dehydrogenase Gene LDH-C1, a New Molecular Marker for Phylogenetic Analysis of Salmonid Fishes (Salmoniformes: Salmonidae). Russ. J. Mar. Biol. 2017, 43, 499–502. [Google Scholar] [CrossRef]
  314. Yamamoto, S.; Morita, K.; Sahashi, G.; Maekawa, K.; Oleinik, A.; Bondar, E.; Brykov, V. Introgressive Hybridization between Southern Asian Dolly Varden, Salvelinus curilus, and Northern Dolly Varden, S. malma malma, on Sakhalin Island. Russ. J. Genet. 2021, 57, 361–370. [Google Scholar] [CrossRef]
  315. Taylor, E.B. The Arctic Char (Salvelinus alpinus) “Complex” in North America Revisited. Hydrobiologia 2016, 783, 283–293. [Google Scholar] [CrossRef]
  316. Huson, D.H.; Bryant, D. Application of Phylogenetic Networks in Evolutionary Studies. Mol. Biol. Evol. 2006, 23, 254–267. [Google Scholar] [CrossRef]
  317. Miya, M.; Nishida, M. The Mitogenomic Contributions to Molecular Phylogenetics and Evolution of Fishes: A 15-Year Retrospect. Ichthyol. Res. 2015, 62, 29–71. [Google Scholar] [CrossRef] [Green Version]
  318. Oleinik, A.G.; Bondar, E.I.; Kukhlevsky, A.D.; Skurikhina, L.A.; Kovpak, N.E. Introgressive Hybridization between Two Phylogenetic Lineages of Charrs (Salvelinus: Salmonidae) in Northeastern Asia. J. Zool. Syst. Evol. Res. 2021, 59, 2119–2133. [Google Scholar] [CrossRef]
  319. Phillips, R.B.; Matsuoka, M.P.; Reed, K.M. Characterization of Charr Chromosomes Using Fluorescence in Situ Hybridization. Environ. Biol. Fishes 2002, 64, 223–228. [Google Scholar] [CrossRef]
  320. Østbye, K.; Amundsen, P.-A.; Bernatchez, L.; Klemetsen, A.; Knudsen, R.; Kristoffersen, R.; Naesje, T.F.; Hindar, K. Parallel Evolution of Ecomorphological Traits in the European Whitefish Coregonus lavaretus (L.) Species Complex during Postglacial Times. Mol. Ecol. 2006, 15, 3983–4001. [Google Scholar] [CrossRef]
  321. Landry, L.; Bernatchez, L. Role of Epibenthic Resource Opportunities in the Parallel Evolution of Lake Whitefish Species Pairs (Coregonus sp.). J. Evol. Biol. 2010, 23, 2602–2613. [Google Scholar] [CrossRef]
  322. Derome, N.; Duchesne, P.; Bernatchez, L. Parallelism in Gene Transcription among Sympatric Lake Whitefish (Coregonus Clupeaformis Mitchill) Ecotypes. Mol. Ecol. 2006, 15, 1239–1249. [Google Scholar] [CrossRef] [PubMed]
  323. Eshenroder, R.L.; Olds, C.M.; Kao, Y.-C.; Davis, C.L.; Kinney, D.N.; Muir, A.M. Status of Cisco (Coregonus Artedi) Ecomorphs in Lake Huron, 1917–2016, with Speculations about Phenotypic Plasticity in Shorthead Cisco. Adv. Limnol. 2021, 66, 383–402. [Google Scholar] [CrossRef]
  324. 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, 1491–1505. [Google Scholar] [PubMed]
  325. Bernatchez, L.; Vuorinen, J.A.; Bodaly, R.A.; Dodson, J.J. Genetic evidence for reproductive isolation and multiple origins of sympatric trophic ecotypes of whitefish (Coregonus). Evolution 1996, 50, 624–635. [Google Scholar]
  326. Kahilainen, K.K.; Østbye, K.; Harrod, C.; Shikano, T.; Malinen, T.; Merilä, J. Species Introduction Promotes Hybridization and Introgression in Coregonus: Is There Sign of Selection against Hybrids? Mol. Ecol. 2011, 20, 3838–3855. [Google Scholar] [CrossRef]
  327. Winkler, K.A.; Pamminger-Lahnsteiner, B.; Wanzenböck, J.; Weiss, S. Hybridization and Restricted Gene Flow between Native and Introduced Stocks of Alpine Whitefish (Coregonus sp.) across Multiple Environments. Mol. Ecol. 2011, 20, 456–472. [Google Scholar] [CrossRef] [Green Version]
  328. Harris, L.N.; Taylor, E.B. Genetic Population Structure of Broad Whitefish, Coregonus Nasus, from the Mackenzie River, Northwest Territories: Implications for Subsistence Fishery Management. Can. J. Fish. Aquat. Sci. 2010, 67, 905–918. [Google Scholar] [CrossRef] [Green Version]
  329. Bochkarev, N.A.; Zuykova, E.I.; Abramov, S.A.; Podorozhnyuk, E.V.; Politov, D.V. The Sympatric Whitefishes Coregonus Ussuriensis and C. chadary from the Amur River Basin: Morphology, Biology and Genetic Diversity. Fundam. Appl. Limnol. 2017, 189, 193–207. [Google Scholar]
  330. Skrodenytė-Arbačiauskienė, V.; Sruoga, A.; Butkauskas, D.; Skrupskelis, K. Phylogenetic Analysis of Intestinal Bacteria of Freshwater Salmon Salmo salar and Sea Trout Salmo trutta trutta and Diet. Fish. Sci. 2008, 74, 1307–1314. [Google Scholar] [CrossRef]
  331. Gaikalov, I.V.; Ilyina, O.v.; Kirilchik, S.V.; Sukhanova, L.V. Description of Three Microsatellite Loci of the Baikal Omul Coregonus Migratorius (Georgi). Russ. J. Genet. 2008, 44, 365–368. [Google Scholar] [CrossRef]
  332. Sendek, D.S.; Bochkarev, N.A.; Zuykova, E.I.; Politov, D.V.; Wanzenböck, J.; Himberg, M.; Titov, S.F. Signs of Introgression of Baikal Omul (Coregonus migratorius) or Arctic Cisco (C. autumnalis) into European Whitefish (C. lavaretus) in the Eastern Baltic Sea. Fundam. Appl. Limnol. 2017, 189, 209–225. [Google Scholar] [CrossRef]
  333. Weiss, S.J.; Gonçalves, D.V.; Secci-Petretto, G.; Englmaier, G.K.; Gomes-Dos-Santos, A.; Denys, G.P.J.; Persat, H.; Antonov, A.; Hahn, C.; Taylor, E.B.; et al. Global Systematic Diversity, Range Distributions, Conservation and Taxonomic Assessments of Graylings (Teleostei: Salmonidae; Thymallus spp.). Org. Divers. Evol. 2021, 21, 25–42. [Google Scholar] [CrossRef]
  334. Bernatchez, L.; Edge, T.A.; Dodson, J.J.; Qadri, S.U. Mitochondrial DNA and Isozyme Electrophoretic Analyses of the Endangered Acadian Whitefish, Coregonus huntsmani Scott, 1987. Can. J. Zool. 1991, 69, 311–316. [Google Scholar] [CrossRef]
  335. Bernatchez, L.; Colombani, F.; Dodson, J.J. Phylogenetic Relationships among the Subfamily Coregoninae as Revealed by Mitochondrial DNA Restriction Analysis. J. Fish Biol. 1991, 39, 283–290. [Google Scholar] [CrossRef]
  336. Vuorinen, J.A.; Bodaly, R.A.; Reist, J.D.; Luczynski, M. Phylogeny of Five Prosopium Species with Comparisons with Other Coregonine Fishes Based on Isozyme Electrophoresis*. J. Fish Biol. 1998, 53, 917–927. [Google Scholar] [CrossRef]
  337. Bochkarev, N.A.; Zuykova, E.I.; Politov, D.V. Taxonomic Status and Origin of Some Ecological Forms of Whitefish Coregonus lavaretus (L.) from Water Bodies of Siberia. Russ. J. Genet. 2017, 53, 875–884. [Google Scholar] [CrossRef]
  338. Lachance, H.; Ackiss, A.S.; Larson, W.A.; Vinson, M.R.; Stockwell, J.D. Genomics Reveals Identity, Phenology and Population Demographics of Larval Ciscoes (Coregonus artedi, C. hoyi, and C. kiyi) in the Apostle Islands, Lake Superior. J. Great Lakes Res. 2021, 47, 1849–1857. [Google Scholar]
  339. Overdyk, L.M.; Braid, H.E.; Naaum, A.M.; Crawford, S.S.; Hanner, R.H. Real-Time PCR Identification of Lake Whitefish Coregonus clupeaformis in the Laurentian Great Lakes. J. Fish Biol. 2016, 88, 1460–1474. [Google Scholar] [CrossRef]
  340. Turgeon, J.; Estoup, A.; Bernatchez, L. Species flock in the North American Great Lakes: Molecular ecology of lake Nipigon ciscoes (Teleostei: Coregonidae: Coregonus). Evolution 1999, 53, 1857–1871. [Google Scholar] [CrossRef]
  341. Sukhanova, L.V.; Smirnov, V.V.; Smirnova-Zalumi, N.S.; Belomestnykh, T.V.; Kirilchik, S.V. Molecular Phylogeography of Lake Baikal Coregonid Fishes. Adv. Limnol. 2012, 63, 261–283. [Google Scholar] [CrossRef]
  342. Teterina, V.; Sukhanova, L.; Smirnov, V.; Smirnova, N.; Kirilchik, S.; Sapozhnikova, Y.; Glizina, O.; Yakhnenko, V.; Tyagun, M.; Sidorova, T. Complete Mitochondrial Genomes of Baikal Endemic Coregonids: Omul and Lacustrine Whitefish (Salmonidae: Coregonus sp.). Mitochondrial DNA Part B 2020, 5, 414–416. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  343. Østbye, K.; Bernatchez, L.; Naesje, T.F.; Himberg, K.-J.M.; Hindar, K. Evolutionary History of the European Whitefish Coregonus lavaretus (L.) Species Complex as Inferred from MtDNA Phylogeography and Gill-Raker Numbers. Mol. Ecol. 2005, 14, 4371–4387. [Google Scholar] [CrossRef] [PubMed]
  344. Østbye, K.; Naesje, T.F.; Bernatchez, L.; Sandlund, O.T.; Hindar, K. Morphological Divergence and Origin of Sympatric Populations of European Whitefish (Coregonus lavaretus L.) in Lake Femund, Norway. J. Evol. Biol. 2005, 18, 683–702. [Google Scholar] [CrossRef]
  345. Borovikova, E.A.; Budin, Y.V. Morphological and Genetic Diversity of Two Forms of Muksun Coregonus Muksun (Salmonidae) of the Khatanga River Basin as a Key for Understanding the Phylogenetic Relationships between Muksun and Whitefish C. lavaretus. J. Ichthyol 2020, 60, 909–922. [Google Scholar] [CrossRef]
  346. Jankun, M.; Martinez, P.; Pardo, B.G.; Kirtiklis, L.; Rab, P.; Rabova, M.; Sanchez, L. Ribosomal Genes in Coregonid Fishes (Coregonus lavaretus, C. albula and C. peled) (Salmonidae): Single and Multiple Nucleolus Organizer Regions. Heredity 2001, 87, 672–679. [Google Scholar] [CrossRef] [PubMed]
  347. Borovikova, E.A.; Makhrov, A.A. Detection of the Mitochondrial DNA Haplotype Characteristic of the Least Cisco (Coregonus Sardinella, Valenciennes, 1848) in the Vendace (C. Albula, Linnaeus, 1758) Population of Vodlozero (the Baltic Sea Basin). Biol. Bull. 2009, 36, 80–83. [Google Scholar] [CrossRef]
  348. Christensen, H.T.; Rigét, F.; Backe, M.B.; Saha, A.; Johansen, T.; Hedeholm, R.B. Comparison of Three Methods for Identification of Redfish (Sebastes mentella and S. norvegicus) from the Greenland East Coast. Fish. Res. 2018, 201, 11–17. [Google Scholar] [CrossRef]
  349. Romanov, V.I.; Polyaeva, K.V. Morphological Features and Endoparasite Fauna of Some Whitefishes and Grayling Fishes from the Eastern Part of Lake Sobachye (the Putorana Plateau). J. Sib. Fed. Univ. Biol. 2019, 12, 410–429. [Google Scholar] [CrossRef]
  350. Artamonova, V.S.; Makhrov, A.A.; Karabanov, D.P.; Rolskiy, A.Y.; Bakay, Y.I.; Popov, V.I. Hybridization of Beaked Redfish (Sebastes mentella) with Small Redfish (Sebastes viviparus) and Diversification of Redfish (Actinopterygii: Scorpaeniformes) in the Irminger Sea. J. Nat. Hist 2013, 47, 1791–1801. [Google Scholar] [CrossRef]
  351. Schlei, O.L.; Crête-Lafrenière, A.; Whiteley, A.R.; Brown, R.J.; Olsen, J.B.; Bernatchez, L.; Wenburg, J.K. DNA Barcoding of Eight North American Coregonine Species. Mol. Ecol. Resour 2008, 8, 1212–1218. [Google Scholar] [CrossRef]
  352. Dulmaa, A.; Slynko, Y.V.; Gordon, N.Y.; Stolbunova, V.V.; Politov, D.V. Coregonus Peled (Gmelin) Transplanted into Ulaagchny Khar Lake (Western Mongolia) Showed No Evidence of Hybridization with Other Introduced Coregonus Species. Contemp. Probl. Ecol. 2016, 9, 172–176. [Google Scholar] [CrossRef]
  353. Balakirev, E.S.; Romanov, N.S.; Ayala, F.J. Complete Mitochondrial Genome of the Kamchatka Grayling Thymallus mertensii (Salmoniformes, Salmonidae). Mitochondrial DNA Part. A 2017, 28, 135–136. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  354. Liu, Y.-G.; Li, Y.-Y.; Meng, W.; Liu, L.-X. The Complete Mitochondrial DNA Sequence of Xinjiang Arctic Grayling Thymallus arcticus grubei. Mitochondrial DNA Part B 2016, 1, 724–725. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  355. Gratton, P.; Allegrucci, G.; Sbordoni, V.; Gandolfi, A. The Evolutionary Jigsaw Puzzle of the Surviving Trout (Salmo trutta L. Complex) Diversity in the Italian Region. A Multilocus Bayesian Approach. Mol. Phylogenet. Evol. 2014, 79, 292–304. [Google Scholar] [CrossRef] [PubMed]
  356. Ogden, R. Fisheries Forensics: The Use of DNA Tools for Improving Compliance, Traceability and Enforcement in the Fishing Industry. Fish Fish. 2008, 9, 462–472. [Google Scholar] [CrossRef]
  357. Martinsohn, J.T.; Raymond, P.; Knott, T.; Glover, K.A.; Nielsen, E.E.; Eriksen, L.B.; Ogden, R.; Casey, J.; Guillen, J. DNA-analysis to Monitor Fisheries and Aquaculture: Too Costly? Fish Fish. 2019, 20, 391–401. [Google Scholar] [CrossRef] [Green Version]
  358. Thompson, N.F.; Anderson, E.C.; Clemento, A.J.; Campbell, M.A.; Pearse, D.E.; Hearsey, J.W.; Kinziger, A.P.; Garza, J.C. A complex phenotype in salmon controlled by a simple change in migratory timing. Science 2020, 370, 609–613. [Google Scholar] [CrossRef]
  359. Papa, Y.; Oosting, T.; Valenza-Troubat, N.; Wellenreuther, M.; Ritchie, P.A. Genetic stock structure of New Zealand fish and the use of genomics in fisheries management: An overview and outlook. N. Z. J. Zool. 2021, 48, 1–31. [Google Scholar] [CrossRef]
  360. Benestan, L. Population Genomics Applied to Fishery Management and Conservation; Springer: Cham, Switzerland, 2019; pp. 399–421. [Google Scholar]
  361. Bradbury, I.R.; Wringe, B.F.; Watson, B.; Paterson, I.; Horne, J.; Beiko, R.; Lehnert, S.J.; Clément, M.; Anderson, E.C.; Jeffery, N.W.; et al. Genotyping-by-Sequencing of Genome-Wide Microsatellite Loci Reveals Fine-Scale Harvest Composition in a Coastal Atlantic Salmon Fishery. Evol. Appl. 2018, 11, 918–930. [Google Scholar] [CrossRef]
Publisher’s Note: MDPI stays neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Share and Cite

MDPI and ACS Style

Dysin, A.P.; Shcherbakov, Y.S.; Nikolaeva, O.A.; Terletskii, V.P.; Tyshchenko, V.I.; Dementieva, N.V. Salmonidae Genome: Features, Evolutionary and Phylogenetic Characteristics. Genes 2022, 13, 2221. https://doi.org/10.3390/genes13122221

AMA Style

Dysin AP, Shcherbakov YS, Nikolaeva OA, Terletskii VP, Tyshchenko VI, Dementieva NV. Salmonidae Genome: Features, Evolutionary and Phylogenetic Characteristics. Genes. 2022; 13(12):2221. https://doi.org/10.3390/genes13122221

Chicago/Turabian Style

Dysin, Artem P., Yuri S. Shcherbakov, Olga A. Nikolaeva, Valerii P. Terletskii, Valentina I. Tyshchenko, and Natalia V. Dementieva. 2022. "Salmonidae Genome: Features, Evolutionary and Phylogenetic Characteristics" Genes 13, no. 12: 2221. https://doi.org/10.3390/genes13122221

Note that from the first issue of 2016, this journal uses article numbers instead of page numbers. See further details here.

Article Metrics

Back to TopTop