Next Article in Journal
Genome–Transcriptome Transition Approaches to Characterize Anthocyanin Biosynthesis Pathway Genes in Blue, Black and Purple Wheat
Previous Article in Journal
Molecular Design-Based Breeding: A Kinship Index-Based Selection Method for Complex Traits in Small Livestock Populations
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Review

Microsatellites as Molecular Markers with Applications in Exploitation and Conservation of Aquatic Animal Populations

Institute of Oceanology, Polish Academy of Sciences, Powstańców Warszawy 55, 81-712 Sopot, Poland
Genes 2023, 14(4), 808; https://doi.org/10.3390/genes14040808
Submission received: 20 December 2022 / Revised: 28 February 2023 / Accepted: 17 March 2023 / Published: 27 March 2023
(This article belongs to the Section Population and Evolutionary Genetics and Genomics)

Abstract

:
A large number of species and taxa have been studied for genetic polymorphism. Microsatellites have been known as hypervariable neutral molecular markers with the highest resolution power in comparison with any other markers. However, the discovery of a new type of molecular marker—single nucleotide polymorphism (SNP) has put the existing applications of microsatellites to the test. To ensure good resolution power in studies of populations and individuals, a number of microsatellite loci from 14 to 20 was often used, which corresponds to about 200 independent alleles. Recently, these numbers have tended to be increased by the application of genomic sequencing of expressed sequence tags (ESTs) and the choice of the most informative loci for genotyping depends on the aims of research. Examples of successful applications of microsatellite molecular markers in aquaculture, fisheries, and conservation genetics in comparison to SNPs are summarized in this review. Microsatellites can be considered superior markers in such topics as kinship and parentage analysis in cultured and natural populations, the assessment of gynogenesis, androgenesis and ploidization. Microsatellites can be coupled with SNPs for mapping QTL. Microsatellites will continue to be used in research of genetic diversity in cultured stocks, and also in natural populations as an economically advantageous genotyping technique.

1. Introduction

The global production of aquatic animals for consumption was almost 178 million tonnes in 2020, including capture fisheries of 90.3 million, and aquaculture production of 87.5 million tonnes [1]. One of the possibilities for further increasing production that has been used in recent decades is the application of scientific research in the fields of biotechnology and genetics [2]. The development of genetics has enabled the discovery of molecular markers suitable for the study of a large number of individuals in natural and breeding populations. A few types of molecular markers used in research associated with aquaculture and fisheries have been described over the last few decades decades [3,4,5,6,7]. Their applications in uncovering the differentiation of natural populations for the purposes of improved fishery management, population genetic structures and local adaptations for conservation biology purposes, identification of selected lines in aquaculture, analysis of kinship in the stocks and identification of individuals, finding quantitative trait loci (QTL) for the purposes of aquaculture, identification of taxa, food products, and forensics have been reported in a large number of papers [8,9,10,11,12,13,14,15,16,17,18,19,20,21,22].
Proteins and allozymes have been continuously used in analyses of natural and hatchery populations since the 1960s and have provided a vast amount of information about their taxonomy, hybridization, genetic polymorphism and spatial differentiation in relation to aquatic environments (Table 1).
Allozymes were analyzed under assumptions of models based on neutrality, whereas more recently, they have been considered molecular markers operating under natural selection [74,75,76]. Differences between alleles of allozyme loci have been classified as amino acid variants caused by nucleotide substitutions, which often reveal less polymorphisms in comparison to DNA markers such as single nucleotide polymorphisms (SNPs) [77]. The resolution power in population genetic analyses usually confirms a lower diagnostic level of allozymes in finding differentiation compared with mtDNA and microsatellites [78]. Allozymes are rarely but still used successfully as markers for identification of taxa and populations of aquatic organisms, and improvement of hatchery stocks [65,79,80,81].
Mitochondrial DNA (mtDNA) as a molecular marker differs from nuclear DNA due to such properties as maternal inheritance in most animal species and a higher mutation rate due to relaxed selection [82]. A few exceptions, mainly in marine and freshwater mussels, include biparental and doubly uniparental inheritance and very rarely recombination [83,84,85,86,87,88]. MtDNA has been used in studies related to fisheries, aquaculture, conservation genetics, estimation of introgression from restocking, introductions and invasions, phylogenetic and phylogeographic analyses, and seafood testing and analysis (Table 2).
The marker MtDNA cytochrome c oxidase 1 (COI) has been commonly used for aquatic species identification—barcoding [149,150,151]. However, mtDNA introgression from one species to another has been reported in populations in which hybridization occurs [152,153,154,155]. In studies of population differentiation, mtDNA often shows incongruent results in comparison to allozymes and nuclear DNA markers, which limits its applications as a versatile molecular marker. Nevertheless, mtDNA variation in population studies can be used as one of the tested markers or to solve specific questions concerning the identification of taxa and female line genealogy.
Nuclear DNA markers have increasingly been used since the 1990s. They were originally based on the detection of single changes in the genome: point mutations or rearrangements in a very limited number of genome sites recognized by cuts with restriction enzymes. DNA marker technologies and their applications in aquaculture genetics have been reviewed several times [82,156,157,158]. Popular genetic markers used in aquaculture-related research include restriction fragment length polymorphism (RFLP), randomly amplified polymorphic DNA (RAPD), amplified fragment length polymorphism (AFLP), simple sequence repeats (SSRs—microsatellites), single nucleotide polymorphism (SNP), and expressed sequence tags (ESTs). The nuclear DNA markers have enabled progress in the assessment of genetic variability and inbreeding, parentage assignments, species and strain identification, and the construction of high-resolution genetic linkage maps for aquaculture species; AFLP can be source of irreproducible results leading to uncertain conclusions, and their use is not recommended for routine population genetics assays. RAPD markers are dominant and do not distinguish between homozygotes and heterozygotes. However, AmpFLP coupled with sequencing is an efficient way of obtaining reliable information about genetic relationships between studied populations. rDNA-ITS markers have been commonly used for species identification–barcoding [159,160,161,162]. Restriction-site associated DNA sequencing (RAD-Seq) and its latest variants coupled with NGS has been extensively applied to generate population-level SNP genotype data [163]. The number of SNPs discovered through the application of high-throughput sequencing for a chosen non-model species can be large (exceeding hundreds of thousands). SNP arrays have been designed for genotyping such a large number of loci [164,165,166,167,168]. However, in studies of populations, a subset of diagnostic SNPs can be employed for cost-effective genotyping in individual laboratories. SNPs and microsatellites are regarded as high-resolution molecular markers. Comparisons of the results of SNP-based and microsatellite-based population studies has often demonstrated the higher accuracy of SNPs. Nevertheless, microsatellites are also effective in the discovery of population processes and kinship analysis and are considered relevant markers in population genetic research [14,169,170,171].
Microsatellites are highly polymorphic, neutral and co-dominant DNA markers based on a variable numbers of short, usually 2–4 bp, nucleotide repeats [172]. Microsatellites have been most widely applied in research related to populations, fisheries and aquaculture beginning in the early 1990s [173,174,175,176]. The next generation sequencing (NGS) genome assay enables the identification of many thousands of microsatellites of which a dozen or so highly polymorphic loci are usually sufficient for population genetic and aquaculture applications, e.g., in [177]. This review is an attempt to summarize the applications of microsatellite DNA markers to studies of natural populations, aquaculture stocks, fisheries and conservation genetics of aquatic animals (Figure 1).

2. Genetic Structure of Wild Populations

Microsatellites have been most useful in terms of genetic characterisations and in the structure analysis of natural populations of many species. Examples of aquatic species for which population genetic structure has been successfully identified are presented in Table 3.
The list includes freshwater, diadromous and marine species. For the study of the genetic structure of the European hake, Pita et al. [177] distinguished several groups of microsatellite markers. The most commonly used sets of microsatellites are neutral. Microsatellites detected in ESTs are characterized by high allelic diversity in regions of functional importance. This group of microsatellites includes neutral markers with non-functional (non-adaptive) polymorphisms, purified (fossil-adaptive) polymorphism, and stabilizing (adaptive) polymorphism located in the adaptive regions of the genome [177]. In the latter group of markers, outliers can occur. These types of microsatellite markers give different pictures of the genetic diversity of populations, as in the case of the European hake. However, for some marine species, no genetic structuring has been observed. Genetic differentiation has not been found with microsatellites in some migrating oceanic species, such as the squid species Loligo reynaudii and Doryteuthis (Amerigo) pealeii [213]. Three different marker types, mitochondrial DNA, microsatellites and SNPs, were used in the analysis of 85 archaeological herring bones in an attempt to reconstruct the genetic diversity and population structure of ancient Pacific herring (Clupea pallasi) populations from the west coast of North America [214]. MtDNA revealed high haplotypic diversity, which is also present in contemporary populations, but no differences between populations. Microsatellite DNA data quality of ancient samples was very poor due to high allele drop-out and stuttering. SNP data had low error rates and were suitable for finding genetic differentiation. However, a recent study of the tiger shark, Galeocerdo cuvier, with microsatellites revealed differentiation between populations from the Atlantic and Indo-Pacific including contemporary and archival samples [215].
A simulation study of the accuracy of assigning individuals to closely related populations of chum salmon, O. keta, with 15 microsatellite loci exhibiting 349 independent alleles compared with 61 SNPs exhibiting 66 independent alleles, revealed that the SNP baseline performed considerably better than the microsatellite baseline [216]. An equivalent level of individual assignment to populations of chinook salmon, O. tshawytscha, in 60 populations from British Columbia obtained with 16 highly polymorphic microsatellite loci was projected to require 179 SNPs [217]. A population-specific estimation for this same species in the Yukon River showed that the nine-SNP baseline was approximately equivalent to a single microsatellite locus with 17–22 alleles [218]. Microsatellites with more alleles provided more accurate estimates of stock composition than those with fewer alleles. Microsatellite DNA markers as neutral markers do not reveal all occurring differences between natural populations, in which adaptive genetic markers, e.g., three circadian clock genes (OtsClock1b, OmyFbxw11, and Omy1009UW) discriminate temporally divergent migratory runs of Chinook salmon in the Feather River [219]. The resolution powers of two sets of SNPs-, a RAD-seq generated SNP panel and SNP (5568 loci) array developed by the Centre for Integrative Genetics (CIGENE, Norway), and a microsatellite panel for the assessment of differences among transatlantic populations of Atlantic salmon, S. salar, have been compared [220]. Both SNP sets and the microsatellite panel confirmed genetic divergence between the east and west Atlantic populations. Evidence consistent with introgression among the east and west Atlantic groups was found in the SNP data sets but not in the microsatellite data. That work highlighted the usefulness of multiple marker comparisons in identifying introgression. However, the costs of genotyping a large number of SNPs in populations exceed those of microsatellites. The application of high-throughput sequencing facilitates the discovery and further use of microsatellite markers in studies of genetic structures of wild populations, e.g., [221].

3. Population Genetics of Invasive Species

Invasions of single or groups of aquatic species have been caused by natural events or human-mediated activities, such as shipping, aquaculture, recreational activities or attempts to enrich local ecosystems, including increasing fishing catches. Despite the understanding of the negative consequences of uncontrolled introductions of species into new areas, their numbers continue to increase, and the economic impact is increasing. There are examples of banning the entry of large ships with hulls contaminated by sessile exotic marine organisms to certain countries, e.g., New Zealand. The use of molecular markers makes it possible to identify invasive species, reconstruct their migration routes and source populations, reveal their hybridization with local congeners, estimate the economic and ecological impacts, predict their further spread and facilitate preventive actions.
Population genetics of invasive species have been examined by studies of microsatellites in many species, including the evolution of brown trout populations originating from Poland and introduced to virgin rivers systems of the subantarctic Kerguelen Islands [222], smallmouth bass, Micropterus dolomieu, introduced to the range of Guadalupe bass, M. treculii in central Texas and its introgressive hybridization in a few rivers [223], Asian black carp, M. piceus, imported to U.S. aquaculture farms in the 1980 and present in the Mississippi River basin since the early 1990s [224], bighead carp, Hypophthalmichthys molitrix, and silver carp, H. nobilis, in Hungary, imported from China [225], and signal crayfish, Pacifastacus leniusculus, introduced to Europe from North American lakes for hatchery purposes [226].
Microsatellites have extended the knowledge of invasive populations of estuarine and marine species, such as rainbow trout, O. mykiss, native to the Pacific coast of North America and brown trout, S. trutta, from Europe, both introduced from Germany to South America over 100 years ago [227]. They have revealed the successful introduction of sockeye salmon, O. nerka, into a new environment, Frazer Lake, Kodiak Island, Alaska [228], estuarine fishes, dusky flathead, Platycephalus fuscus, and sand whiting, Sillago ciliata, objects of recreational and commercial fishers on the east coast of Australia [229], and the origin (source population) of invasive American brine shrimp, Artemia franciscana, in the Mediterranean Sea region [230], Manila clam, Ruditapes philippinarum, in the Mediterranean introduced for aquaculture purposes from the Indo-Pacific in 1983 [231,232], invasive brown mussel, Perna perna, in the Gulf of Mexico [233], and American Pacific oyster, C. gigas, introduced for aquaculture and present as a wild populations in Ireland, France, The Netherlands [234,235] and many other countries.

4. Conservation Genetics

Despite the great understanding of the need to protect entire aquatic ecosystems, actions to protect and restore individual species often remain difficult. One way to support these activities is to plan and manage endangered populations, taking into account their genetic characteristics. Based on the results of population genetics research, it is possible to define evolutionary conservation units [236]. Significant evolutionary units that are communities of individuals of the same species that are adapted to local conditions and reproductively self-sustaining. Supportive breeding and restocking is one way to prevent the decline of populations of aquatic animals [237].
Microsatellites have become a tool in the increasing the development and application of conservation genetics, and a few examples of its uses can be listed: analysis of populations of endemic California Paiute cutthroat trout, O. clarki seleniris, threatened by hybridization with introduced rainbow trout, O. mykiss [12], maintenance of the genetic diversity of steelhead, O. mykiss, tributary populations from the Bulkley-Morice River, British Columbia [238], analysis of the genetic variation of rainbow trout, O. mykiss, below and above natural barriers and man-made dams in rivers in California [239], description of population subdivision and conservation implications of westslope cutthroat trout (O. clarki lewisi) on the northern periphery of its range [240], and the effects of stocking in populations of Chinook salmon, O. tshawytscha, in the North Fork Stillaguamish River, WA, USA [241]. Significant genetic differences between seasonal runs of sockeye salmon revealed by microsatellite DNA analysis provide support for the management strategy that has been employed for nearly 20 years to protect the genetic diversity of this species in Bear Lake, Alaska [242]. Microsatellites were used for analysis of native white-spotted charr populations of Salvelinus leucomaenis threatened by hybridization with non-native brook trout, S. fontinalis, in Japan [243] and analysis of genetic structures among stocked and native populations of the European grayling, Thymallus thymallus, in Europe [244].
The Fraser River system consisted of five white sturgeon (Acipenser transmontanus) management units, two of which were listed as endangered populations under Canada’s Species at Risk Act, which were verified with microsatellite markers [245]. Microsatellites were applied in the assessment of using barriers for the conservation of native salmonid populations threatened by non-native migratory fish invasions [246], analysis of population genetic structure of endangered freshwater pearl mussel (Margaritifera margaritifera L.) in Europe, elaboration of the recommendation concerning conservation units [247] and study of population genetics of the exploited oyster, Crassostrea rhizophorae, in Brazil [248].
In order to monitor spawning success in earthen ponds, individual broodfish of the channel catfish, Ictalurus punctatus were identified prior to stocking, by genotype analysis with polymorphic microsatellite DNA markers [249]. Effective population size and demographic rate estimation can be performed with microsatellites, e.g., in wild Atlantic sturgeon, Acipenser oxyrinchus oxyrinchus [250]. Translocations and reintroductions have been used to prevent the extinction of freshwater fish populations of the hardyhead, C. fluviatilis, in The Murray–Darling Basin in south-eastern Australia [203]. The guided artificial gene flow strategy was based on genetic analyses of 14 microsatellite loci and enabled the rescue of this species from extinction. All 46 species of seahorses, Hippocampus, are listed as subjects of different types of protection from over-exploitation and are of particular conservation significance [169]. A set of up to 24 microsatellites has been constructed for cross-species amplification of 15 of species, which will facilitate their further conservation activities.

5. Identification of Management Units and Mixed Stock Analysis

Fishery management aims to determine the maximum catch size, a level that allows exploited populations of aquatic animals to be restored for sustainable harvesting. Marine waters are divided into geographical and administrative areas (management areas) inhabited by theoretically separate stocks. The stocks correspond to spatial management units. The catch limits in Europe (total allowable catches—TACs) are set annually for each management area. There is a growing understanding among fishery managers that the management units should be delineated using not only environmental but also biological criteria, including the genetic structure of stocks and populations. Mismatches between management units and population genetic structure have often been reported [251]. Microsatellites have been used successfully for identification of management units of exploited species, e.g., four species of Pacific salmon [59], freshwater fish, Piaractus brachypomus, in the Orinoco and the Amazon basin in South America [116], Korean rockfish, Sebastes schlegeli [252], common snook, Centropomus undecimalis [253], geoduck clams, Panopea abrupta and red sea urchins, Stronglocentrotus franciscanus, in British Columbia [254], analysis of relatedness in natural populations of brown trout, S. trutta, in Denmark [255,256], parentage analysis and reproductive success at spawning sites of sedentary brown trout, S. trutta [257], and impacts of fishery on populations of targeted and by-catch species such as the cod, G. morhua, population in the Flamborough Head area, the North Sea [258]. Studies of reproductive success in Atlantic cod revealed that females and males achieved their highest reproductive success when breeding with mates larger than themselves. Therefore, size-selective harvesting may have negative consequences for population recovery due to reductions in the mean body size of commercially exploited marine fishes [259].
Overfishing effects on dusky kob, A. japonicus, in South Africa [260], stock composition in mixed stock fisheries as in Fraser River (British Columbia, Canada) sockeye salmon, O. nerka [261], mixed-stock analysis of Chinook salmon, O. tshawytscha, from the Yukon River, Alaska [218], mixed stock analysis of Lake Michigan’s lake whitefish, Coregonus clupeaformis [262], mixed-stock analysis of American shad, Alosa sapidissima, in two Atlantic coast fisheries, Delaware Bay, USA, and Inner Bay of Fundy, Canada [263], mixed-stock analysis of brown trout, S. trutta from the Gulf of Finland [264] have all been studied with microsatellites. Genetic homogenization between the wild Vindelalven salmon population and hatchery stocks of the Angermanalven and Lulealven was observed over 1985–2003, confirmed extensive straying from geographically distant hatchery releases into the wild salmon population and indicated genetic risks associated with large-scale stocking practices in the Baltic Sea [265]. Only rare examples of published reports on selective angling can be found: angling captures more hatchery released and hybrid brown trout in comparison with wild individuals from stocked populations in the Doubs River, Switzerland [266].

6. Population Genetic Structure over Time

Stability in the genetic structure of populations throughout the age cohorts (juveniles and adults) has been related to a sufficiently large effective population size, which prevents genetic drift over generations, e.g., that detected with microsatellites and mtDNA in wild populations of the Antarctic toothfish, Dissostichus mawsoni [267]. Based on microsatellite detection, the following have been reported for some other species: the long-term stability of the genetic diversity of the declined population of the Japanese eel, Anguilla japonica, in the north of Taiwan from 1986 to 2007 [268], the temporal stability over the last 45 years of a pike, Esox lucius L., stocked population in Stege Nor, Denmark [269], the stability of Atlantic herring, C. harengus, in the Baltic Sea and Skagerrak waters over a 24-year period [28], in the Gulf of St. Lawrence in Canada over a 80-year period despite intensive fishing [270], and in wild Australian populations of barramundi, Lates calcarifer, over 25 years [271], the genetic composition of cod, G. morhua, in the Western Bank over a few years [272], and gilthead sea bream, Sparus aurata, in wild samples from the Aegean and Ionian Seas [273]. Historic angling records suggest the occurrence of a drastic decline in the Atlantic salmon population size in the River Eo, Asturia, Spain, during the past two decades; but high levels of diversity found with microsatellites suggest that the population has not been greatly affected by the historical declines and can be expected to recover in the future [274].
On the other hand, temporal changes in the genetic structure of wild populations have been reported for brown trout, Salmo trutta, before and after population decline and for its stocking with non-local strains of hatchery trout in rivers in Denmark over 60 years [275,276,277,278,279], and for populations of Lake trout, Salvelinus namaycush, in the upper Laurentian Great Lakes of North America after their substantial decline in abundance and distribution during the mid-twentieth century and following their recovery with the partial contribution of enhancement from hatcheries [280]. Historical analysis of genetic variation reveals the low effective population size of a northern pike, Esox lucius, population and the 8% loss of its heterozygosity over a 32-year period [281], the temporal changes in populations of Atlantic salmon in northern Spain over 20 years [282] and in Denmark for 60 years [283], the introgression of introduced Scottish strains in wild Atlantic salmon populations in southern France assessed through historic scale collections from 1970–1997, and the changes in the genetic composition of in wild rainbow trout populations after the chemical spill in the upper Sacramento River, which generated significant effects over time (1993–1996) on the genetic population structure of rainbow trout throughout the entire upper river basin [284].

7. Stocking Effects and Restoration of Wild Populations

Hatchery populations may undergo genome-wide selective sweeps that can affect their fitness and linked neutral loci, such that individuals destined to be released to the wild, should be modified to minimize genetic adaptation to captivity [285]. Temporal genetic variation in the endangered eastern freshwater cod, Maccullochella ikei, has been found in the Clarence River system, eastern Australia, using microsatellite DNA markers [286]. Comparison between historical extirpated and restocked populations revealed a significant loss of heterozygosity and allelic richness. Released hatchery-produced material has contributed to the genetic decline in the largest wild M. ikei population. This observation demonstrates the adverse effects of stocking programs and the necessity of support from genetic analysis in the design of the management of breeding and stocking strategies, particularly for threatened species [286]. Microsatellite markers revealed very low levels of genetic diversity in the Kootenai River white sturgeon population. The conservation aquaculture program captured 96% of the population’s microsatellite diversity in hatchery-released progeny in only 10 years by using high numbers of broodstock. A panel of 18 microsatellite loci has been validated for parentage analysis [287].
Effects of stocking on the genetic integrity of Arctic charr, Salvelinus, populations have been found in two lakes in the Bavarian Alpine region [288]. A loss of genetic integrity has been observed in stocked populations of lake trout (S. namaycush) from 72 unstocked and stocked lakes in Canada, in which an increase in genetic diversity and a twofold decrease in the extent of genetic differentiation among stocked populations compared to that among unstocked populations has been found [289]. Possible changes in the effective population size of brook charr, S. fontinalis, in Québec, Canada, related to long term stocking have been reported [290].
Stocked populations were characterized by significant admixture at both population and individual levels, in populations of brown trout, Salmo trutta in the Borne River in the Northern French Alps [291], stocked populations on Funen Island, Denmark [292] and populations in tributaries to the Limfjord, Denmark [256]. Admixture was found in populations in which population structure was highly affected by multiple stocking and river diversion in a high mountain national park in Norway [293], in restocked populations in Asturias, Spain [294], and as impact of supportive breeding to enhance populations of salmon in Asturias, Spain [282], sea trout S. t. m. trutta in Polish rivers, southern Baltic [295,296], amago salmon, Oncorhynchus masou ishikawae in Japan [297], Japanese chum salmon, O. keta [298] and in striped bass, Morone saxatilis populations in the south eastern USA [299].
Enhancement of wild populations has generated changes such as the impact of straying of sea ranched hatchery-reared Atlantic salmon, S. salar, on the genetic composition of populations within the small Ellidaar river system in SW Iceland [300], differences in the reproductive success of released natural and hatchery salmon in the Swedish river Dalalven [301,302], differentiation of admixture rates in samples of salmon collected between 1998 and 2006 compared to samples from 1965 to 1987 in France suggesting the similar rising, long-lasting or short-term impacts of stocking with captive-bred fish [303,304,305], and the enhancement of populations of Atlantic salmon in the Connecticut River and Penobscot River, USA [306]. Alleles originating from stocking with non-native, mainly Scottish fishes performed in the 1970s–1990s are present in the contemporary populations of salmon in the River Sella, Spain, which confirms the long term effects of introgression into pristine populations [307].
Microsatellite DNA polymorphism evaluation of hatchery-based stock enhancement of black sea bream, Acanthopagrus schlegelii, in the South China Sea revealed the significant genetic differentiation of enhanced populations from native populations and the lower genetic diversity of the recaptured released groups of individuals [308]. It has been concluded that the release of cultured juveniles with lowered genetic quality is potentially harmful to the conservation of wild genotypes in native populations. The impact of releases of hatchery-reared fish on natural populations of red sea bream, Pagrus major, in Sagami Bay and Tokyo Bay in the Kanagawa Prefecture [309], Kagoshima Bay, Kiusiu [310], Shikoku Island [311], and black sea bream, A. schlegelii, in Hiroshima Bay, Japan [312], and the genetic effects of nearly three decades of Murray cod (Maccullochella peelii peelii) stocking in five river catchments in southern Australia [313] have been reported. Changes in wild populations have also been caused by releases of Japanese flounder, Paralichthys olivaceus [314]. Restocking programs have been developed as a conservation method for a tropical fish, the pacu, Piaractus mesopotamicus, because of declines in the number of wild populations in the Tiete and Grande rivers, Brazil, to be accompanied by the genetic monitoring of populations and broodstock to ensure the viability of such programs [315].
Microsatellite analysis has contributed to the identification of the sturgeon Acipenser oxyrinchus in North America (as opposed to the European A. sturio) as a native extinct species in the Baltic Sea and to the genetic control of its imported fry from Canada before the release into the wild in attempts to achieve the restoration of its population in drainages [316,317], restoration of endangered populations of the Adriatic sturgeon (A. naccarii) endemic to the North Adriatic region, in the Ticino River Park, Italy [115], and the restocking of endangered populations of dusky grouper, Epinephelus marginatus, in the Mediterranean [318], and of white seabream, Diplodus sargus, in a fishery reserve in Sicily, Italy [319]. Microsatellites have been used for the estimation of the aquaculture potential of new species such as that of the Korean kelp grouper, Epinephelus bruneus [320], yellowtail amberjack, Seriola lalandi, in Chile [321] and a flatfish Senegalese sole, Solea senegalensis [322].
The recovery rate of hatchery-released red drum, Sciaenops ocellatus, in some Texas bays and estuaries has been studied [323], as well as that of rock carp, Procypris rabaudi— an endemic fish in the upper Yangtze River, China, supplemented with hatchery-produced fish [324]. Assessment of the impact of releasing hatchery-reared juveniles of Pacific abalone, H. discus [325], and using molecular pedigree reconstruction to evaluate the long-term survival of out-planted hatchery-reared larval and juvenile northern abalone, Haliotis kamtschatkana [326], and assessment of the negative effects of supplementing natural populations of the grooved carpet shell, Ruditapes decussatus, on Atlantic coasts of northern Spain using seeds produced in hatcheries [327] have been performed.

8. Escapees’ Impact on Natural Populations

To assess the impact of escapees from cultures and hatcheries, the fitness consequences of the introgression of fast-growing domesticated fish into a wild population were tested [328]. Fry from wild and domesticated rainbow trout (Oncorhynchus mykiss) crosses were released into two natural lakes. Parentage analysis was performed using microsatellite loci. The results indicated that domesticated fish can survive as well as wild fish. During the first summer, the fastest-growing crosses had the highest survival, but this trend was reversed after one winter and another summer. The experiment confirms the multigenerational risk of domesticated fish escaping or being released in the case of interbreeding with wild fish. Nile tilapia, Oreochromis niloticus, has been introduced throughout Africa, outside its native range for aquaculture purposes, and escapees hybridize with native populations of Oreochromis species, which result in negative effects on the conservation of fish biodiversity, aquaculture and capture fisheries in fresh water bodies [329,330]. Microsatellites have been used for identification of escapees of the Oujiang color common carp, C. carpio var. color, in China [331], analysis of natural hybridization between two species of Andean pupfishes (Cyprinodontidae; Orestias agassizii and O. luteus) mainly in the Lake Titicaca with implications for local fisheries, stocking and conservation [332], and identification of the occurrence and postulated hatchery origin of hybrids of Pseudoplatystoma corruscans and P. reticulatum in the Upper Paraná River, in South America [333]. However, modelling gene flow caused by escapees from a few farms simultaneously, revealed that changes detected in a wild population were lower when gene flow was simulated from one farm strain only [334].
In the marine environments, net-cage aquaculture poses a risk of the escape of a large number of fish in the case of mechanical damage caused by natural factors (e.g., tsunamis or typhoons) or related to human activities (e.g., shipping and fishing). In Japan, in the mariculture areas, the frequency of gilthead sea bream, S. aurata, escapes was estimated to be from 14.1% to 30.2% using microsatellites [335]. Hybridization of escapees in natural populations has been observed. Microsatellites were also used in the identification of escaped farmed gilthead sea bream in the Mediterranean area [273,336,337,338]. Escaped farmed cod, G. morhua, were identified in wild populations in Norway [339,340]. Farmed cod from genetically diverse populations grown outside their native range pose the threat of outbreeding depression if they escape and interbreed with wild fish [341]. Microsatellites were used for the assessment of the genetic impact of domesticated farmed escapees on native Atlantic salmon, S. salar, populations in Norway [342,343] and Iceland [300], for the assessment of the impact of European Atlantic salmon escapees from hatcheries in Nova Scotia and New Brunswick, Canada on native American, S. salar, populations [344], in studies of trophic and epidemiological interactions between salmon farms and the receiving ecosystem including cod preying [345], and for the assessment of the impact of escapees of hatchery European seabass, Dicentrarchus labrax, on natural populations in waters around Cyprus [346], and increased relatedness and possible inbreeding in wild populations because of escapees of tropical fish, barramundi L. calcarifer, from a sea-cage facility in northern Australia [347].

9. Comparison of Wild and Hatchery Stocks

Comparisons of feral populations and hatchery stocks using microsatellites have been published for many species (Table 4).
The lower genetic diversity of cultured populations in comparison with wild populations has been reported for many species, including fish, prawns, sea urchin, oysters, green mussels and abalones [371,372,373,374,375,376,377,378,379,380,381].
A small loss of genetic variation in comparison to wild populations was found for Atlantic salmon in a hatchery in Canada [382], and for hatchery brown trout stocks in Finland [383] and in Hungary [384]. Introduced to a Western Australian hatchery, stocks of rainbow trout (O. mykiss) are derived from imports from New Zealand, the latter being largely derived from Californian imports in 1883 and also having lower diversity in comparison with wild populations in the north Pacific [385]. Similarly, Tasmanian cultured Atlantic salmon had lower diversity when compared to the progenitor Canadian population [386,387]. A loss of genetic variation was reported in the hatchery stocks of Bleeker’s sheatfish, Phalacronotus bleekeri [388], and a loss of genetic diversity (rare alleles) was reported in the cultured stocks of the large yellow croaker, L. crocea [389], and starry flounder, Platichthys stellatus, in Korea [390], Florida bass, Micropterus salmoides floridanus [391], and channel catfish, I. punctatus, in farms in Mexico [392], Siberian sturgeon, Acipenser baeri [393], A. gueldenstaedti and A. ruthens in a farm in Poland [394], American paddlefish, Polyodon spathula, in Poland [395], barramundi, L. calcarifer, hatcheries in Australia [396,397], redclaw crayfish, Cherax quadricarinatus, introduced from Australia to culture in China [398], and common carp, C. carpio, in the Czech Republic [399]. A loss of genetic variation has also been found in Greek hatchery stocks of the European sea bass, D. labrax [400], three tilapia, Oreochromis, species cultured in Mexico [401], in hatchery strains of the Pacific abalone, H. discus hannai [402], and in cultured Pacific bivalve geoducks (Panopea generosa) in Washington state, USA [403].
The genetic analysis of pacu broodstocks, P. mesopotamicus, used in the stocking program of the Paranapanema River, Brazil, did not confirm reduction in genetic diversity [404]. Sriphairoj et al. [405] concluded from their study that at least 100 brooders (Ne) should be used in practices of managing the critically endangered Mekong giant catfish, Pangasianodon gigas, in Thailand. Studies of the banana shrimp, Fenneropenaeus merguiensis, in Australia demonstrated the loss of alleles in the mass selection program carried out, even with over 1000 broodstocks being compared with similarly selected but outbred stocks [406]. It is recommended to maintain different and independent lines instead of one line. A high level of genetic variability among Pacific white shrimp, L. vannamei, in Pernambuco, Brazil, is sustained by the exchange of breeders between marine shrimp hatcheries [407]. In an attempt to reduce the exploitation of the humpback grouper, Cromileptes altivelis, captive breeding has been performed, and the recommendation has been made to increase the effective population size with wild fingerlings in order to avoid diversity reduction detected by microsatellites [408]. Sex-linked microsatellites were found in Oncorhynchus [409].
Population parameters and the power of 16 microsatellites and 26 SNPs to assign single individuals to their sampling population in wild and farmed stocks of Atlantic salmon (S. salar) in Norway were estimated [410]. Microsatellite strain-specific alleles were found. The effectiveness of genetic assignment analysis of populations was almost the same for microsatellites and SNPs (96% of the individuals). The results of analysis of two wild and three cultured Pacific oyster (C. gigas) populations with a set of 18 microsatellites (8 genomic simple sequence repeat—SSR; 10 expressed sequences tag (EST)-derived SSR) and 10 EST-derived SNP markers suggest that genomic SSRs and EST-SSRs are more suitable for population genetic analysis than are SNPs [411].
Microsatellites are continuously used in studies of genetic diversity in cultured stocks [22,412,413].

10. Kinship Analysis of Aquacultured Stocks

Lowered genetic diversity in cultured stocks is usually caused by a low effective population size (a reduced number of breeders transmitting their genes to progeny). A loss of genetic diversity decreases the fitness and adaptive potential of the progeny. If the process lasts for some generations, it results in increased relatedness (kinship) among individuals and the effect of inbreeding. Analysis of relatedness has been performed for different strains of rainbow trout, O. mykiss, and can be used for the prevention of diversity loss in cultured stocks [414]. Effects of captivity rearing on fitness-correlated traits have been studied in endangered Atlantic salmon, S. salar, from the inner Bay of Fundy [415], and on individual reproductive success for fish from the Ste-Marguerite River, QC, Canada [416]. Reproductive success analysis and parentage assignment in culture have been conducted for the optimization of breeding protocols (e.g., by controlled mixing of gamete portions) for white seabass, Atractoscion nobilis, and red drum, Sciaenops ocellatus, for conservation purposes through stock enhancement programs in California [417].
Microsatellite loci have been applied in pedigree tracing of a hatchery strain of Japanese flounder, P. olivaceus, to be stocked in natural sea areas [418], parentage analysis and paternity success assessment of turbot, S. maximus L. in hatcheries in Spain [419,420], parentage assignment of turbot and rainbow trout in France [421], and Atlantic salmon, S. salar, in Ireland [422], parentage analysis and paternity success assessment of cod in Denmark [423], parentage analysis and identification of trait differences in survival and growth among a harvest of communally reared families of Atlantic cod, G. morhua, in Canada [424], identification of relatedness and differentiation of hatchery populations of Asian seabass (L. calcarifer) broodstock in Thailand [425], parentage analysis of and identification of spawning frequency and timing of brood dams and sires of red drum, S. ocellatus, in a marine fish stock-enhancement hatchery in the USA [426], and of European anchovy, E. encrasicolus, under a pilot project for aquaculture and enhancement of native populations in Spain [427], analysis of increased genetic relatedness in a hatchery stock in comparison with a wild Senegal sole, S. senegalensis, population [428], pedigree classification for giant grouper, Epinephelus lanceolatus, broodstock management in Taiwan [429], identification of parental relatedness in a naturalized population of Pacific oysters, Crassostrea gigas in Dabob Bay, Washington, USA [430], parentage analysis in Asian seabass for hatchery purposes [431], parentage assessment of blunt snout bream, Megalobrama amblycephala, crosses for a freshwater polyculture system in China [432], and greater amberjack, S. dumerili [433], estimation of parentage and relatedness in the polyploid white sturgeon, A. transmontanus [434], and experimental assessment of genetic tagging with multiplexed microsatellite markers and founder representation in hatchery-reared red drum (S. ocellatus) fingerlings used in stock enhancement [435]. A microsatellite-based multiplex PCR panel was constructed that allowed 95% of the offspring to be assigned to a single pair of parents of the meagre Argyrosomus regius to support the breeding program in the Mediterranean aquaculture [436].
Microsatellites have been employed in shellfish aquaculture for kinship analysis and genetic variation monitoring in a whiteleg shrimp, Litopenaeus vannamei, breeding program [437], identification of parentage markers in the swimming crab, Portunus trituberculatus [438], parentage analysis of H. discus hannai abalone mixed family farming [439] and in a South African hatchery of H. midae abalone [440], genotyping of individual H. asinina abalone larvae for parentage assignment in aquaculture in Australia in order to maintain the level of genetic diversity [441], genetic improvement in the clam, Meretrix meretrix, by crosses and parentage assignment [442,443], parentage analysis and identification of variation in reproductive success of Pacific oyster, C. gigas, in a hatchery in France [444,445], and parentage analysis of different color lineages of scallop, Patinopecten yessoensis [446].
Genetic analysis of broodstock and progeny of the European sea bass, Dicentrarchus labrax, with microsatellies, in aquaculture in the larval stage, when families of progeny had been mixed to start the production cycle, has been conducted to help attain balanced parental contribution [447]. The results from an analysis of spotted seatrout, Cynoscion nebulosus, dams and sires in two restoration enhancement facilities in Texas were assessed throughout a spawning year by using parentage analysis based on 12 variable microsatellite loci. That and other studies indicate that reductions in Ne of hatchery- or farm-raised progeny stem primarily from non-contributing dams, suggesting that periodic identification and removal of low-contributing dams from broodfish stocks constitute a critical step toward maximizing the Ne levels of hatchery offspring used in restoration enhancement [448]. The distribution of the F-1-selected breeders into spawning batches should be designed using co-ancestry data, in order to maintain optimal levels of genetic variability in the next generation. This procedure should be repeated for each generation [447]. Based on a simulation study Villanueva et al. [449] reported that, for a set of Atlantic salmon, highly polymorphic microsatellites show, in simulations, that the four most informative loci are sufficient to assign at least 99% of the offspring to the correct parental pair with 100 crosses involving 100 males and 100 females. An additional locus was required for correctly assigning 99% of the offspring when the 100 crosses were produced with 10 males and 10 females. The possibility of selective recovery of founder genetic diversity in aquacultural fish broodstocks has been pointed out [450]. In the case of a limited number of breeders in culture stock, breeding pairs can be matched based on their genetic profiles, obtained with microsatellite loci, in order to assure that only genetically diverse fish are mated [451,452].
Statistical analysis of parentage assignment has been conducted for artificially propagated hatchery fish (removed adipose fin) with parents originating from wild salmon (preset adipose fin) in a population of Chinook salmon, O. tshawytscha, from the Wenatchee River, Washington [453]. Simulations demonstrated a lower number of identified parents for hatchery fish in comparison with the number of wild parents. Wild populations of Arctic grayling, Thymallus arcticus, from the Lubbock River, Yukon, were sampled for adults and young-of-the-year independently in order to enable the identification of parent–offspring pairs [454]. The genotyping of samples with 38 microsatellites confirmed that a small number of families over-dominated the global number of full-sibs, a phenomenon that is well-known from hatchery stocks.
A comparison of SNP and microsatellite applicability in parentage and kinship assignment of a wild sockeye salmon (O. nerka) population in Alaska demonstrated that the assignment success of 80 SNPs (80 independent alleles) was higher than that of 11 microsatellites (192 independent alleles) but the identification of full-sib groups without parental information from relatedness measures was possible using both marker systems [14]. In a study of hatchery steelhead in the Snake River basin, it was confirmed that a panel of 17 microsatellites was comparable in accuracy in conducting parentage-based tagging (PBT) to a panel of 95 SNPs, and matched that using traditional coded-wire tags (CWT) [455]. The advantages of using microsatellites and SNPs in parentage assignment have been reviewed with an indication that the SNP-based method can benefit from the development of genomics [456]. The parentage analysis with the close-kin mark–recapture (CKMR) method based usually on a large number of SNP loci, has recently been worked out for a few species, such as salmon, thornback ray, Raja clavate, and the Pacific white shrimp, L. vannamei, with high potential for further applications [457,458,459]. Nevertheless, microsatellites remain the most accessible marker system in the preset day kinship analyses of cultured and wild populations [460,461,462].

11. Selection of Characters of Choice and Heritability

Selection for a breeding date has been conducted for a coho salmon, O. kisutch, population colonizing a new habitat, made accessible by the modification of the Landsburg Diversion Dam, in the Cedar River, Washington, USA [463]. Microsatellites were used for parentage analysis. The offspring of fish arriving earlier to the spawning ground were larger. Larger fish produced more offspring. The hypothesis of ‘pathogen-driven selection in the wild by means of frequency-dependent selection or change in selection through time and space’ has been confirmed by the results of studies of the correlation between the major histocompatibility complex (MHC) class II beta variation and the pathogen infection levels in wild populations of Atlantic salmon, S. salar, in Quebec, Canada [464]. Microsatellites were used for monitoring changes in the studied populations. Studies on disease resistance and polymorphisms of major histocompatibility genes, genotypes frequencies and control microsatellite loci in the parr and migrant stages in the wild in Atlantic salmon S. salar in Ireland revealed that the additive allelic effects were more likely to determine survival, which highlights the importance of preserving genetic diversity in the wild [465]. White spot disease caused by white spot syndrome virus (WSSV) infection affected shrimp culture throughout the world. Resistant P. monodon shrimp identified by microsatellite DNA markers in different seasons were collected from natural populations along the entire east coast of India with the aim to obtain broodstock to prevent a repeated outbreak of white spot disease in a hatchery [466,467].
Selection programs have been executed with the aim to improve the quality of cultured fish and shellfish for releases to the wild or for consumption. Microsatellites were employed in selection in the larval stage for the faster growth of adults of Asian sea bass (barramundi), L. calcarifer [468], selection for growth in the European sea bass, D. labrax [469], genetic comparison of different hatchery strains of rainbow trout, O. mykiss, from the Northwest in the USA selected for growth and immunological response [470], commercial selection using DNA parentage assignment in rainbow trout aquaculture [471], confirmation of the ‘hypothesis of a dominant mutation mechanism’ as a possible cause of rib and vertebral deformities found in farmed rainbow trout [472], and a study of the impact of domestication on the stress response and immune modulation in Eurasian perch, P. fluviatilis [473]. Selection for performance in salt water and S. salar salmon broodstock development for hatcheries in New Brunswick, Newfoundland and Labrador in Canada, included genetic analysis, an estimation of heritability of bacterial kidney disease, sea lice, growth, fillet yield and deformities [474]. Immune-related loci were identified as FST outliers in pairwise comparisons of samples at a 10-fold higher frequency than that of neutral loci in Atlantic salmon, which means that well characterized immune-related loci as well as neutral loci (microsatellites) in cultured species can be useful when disease control and prevention is a goal [475]. Assessment of the genotype by environment interaction for the growth of sole (S. solea) in an intensive recirculation aquaculture system (RAS) and in a semi-natural outdoor pond (POND) has shown low genetic correlations for growth between environments, which implies that the best genotypes in an intensive aquaculture environment are not necessarily the best genotypes in more natural environments such as ponds [476].
Microsatellites were used for parentage identification in studies of heritability and genotype by diet interactions of European sea bass (D. labrax) affecting fish weight and size at age [477] and flesh quality, studies of the heritability of cold tolerance under progressive temperature decrease, and of acute cold-stress tolerance in red drum, S. ocellatus, in southern USA [478,479]. Arctic charr, S. alpinus, eggs were obtained for research from two North American sources, an eastern (Fraser River, Canada) and western (Bristol Bay, Alaska, USA) stock and fishes were mixed and grown for 2 years in tanks [480]. Genetic correlations between body size traits were highly positive and significant. The genetic correlation of fillet fat with fillet color was positive and significant. The eastern stock was composed of an admixture of two sources; the commercial stock was composed of three different sources, and the western stock was composed of three to four source populations as inferred from 480 microsatellites. The heritability of Atlantic salmon flesh color traits was low to medium with carotenoid content in the flesh exhibiting the lowest additive genetic variation [481]. The heritability of morphological abnormalities in gilthead seabream, Spares aurata, has been shown to be significant [482]. Microsatellite-based parentage pedigree analysis is continuously used in the present day research of the heritability of different traits in fish and shellfish species [483,484,485].

12. Gynogenesis Assessment

Gynogenesis involves fertilization of eggs with UV-irradiated sperm (with DNA deactivated) exposed to the cold or chemical shock followed by pressure treatment [486,487]. Inhibition of the mitotic cleavage results in homozygous double haploid (mitotic) gynogenetic progeny. Maturing female eggs fertilized with irradiated sperm and interrupted meiosis (meiotic gynogenesis) are clones of the same female sex. Gynogenesis occurs very rarely in wild populations, but is used for research and culture purposes to obtain all-female specimens. Such monosex females are more economical in commercial culture because of their faster growth. Microsatellites have often been used for genogenetic assessment, e.g., in the case of African catfish, Clarias gariepinus [488], in the case of gynogenetic diploids being generated in order to map centromeres of walking catfish, Clarias microcephalus [489], large yellow croaker, Pseudosciaena crocea [490] (Miao et al., 2015), turbot, S. maximus, in Spain [491], half-smooth tongue sole, Cynoglossus semilaevis, [492], shortnose sturgeon, Acipenser brevirostrum Lesuere [493], Siberian sturgeon, A. baeri Brandt [494,495], and starlet, A. ruthens [496], in studies of the sex determination system in ship sturgeon, Acipenser nudiventris, using meiotic gynogenesis [497], in the assessment of gynogenesis in stellate sturgeon, Acipenser stellatus [498], wels catfish, Silurus glanis [499], red crucian carp [500], and Japanese flounder, P. olivaceus [501], and in the analysis of gynogenetic diploids induced by heterologous sperm in Chlamys farreri [502].
Microsatellites were used for assessment of androgenesis (originating from males - all-males), e.g., in the loach, Misgurnus anguillicaudatus [503], amago salmon, O. masou ishikawae [504], and in large yellow croaker, P. crocea [505]. Microsatellites were useful in the analysis of ploidization, including evolutionary polyploidy, which occurred through the hybridization of common carp, C. carpio, approximately 12 MYA [506], functional hexaploidy in the shortnose sturgeon, A. brevirostrum, ‘functional genome reduction’ in other species of sturgeons [507], natural ploidization and induced gynogenesis in the loach, M. anguillicaudatus, and blunt snout bream, M. amblycephala [508,509], parental assignment of natural triploids, diploids and laboratory induced gynogensis of loaches, M. anguillicaudatus, in the Hokkaido island, Japan [510], analysis for laboratory and aquaculture purposes of ploidy levels in Acipenser hybrid larvae [511] and tests of Mendelian segregation in bester—a hybrid of beluga, Huso huso L., and sterlet, Acipenser ruthenus L., in the fourth generation [512].
Microsatellites have been found to have applications in research concerning the induction of triploids in cod, G. morhua [513], induction of triploids and gynogenesis in Senegalese sole, S. senegalensis, for aquaculture [322], pressure and cold shock induction of meiotic gynogenesis and triploidy in the European sea bass, D. labrax [514], parental assignment in triploids of Pacific oysters, C. gigas, for improved selection for fast growth in aquaculture [515], production of tetraploids [73], characteristics of spotted mandarin fish, Siniperca scherzeri, and X (female) mandarin fish, S. chuatsi (sic), F1 and F2 hybrids [516] (Li et al., 2014), assessment of hybrids of Haliotis rufescens with H. discus hannai produced in Chile [517] (Lafarga de la Cruz et al., 2010) and H. discus hannai Ino with Haliotis gigantea in China [518] (Luo et al., 2010), and the hybridization of the swimming crab, P. trituberculatus, distributed in the coastal waters of Asia-Pacific with the aim to increase its performance in hatchery [519]. Sex-linked microsatellite alleles were found in gynogenetic individuals of turbot, S. maximus [520]. Microsatellites are applied in assessment in the most recent research concerning gynogenesis, androgenesis and ploidization [521,522,523].

13. QTL Identification

Tolerance to environmental stress factors such as temperature, salinity, low oxygen (hypoxia), changes in pH, food availability and biological factors, resistance to viral and bacterial pathogens and parasites determine survival under natural conditions and productivity under aquaculture aconditions. Genomic regions associated with the trait of interest can be mapped in order to identify quantitative trait loci (QTL). Microsatellites have been commonly used successfully for finding the linkage of molecular markers with QTL. Subsequently, such markers can be used in marker-assisted selection (MAS) in aquaculture. Analysis of genomic sequences and genetic differentiation of associated tandem repeat markers (microsatellites) in the growth hormone somatolactin and insulin-like growth factor-1 genes of the sea bass, D. labrax, proved that gene-associated markers are more efficient than formerly used anonymous microsatellite loci at providing a clear picture of genetic differentiation [524]. Genetic linkage analysis is an effective method for identifying quantitative trait loci (QTL) associated with resistance to a disease [525]. Microsatellies have been intensively applied in the construction of genetic linkage maps, e.g., for salmon, S. salar [526,527], for Arctic char (S. alpinus) using two backcrosses between genetically divergent strains [528], a common carp, C. carpio [529,530], European sea bass, D. labrax [531], gilthead sea bream, S. aurata [532], and Pacific abalone H. discus hannai [533].
A genetic map was developed for a population of European sea bass with the use of ove 90 microsatellites and enabled the finding of two QTL for body weight, six QTL for morphometric traits and three suggestive QTL for stress response [534]. Genetic linkage maps have been constructed for red drum, S. ocellatus [535,536], barfin flounder, Verasper moseri, spotted halibut, Verasper variegatus [537], Atlantic halibut broodstock management in Canada including tentative QTL for pigmentation, body size and eye migration [538], and for turbot, S. maximus, and one possible QTL associated with body length was detected [539]; however, the observation of a high mean variation between traits among families made it difficult to evaluate QTL effects [540]. Microsatellite genetic linkage maps were elaborated for brill, S. rhombus, for a preliminary study on growth-related QTL for body weight, length and Fulton’s condition factor [541] and for four tilapia species [542]. Located on the maps were QTL found to be related to sex determination in Mozambique tilapia [543], QTL for the spawning date of Coho salmon, O. kisutch [544], QTL influencing early maturation [545] and upper thermal tolerance in outbred strains of rainbow trout, O. mykiss [546]. Thirteen QTL markers for spawning time representing seven linkage groups were found, and eight markers from five linkage groups showed consistent effects in two sampling years, which suggests this trait is highly polygenic [547]. for a high heritability of body mass and the condition factor and a moderate heritability of the age of sexual maturity of males was found in two cultured strains (Rainbow Springs and Spring Valley) of rainbow trout [548]. Faster growing individuals were more likely to mature at two years of age than slower growing individuals. The location of microsatellite markers of body mass QTL in linkage groups was reconfirmed and new ones were detected. Seven tentative and three significant QTL were detected in families that exhibited high or low plasma cortisol concentrations in response to crowding stress in rainbow trout culture production.
QTL for morphometric traits in gilthead seabream, S. aurata [549,550,551], and QTL for short-duration vigorous swimming movements in common carp (C. carpio) based on LDH Activity [552] have been found. A first-generation microsatellite-based linkage map was created for the Chinese mitten crab, Eriocheir sinensis, and used for QTL detection, and nine growth-related QTL for body length, width and weight were mapped to seven linkage groups [553]. Additionally, two QTL were identified to be associated with sexual precocity. The linkage maps and the identified QTL will be valuable for marker-assisted selection breeding programs. Genetic linkage maps were elaborated for the Pacific lion-paw scallop, Nodipecten subnodosus, and putative QTL were identified for morphometric traits [554], QTL for pearl quality traits in the freshwater triangle pearl mussel, Hyriopsis cumingii in China [555], and for the growth rate in the blacklip abalone, H. rubra from Australia [556] were found, QTL for size in Bay scallop, Argopecten irradians [557], was found, and a linkage map for Chinese shrimp, Penaeus (Fenneropaeneus) chinensis [558] was constructed. A total of 159 microsatellite markers were selected from genetic linkage maps of Japanese flounder, P. olivaceus, and F-1 progeny of crosses between disease-resistant and disease-susceptible parents were used for the detection of QTL associated with resistance to Streptococcal disease (streptococcosis) caused by Streptococcus iniae [525]. These candidate QTL regions have been found. QTL with significant effects on infectious pancreatic necrosis (IPN) resistance in Atlantic salmon, S. salar, using a genome scan [559], and in rainbow trout, O. mykiss [560] (Ozaki et al., 2001), were found, and candidate QTL for infectious salmon anemia (ISA) resistance in Atlantic salmon, S. salar [561], were found. Bacterial cold water disease (BCWD) causes significant economic loss in salmonid aquaculture, and previously, genetic variation was detected in survivors following challenge with Flavobacterium psychrophilum, the causative agent of BCWD in rainbow trout, O. mykiss [562]. The nine major QTL identified in that study are candidates for fine mapping to identify new markers that are tightly linked to disease resistance loci for use in marker-assisted selection strategies.
Backcrosses of rainbow trout and steelhead (O. mykiss) were used to construct a linkage map and to identify associations between molecular markers, QTL determining resistance to infectious hematopoietic necrosis virus (IHNV) and associations were observed for four of the markers [563]. Candidate QTL were found for resistance to viral nervous necrosis disease (VNN) in Asian seabass [564], tentative QTL were found for resistance of gilthead sea bream to fish pasteurellosis caused by a bacterial pathogen, Photobacterium damselae, subsp. Piscicida [565], microsatellite loci were associated with growth-related traits in the Manila clam R. philippinarum [566], QTL connected with stage-specific inbreeding depression were found in the Pacific oyster, C. gigas [567], and QTL were found in the flat oyster, O. edulis, for resistance to a parasite, Bonamia ostreae [568]. The discovery of QTL for resistance to summer mortality and OsHV-1 infection in the Pacific oyster (C. gigas) opens new possibilities of selection for resistance to oyster herpesvirus, OsHV-1 [569], and applications in the epidemiology of livestock species, such as in flat oysters, O. edulis, in France and around the world, with worldwide mortality being observed since 2008 [570,571,572]. A characterization of novel EST-SSR markers and their correlations with growth and nacreous secretion traits has been performed for the Pinctada martensii pearl oyster, the primary cultured species of marine pearls in southern China [573]. Twenty-nine novel polymorphic microsatellite markers were developed to facilitate marker-assisted selection in the genetic improvement of this species.
Several QTL related to phenotypic variation have been identified in the salmon, S. salar, genome. Population differentiation and assessment of linkage disequilibrium in chromosomes containing QTL for body weight, infectious pancreatic necrosis resistance and infectious salmon anaemia resistance to detect the selection history at the genomic level in Atlantic salmon have been used by Martinez et al. [574]. They demonstrated that the body weight QTL (marker SSA0343BSFU on chromosome 3) has been under directional selection. This marker is physically mapped near the coding sequence of DVL2 (for segment polarity protein disheveled homolog DVL-2) and is a good candidate gene related to the quick response to selection for growth. However, only low diversifying selection was found in the QTL associated with infectious pancreatic necrosis and infectious salmon anemia resistance. Due to their rather high selection intensity, individual loci may undergo indirect selection and increased inbreeding. Therefore, it can be concluded that artificial selection has inflicted significant changes to the Atlantic salmon genome, validating the QTL in cultured salmon populations used in industry production resulting from the recent selection history [574]. This conclusion can be ubiquitously applied to other cultured aquatic species.
Microsatellites have been used also in studies of phenotypic changes in transgenetic C. carpiovar var. Jian [575]. Microsatellites can also be useful markers for the identification of novel transposable elements, such as the major histocompatibility complex (MHC) class I region of a teleost, medaka, Oryzias latipes [576]. Population genetic differentiation has been found in parasitic sea lice (Lepeophtheirus salmonis) on Atlantic and Pacific salmonids from wild and farmed fish, by analyses of microsatellite DNA [577]. Microsatellite identification coupled with SNP discovery have been used recently for mapping QTL for such traits as cold-tolerance and disease resistance of farmed tilapia [578,579], and for identifying the occurrence of skeletal abnormalities in gilthead seabream [580]. Nevertheless, it should be noted that SNP markers are increasingly used for research related to QTL.

14. Conclusions

Simple sequence repeats (SSR—microsatellites) have been found to have a variety of applications in genetic research related to natural populations and cultured stocks over the past 30 years. Microsatellites have been known as hypervariable neutral molecular markers with the highest resolution power in comparison with any other markers. The largest number of publications concerned the detection of the genetic structure of pristine populations in natural conditions, identification of invasive species, detection of evolutionary management units for conservation genetics, identification of fishery management units and mixed stock analysis, effects of stocking or escapes of farmed fish and their interactions with natural populations, analysis of hatchery stocks including kinship and parentage analysis, assessment of gynogenesis, selection for characteristics of choice and heritability in the wild and identification of quantitative trait loci (QTL). However, the discovery of a new type of molecular marker—single nucleotide polymorphism (SNP) has put the existing applications of microsatellites to the test. Both types of markers have become more available with the development of genomics. In population related studies, several (14–20) microsatellite loci correspond to about 200 independent alleles, which can ensure good accuracy when distinguishing populations. Recently, these numbers tend to be increased by the application of genomic sequencing of expressed sequence tags (ESTs), and using only the most informative loci for genotyping, depending on the aims of research. However, the number of SNP loci can be increased to thousands, which makes it possible to obtain more detailed and precise information on populations. Because SNPs are neutral or outliers (possibly under natural selection), they are able to uncover more effectively both differentiation among populations and the various processes taking place in them, such as introgression, hybridization or the reconstruction of demographic changes. It is noticeable that the number of recent papers related to population level processes associated with microsatellites has decreased, whereas of those using SNPs has increased. Nevertheless, microsatellites can be considered to be superior markers in such topics as kinship and parentage analysis in cultured and natural populations, assessment of gynogenesis, androgenesis and ploidization. For other purposes, microsatellites coupled with SNPs for mapping QTL will remain feasible. Microsatellites will continue to be used in studies of genetic diversity in cultured stocks, and also in the research of natural populations as an economically advantageous research technique.

Funding

This research received no external funding.

Institutional Review Board Statement

Not applicable.

Informed Consent Statement

Not applicable.

Data Availability Statement

Not applicable.

Acknowledgments

This study was partially funded by statutory topic IV.1 in the IO PAS Sopot.

Conflicts of Interest

The author declares no conflict of interest.

References

  1. FAO. The State of World Fisheries and Aquaculture towards Blue Transformation; FAO: Rome, Italy, 2022; pp. 1–266. [Google Scholar] [CrossRef]
  2. Beaumont, A.; Boudry, P.; Hoare, K. Biotechnology and Genetics in Fisheries and Aquaculture; John Wiley & Sons: Hoboken, NJ, USA, 2010; pp. 1–216. [Google Scholar]
  3. Chauhan, T.; Rajiv, K. Molecular markers and their applications in fisheries and aquaculture. Adv. Biosci. Biotechnol. 2010, 1, 281–291. [Google Scholar] [CrossRef]
  4. Cuéllar-Pinzón, J.; Presa, P.; Hawkins, S.J.; Pita, A. Genetic markers in marine fisheries: Types, tasks and trends. Fish. Res. 2016, 173, 194–205. [Google Scholar] [CrossRef]
  5. Dunham, R.A.; Taylor, J.F.; Rise, M.L.; Liu, Z. Development of strategies for integrated breeding, genetics and applied genomics for genetic improvement of aquatic organisms. Aquaculture 2014, 420, S121–S123. [Google Scholar] [CrossRef]
  6. Kucuktas, H.; Liu, Z. Allozyme and Mitochondrial DNA Markers. In Aquaculture Genome Technologies; Blackwell Publishing: Hoboken, NJ, USA, 2007; pp. 73–85. [Google Scholar]
  7. Ward, R.D. Genetics in fisheries management. Hydrobiologia 2000, 420, 191–201. [Google Scholar] [CrossRef]
  8. Alarcon, J.A.; Magoulas, A.; Georgakopoulos, T.; Zouros, E.; Alvarez, M.C. Genetic comparison of wild and cultivated European populations of the gilthead sea bream (Sparus aurata). Aquaculture 2004, 230, 65–80. [Google Scholar] [CrossRef]
  9. Apostolidis, A.; Moutou, K.; Stamatis, C.; Mamuris, Z. Genetic structure of three marine fishes from the Gulf of Pagasitikos (Greece) based on allozymes, RAPD, and mtDNA RFLP markers. Biologia 2009, 64, 1005–1010. [Google Scholar] [CrossRef]
  10. Berrebi, P.; Poteaux, C.; Fissier, M.; Cattaneo-Berrebi, G. Stocking impact and allozyme diversity in brown trout from Mediterranean southern France. J. Fish Biol. 2000, 56, 949–960. [Google Scholar] [CrossRef]
  11. Bylemans, J.; Maes, G.E.; Diopere, E.; Cariani, A.; Senn, H.; Taylor, M.I.; Helyar, S.; Bargelloni, L.; Bonaldo, A.; Carvalho, G.; et al. Evaluating genetic traceability methods for captive-bred marine fish and their applications in fisheries management and wildlife forensics. Aquac. Environ. Interact. 2016, 8, 131–145. [Google Scholar] [CrossRef]
  12. Cordes, J.F.; Israel, J.A.; May, B. Conservation of Paiute cutthroat trout: The genetic legacy of population transplants in an endemic California salmonid. Calif. Fish Game 2004, 90, 101–118. [Google Scholar]
  13. Giantsis, I.A.; Abatzopoulos, T.J.; Angelidis, P.; Apostolidis, A.P. Mitochondrial control region variability in Mytilus galloprovincialis populations from the central-Eastern Mediterranean Sea. Int. J. Mol. Sci. 2014, 15, 11614–11625. [Google Scholar] [CrossRef]
  14. Hauser, L.; Baird, M.; Hilborn, R.; Seeb, L.W.; Seeb, J.E. An empirical comparison of SNPs and microsatellites for parentage and kinship assignment in a wild sockeye salmon (Oncorhynchus nerka) population. Mol. Ecol. Resour. 2011, 11 (Suppl. 1), 150–161. [Google Scholar] [CrossRef] [PubMed]
  15. Layton, K.K.S.; Dempson, B.; Snelgrove, P.V.R.; Duffy, S.J.; Messmer, A.M.; Paterson, I.G.; Jeffery, N.W.; Kess, T.; Horne, J.B.; Salisbury, S.J.; et al. Resolving fine-scale population structure and fishery exploitation using sequenced microsatellites in a northern fish. Evol. Appl. 2020, 13, 1055–1068. [Google Scholar] [CrossRef]
  16. Liu, Z.J. A review of catfish genomics: Progress and perspectives. Comp. Funct. Genom. 2011, 4, 259–265. [Google Scholar] [CrossRef] [PubMed]
  17. Napora-Rutkowski, Ł.; Rakus, K.; Nowak, Z.; Szczygieł, J.; Pilarczyk, A.; Ostaszewska, T.; Irnazarow, I. Genetic diversity of common carp (Cyprinus carpio L.) strains breed in Poland based on microsatellite, AFLP, and mtDNA genotype data. Aquaculture 2017, 473, 433–442. [Google Scholar] [CrossRef]
  18. Okumus, I.; Ciftci, Y. Fish population genetics and molecular markers: II—Molecular markers and their applications in fisheries and aquaculture. Turk. J. Fish. Aquat. Sci. 2003, 3, 51–79. [Google Scholar]
  19. Paaver, T.; Gross, R.; Ilves, P. Growth rate, maturation level and flesh quality of three strains of large rainbow trout (Oncorhynchus mykiss) reared in Estonia. Aquac. Int. 2004, 12, 33–45. [Google Scholar] [CrossRef]
  20. Shaklee, J.; Bentzen, P. Genetic Identification of Stocks of Marine Fish and Shellfish. Bull. Mar. Sci. 1998, 62, 589–621. [Google Scholar]
  21. Wenne, R.; Boudry, P.; Hemmer-Hansen, J.; Lubieniecki, K.P.; Was, A.; Kause, A. What role for genomics in fisheries management and aquaculture? Aquat. Living Resour. 2007, 20, 241–255. [Google Scholar] [CrossRef]
  22. Zhang, L.; Mou, C.; Zhou, J.; Ye, H.; Wei, Z.; Ke, H.; Huang, Z.; Duan, Y.; Zhao, Z.; Zhao, H. Genetic Diversity of Chinese Longsnout Catfish (Leiocassis longirostris) in Four Farmed Populations Based on 20 New Microsatellite DNA Markers. Diversity 2022, 14, 654. [Google Scholar] [CrossRef]
  23. Bembo, D.G.; Carvalho, G.R.; Cingolani, N.; Arneri, E.; Giannetti, G.; Pitcher, T.J. Allozymic and morphometric evidence for two stocks of the European anchovy Engraulis encrasicolus in Adriatic waters. Mar. Biol. 1996, 126, 529–538. [Google Scholar] [CrossRef]
  24. Hovgaard, K.; Skaala, O.; Naevdal, G. Genetic differentiation among sea trout, Salmo trutta L. populations from western Norway. J. Appl. Ichthyol. 2006, 22, 57–61. [Google Scholar] [CrossRef]
  25. Luczynski, M.; Bartel, R.; Vuorinen, J.A.; Domagala, J.; Zolkiewicz, L.; Brzuzan, P. Biochemical genetic characteristics of four Polish sea trout (Salmo trutta trutta L.) populations. Pol. Arch. Hydrobiol. 2000, 47, 21–28. [Google Scholar]
  26. Jordan, W.C.; Cross, T.F.; Crozier, W.W.; Ferguson, A.; Galvin, P.; Hurrell, R.H.; McGinnity, P.; Martin, S.A.M.; Moffett, I.J.J.; Price, D.J.; et al. Allozyme variation in Atlantic salmon from the British Isles: Associations with geography and the environment. J. Fish Biol. 2005, 67, 146–168. [Google Scholar] [CrossRef]
  27. Verspoor, E.; Beardmore, J.A.; Consuegra, S.; De Leaniz, C.G.; Hindar, K.; Jordan, W.C.; Koljonen, M.L.; Mahkrov, A.A.; Paaver, T.; Sánchez, J.A.; et al. Population structure in the Atlantic salmon: Insights from 40 years of research into genetic protein variation. J. Fish Biol. 2005, 67, 3–54. [Google Scholar] [CrossRef]
  28. Larsson, L.C.; Laikre, L.; Andre, C.; Dahlgren, T.G.; Ryman, N. Temporally stable genetic structure of heavily exploited Atlantic herring (Clupea harengus) in Swedish waters. Heredity 2010, 104, 40–51. [Google Scholar] [CrossRef]
  29. Mamuris, Z.; Apostolidis, A.P.; Triantaphyllidis, C. Genetic protein variation in red mullet (Mullus barbatus) and striped red mullet (M-surmuletus) populations from the Mediterranean Sea. Mar. Biol. 1998, 130, 353–360. [Google Scholar] [CrossRef]
  30. Turan, C. Phylogenetic relationships of Mediterranean Mullidae species (Perciformes) inferred from genetic and morphologic data. Sci. Mar. 2006, 70, 311–318. [Google Scholar] [CrossRef]
  31. Turan, C.; Caliskan, M.; Kucuktas, H. Phylogenetic relationships of nine mullet species (Mugilidae) in the Mediterranean Sea. Hydrobiologia 2005, 532, 45–51. [Google Scholar] [CrossRef]
  32. Papasotiropoulos, V.; Klossa-Kilia, E.; Kilias, G.; Alahiotis, S. Genetic divergence and phylogenetic relationships in grey mullets (Teleostei Mugilidae) using allozyme data. Biochem. Genet. 2001, 39, 155–168. [Google Scholar] [CrossRef]
  33. Sujatha, K.; Deepti, V.A.I. Morphometric and allozyme electrophoretic studies of seven species of tunas off north Andhra region. J. Exp. Zool. India 2013, 16, 567–581. [Google Scholar]
  34. Seeb, L.W.; Crane, P.A. High genetic heterogeneity in chum salmon in western Alaska, the contact zone between northern and southern lineages. Trans. Am. Fish. Soc. 1999, 128, 58–87. [Google Scholar] [CrossRef]
  35. Sato, S.; Urawa, S. Genetic structure of chum salmon populations in Japan. Bull. Fish. Res. Agency 2015, 39, 21–47. [Google Scholar]
  36. Vuorinen, J.; Piironen, J. Electrophoretic identification of Atlantic salmon (Salmo salar), brown trout (Salmo trutta), and their hybrids. Can. J. Fish. Aquat. Sci. 2015, 41, 1834–1837. [Google Scholar] [CrossRef]
  37. Winans, G.A.; Urawa, S. Allozyme variability of Oncorhynchus nerka in Japan. Ichthyol. Res. 2000, 47, 343–352. [Google Scholar] [CrossRef]
  38. Ferguson, M.M.; Danzmann, R.G.; Hutchings, J.A. Incongruent estimates of population differentiation among brook charr, Salvelinus fontinalis, from Cape Race, Newfoundland, Canada, based upon allozyme and mitochondrial DNA variation. J. Fish Biol. 1991, 39, 79–85. [Google Scholar] [CrossRef]
  39. Dahle, G.; Jorstad, K.E.; Rusaas, H.E.; Ottera, H. Genetic characteristics of broodstock collected from four Norwegian coastal cod (Gadus morhua) populations. ICES J. Mar. Sci. 2006, 63, 209–215. [Google Scholar] [CrossRef]
  40. Apostolidis, A.; Karakousis, Y.; Triantaphyllidis, C. Genetic divergence and phylogenetic relationships among Salmo trutta L (brown trout) populations from Greece and other European countries. Heredity 1996, 76, 551–560. [Google Scholar] [CrossRef]
  41. Cimmaruta, R.; Bondanelli, P.; Ruggi, A.; Nascetti, G. Genetic structure and temporal stability in the horse mackerel (Trachurus trachurus). Fish. Res. 2008, 89, 114–121. [Google Scholar] [CrossRef]
  42. Lowe, S.A.; Van Doornik, D.M.; Winans, G.A. Geographic variation in genetic and growth patterns of Atka mackerel, Pleurogrammus monopterygius (Hexagrammidae), in the Aleutian archipelago. Fish. Bull. 1998, 96, 502–515. [Google Scholar]
  43. Blanquer, A.; Alayse, J.P.; Berradarkhami, O.; Berrebi, P. Allozyme variation in turbot (Psetta maxima) and brill (Scophthalmus rhombus) (Ostechthyes, Pleuronectiformes, Scophthalmidae) throughout their range in Europe. J. Fish Biol. 1992, 41, 725–736. [Google Scholar] [CrossRef]
  44. Grant, W.S.; Merkouris, S.E.; Kruse, G.H.; Seeb, L.W. Low allozyme heterozygosity in North Pacific and Bering Sea populations of red king crab (Paralithodes camtschaticus): Adaptive specialization, population bottleneck, or metapopulation structure? ICES J. Mar. Sci. 2011, 68, 499–506. [Google Scholar] [CrossRef]
  45. Stamatis, C.; Triantafyllidis, A.; Moutou, K.A.; Mamuris, Z. Allozymic variation in Northeast Atlantic and Mediterranean populations of Norway lobster, Nephrops norvegicus. ICES J. Mar. Sci. 2011, 63, 875–882. [Google Scholar] [CrossRef]
  46. Meehan, B.W.; Carlton, J.T.; Wenne, R. Genetic affinities of the bivalve Macoma balthica from the Pacific coast of North America: Evidence for recent introduction and historical distribution. Mar. Biol. 1989, 102, 235–241. [Google Scholar] [CrossRef]
  47. Wenne, R. Enzyme electrophoretic variation of the coot clam (Mulinia lateralis, Bivalvia) along the Atlantic coast of the U.S.A. Genet. Pol. 1992, 33, 131–139. [Google Scholar]
  48. Koehn, R.K.; Milkman, R.; Mitton, J.B. Population genetics of marine pelecypods. IV. Selection, migration and genetic differentiation in the blue mussel Mytilus edulis. Evolution 1976, 30, 2–32. [Google Scholar] [CrossRef] [PubMed]
  49. McDonald, J.H.; Seed, R.; Koehn, R.K. Allozymes and morphometric characters of three species of Mytilus in the northern and southern hemispheres. Mar. Biol. 1991, 111, 323–333. [Google Scholar] [CrossRef]
  50. Crane, P.A.; Seeb, L.W.; Seeb, J.E. Genetic Relationships among Salvelinus Species Inferred from Allozyme Data. Can. J. Fish. Aquat. Sci. 1994, 51, 182–197. [Google Scholar] [CrossRef]
  51. Hanfling, B.; Brandl, R. Phylogenetics of European cyprinids: Insights from allozymes. J. Fish Biol. 2000, 57, 265–276. [Google Scholar] [CrossRef]
  52. Roldan, M.I.; Pla, C. Species identification of two sympatric hakes by allozymic markers. Sci. Mar. 2001, 65, 81–84. [Google Scholar] [CrossRef]
  53. PerezLosada, M.; Guerra, A.; Sanjuan, A. Allozyme electrophoretic technique and phylogenetic relationships in three species of Sepia (Cephalopoda: Sepiidae). Comp. Biochem. Physiol. B-Biochem. Mol. Biol. 1996, 114, 11–18. [Google Scholar] [CrossRef]
  54. Ayllon, F.; Moran, P.; Garcia-Vazquez, E. Maintenance of a small anadromous subpopulation of brown trout (Salmo trutta L.) by straying. Freshw. Biol. 2006, 51, 351–358. [Google Scholar] [CrossRef]
  55. Seeb, L.W.; Habicht, C.; Templin, W.D.; Tarbox, K.E.; Davis, R.Z.; Brannian, L.K.; Seeb, J.E. Genetic diversity of sockeye salmon of Cook Inlet, Alaska, and its application to management of populations affected by the Exxon Valdez oil spill. Trans. Am. Fish. Soc. 2000, 129, 1223–1249. [Google Scholar] [CrossRef]
  56. Utter, F.M.; Seeb, J.E.; Seeb, L.W. Complementary uses of ecological and biochemical genetic data in identifying and conserving salmon populations. Fish. Res. 1993, 18, 59–76. [Google Scholar] [CrossRef]
  57. Koljonen, M.L.; Pella, J.J. The advantage of using smolt age with allozymes for assessing wild stock contributions to Atlantic salmon catches in the Baltic Sea. ICES J. Mar. Sci. 1997, 54, 1015–1030. [Google Scholar] [CrossRef]
  58. Scribner, K.T.; Crane, P.A.; Spearman, W.J.; Seeb, L.W. DNA and allozyme markers provide concordant estimates of population differentiation: Analyses of US and Canadian populations of Yukon River fall-run chum salmon (Oncorhynchus keta). Can. J. Fish. Aquat. Sci. 1998, 55, 1748–1758. [Google Scholar] [CrossRef]
  59. Shaklee, J.B.; Beacham, T.D.; Seeb, L.; White, B.A. Managing fisheries using genetic data: Case studies from four species of Pacific salmon. Fish. Res. 1999, 43, 45–78. [Google Scholar] [CrossRef]
  60. Bartley, D.M.; Gall, G.A.E. Genetic identification of native cutthroat trout (Oncorhynchus clarki) and intrograssive hybridisation with introduced rainbow trout (O. mykiss) in streams associated with the Alvord Basin, Oregon and Nevada. Copeia 1991, 1991, 854–859. [Google Scholar] [CrossRef]
  61. Poteaux, C.; Beaudou, D.; Berrebi, P. Temporal variations of genetic introgression in stocked brown trout populations. J. Fish Biol. 1991, 53, 701–713. [Google Scholar]
  62. Farrington, L.W.; Austin, C.M.; Coutin, P.C. Allozyme variation and stock structure in the black bream, Acanthopagrus butcheri (Munro) (Sparidae) in southern Australia: Implications for fisheries management, aquaculture and taxonomic relationship with Acanthopagrus australis (Günther). Fish. Manag. Ecol. 2000, 7, 265–279. [Google Scholar] [CrossRef]
  63. Crozier, W.W. Evidence of genetic interaction between escaped farmed salmon and wild Atlantic salmon (Salmo salar L.) in a Northern Irish river. Aquaculture 1993, 113, 19–29. [Google Scholar] [CrossRef]
  64. Ferguson, A.; Taggart, J.B.; Prodohl, P.A.; McMeel, O.; Thompson, C.; Stone, C.; McGinnity, P.; Hynes, R.A. The application of molecular markers to the study and conservation of fish populations, with special reference to Salmo. J. Fish Biol. 1995, 47, 103–126. [Google Scholar] [CrossRef]
  65. Mezherin, S.; Pukhtaevych, P.; Kokodiy, S. Polyclone structure of settlements of unisexual european crucian carp (Carassius gibelio (Bloch, 1782)) in northern ukraine: Allozyme markers’ and chromosome number’s analysis. Tsitologiya Genet. 2020, 54, 71–79. [Google Scholar]
  66. Bouza, C.; Sanchez, L.; Martinez, P. Gene diversity analysis in natural populations and cultured stocks of turbot (Scophthalmus maximus L). Anim. Genet. 1997, 28, 28–36. [Google Scholar] [CrossRef]
  67. Crozier, W.W. Genetic implications of hatchery rearing in Atlantic salmon: Effects of rearing environment on genetic composition. J. Fish Biol. 1998, 52, 1014–1025. [Google Scholar] [CrossRef]
  68. Hallerman, E.M.; Dunham, R.A.; Smitherman, R.O. Selection or Drift–Isozyme Allele Frequency Changes among Channel Catfish Selected for Rapid Growth. Trans. Am. Fish. Soc. 1986, 115, 60–68. [Google Scholar] [CrossRef]
  69. Seeb, J.E.; Thorgaard, G.H.; Utter, F.M. Survival and allozyme expression in diploid and triploid hybrids between chum, chinook, and coho salmon. Aquaculture 1988, 72, 31–48. [Google Scholar] [CrossRef]
  70. Vuorinen, J. Reduction of genetic variability in a hatchery stock of brown trout, Salmo trutta L. J. Fish Biol. 1984, 24, 339–348. [Google Scholar] [CrossRef]
  71. Winkler, F.M.; Bartley, D.; Diaz, N.F. Genetic differences among year classes in a hatchery population of coho salmon (Oncorhynchus kisutch (Walbaum, 1792)) in Chile. Aquaculture 1999, 173, 425–433. [Google Scholar] [CrossRef]
  72. English, L.J.; Maguire, G.B.; Ward, R.D. Genetic variation of wild and hatchery populations of the Pacific oyster, Crassostrea gigas (Thunberg), in Australia. Aquaculture 2000, 187, 283–298. [Google Scholar] [CrossRef]
  73. Ward, R.D.; English, L.J.; McGoldrick, D.J.; Maguire, G.B.; Nell, J.A.; Thompson, P.A. Genetic improvement of the Pacific oyster Crassostrea gigas (Thunberg) in Australia. Aquac. Res. 2000, 31, 35–44. [Google Scholar] [CrossRef]
  74. Allendorf, F.W.; Seeb, L.W. Concordance of genetic divergence among sockeye salmon populations at allozyme, nuclear DNA, and mitochondrial DNA markers. Evolution 2000, 54, 640–651. [Google Scholar] [PubMed]
  75. Maes, G.E.; Volckaert, F.A.M. Clinal genetic variation and isolation by distance in the European eel Anguilla anguilla (L.). Biol. J. Linn. Soc. 2002, 77, 509–521. [Google Scholar] [CrossRef]
  76. Thelen, G.C.; Allendorf, F.W. Heterozygosity-fitness correlations in rainbow trout: Effects of allozyme loci or associative overdominance? Evolution 2001, 55, 1180–1187. [Google Scholar] [PubMed]
  77. Taggart, J.B.; Leaver, M.J.; Bekaert, M. DNA polymorphism underlying allozyme variation at a malic enzyme locus (mMEP-2*) in Atlantic salmon (Salmo salar L.). J. Fish. Biol. 2022, 101, 1371–1374. [Google Scholar] [CrossRef] [PubMed]
  78. Gharrett, A.J.; Mecklenburg, C.W.; Seeb, L.W.; Li, Z.; Matala, A.P.; Gray, A.K.; Heifetz, J. Do genetically distinct rougheye rockfish sibling species differ phenotypically? Trans. Am. Fish. Soc. 2001, 135, 792–800. [Google Scholar] [CrossRef]
  79. Deli, T.; Guizeni, S.; Ben Abdallah, L.; Said, K.; Chatti, N. Chaotic genetic patchiness in the pelagic teleost fish Sardina pilchardus across the Siculo-Tunisian Strait. Mar. Biol. Res. 2020, 16, 280–298. [Google Scholar] [CrossRef]
  80. Khan, M.H.K.; Islam, M.F.; Sarder, M.R.I.; Mollah, M.F.A. Genetic quality deterioration in captive developed broodstocks of Macrobrachium rosenbergii revealed by allozyme electrophoresis. Bangladesh J. Fish. 2019, 31, 17–29. [Google Scholar]
  81. Singh, I.J.; Pandey, N.N.; Pandey, A.; Pandey, A. Featherback, Notopterus notopterus (Pallas) from different reservoirs of Uttarakhand, India using allozyme marker. J. Exp. Zool. India 2020, 23, 131–134. [Google Scholar]
  82. Liu, Z.J.; Cordes, J.F. DNA marker technologies and their applications in aquaculture genetics. Aquaculture 2004, 238, 1–37. [Google Scholar] [CrossRef]
  83. Guerra, D.; Plazzi, F.; Stewart, D.T.; Bogan, A.E.; Hoeh, W.R.; Breton, S. Evolution of sex-dependent mtDNA transmission in freshwater mussels (Bivalvia: Unionida). Sci. Rep. 2017, 7, 1551. [Google Scholar] [CrossRef]
  84. Filipowicz, M.; Burzynski, A.; Smietanka, B.; Wenne, R. Recombination in Mitochondrial DNA of European Mussels Mytilus. J. Mol. Evol. 2008, 67, 377–388. [Google Scholar] [CrossRef] [PubMed]
  85. Smietanka, B.; Wenne, R.; Burzynski, A. Complete male mitochondrial genomes of European Mytilus edulis mussels. Mitochondrial DNA 2016, 27, 1634–1635. [Google Scholar] [PubMed]
  86. Zbawicka MSkibinski, D.O.F.; Wenne, R. Doubly uniparental transmission of mitochondrial DNA length variants in the mussel Mytilus trossulus. Mar. Biol. 2003, 142, 455–460. [Google Scholar] [CrossRef]
  87. Zbawicka, M.; Burzynski, A.; Wenne, R. Complete sequences of mitochondrial genomes from the Baltic mussel Mytilus trossulus. Gene 2007, 406, 191–198. [Google Scholar] [CrossRef]
  88. Zbawicka, M.; Burzynski, A.; Skibinski, D.; Wenne, R. Scottish Mytilus trossulus mussels retain ancestral mitochondrial DNA: Complete sequences of male and female mtDNA genomes. Gene 2010, 456, 45–53. [Google Scholar] [CrossRef]
  89. Ren, J.; Hou, Z.; Wang, H.; Sun, M.-a.; Liu, X.; Liu, B.; Guo, X. Intraspecific Variation in Mitogenomes of Five Crassostrea Species Provides Insight into Oyster Diversification and Speciation. Mar. Biotechnol. 2016, 18, 242–254. [Google Scholar] [CrossRef]
  90. Smietanka, B.; Burzynski, A.; Hummel, H.; Wenne, R. Glacial history of the European marine mussels Mytilus, inferred from distribution of mitochondrial DNA lineages. Heredity 2014, 113, 250–258. [Google Scholar] [CrossRef]
  91. Liang, Y.; He, D.; Jia, Y.; Sun, H.; Chen, Y. Phylogeographic studies of schizothoracine fishes on the central Qinghai-Tibet Plateau reveal the highest known glacial microrefugia. Sci. Rep. 2017, 7, 10983. [Google Scholar] [CrossRef]
  92. Purcell, C.M.; Chabot, C.L.; Craig, M.T.; Martinez-Takeshita, N.; Allen, L.G.; Hyde, J.R. Developing a genetic baseline for the yellowtail amberjack species complex, Seriola lalandi sensu lato, to assess and preserve variation in wild populations of these globally important aquaculture species. Conserv. Genet. 2015, 16, 1475–1488. [Google Scholar] [CrossRef]
  93. Bernatchez, L. The evolutionary history of brown trout (Salmo trutta L.) inferred from phylogeographic, nested clade, and mismatch analyses of mitochondrial DNA variation. Evolution 2001, 55, 351–379. [Google Scholar]
  94. 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]
  95. Diringer, B.; Pretell, K.; Avellan, R.; Chanta, C.; Cedeno, V.; Gentile, G. Genetic structure, phylogeography, and demography of Anadara tuberculosa (Bivalvia) from East Pacific as revealed by mtDNA: Implications to conservation. Ecol. Evol. 2019, 9, 4392–4402. [Google Scholar] [CrossRef] [PubMed]
  96. Yang, T.; Wang, Z.; Liu, Y.; Gao, T. Population genetics and molecular phylogeography of Thamnaconus modestus (Tetraodontiformes, Monachanthidae) in Northwestern Pacific inferred from variation of the mtDNA control region. Aquat. Living Resour. 2019, 32, 18. [Google Scholar] [CrossRef]
  97. Habib, K.A.; Nam, K.; Xiao, Y.; Sathi, J.; Islam, M.N.; Panhwar, S.K.; Habib, A.H.M.S. Population structure, phylogeography and demographic history of Tenualosa ilisha populations in the Indian Ocean region inferred from mitochondrial DNA sequence variation. Reg. Stud. Mar. Sci. 2022, 54, 102478. [Google Scholar] [CrossRef]
  98. Segvic-Bubic, T.; Marrone, F.; Grubisic, L.; Izquierdo-Gomez, D.; Katavic, I.; Arculeo, M.; Lo Brutto, S. Two seas, two lineages: How genetic diversity is structured in Atlantic and Mediterranean greater amberjack Seriola dumerili Risso, 1810 (Perciformes, Carangidae). Fish. Res. 2013, 179, 271–279. [Google Scholar] [CrossRef]
  99. Sato, S.; Kojima, H.; Ando, J.; Ando, H.; Wilmot, R.L.; Seeb, L.W.; Efremov, V.; LeClair, L.; Buchholz, W.; Jin, D.-H.; et al. Genetic population structure of chum salmon in the Pacific Rim inferred from mitochondrial DNA sequence variation. Environ. Biol. Fishes 2004, 69, 37–50. [Google Scholar] [CrossRef]
  100. Yoon, M.; Sato, S.; Seeb, J.E.; Brykov, V.; Seeb, L.W.; Varnavskaya, N.V.; Wilmot, R.L.; Jin, D.H.; Urawa, S.; Urano, A.; et al. Mitochondrial DNA variation and genetic population structure of chum salmon Oncorhynchus keta around the Pacific Rim. J. Fish Biol. 2008, 73, 1256–1266. [Google Scholar] [CrossRef]
  101. Schenekar, T.; Lerceteau-Köhler, E.; Weiss, S. Fine-scale phylogeographic contact zone in Austrian brown trout Salmo trutta reveals multiple waves of post-glacial colonization and a pre-dominance of natural versus anthropogenic admixture. Conserv. Genet. 2014, 15, 561–572. [Google Scholar] [CrossRef]
  102. Wlodarczyk, E.; Wenne, R. Mitochondrial DNA variation in sea trout from coastal rivers in the southern Baltic region. ICES J. Mar. Sci. 2001, 58, 230–237. [Google Scholar] [CrossRef]
  103. Zbawicka, M.; Wenne, R.; Skibinski, D.O.F. Mitochondrial DNA variation in populations of the mussel Mytilus trossulus from the Southern Baltic. Hydrobiologia 2003, 499, 1–12. [Google Scholar] [CrossRef]
  104. Kijewski, T.K.; Zbawicka, M.; Vainola, R.; Wenne, R. Introgression and mitochondrial DNA heteroplasmy in the Baltic populations of mussels Mytilus trossulus and Mytilus edulis. Mar. Biol. 2006, 149, 1371–1385. [Google Scholar] [CrossRef]
  105. Bembo, D.G.; Carvalho, G.R.; Snow, M.; Cingolani, N.; Pitcher, T.J. Stock discrimination among European anchovies, Engraulis encrasicolus, by means of PCR-amplified mitochondrial DNA analysis. Fish. Bull. 1996, 94, 31–40. [Google Scholar]
  106. Avise, J.C.; Helfman, G.S.; Saunders, N.C.; Hales, L.S. Mitochondrial-DNA differentiation in North-Atlantic eels—Population genetic consequences of an unusual life-history pattern. Proc. Natl. Acad. Sci. USA 1986, 83, 4350–4354. [Google Scholar] [CrossRef] [PubMed]
  107. Saunders, N.C.; Kessler, L.G.; Avise, J.C. Genetic-variation and geographic differentiation in mitochondrial DNA of the horseshoe crab, Limulus polyphemus. Genetics 1986, 112, 613–627. [Google Scholar] [CrossRef]
  108. Wilson, G.M.; Thomas, W.K.; Beckenbach, A.T. Mitochondrial DNA Analysis of Pacific Northwest Populations of Oncorhynchus tshawytscha. Can. J. Fish. Aquat. Sci. 1987, 44, 1301–1305. [Google Scholar] [CrossRef]
  109. Kijewska, A.; Burzynski, A.; Wenne, R. Variation in the copy number of tandem repeats of mitochondrial DNA in the North-East Atlantic cod populations. Mar. Biol. Res. 2009, 5, 186–192. [Google Scholar] [CrossRef]
  110. Bert, T.M.; Arnold, W.S.; McMillen-Jackson, A.L.; Wilbur, A.E.; Crawford, C. Natural and Anthropogenic Forces Shape the Population Genetics and Recent Evolutionary History of Eastern United States Bay Scallops (Argopecten irradians). J. Shellfish. Res. 2011, 30, 583–608. [Google Scholar] [CrossRef]
  111. Bert, T.M.; Arnold, W.S.; Wilbur, A.E.; Seyoum, S.; McMillen-Jackson, A.L.; Stephenson, S.P.; Weisberg, R.H.; Yarbro, L.A. Florida Gulf Bay Scallop (Argopecten irradians concentricus) Population Genetic Structure: Form, Variation, and Influential Factors. J. Shellfish. Res. 2014, 33, 99–136. [Google Scholar] [CrossRef]
  112. Laakkonen, H.M.; Strelkov, P.; Lajus, D.L.; Väinölä, R. Introgressive hybridization between the Atlantic and Pacific herrings (Clupea harengus and C. pallasii) in the north of Europe. Mar. Biol. 2014, 162, 39–54. [Google Scholar] [CrossRef]
  113. Zhang, C.; Li, Q.; Wu, X.; Liu, Q.; Cheng, Y. Genetic diversity and genetic structure of farmed and wild Chinese mitten crab (Eriocheir sinensis) populations from three major basins by mitochondrial DNA COI and Cyt b gene sequences. Mitochondrial DNA Part A 2018, 29, 1081–1089. [Google Scholar] [CrossRef]
  114. Bernini, G.; Bellati, A.; Pellegrino, I.; Negri, A.; Ghia, D.; Fea, G.; Sacchi, R.; Nardi, P.A.; Fasola, M.; Galeotti, P. Complexity of biogeographic pattern in the endangered crayfish Austropotamobius italicus in northern Italy: Molecular insights of conservation concern. Conserv. Genet. 2015, 17, 141–154. [Google Scholar] [CrossRef]
  115. Boscari, E.; Congiu, L. The need for genetic support in restocking activities andex situconservation programmes: The case of the Adriatic sturgeon (Acipenser naccarii Bonaparte, 1836) in the Ticino River Park. J. Appl. Ichthyol. 2014, 30, 1416–1422. [Google Scholar] [CrossRef]
  116. Escobar, M.D.; Andrade-Lopez, J.; Farias, I.P.; Hrbek, T. Delimiting Evolutionarily Significant Units of the Fish, Piaractus brachypomus (Characiformes: Serrasalmidae), from the Orinoco and Amazon River Basins with Insight on Routes of Historical Connectivity. J. Hered. 2015, 106 (Suppl. 1), 428–438. [Google Scholar] [CrossRef] [PubMed]
  117. Kikko, T.; Okamoto, H.; Ujiie, M.; Usuki, T.; Nemoto, M.; Saegusa, J.; Ishizaki, D.; Fujioka, Y.; Kai, Y.; Nakayama, K. Genetic evaluation of hatchery stocks of Honmoroko Gnathopogon caerulescens by mitochondrial DNA sequence for stock enhancement. Fish. Sci. 2015, 82, 269–278. [Google Scholar] [CrossRef]
  118. Su, J.; Ji, W.; Zhang, Y.; Gleeson, D.M.; Lou, Z.; Ren, J.; Wei, Y. Genetic diversity and demographic history of the endangered and endemic fish (Platypharodon extremus): Implications for stock enhancement in Qinghai Tibetan Plateau. Environ. Biol. Fishes 2014, 98, 763–774. [Google Scholar] [CrossRef]
  119. Snoj, A.; Bravničar, J.; Zabric, D.; Sušnik Bajec, S. Conservation genetics study of huchen in Slovenia recommends river system-based management and indicates self-sustainability of the middle Sava population. Aquat. Conserv. Mar. Freshw. Ecosyst. 2022, 32, 1171–1183. [Google Scholar] [CrossRef]
  120. Fernández-Pérez, J.; Nantón, A.; Arias-Pérez, A.; Freire, R.; Martínez-Patiño, D.; Méndez, J. Mitochondrial DNA analyses of Donax trunculus (Mollusca: Bivalvia) population structure in the Iberian Peninsula, a bivalve with high commercial importance. Aquat. Conserv. Mar. Freshw. Ecosyst. 2018, 28, 1139–1152. [Google Scholar] [CrossRef]
  121. Kijewska, A.; Burzynski, A.; Wenne, R. Molecular identification of European flounder (Platichthys flesus) and its hybrids with European plaice (Pleuronectes platessa). ICES J. Mar. Sci. 2009, 66, 902–906. [Google Scholar] [CrossRef]
  122. Gu, D.E.; Mu, X.D.; Xu, M.; Luo, D.; Wei, H.; Li, Y.Y.; Zhu, Y.J.; Luo, J.R.; Hu, Y.C. Identification of wild tilapia species in the main rivers of south China using mitochondrial control region sequence and morphology. Biochem. Syst. Ecol. 2016, 65, 100–107. [Google Scholar] [CrossRef]
  123. Panicz, R.; Keszka, S. First occurrence of thinlip grey mullet, Liza ramada (Risso, 1827) in the Odra River estuary (NW Poland): Genetic identification. Oceanologia 2016, 58, 196–200. [Google Scholar] [CrossRef]
  124. Madeira, M.J.; Gomez-Moliner, B.J.; Machordom, A. Genetic introgression on freshwater fish populations caused by restocking programmes. Biol. Invasions 2005, 7, 117–125. [Google Scholar] [CrossRef]
  125. Splendiani, A.; Ruggeri, P.; Giovannotti, M.; Caputo Barucchi, V. Role of environmental factors in the spread of domestic trout in Mediterranean streams. Freshw. Biol. 2013, 58, 2089–2101. [Google Scholar] [CrossRef]
  126. Hamasaki, K.; Toriya, S.; Shishidou, H.; Sugaya, T.; Kitada, S. Genetic effects of hatchery fish on wild populations in red sea bream Pagrus major (Perciformes, Sparidae) inferred from a partial sequence of mitochondrial DNA. J. Fish. Biol. 2010, 77, 2123–2136. [Google Scholar] [CrossRef] [PubMed]
  127. Nakajima, K.; Kitada, S.; Habara, Y.; Sano, S.; Yokoyama, E.; Sugaya, T.; Iwamoto, A.; Kishino, H.; Hamasaki, K. Genetic effects of marine stock enhancement: A case study based on the highly piscivorous Japanese Spanish mackerel. Can. J. Fish. Aquat. Sci. 2014, 71, 301–314. [Google Scholar] [CrossRef]
  128. Lallias, D.; Boudry, P.; Batista, F.M.; Beaumont, A.; King, J.W.; Turner, J.R.; Lapègue, S. Invasion genetics of the Pacific oyster Crassostrea gigas in the British Isles inferred from microsatellite and mitochondrial markers. Biol. Invasions 2015, 17, 2581–2595. [Google Scholar] [CrossRef]
  129. Louati, M.; Kohlmann, K.; Ben Hassine, O.K.; Kersten, P.; Poulet, N.; Bahri-Sfar, L. Genetic characterization of introduced Tunisian and French populations of pike-perch (Sander lucioperca) by species-specific microsatellites and mitochondrial haplotypes. Czech J. Anim. Sci. 2016, 61, 159–171. [Google Scholar] [CrossRef]
  130. Fatsi, P.S.K.; Hashem, S.; Kodama, A.; Appiah, E.K.; Saito, H.; Kawai, K. Population genetics and taxonomic signatures of wild Tilapia in Japan based on mitochondrial DNA control region analysis. Hydrobiologia 2020, 847, 1491–1504. [Google Scholar] [CrossRef]
  131. Laikre, L.; Palm, S.; Ryman, N. Genetic Population Structure of Fishes: Implications for Coastal Zone Management. AMBIOA J. Hum. Environ. 2005, 34, 111–119. [Google Scholar] [CrossRef]
  132. Bineesh, K.K.; Gopalakrishnan, A.; Akhilesh, K.V.; Sajeela, K.A.; Abdussamad, E.M.; Pillai, N.G.K.; Basheer, V.S.; Jena, J.K.; Ward, R.D. DNA barcoding reveals species composition of sharks and rays in the Indian commercial fishery. Mitochondrial DNA Part A 2017, 28, 458–472. [Google Scholar] [CrossRef]
  133. Vella, A.; Vella, N.; Schembri, S. A molecular approach towards taxonomic identification of elasmobranch species from Maltese fisheries landings. Mar. Genom. 2017, 36, 17–23. [Google Scholar] [CrossRef]
  134. Palva, T.K.; Palva, E.T. Restriction site polymorphism in mitochondrial DNA of rainbow-trout, Salmo gairdneri Richardson, stocks in Finland. Aquaculture 1987, 67, 283–289. [Google Scholar] [CrossRef]
  135. Ahmed, F.; Koike, Y.; Strüssmann, C.A.; Yamasaki, I.; Yokota, M.; Watanabe, S. Genetic characterization and gonad development of artificially produced interspecific hybrids of the abalones, Haliotis discus discus Reeve, Haliotis gigantea Gmelin and Haliotis madaka Habe. Aquac. Res. 2008, 39, 532–541. [Google Scholar] [CrossRef]
  136. Hai Nguyen, T.; Liu, Q.; Zhao, L.; Zhang, H.; Liu, J.; Dang Nguyen, H. Genetic diversity of cultured populations of giant freshwater prawn (Macrobrachium rosenbergii) in China using mtDNA COI and 16S rDNA markers. Biochem. Syst. Ecol. 2015, 62, 261–269. [Google Scholar] [CrossRef]
  137. Tóth, B.; Ashrafzadeh, M.R.; Khosravi, R.; Bagi, Z.; Fehér, M.; Bársony, P.; Kovács, G.; Kusza, S. Insights into mitochondrial DNA variation of common carp Cyprinus carpio strains in the Centre of Carpathian Basin. Aquaculture 2022, 554, 738116. [Google Scholar] [CrossRef]
  138. Guo, D.-D.; Liu, F.; Niu, B.-L.; Zhan, W.; Xie, Q.-P.; Zhang, Y.; Lou, B. Establishment of diploid hybrid strains derived from female Larimichthys crocea × male Larimichthys polyactis and transmission of parental mtDNA in hybrid progenies. Aquaculture 2022, 561, 738693. [Google Scholar] [CrossRef]
  139. Hoelzel, A.R. Shark fishing in fin soup. Conserv. Genet. 2001, 2, 69–72. [Google Scholar] [CrossRef]
  140. Bénard-Capelle, J.; Guillonneau, V.; Nouvian, C.; Fournier, N.; Le Loët, K.; Dettai, A. Fish mislabelling in France: Substitution rates and retail types. PeerJ 2015, 2, e714. [Google Scholar] [CrossRef]
  141. Clark, L.F. The current status of DNA barcoding technology for species identification in fish value chains. Food Policy 2015, 54, 85–94. [Google Scholar] [CrossRef]
  142. Griffiths, A.M.; Sotelo, C.G.; Mendes, R.; Pérez-Martín, R.I.; Schröder, U.; Shorten, M.; Silva, H.A.; Verrez-Bagnis, V.; Mariani, S. Current methods for seafood authenticity testing in Europe: Is there a need for harmonisation? Food Control 2014, 45, 95–100. [Google Scholar] [CrossRef]
  143. Kappel, K.; Schröder, U. Species identification of fishery products in Germany. J. Verbrauch. Lebensm. 2014, 10, 31–34. [Google Scholar] [CrossRef]
  144. Kappel, K.; Schröder, U. Substitution of high-priced fish with low-priced species: Adulteration of common sole in German restaurants. Food Control 2016, 59, 478–486. [Google Scholar] [CrossRef]
  145. Lamendin, R.; Miller, K.; Ward, R.D. Labelling accuracy in Tasmanian seafood: An investigation using DNA barcoding. Food Control 2015, 47, 436–443. [Google Scholar] [CrossRef]
  146. Ludwig, A.; Lieckfeldt, D.; Jahrl, J. Mislabeled and counterfeit sturgeon caviar from Bulgaria and Romania. J. Appl. Ichthyol. 2015, 31, 587–591. [Google Scholar] [CrossRef]
  147. Vandamme, S.G.; Griffiths, A.M.; Taylor, S.A.; Di Muri, C.; Hankard, E.A.; Towne, J.A.; Watson, M.; Mariani, S. Sushi barcoding in the UK: Another kettle of fish. PeerJ 2016, 4, e1891. [Google Scholar] [CrossRef]
  148. Poniente, J.A.; Pereda, J.M.R.; Dela Peña, J.T.; Ventolero, M.F.H.; Santos, M.D. Mitochondrial DNA-based species testing of confiscated aquatic wildlife in the Philippines. Forensic Sci. Int. Anim. Environ. 2022, 2, 100051. [Google Scholar] [CrossRef]
  149. Griffiths, A.M.; Miller, D.D.; Egan, A.; Fox, J.; Greenfield, A.; Mariani, S. DNA barcoding unveils skate (Chondrichthyes: Rajidae) species diversity in ‘ray’ products sold across Ireland and the UK. PeerJ 2013, 1, e129. [Google Scholar] [CrossRef]
  150. Elías-Gutiérrez, M.; Hubert, N.; Collins, R.A.; Andrade-Sossa, C. Aquatic Organisms Research with DNA Barcodes. Diversity 2021, 13, 306. [Google Scholar] [CrossRef]
  151. Kijewski, T.; Smietanka, B.; Zbawicka, M.; Gosling, E.; Hummel, H.; Wenne, R. Distribution of Mytilus taxa in European coastal areas as inferred from molecular markers. J. Sea Res. 2011, 65, 224–234. [Google Scholar] [CrossRef]
  152. Wang, S.; Jiao, N.; Zhao, L.; Zhang, M.; Zhou, P.; Huang, X.; Hu, F.; Yang, C.; Shu, Y.; Li, W.; et al. Evidence for the paternal mitochondrial DNA in the crucian carp-like fish lineage with hybrid origin. Sci. China Life Sci. 2020, 63, 102–115. [Google Scholar] [CrossRef]
  153. Zbawicka, M.; Wenne, R.; Burzynski, A. Mitogenomics of recombinant mitochondrial genomes of Baltic Sea Mytilus mussels. Mol. Genet. Genom. 2014, 289, 1275–1287. [Google Scholar] [CrossRef]
  154. Quesada, H.; Wenne, R.; Skibinski, D.O.F. Differential introgression of mitochondrial DNA across species boundaries within the marine mussel genus Mytilus. Proc. R. Soc. B-Biol. Sci. 1995, 262, 51–56. [Google Scholar]
  155. Quesada, H.; Wenne, R.; Skibinski, D.O.F. Interspecies transfer of female mitochondrial DNA is coupled with role-reversals and departure from neutrality in the mussel Mytilus trossulus. Mol. Biol. Evol. 1999, 16, 655–665. [Google Scholar] [CrossRef] [PubMed]
  156. Lo Presti RLisa, C.; Di Stasio, L. Molecular genetics in aquaculture. Ital. J. Anim. Sci. 2009, 8, 299–313. [Google Scholar] [CrossRef]
  157. Maqsood, H.M.; Ahmad, S.M. Advances in Molecular Markers and Their Applications in Aquaculture and Fisheries. Genet. Aquat. Org. 2017, 1, 27–41. [Google Scholar]
  158. Corse, E.; Costedoat, C.; Chappaz, R.; Pech, N.; Martin, J.F.; Gilles, A. A PCR-based method for diet analysis in freshwater organisms using 18S rDNA barcoding on faeces. Mol. Ecol. Resour. 2010, 10, 96–108. [Google Scholar] [CrossRef] [PubMed]
  159. Kijewski, T.; Wijsman, J.W.M.; Hummel, H.; Wenne, R. Genetic composition of cultured and wild mussels Mytilus from The Netherlands and transfers from Ireland and Great Britain. Aquaculture 2009, 287, 292–296. [Google Scholar] [CrossRef]
  160. Vinas, J.; Tudela, S. A Validated Methodology for Genetic Identification of Tuna Species (Genus Thunnus). PLoS ONE 2009, 4, e7606. [Google Scholar] [CrossRef]
  161. Wang, X.C.; Liu, C.; Huang, L.; Bengtsson-Palme, J.; Chen, H.M.; Zhang, J.H.; Cai, D.Y.; Li, J.Q. ITS1, a DNA barcode better than ITS2 in eukaryotes? Mol. Ecol. Resour. 2009, 15, 573–586. [Google Scholar] [CrossRef]
  162. Robledo, D.; Palaiokostas, C.; Bargelloni, L.; Martínez, P.; Houston, R. Applications of genotyping by sequencing in aquaculture breeding and genetics. Rev. Aquac. 2017, 10, 670–682. [Google Scholar] [CrossRef]
  163. Lapegue, S.; Harrang, E.; Heurtebise, S.; Flahauw, E.; Donnadieu, C.; Gayral, P.; Ballenghien, M.; Genestout, L.; Barbotte, L.; Mahla, R.; et al. Development of SNP-genotyping arrays in two shellfish species. Mol. Ecol. Resour. 2014, 14, 820–830. [Google Scholar] [CrossRef]
  164. Villanueva, B.; Fernández, A.; Peiró-Pastor, R.; Peñaloza, C.; Houston, R.D.; Sonesson, A.K.; Tsigenopoulos, C.S.; Bargelloni, L.; Gamsız, K.; Karahan, B.; et al. Population structure and genetic variability in wild and farmed Mediterranean populations of gilthead seabream and European seabass inferred from a 60K combined species SNP array. Aquac. Rep. 2022, 24, 101145. [Google Scholar] [CrossRef]
  165. Pocwierz-Kotus, A.; Bernas, R.; Kent, M.P.; Lien, S.; Leliuna, E.; Debowski, P.; Wenne, R. Restitution and genetic differentiation of salmon populations in the southern Baltic genotyped with the Atlantic salmon 7K SNP array. Genet. Sel. Evol. 2015, 47, 39. [Google Scholar] [CrossRef] [PubMed]
  166. Zeng, Q.; Fu, Q.; Li, Y.; Waldbieser, G.; Bosworth, B.; Liu, S.; Yang, Y.; Bao, L.; Yuan, Z.; Li, N.; et al. Development of a 690 K SNP array in catfish and its application for genetic mapping and validation of the reference genome sequence. Sci. Rep. 2017, 7, 40347. [Google Scholar] [CrossRef] [PubMed]
  167. Wenne, R. Single nucleotide polymorphism markers with applications in aquaculture and assessment of its impact on natural populations. Aquat. Living Resour. 2018, 31, 2. [Google Scholar] [CrossRef]
  168. Wenne, R. Single Nucleotide Polymorphism Markers with Applications in Conservation and Exploitation of Aquatic Natural Populations. Animals 2023, 13, 1089. [Google Scholar] [CrossRef]
  169. Wilson, A.B.; Ashe, J.; Padron, M.; Hamilton, H. Comprehensive genus-wide screening of seahorse microsatellite loci identifies priority species for conservation assessment. Conserv. Genet. Resour. 2021, 13, 221–230. [Google Scholar] [CrossRef]
  170. Chistiakov, D.A.; Hellemans, B.; Volckaert, F.A.M. Microsatellites and their genomic distribution, evolution, function and applications: A review with special reference to fish genetics. Aquaculture 2006, 255, 1–29. [Google Scholar] [CrossRef]
  171. Oreilly, P.; Wright, J.M. The evolving technology of DNA fingerprinting and its application to fisheries and aquaculture. J. Fish Biol. 1995, 47, 29–55. [Google Scholar] [CrossRef]
  172. Wright, J.M.; Bentzen, P. Microsatellites: Genetic markers for the future. Rev. Fish Biol. Fish. 1994, 4, 384–388. [Google Scholar] [CrossRef]
  173. Askari, G.; Shabani, A.; Miandare, H.K. Application of molecular markers in fish. Sci. J. Anim. Sci. 2013, 2, 82–88. [Google Scholar]
  174. Khatei, A.; Tripathy, P.S.; Parhi, J. Molecular Markers in Aquaculture. In Advances in Fisheries Biotechnology; Pandey, P.K., Parhi, J., Eds.; Springer Nature Singapore: Singapore, 2021; p. 165. [Google Scholar]
  175. Qiu, B.; Fang, S.; Ikhwanuddin, M.; Wong, L.; Ma, H. Genome survey and development of polymorphic microsatellite loci for Sillago sihama based on Illumina sequencing technology. Mol. Biol. Rep. 2020, 47, 3011–3017. [Google Scholar] [CrossRef]
  176. Xu, P.; Lu, C.; Sun, Z.; Kuang, Y.; Cao, D.; Huo, T.; Li, C.; Jin, H.; Zheng, X. In Silico Screening and Development of Microsatellite Markers for Genetic Analysis in Perca fluviatilis. Animals 2022, 12, 1809. [Google Scholar] [CrossRef]
  177. Pita, A.; Fernandez-Miguez, M.; Presa, P. EST-Microsatellite Types and Structural Scenarios in European Hake Fisheries. Animals 2022, 12, 1462. [Google Scholar] [CrossRef]
  178. Winans, G.A.; McHenry, M.L.; Baker, J.; Elz, A.; Goodbla, A.; Iwamoto, E.; Kuligowski, D.; Miller, K.M.; Small, M.P.; Spruell, P.; et al. Genetic inventory of anadromous pacific salmonids of the Elwha River prior to dam removal. Northwest Sci. 2008, 82, 128–141. [Google Scholar] [CrossRef]
  179. Angers, B.; Bernatchez, L.; Angers, A.; Desgroseillers, L. Specific microsatellite loci for brook charr reveal strong population subdivision on a microgeographic scale. J. Fish Biol. 2015, 47, 177–185. [Google Scholar] [CrossRef]
  180. Estoup, A.; Presa, P.; Krieg, F.; Vaiman, D.; Guyomard, R. (CT)n and (GT)n microsatellites: A new class of genetic markers for Salmo trutta L. (brown trout). Heredity 1993, 71, 488–496. [Google Scholar] [CrossRef] [PubMed]
  181. Olafsson, K.; Pampoulie, C.; Hjorleifsdottir, S.; Gudjonsson, S.; Hreggvidsson, G.O. Present-Day Genetic Structure of Atlantic Salmon (Salmo salar) in Icelandic Rivers and Ice-Cap Retreat Models. PLoS ONE 2014, 9, e86809. [Google Scholar] [CrossRef] [PubMed]
  182. McConnell, S.K.; Oreilly, P.; Hamilton, L.; Wright, J.N.; Bentzen, P. Polymorphic microsatellite loci from Atlantic salmon (Salmo salar): Genetic differentiation of North American and European populations. Can. J. Fish. Aquat. Sci. 1995, 52, 1863–1872. [Google Scholar] [CrossRef]
  183. McConnell, S.; Hamilton, L.; Morris, D.; Cook, D.; Paquet, D.; Bentzen, P.; Wright, J. Isolation of salmonid microsatellite loci and their application to the population genetics of Canadian east coast stocks of Atlantic salmon. Aquaculture 1995, 137, 19–30. [Google Scholar] [CrossRef]
  184. 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, 905–918141. [Google Scholar] [CrossRef]
  185. Lundrigan, T.A.; Reist, J.D.; Ferguson, M.M. Microsatellite genetic variation within and among Arctic charr (Salvelinus alpinus) from aquaculture and natural populations in North America. Aquaculture 2005, 244, 63–75. [Google Scholar] [CrossRef]
  186. Archangi, B.; Chand, V.; Mather, P.B. Isolation and characterization of 15 polymorphic microsatellite DNA loci from Argyrosomus japonicus (mulloway), a new aquaculture species in Australia. Mol. Ecol. Resour. 2009, 9, 412–414. [Google Scholar] [CrossRef] [PubMed]
  187. Ferreira, D.G.; Souza-Shibatta, L.; Shibatta, O.A.; Sofia, S.H.; Carlsson, J.; Dias, J.H.P.; Makrakis, S.; Makrakis, M.C. Genetic structure and diversity of migratory freshwater fish in a fragmented Neotropical river system. Rev. Fish Biol. Fish. 2017, 27, 209–231. [Google Scholar] [CrossRef]
  188. Coimbra, M.R.M.; Dantas, H.L.; Luna, M.M.S.; Lima, M.A.; Sales, M.; da Silva, B.C.N.R.; Lima, A.P.S. High gene flow in two migratory Neotropical fish species, Salminus franciscanus and Brycon orthotaenia, and implications for conservation aquaculture Aquat. Conserv. Mar. Freshw. Ecosyst. 2020, 30, 1063–1073. [Google Scholar] [CrossRef]
  189. Gold, J.R.; Pak, E.; Richardson, L.R. Microsatellite variation among red snapper (Lutjanus campechanus) from the Gulf of Mexico. Mar. Biotechnol. 2001, 3, 293–304. [Google Scholar] [CrossRef]
  190. Castillo, A.G.F.; Martinez, J.L.; Garcia-Vazquez, E. Fine spatial structure of Atlantic hake (Merluccius merluccius) stocks revealed by variation at microsatellite loci. Mar. Biotechnol. 2004, 6, 299–306. [Google Scholar] [CrossRef] [PubMed]
  191. Hutchinson, W.F.; Carvalho, G.R.; Rogers, S.I. Marked genetic structuring in localised spawning populations of cod Gadus morhua in the North Sea and adjoining waters, as revealed by microsatellites. Mar. Ecol. Prog. Ser. 2001, 223, 251–260. [Google Scholar] [CrossRef]
  192. Skarstein, T.H.; Westgaard, J.I.; Fevolden, S.E. Comparing microsatellite variation in north-east Atlantic cod (Gadus morhua L.) to genetic structuring as revealed by the pantophysin (Pan I) locus. J. Fish Biol. 2007, 70, 271–290. [Google Scholar] [CrossRef]
  193. Bentzen, P.; Taggart, C.T.; Ruzzante, D.E.; Cook, D. Microsatellite polymorphism and the population structure of Atlantic cod (Gadus morhua) in the northwest Atlantic. Can. J. Fish. Aquat. Sci. 1996, 53, 2706–2721. [Google Scholar] [CrossRef]
  194. O’Leary, D.B.; Coughlan, J.; Dillane, E.; McCarthy, T.V.; Cross, T.F. Microsatellite variation in cod Gadus morhua throughout its geographic range. J. Fish Biol. 2007, 70, 310–335. [Google Scholar] [CrossRef]
  195. O’Connell, M.; Dillon, M.C.; Wright, J.M.; Bentzen, P.; Merkouris, S.; Seeb, J. Genetic structuring among Alaskan Pacific herring populations identified using microsatellite variation. J. Fish Biol. 1998, 53, 150–163. [Google Scholar] [CrossRef]
  196. Gíslason, D.; Helyar, S.J.; Óskarsson, G.J.; Ólafsdóttir, G.; Slotte, A.; Jansen, T.; Jacobsen, J.A.; Ólafsson, K.; Skirnisdottir, S.; Dahle, G.; et al. The genetic composition of feeding aggregations of the Atlantic mackerel (Scomber scombrus) in the central north Atlantic: A microsatellite loci approach. ICES J. Mar. Sci. 2020, 77, 604–612. [Google Scholar] [CrossRef]
  197. Aguila, R.D.; Perez, S.K.; Catacutan, B.J.; Lopez, G.V.; Barut, N.C.; Santos, M.D. Distinct Yellowfin Tuna (Thunnus albacares) Stocks Detected in Western and Central Pacific Ocean (WCPO) Using DNA Microsatellites. PLoS ONE 2015, 10, e0138292. [Google Scholar]
  198. Phinchongsakuldit, J.; Chaipakdee, P.; Collins, J.F.; Jaroensutasinee, M.; Brookfield, J.F.Y. Population genetics of cobia (Rachycentron canadum) in the Gulf of Thailand and Andaman Sea: Fisheries management implications. Aquac. Int. 2013, 21, 197–217. [Google Scholar] [CrossRef]
  199. White, S.L.; Kazyak, D.C.; Darden, T.L.; Farrae, D.J.; Lubinski, B.A.; Johnson, R.L.; Eackles, M.S.; Balazik, M.T.; Brundage, H.M.; Fox, A.G.; et al. Establishment of a microsatellite genetic baseline for North American Atlantic sturgeon (Acipenser o. oxyrhinchus) and range-wide analysis of population genetics. Conserv. Genet. 2021, 22, 977–992. [Google Scholar] [CrossRef]
  200. Moghim, M.; Javanmard, A.; Pourkazemi, M.; Tan, S.G.; Panandam, J.M.; Kor, D.; Laloei, F. Application of microsatellite markers for genetic conservation and management of Persian sturgeon (Acipenser persicus, Borodin, 1897) in the Caspian Sea. J. Appl. Ichthyol. 2013, 29, 696–703. [Google Scholar] [CrossRef]
  201. Zhou, Y.; Tong, J.; Wang, J.; Yu, X. Development of microsatellite markers and genetic diversity in wild and cultured populations of black carp (Mylopharyngodon piceus) along the Yangtze River. Aquac. Int. 2020, 28, 1867–1882. [Google Scholar] [CrossRef]
  202. Nyingi, D.; De Vos, L.; Aman, R.; Agnese, J.-F. Genetic characterization of an unknown and endangered native population of the Nile tilapia Oreochromis niloticus (Linnaeus, 1758) (Cichlidae; Teleostei) in the Loboi Swamp (Kenya). Aquaculture 2009, 297, 57–63. [Google Scholar] [CrossRef]
  203. Thiele, S.; Adams, M.; Hammer, M.; Wedderburn, S.; Whiterod, N.S.; Unmack, P.J.; Sasaki, M.; Beheregaray, L.B. Range-wide population genetics study informs on conservation translocations and reintroductions for the endangered Murray hardyhead (Craterocephalus fluviatilis). Aquat. Conserv. Mar. Freshw. Ecosyst. 2020, 30, 1959–1974. [Google Scholar] [CrossRef]
  204. García-Castro, K.L.; Rangel-Medrano, J.D.; Landínez-García, R.M.; Márquez, E.J. Population genetics of the endangered catfish Pseudoplatystoma magdaleniatum (Siluriformes: Pimelodidae) based on species-specific microsatellite loci. Neotrop. Ichthyol. 2021, 19, e200120. [Google Scholar] [CrossRef]
  205. Prado, F.D.; Fernandez-Cebrián, R.; Hashimoto, D.T.; Senhorini, J.A.; Foresti, F.; Martínez, P.; Porto-Foresti, F. Hybridization and genetic introgression patterns between two South American catfish along their sympatric distribution range. Hydrobiologia 2017, 788, 319–343. [Google Scholar] [CrossRef]
  206. Guimarães-Marques, G.M.; de Jesus Bentes, A.; Formiga, K.M.; da Silva Batista, J. Microsatellite DNA markers for the commercially important fish Hypophthalmus donascimientoi (Pisces: Siluriformes: Pimelodidae) in the Amazon basin: Isolation, characterization and amplification in congeneric species. Mol. Biol. Rep. 2023. [Google Scholar] [CrossRef] [PubMed]
  207. Vargas-Caro, C.; Bustamante, C.; Bennett, M.B.; Ovenden, J.R. Towards sustainable fishery management for skates in South America: The genetic population structure of Zearaja chilensis and Dipturus trachyderma (Chondrichthyes, Rajiformes) in the south-east Pacific Ocean. PLoS ONE 2017, 12, e0172255. [Google Scholar] [CrossRef]
  208. Buresch, K.C.; Gerlach, G.; Hanlon, R.T. Multiple genetic stocks of longfin squid Loligo pealeii in the NW Atlantic: Stocks segregate inshore in summer, but aggregate offshore in winter. Mar. Ecol. Prog. Ser. 2006, 310, 263–270. [Google Scholar] [CrossRef]
  209. Henriques, F.F.; Serrão, E.A.; González-Wangüemert, M. Novel polymorphic microsatellite loci for a new target species, the sea cucumber Holothuria mammata. Biochem. Syst. Ecol. 2016, 66, 109–113. [Google Scholar] [CrossRef]
  210. Selvamani, M.J.P.; Degnan, S.M.; Paetkau, D.; Degnan, B.M. Highly polymorphic microsatellite loci in the Heron Reef population of the tropical abalone Haliotis asinina. Mol. Ecol. 2000, 9, 1184–1186. [Google Scholar] [CrossRef]
  211. Beaumont, A.; Garcia, M.T.; Honig, S.; Low, P. Genetics of Scottish populations of the native oyster, Ostrea edulis: Gene flow, human intervention and conservation. Aquat. Living Resour. 2006, 19, 389–402. [Google Scholar] [CrossRef]
  212. Larrain, A.M.; Diaz, N.F.; Lamas, C.; Uribe, C.; Jilberto, F.; Araneda, C. Heterologous microsatellite-based genetic diversity in blue mussel (Mytilus chilensis) and differentiation among localities in southern Chile. Lat. Am. J. Aquat. Res. 2015, 43, 998–1010. [Google Scholar] [CrossRef]
  213. Shaw, P.W.; Hendrickson, L.; McKeown, N.J.; Stonier, T.; Naud, M.J.; Sauer, W.H.H. Discrete spawning aggregations of loliginid squid do not represent genetically distinct populations. Mar. Ecol. Prog. Ser. 2020, 408, 117–127. [Google Scholar] [CrossRef]
  214. Speller, C.F.; Hauser, L.; Lepofsky, D.; Moore, J.; Rodrigues, A.T.; Moss, M.L.; McKechnie, I.; Yang, D.Y. High Potential for Using DNA from Ancient Herring Bones to Inform Modern Fisheries Management and Conservation. PLoS ONE 2012, 7, e51122. [Google Scholar] [CrossRef]
  215. Sort, M.; Manuzzi, A.; Jiménez-Mena, B.; Ovenden, J.R.; Holmes, B.J.; Bernard, A.M.; Shivji, M.S.; Meldrup, D.; Bennett, M.B.; Nielsen, E.E. Come together: Calibration of tiger shark (Galeocerdo cuvier) microsatellite databases for investigating global population structure and assignment of historical specimens. Conserv. Genet. Resour. 2021, 13, 209–220. [Google Scholar] [CrossRef]
  216. Smith, C.T.; Seeb, L.W. Number of alleles as a predictor of the relative assignment accuracy of short tandem repeat (STR) and single-nucleotide-polymorphism (SNP) baselines for chum salmon. Trans. Am. Fish. Soc. 2008, 137, 751–762. [Google Scholar] [CrossRef]
  217. Beacham, T.D.; Jonsen, K.; Wallace, C. A Comparison of Stock and Individual Identification for Chinook Salmon in British Columbia Provided by Microsatellites and Single-Nucleotide Polymorphisms. Mar. Coast. Fish. 2012, 4, 1–22. [Google Scholar] [CrossRef]
  218. Beacham, T.D.; Wetklo, M.; Wallace, C.; Olsen, J.B.; Flannery, B.G.; Wenburg, J.K.; Templin, W.D.; Antonovich, A.; Seeb, L.W. The application of Microsatellites for stock identification of Yukon River Chinook salmon. N. Am. J. Fish. Manag. 2012, 28, 283–295. [Google Scholar] [CrossRef]
  219. O’Malley, K.G.; Jacobson, D.P.; Kurth, R.; Dill, A.J.; Banks, M.A. Adaptive genetic markers discriminate migratory runs of Chinook salmon (Oncorhynchus tshawytscha) amid continued gene flow. Evol. Appl. 2013, 6, 1184–1194. [Google Scholar] [CrossRef]
  220. Bradbury, I.R.; Hamilton, L.C.; Dempson, B.; Robertson, M.J.; Bourret, V.; Bernatchez, L.; Verspoor, E. Transatlantic secondary contact in Atlantic Salmon, comparing microsatellites, a single nucleotide polymorphism array and restriction-site associated DNA sequencing for the resolution of complex spatial structure. Mol. Ecol. 2015, 24, 5130–5144. [Google Scholar] [CrossRef] [PubMed]
  221. Gallagher, J.; Lordan, C.; Hughes, G.M.; Jonasson, J.P.; Carlsson, J. Microsatellites obtained using high throughput sequencing and a novel microsatellite genotyping method reveals population genetic structure in Norway Lobster, Nephrops norvegicus. J. Sea Res. 2022, 179, 102139. [Google Scholar] [CrossRef]
  222. Ayllon, F.; Davaine, P.; Beall, E.; Garcia-Vazquez, E. Dispersal and rapid evolution in brown trout colonizing virgin Subantarctic ecosystems. J. Evol. Biol. 2006, 19, 1352–1358. [Google Scholar] [CrossRef]
  223. Bean, P.T.; Lutz-Carrillo, D.J.; Bonner, T.H. Rangewide Survey of the Introgressive Status of Guadalupe Bass: Implications for Conservation and Management. Trans. Am. Fish. Soc. 2013, 142, 681–689. [Google Scholar] [CrossRef]
  224. Hunter, M.E.; Nico, L.G. Genetic analysis of invasive Asian Black Carp (Mylopharyngodon piceus) in the Mississippi River Basin: Evidence for multiple introductions. Biol. Invasions 2015, 17, 99–114. [Google Scholar] [CrossRef]
  225. Molnár, T.; Lehoczky, I.; Edviné Meleg, E.; Boros, G.; Specziár, A.; Mozsár, A.; Vitál, Z.; Józsa, V.; Allele, W.; Urbányi, B. Comparison of the Genetic Structure of Invasive Bigheaded Carp (Hypophthalmichthys spp.) Populations in Central-European Lacustrine and Riverine Habitats. Animals 2021, 11, 2018. [Google Scholar] [CrossRef] [PubMed]
  226. Froufe, E.; Varandas, S.; Teixeira, A.; Sousa, R.; Filipova, L.; Petrusek, A.; Edsman, L.; Lopes-Lima, M. First results on the genetic diversity of the invasive signal crayfish Pacifastacus leniusculus (Dana, 1852) in Europe using novel microsatellite loci. J. Appl. Genet. 2015, 56, 375–380. [Google Scholar] [CrossRef] [PubMed]
  227. Monzon-Arguello, C.; Consuegra, S.; Gajardo, G.; Marco-Rius, F.; Fowler, D.M.; DeFaveri, J.; Garcia de Leaniz, C. Contrasting patterns of genetic and phenotypic differentiation in two invasive salmonids in the southern hemisphere. Evol. Appl. 2014, 7, 921–936. [Google Scholar] [CrossRef] [PubMed]
  228. Burger, C.V.; Scribner, K.T.; Spearman, W.J.; Swanton, C.O.; Campton, D.E. Genetic contribution of three introduced life history forms of sockeye salmon to colonization of Frazer Lake, Alaska. Can. J. Fish. Aquat. Sci. 2000, 57, 2096–2111. [Google Scholar] [CrossRef]
  229. Roberts, D.G.; Gray, C.A.; Ayre, D.J. Microsatellite primers for Australian recreationally and commercially important estuarine fishes. J. Fish Biol. 2014, 84, 273–281. [Google Scholar] [CrossRef]
  230. Munoz, J.; Gomez, A.; Figuerola, J.; Amat, F.; Rico, C.; Green, A.J. Colonization and dispersal patterns of the invasive American brine shrimp Artemia franciscana (Branchiopoda: Anostraca) in the Mediterranean region. Hydrobiologia 2014, 726, 25–41. [Google Scholar] [CrossRef]
  231. Chiesa, S.; Marzano, F.N.; Minervini, G.; De Lucrezia, D.; Baccarani, G.; Bordignon, G.; Poli, I.; Ravagnan, G.; Argese, E. The invasive Manila clam Ruditapes philippinarum (Adams and Reeve, 1850) in Northern Adriatic Sea: Population genetics assessed by an integrated molecular approach. Fish. Res. 2011, 110, 259–267. [Google Scholar] [CrossRef]
  232. Arias-Perez, A.; Cordero, D.; Borrell, Y.; Sanchez, J.A.; Blanco, G.; Freire, R.; Insua, A.; Saavedra, C. Assessing the geographic scale of genetic population management with microsatellites and introns in the clam Ruditapes decussatus. Ecol. Evol. 2016, 6, 3380–3404. [Google Scholar] [CrossRef]
  233. Holland, B.S. Invasion without a bottleneck: Microsatellite variation in natural and invasive populations of the brown mussel Perna perna (L). Mar. Biotechnol. 2001, 3, 407–415. [Google Scholar] [CrossRef]
  234. Kochmann, J.; Carlsson, J.; Crowe, T.P.; Mariani, S. Genetic Evidence for the Uncoupling of Local Aquaculture Activities and a Population of an Invasive Species-A Case Study of Pacific Oysters (Crassostrea gigas). J. Hered. 2012, 103, 661–671. [Google Scholar] [CrossRef]
  235. Meistertzheim, A.-L.; Arnaud-Haond, S.; Boudry, P.; Thebault, M.-T. Genetic structure of wild European populations of the invasive Pacific oyster Crassostrea gigas due to aquaculture practices. Mar. Biol. 2013, 160, 453–463. [Google Scholar] [CrossRef]
  236. Waples, R.S.; Punt, A.E.; Cope, J.M. Integrating genetic data into management of marine resources: How can we do it better? Fish Fish. 2008, 9, 423–449. [Google Scholar] [CrossRef]
  237. Stoeckle, B.C.; Mueller, M.; Nagel, C.; Kuehn, R.; Geist, J. A conservation genetics perspective on supportive breeding: A case study of the common nase (Chondrostoma nasus). Aquat. Conserv. Mar. Freshw. Ecosyst. 2022, 32, 1596–1605. [Google Scholar] [CrossRef]
  238. Wellband, K.W.; Atagi, D.Y.; Koehler, R.A.; Heath, D.D. Fine-Scale Population Genetic Structure and Dispersal of Juvenile Steelhead in the Bulkley-Morice River, British Columbia. Trans. Am. Fish. Soc. 2012, 141, 392–401. [Google Scholar] [CrossRef]
  239. Deiner, K.; Garza, J.C.; Coey, R.; Girman, D.J. Population structure and genetic diversity of trout (Oncorhynchus mykiss) above and below natural and man-made barriers in the Russian River, California. Conserv. Genet. 2007, 8, 437–454. [Google Scholar] [CrossRef]
  240. Taylor, E.B.; Stamford, M.D.; Baxter, J.S. Population subdivision in westslope cutthroat trout (Oncorhynchus clarki lewisi) at the northern periphery of its range: Evolutionary inferences and conservation implications. Mol. Ecol. 2003, 12, 2609–2622. [Google Scholar] [CrossRef]
  241. Eldridge, W.H.; Killebrew, K. Genetic diversity over multiple generations of supplementation: An example from Chinook salmon using microsatellite and demographic data. Conserv. Genet. 2008, 9, 13–28. [Google Scholar] [CrossRef]
  242. Ramstad, K.M.; Foote, C.J.; Olsen, J.B. Genetic and phenotypic evidence of reproductive isolation between seasonal runs of Sockeye salmon in Bear Lake, Alaska. Trans. Am. Fish. Soc. 2003, 132, 997–1013. [Google Scholar] [CrossRef]
  243. Kitano, S.; Ohdachi, S.; Koizumi, I.; Hasegawa, K. Hybridization between native white-spotted charr and nonnative brook trout in the upper Sorachi River, Hokkaido, Japan. Ichthyol. Res. 2014, 61, 1–8. [Google Scholar] [CrossRef]
  244. Dawnay, N.; Dawnay, L.; Hughes, R.N.; Cove, R.; Taylor, M.I. Substantial genetic structure among stocked and native populations of the European grayling (Thymallus thymallus, Salmonidae) in the United Kingdom. Conserv. Genet. 2011, 12, 731–744. [Google Scholar] [CrossRef]
  245. Schreier, A.D.; Mahardja, B.; May, B. Hierarchical patterns of population structure in the endangered Fraser River white sturgeon (Acipenser transmontanus) and implications for conservation. Can. J. Fish. Aquat. Sci. 2012, 69, 1968–1980. [Google Scholar] [CrossRef]
  246. Muhlfeld, C.C.; D’Angelo, V.; Kalinowski, S.T.; Landguth, E.L.; Downs, C.C.; Tohtz, J.; Kershner, J.L. A Fine-scale Assessment of Using Barriers to Conserve Native Stream Salmonids: A Case Study in Akokala Creek, Glacier National Park, USA. Open Fish Sci. J. 2012, 5, 9–20. [Google Scholar] [CrossRef]
  247. Geist, J.; Kuehn, R. Genetic diversity and differentiation of central European freshwater pearl mussel (Margaritifera margaritifera L.) populations: Implications for conservation and management. Mol. Ecol. 2005, 14, 425–439. [Google Scholar] [CrossRef] [PubMed]
  248. Nunes Galvao, M.S.; Silva Hilsdorf, A.W. Assessing the genetic diversity of the mangrove oyster Crassostrea rhizophorae (Bivalvia, Ostreidae) by microsatellite markers in southeastern Brazil. Mar. Biol. Res. 2015, 11, 944–954. [Google Scholar] [CrossRef]
  249. Waldbieser, G.C.; Wolters, W.R. Application of polymorphic microsatellite loci in a channel catfish Ictalurus punctatus breeding program. J. World Aquac. Soc. 1999, 30, 256–262. [Google Scholar] [CrossRef]
  250. Moyer, G.R.; Sweka, J.A.; Peterson, D.L. Past and Present Processes Influencing Genetic Diversity and Effective Population Size in a Natural Population of Atlantic Sturgeon. Trans. Am. Fish. Soc. 2012, 141, 56–67. [Google Scholar] [CrossRef]
  251. Reiss, H.; Hoarau, G.; Dickey-Collas, M.; Wolff, W.J. Genetic population structure of marine fish: Mismatch between biological and fisheries management units. Fish Fish. 2009, 10, 361–395. [Google Scholar] [CrossRef]
  252. An, H.; Kim, M.-J.; Park, K.; Cho, K.; Bae, B.; Kim, J.; Myeong, J.-I. Genetic Diversity and Population Structure in the Heavily Exploited Korean Rockfish, Sebastes schlegeli, in Korea. J. World Aquac. Soc. 2012, 43, 73–83. [Google Scholar] [CrossRef]
  253. Tringali, M.D.; Seyoum, S.; Wallace, E.M.; Higham, M.; Taylor, R.G.; Trotter, A.A.; Whittington, J.A. Limits to the use of contemporary genetic analyses in delineating biological populations for restocking and stock enhancement. Rev. Fish. Sci. 2008, 16, 111–116. [Google Scholar] [CrossRef]
  254. Miller, K.M.; Supernault, K.J.; Li, S.R.; Withler, R.E. Population structure in two marine invertebrate species (Panopea abrupta and Strongylocentrotus franciscanus) targeted for aquaculture and enhancement in British Columbia. J. Shellfish Res. 2006, 25, 33–42. [Google Scholar] [CrossRef]
  255. Hansen, M.M.; Nielsen, E.E.; Mensberg, K.L.D. The problem of sampling families rather than populations: Relatedness among individuals in samples of juvenile brown trout Salmo trutta L. Mol. Ecol. 1997, 6, 469–474. [Google Scholar] [CrossRef]
  256. Ruzzante, D.E.; Hansen, M.M.; Meldrup, D. Distribution of individual inbreeding coefficients, relatedness and influence of stocking on native anadromous brown trout (Salmo trutta) population structure. Mol. Ecol. 2001, 10, 2107–2128. [Google Scholar] [CrossRef] [PubMed]
  257. Largiader, C.R.; Estoup, A.; Lecerf, F.; Champigneulle, A.; Guyomard, R. Microsatellite analysis of polyandry and spawning site competition in brown trout (Salmo trutta L.). Genet. Sel. Evol. 2001, 33, S205–S222. [Google Scholar] [CrossRef]
  258. Hutchinson, W.F.; van Oosterhout, C.; Rogers, S.I.; Carvalho, G.R. Temporal analysis of archived samples indicates marked genetic changes in declining North Sea cod (Gadus morhua). Proc. R. Soc. B-Biol. Sci. 2003, 270, 2125–2132. [Google Scholar] [CrossRef]
  259. Rowe, S.; Hutchings, J.A.; Skjaeraasen, J.E. Nonrandom mating in a broadcast spawner: Mate size influences reproductive success in Atlantic cod (Gadus morhua). Can. J. Fish. Aquat. Sci. 2007, 64, 219–226. [Google Scholar] [CrossRef]
  260. Mirimin, L.; Macey, B.; Kerwath, S.; Lamberth, S.; Bester-van der Merwe, A.; Cowley, P.; Bloomer, P.; Roodt-Wilding, R. Genetic analyses reveal declining trends and low effective population size in an overfished South African sciaenid species, the dusky kob (Argyrosomus japonicus). Mar. Freshw. Res. 2016, 67, 266–276. [Google Scholar] [CrossRef]
  261. Beacham, T.D.; Lapointe, M.; Candy, J.R.; Miller, K.M.; Withler, R.E. DNA in action: Rapid application of DNA variation to sockeye salmon fisheries management. Conserv. Genet. 2004, 5, 411–416. [Google Scholar] [CrossRef]
  262. VanDeHey, J.A.; Sloss, B.L.; Peeters, P.J.; Sutton, T.M. Determining the efficacy of microsatellite DNA-based mixed stock analysis of Lake Michigan’s Lake Whitefish commercial Fishery. J. Great Lakes Res. 2010, 36, 52–58. [Google Scholar] [CrossRef]
  263. Waldman, J.; Hasselman, D.; Bentzen, P.; Dadswell, M.; Maceda, L.; Wirgin, I. Genetic Mixed-Stock Analysis of American Shad in Two Atlantic Coast Fisheries: Delaware Bay, USA, and Inner Bay of Fundy, Canada. N. Am. J. Fish. Manag. 2014, 34, 1190–1198. [Google Scholar] [CrossRef]
  264. Koljonen, M.L.; Gross, R.; Koskiniemi, J. Wild Estonian and Russian sea trout (Salmo trutta) in Finnish coastal sea trout catches: Results of genetic mixed-stock analysis. Hereditas 2014, 151, 177–195. [Google Scholar] [CrossRef]
  265. Vasemagi, A.; Gross, R.; Paaver, T.; Koljonen, M.L.; Nilsson, J. Extensive immigration from compensatory hatchery releases into wild Atlantic salmon population in the Baltic sea: Spatio-temporal analysis over 18 years. Heredity 2005, 95, 76–83. [Google Scholar] [CrossRef]
  266. Mezzera, M.; Largiader, C.R. Evidence for selective angling of introduced trout and their hybrids in a stocked brown trout population. J. Fish Biol. 2001, 59, 287–301. [Google Scholar] [CrossRef]
  267. Choi, H.-k.; Jang, J.E.; Byeon, S.Y.; Kim, Y.R.; Maschette, D.; Chung, S.; Choi, S.-G.; Kim, H.-W.; Lee, H.J. Genetic Diversity and Population Structure of the Antarctic Toothfish, Dissostichus mawsoni, Using Mitochondrial and Microsatellite DNA Markers. Front. Mar. Sci. 2021, 8, 666417. [Google Scholar]
  268. Han, Y.-S.; Sun, Y.-L.; Liao, Y.-F.; Liao, I.C.; Shen, K.-N.; Tzeng, W.-N. Temporal analysis of population genetic composition in the overexploited Japanese eel Anguilla japonica. Mar. Biol. 2008, 155, 613–621. [Google Scholar] [CrossRef]
  269. Larsen, P.F.; Hansen, M.M.; Nielsen, E.E.; Jensen, L.F.; Loeschcke, V. Stocking impact and temporal stability of genetic composition in a brackish northern pike population (Esox lucius L.), assessed using microsatellite DNA analysis of historical and contemporary samples. Heredity 2005, 95, 136–143. [Google Scholar] [CrossRef] [PubMed]
  270. Therkildsen, N.O.; Nielsen, E.E.; Swain, D.P.; Pedersen, J.S. Large effective population size and temporal genetic stability in Atlantic cod (Gadus morhua) in the southern Gulf of St. Lawrence. Can. J. Fish. Aquat. Sci. 2010, 67, 1585–1595. [Google Scholar] [CrossRef]
  271. Loughnan, S.R.; Smith-Keune, C.; Beheregaray, L.B.; Robinson, N.A.; Jerry, D.R. Population genetic structure of barramundi (Lates calcarifer) across the natural distribution range in Australia informs fishery management and aquaculture practices. Mar. Freshw. Res. 2019, 70, 1533–1542. [Google Scholar] [CrossRef]
  272. Ruzzante, D.E.; Taggart, C.T.; Cook, D. Spatial and temporal variation in the genetic composition of a larval cod (Gadus morhua) aggregation: Cohort contribution and genetic stability. Can. J. Fish. Aquat. Sci. 1996, 53, 2695–2705. [Google Scholar] [CrossRef]
  273. Gkagkavouzis, K.; Karaiskou, N.; Katopodi, T.; Leonardos, I.; Abatzopoulos, T.J.; Triantafyllidis, A. The genetic population structure and temporal genetic stability of gilthead sea bream Sparus aurata populations in the Aegean and Ionian Seas, using microsatellite DNA markers. J. Fish Biol. 2019, 94, 606–613. [Google Scholar] [CrossRef]
  274. Ribeiro, A.; Moran, P.; Caballero, A. Genetic diversity and effective size of the Atlantic salmon Salmo salar L. inhabiting the River Eo (Spain) following a stock collapse. J. Fish Biol. 2008, 72, 1933–1944. [Google Scholar] [CrossRef]
  275. Hansen, M.M.; Fraser, D.J.; Meier, K.; Mensberg, K.-L.D. Sixty years of anthropogenic pressure: A spatio-temporal genetic analysis of brown trout populations subject to stocking and population declines. Mol. Ecol. 2009, 18, 2549–2562. [Google Scholar] [CrossRef] [PubMed]
  276. Hansen, M.M.; Ruzzante, D.E.; Nielsen, E.E.; Mensberg, K.L.D. Microsatellite and mitochondrial DNA polymorphism reveals life-history dependent interbreeding between hatchery and wild brown trout (Salmo trutta L.). Mol. Ecol. 2000, 9, 583–594. [Google Scholar] [CrossRef] [PubMed]
  277. Hansen, M.M. Estimating the long-term effects of stocking domesticated trout into wild brown trout (Salmo trutta) populations: An approach using microsatellite DNA analysis of historical and contemporary samples. Mol. Ecol. 2002, 11, 1003–1015. [Google Scholar] [CrossRef]
  278. Hansen, M.M.; Bekkevold, D.; Jensen, L.F.; Mensberg, K.L.D.; Nielsen, E.E. Genetic restoration of a stocked brown trout Salmo trutta population using microsatellite DNA analysis of historical and contemporary samples. J. Appl. Ecol. 2006, 43, 669–679. [Google Scholar] [CrossRef]
  279. Hansen, M.M.; Ruzzante, D.E.; Nielsen, E.E.; Mensberg, K.L.D. Brown trout (Salmo trutta) stocking impact assessment using microsatellite DNA markers. Ecol. Appl. 2001, 11, 148–160. [Google Scholar] [CrossRef]
  280. Guinand, B.; Scribner, K.T.; Page, K.S.; Burnham-Curtis, M.K. Genetic variation over space and time: Analyses of extinct and remnant lake trout populations in the Upper Great Lakes. Proc. R. Soc. B-Biol. Sci. 2003, 270, 425–433. [Google Scholar] [CrossRef]
  281. Miller, L.M.; Kapuscinski, A.R. Historical analysis of genetic variation reveals low effective population size in a northern pike (Esox lucius) population. Genetics 1997, 147, 1249–1258. [Google Scholar] [CrossRef]
  282. Horreo, J.L.; Machado-Schiaffino, G.; Ayllon, F.; Griffiths, A.M.; Bright, D.; Stevens, J.R.; Garcia-Vazquez, E. Impact of climate change and human-mediated introgression on southern European Atlantic salmon populations. Glob. Chang. Biol. 2011, 17, 1778–1787. [Google Scholar] [CrossRef]
  283. Nielsen, E.E.; Hansen, M.M.; Loeschcke, V. Analysis of microsatellite DNA from old scale samples of Atlantic salmon Salmo salar: A comparison of genetic composition over 60 years. Mol. Ecol. 1997, 6, 487–492. [Google Scholar] [CrossRef]
  284. Nielsen, J.L.; Heine, E.L.; Gan, C.A.; Fountain, M.C. Molecular analysis of population genetic structure and recolonization of rainbow trout following the Cantara Spill. Calif. Fish Game 2000, 86, 21–40. [Google Scholar]
  285. Montgomery, M.E.; Woodworth, L.M.; England, P.R.; Briscoe, D.A.; Frankham, R. Widespread selective sweeps affecting microsatellites in Drosophila populations adapting to captivity: Implications for captive breeding programs. Biol. Conserv. 2010, 143, 1842–1849. [Google Scholar] [CrossRef]
  286. Nock, C.J.; Ovenden, J.R.; Butler, G.L.; Wooden, I.; Moore, A.; Baverstock, P.R. Population structure, effective population size and adverse effects of stocking in the endangered Australian eastern freshwater cod Maccullochella ikei. J. Fish Biol. 2011, 78, 303–321. [Google Scholar] [CrossRef] [PubMed]
  287. Schreier, A.D.; Rodzen, J.; Ireland, S.; May, B. Genetic techniques inform conservation aquaculture of the endangered Kootenai River white sturgeon Acipenser transmontanus. Endanger. Species Res. 2012, 16, 65–75. [Google Scholar] [CrossRef]
  288. Englbrecht, C.C.; Schliewen, U.; Tautz, D. The impact of stocking on the genetic integrity of Arctic charr (Salvelinus) populations from the Alpine region. Mol. Ecol. 2002, 11, 1017–1027. [Google Scholar] [CrossRef]
  289. Valiquette, E.; Perrier, C.; Thibault, I.; Bernatchez, L. Loss of genetic integrity in wild lake trout populations following stocking: Insights from an exhaustive study of 72 lakes from Quebec, Canada. Evol. Appl. 2014, 7, 625–644. [Google Scholar] [CrossRef] [PubMed]
  290. Gossieaux, P.; Bernatchez, L.; Sirois, P.; Garant, D. Impacts of stocking and its intensity on effective population size in Brook Charr (Salvelinus fontinalis) populations. Conserv. Genet. 2019, 20, 729–742. [Google Scholar] [CrossRef]
  291. Caudron, A.; Champigneulle, A.; Guyomard, R. Evidence of two contrasting brown trout Salmo trutta populations spatially separated in the River Borne (France) and shift in management towards conservation of the native lineage. J. Fish Biol. 2009, 74, 1070–1085. [Google Scholar] [CrossRef]
  292. Fritzner, N.G.; Hansen, M.M.; Madsen, S.S.; Kristiansen, K. Use of microsatellite markers for identification of indigenous brown trout in a geographical region heavily influenced by stocked domesticated trout. J. Fish Biol. 2001, 58, 1197–1210. [Google Scholar] [CrossRef]
  293. Thaulow, J.; Borgstrom, R.; Heun, M. Brown trout population structure highly affected by multiple stocking and river diversion in a high mountain national park. Conserv. Genet. 2013, 14, 145–158. [Google Scholar] [CrossRef]
  294. Izquierdo, J.I.; Castillo, A.G.F.; Ayllon, F.; del al Hoz, J.; Garcia-Vazquez, E. Stock transfers in Spanish brown trout populations: A long-term assessment. Environ. Biol. Fishes 2006, 75, 153–157. [Google Scholar] [CrossRef]
  295. Wąs, A.; Bernaś, R. Long-term and seasonal genetic differentiation in wild and enhanced stocks of sea trout (Salmo trutta m. trutta L.) from the Vistula River, in the southern Baltic—Management implications. Fish. Res. 2016, 175, 57–65. [Google Scholar] [CrossRef]
  296. Was, A.; Wenne, R. Microsatellite DNA polymorphism in intensely enhanced populations of sea trout (Salmo trutta) in the Southern Baltic. Mar. Biotechnol. 2003, 5, 234–243. [Google Scholar] [CrossRef] [PubMed]
  297. Kawamura, K.; Furukawa, M.; Kubota, M.; Harada, Y. Effects of stocking hatchery fish on the phenotype of indigenous populations in the amago salmon Oncorhynchus masou ishikawae in Japan. J. Fish Biol. 2012, 81, 94–109. [Google Scholar] [CrossRef] [PubMed]
  298. Kitada, S. Japanese chum salmon stock enhancement: Current perspective and future challenges. Fish. Sci. 2014, 80, 237–249. [Google Scholar] [CrossRef]
  299. Darden, T.L.; Sessions, F.; Denson, M.R. Use of Genetic Microsatellite Markers to Identify Factors Affecting Stocking Success in Striped Bass. In Biology and Management of Inland Striped Bass and Hybrid Striped Bass; Bulak, J.S., Coutant, C.C., Rice, J.A., Eds.; American Fisheries Society: Grand Rapids, MI, USA, 2013; p. 395. [Google Scholar]
  300. Gudmundsson, L.A.; Gudjonsson, S.; Marteinsdottir, G.; Scarnecchia, D.L.; Danielsdttir, A.K.; Pampoulie, C. Spatio-temporal effects of stray hatchery-reared Atlantic salmon Salmo salar on population genetic structure within a 21 km-long Icelandic river system. Conserv. Genet. 2013, 14, 1217–1231. [Google Scholar] [CrossRef]
  301. Dannewitz, J.; Petersson, E.; Dahl, J.; Prestegaard, T.; Lof, A.C.; Jarvi, T. Reproductive success of hatchery-produced and wild-born brown trout in an experimental stream. J. Appl. Ecol. 2004, 41, 355–364. [Google Scholar] [CrossRef]
  302. Dannewitz, J.; Petersson, E.; Prestegaard, T.; Jarvi, T. Effects of sea-ranching and family background on fitness traits in brown trout Salmo trutta reared under near-natural conditions. J. Appl. Ecol. 2003, 40, 241–250. [Google Scholar] [CrossRef]
  303. Perrier, C.; Guyomard, R.; Bagliniere, J.L.; Nikolic, N.; Evanno, G. Changes in the genetic structure of Atlantic salmon populations over four decades reveal substantial impacts of stocking and potential resiliency. Ecol. Evol. 2013, 3, 2334–2349. [Google Scholar] [CrossRef]
  304. Perrier, C.; Bagliniere, J.-L.; Evanno, G. Understanding admixture patterns in supplemented populations: A case study combining molecular analyses and temporally explicit simulations in Atlantic salmon. Evol. Appl. 2013, 6, 218–230. [Google Scholar] [CrossRef]
  305. Perrier, C.; Daverat, F.; Evanno, G.; Pecheyran, C.; Bagliniere, J.-L.; Roussel, J.-M. Coupling genetic and otolith trace element analyses to identify river-born fish with hatchery pedigrees in stocked Atlantic salmon (Salmo salar) populations. Can. J. Fish. Aquat. Sci. 2011, 68, 977–987. [Google Scholar] [CrossRef]
  306. Martinez, J.L.; Gephard, S.; Juanes, F.; Garcia-Vazquez, E. The use of microsatellite DNA loci for genetic monitoring of Atlantic salmon populations. N. Am. J. Aquac. 2000, 62, 285–289. [Google Scholar] [CrossRef]
  307. Almodóvar, A.; Leal, S.; Nicola, G.G.; Hórreo, J.L.; García-Vázquez, E.; Elvira, B. Long-term stocking practices threaten the original genetic diversity of the southernmost European populations of Atlantic salmon Salmo salar. Endanger. Species Res. 2020, 41, 303–317. [Google Scholar] [CrossRef]
  308. Wang, X.; Weng, Z.; Yang, Y.; Hua, S.; Zhang, H.; Meng, Z. Genetic Evaluation of Black Sea Bream (Acanthopagrus schlegelii) Stock Enhancement in the South China Sea Based on Microsatellite DNA Markers. Fishes 2021, 6, 47. [Google Scholar] [CrossRef]
  309. Blanco Gonzalez, E.; Aritaki, M.; Sakurai, S.; Taniguchi, N. Inference of potential genetic risks associated with large-scale releases of red sea bream in Kanagawa prefecture, Japan based on nuclear and mitochondrial DNA analysis. Mar. Biotechnol. 2013, 15, 206–220. [Google Scholar] [CrossRef] [PubMed]
  310. Shishidou, H.; Kitada, S.; Sakamoto, T.; Hamasaki, K. Genetic variability of wild and hatchery-released red sea bream in Kagoshima Bay, Japan, evaluated by using microsatellite DNA analysis. Nippon Suisan Gakkaishi 2008, 74, 183–188. [Google Scholar] [CrossRef]
  311. Perez-Enriquez, R.; Takemura, M.; Tabata, K.; Taniguchi, N. Genetic diversity of red sea bream Pagrus major in western Japan in relation to stock enhancement. Fish. Sci. 2001, 67, 71–78. [Google Scholar] [CrossRef]
  312. Blanco Gonzalez, E.; Umino, T. Fine-scale genetic structure derived from stocking black sea bream, Acanthopagrus schlegelii (Bleeker, 1854), in Hiroshima Bay, Japan. J. Appl. Ichthyol. 2009, 25, 407–410. [Google Scholar] [CrossRef]
  313. Rourke, M.L.; McPartlan, H.C.; Ingram, B.A.; Taylor, A.C. Biogeography and life history ameliorate the potentially negative genetic effects of stocking on Murray cod (Maccullochella peelii peelii). Mar. Freshw. Res. 2010, 61, 918–927. [Google Scholar] [CrossRef]
  314. Sekino, M.; Saitoh, K.; Yamada, T.; Hara, M.; Yamashita, Y. Genetic tagging of released Japanese flounder (Paralichthys olivaceus) based on polymorphic DNA markers. Aquaculture 2005, 244, 49–61. [Google Scholar] [CrossRef]
  315. Ribeiro, R.P.; Lopera-Barrero, N.M.; Alves dos Santos, S.C.; Rodriguez-Rodriguez, M.d.P.; de Oliveira, S.N.; Alexandre Filho, L.; Vargas, L.; Souza dos Reis, G. Genetic diversity of pacu for restocking programs in the Tiete and Grande rivers, Brazil. Semin.-Cienc. Agrar. 2015, 36, 3807–3825. [Google Scholar] [CrossRef]
  316. Gessner, J.; Arndt, G.-M.; Tiedemann, R.; Bartel, R.; Kirschbaum, F. Remediation measures for the Baltic sturgeon: Status review and perspectives. J. Appl. Ichthyol. 2006, 22, 23–31. [Google Scholar] [CrossRef]
  317. Fopp-Bayat, D.; Kuciński, M.; Liszewski, T.; Teodorowicz, T.; Łączyńska, B.; Lebeda, I. Genetic protocol of Atlantic sturgeon Acipenser oxyrinchus (L.) fry for restocking the Vistula river, Poland. J. Surv. Fish. Sci. 2015, 2, 1–10. [Google Scholar] [CrossRef]
  318. De Innocentiis, S.; Longobardi, A.; Marino, G. Molecular tools in a marine restocking program for the endangered dusky grouper, Epinephelus marginatus. Rev. Fish. Sci. 2008, 16, 269–277. [Google Scholar] [CrossRef]
  319. Gonzalez-Wangueemert, M.; Fernandez, T.V.; Perez-Ruzafa, A.; Giacalone, M.; D’Anna, G.; Badalamenti, F. Genetic considerations on the introduction of farmed fish in marine protected areas: The case of study of white seabream restocking in the Gulf of Castellammare (Southern Tyrrhenian Sea). J. Sea Res. 2012, 68, 41–48. [Google Scholar] [CrossRef]
  320. An, H.S.; Yang, S.G.; Moon, T.S.; Park, J.Y.; Hong, C.G.; Hwang, H.K.; Myeong, J.I.; An, C. Comparison of genetic diversity between wild-caught broodstock and hatchery-produced offspring populations of the vulnerable Korean kelp grouper (Epinephelus bruneus) by microsatellites. Genet. Mol. Res. 2014, 13, 9675–9686. [Google Scholar] [CrossRef] [PubMed]
  321. Fernandez, G.; Cichero, D.; Patel, A.; Martinez, V. Genetic structure of Chilean populations of Seriola lalandi for the diversification of the national aquaculture in the north of Chile. Lat. Am. J. Aquat. Res. 2015, 43, 374–379. [Google Scholar] [CrossRef]
  322. Jesus Molina-Luzon, M.; Ramon Lopez, J.; Robles, F.; Navajas-Perez, R.; Ruiz-Rejon, C.; De la Herran, R.; Ignacio Navas, J. Chromosomal manipulation in Senegalese sole (Solea senegalensis Kaup, 1858): Induction of triploidy and gynogenesis. J. Appl. Genet. 2015, 56, 77–84. [Google Scholar] [CrossRef] [PubMed]
  323. Carson, E.W.; Bumguardner, B.W.; Fisher, M.; Saillant, E.; Gold, J.R. Spatial and temporal variation in recovery of hatchery-released red drum (Sciaenops ocellatus) in stock-enhancement of Texas bays and estuaries. Fish. Res. 2014, 151, 191–198. [Google Scholar] [CrossRef]
  324. Cheng, F.; Wu, Q.; Liu, M.; Radhakrishnan, K.V.; Murphy, B.R.; Xie, S. Impacts of hatchery release on genetic structure of rock carp Procypris rabaudi in the upper Yangtze River, China. Fish. Sci. 2011, 77, 765–771. [Google Scholar] [CrossRef]
  325. Hara, M.; Onoue, S.; Taniguchi, N. Assessing the impact of releasing exogenous hatchery-reared juveniles of Pacific abalone, Haliotis discus. Rev. Fish. Sci. 2008, 16, 278–284. [Google Scholar] [CrossRef]
  326. Read, K.D.; Lemay, M.A.; Acheson, S.; Boulding, E.G. Using molecular pedigree reconstruction to evaluate the long-term survival of outplanted hatchery-reared larval and juvenile northern abalone (Haliotis kamtschatkana). Conserv. Genet. 2012, 13, 801–810. [Google Scholar] [CrossRef]
  327. Borrell, Y.J.; Arias-Perez, A.; Freire, R.; Valdes, A.; Sanchez, J.A.; Mendez, J.; Martinez, D.; López, J.; Carleos, C.; Blanco, G.; et al. Microsatellites and multiplex PCRs for assessing aquaculture practices of the grooved carpet shell Ruditapes decussatus in Spain. Aquaculture 2014, 426, 49–59. [Google Scholar] [CrossRef]
  328. Vandersteen, W.; Biro, P.; Harris, L.; Devlin, R. Introgression of domesticated alleles into a wild trout genotype and the impact on seasonal survival in natural lakes. Evol. Appl. 2012, 5, 76–88. [Google Scholar] [CrossRef] [PubMed]
  329. Deines, A.M.; Bbole, I.; Katongo, C.; Feder, J.L.; Lodge, D.M. Hybridisation between native Oreochromis species and introduced Nile tilapia O. niloticus in the Kafue River, Zambia. Afr. J. Aquat. Sci. 2014, 39, 23–34. [Google Scholar] [CrossRef]
  330. Geletu, T.T.; Zhao, J. Genetic resources of Nile tilapia (Oreochromis niloticus Linnaeus, 1758) in its native range and aquaculture. Hydrobiologia 2022. [Google Scholar] [CrossRef]
  331. Ma, Y.; Wang, C.; Wang, J.; Yang, X.; Bi, X.; Xu, L. Genetic differentiation of wild and hatchery Oujiang color common carp: Potential application to the identification of escapees. Fish. Sci. 2011, 77, 591–597. [Google Scholar] [CrossRef]
  332. Esquer-Garrigos, Y.; Hugueny, B.; Ibañez, C.; Zepita, C.; Koerner, K.; Lambourdière, J.; Couloux, A.; Gaubert, P. Detecting natural hybridization between two vulnerable Andean pupfishes (Orestias agassizii and O. luteus) representative of the Altiplano endemic fisheries. Conserv. Genet. 2015, 16, 717–727. [Google Scholar] [CrossRef]
  333. Baggio, R.A.; Moretti, C.B.; Bialetzki, A.; Boeger, W.A. Hybrids between Pseudoplatystoma corruscans and P. reticulatum (Siluriformes: Pimelodidae) previously reported in the Upper Paraná River are likely escapes from aquaculture farms: Evidence from microsatellite markers. Zoologia 2016, 33, e20150200. [Google Scholar] [CrossRef]
  334. Besnier, F.; Glover, K.A.; Skaala, O. Investigating genetic change in wild populations: Modelling gene flow from farm escapees. Aquac. Environ. Interact. 2016, 2, 75–86. [Google Scholar] [CrossRef]
  335. Sawayama, E.; Nakao, H.; Kobayashi, W.; Minami, T.; Takagi, M. Identification and quantification of farmed red sea bream escapees from a large aquaculture area in Japan using microsatellite DNA markers. Aquat. Living Resour. 2019, 32, 26. [Google Scholar] [CrossRef]
  336. Segvic-Bubic, T.; Talijancic, I.; Grubisic, L.; Izquierdo-Gomez, D.; Katavic, I. Morphological and molecular differentiation of wild and farmed gilthead sea bream Sparus aurata: Implications for management. Aquac. Environ. Interact. 2014, 6, 43–54. [Google Scholar] [CrossRef]
  337. Žužul, I.; Šegvić-Bubić, T.; Talijančić, I.; Džoić, T.; Lepen Pleić, I.; Beg Paklar, G.; Ivatek-Šahdan, S.; Katavić, I.; Grubišić, L. Spatial connectivity pattern of expanding gilthead seabream populations and its interactions with aquaculture sites: A combined population genetic and physical modelling approach. Sci. Rep. 2019, 9, 14718. [Google Scholar] [CrossRef] [PubMed]
  338. Žužul, I.; Grubišić, L.; Šegvić-Bubić, T. Genetic discrimination of wild versus farmed gilthead sea bream Sparus aurata using microsatellite markers associated with candidate genes. Aquat. Living Resour. 2022, 35, 8. [Google Scholar] [CrossRef]
  339. Glover, K.A.; Dahle, G.; Jorstad, K.E. Genetic identification of farmed and wild Atlantic cod, Gadus morhua, in coastal Norway. ICES J. Mar. Sci. 2011, 68, 901–910. [Google Scholar] [CrossRef]
  340. Glover, K.A.; Dahle, G.; Westgaard, J.I.; Johansen, T.; Knutsen, H.; Jorstad, K.E. Genetic diversity within and among Atlantic cod (Gadus morhua) farmed in marine cages: A proof-of-concept study for the identification of escapees. Anim. Genet. 2010, 41, 515–522. [Google Scholar] [CrossRef]
  341. Puckrin, O.A.; Purchase, C.F.; Trippel, E.A. Using purposeful inbreeding to reduce outbreeding depression caused by escaped farmed Atlantic cod. Aquac. Environ. Interact. 2013, 4, 207–221. [Google Scholar] [CrossRef]
  342. Glover, K.A.; Quintela, M.; Wennevik, V.; Besnier, F.; Sorvik, A.G.E.; Skaala, O. Three Decades of Farmed Escapees in the Wild: A Spatio-Temporal Analysis of Atlantic Salmon Population Genetic Structure throughout Norway. PLoS ONE 2012, 7, e43129. [Google Scholar] [CrossRef] [PubMed]
  343. Glover, K.A.; Skilbrei, O.T.; Skaala, O. Genetic assignment identifies farm of origin for Atlantic salmon Salmo salar escapees in a Norwegian fjord. ICES J. Mar. Sci. 2012, 65, 912–920. [Google Scholar] [CrossRef]
  344. O’Reilly, P.T.; Carr, J.W.; Whoriskey, F.G.; Verspoor, E. Detection of European ancestry in escaped farmed Atlantic salmon, Salmo salar L.; in the magaguadavic river and Chamcook Stream, New Brunswick, Canada. ICES J. Mar. Sci. 2006, 63, 1256–1262. [Google Scholar] [CrossRef]
  345. Glover, K.A.; Sorvik, A.G.E.; Karlsbakk, E.; Zhang, Z.; Skaala, O. Molecular Genetic Analysis of Stomach Contents Reveals Wild Atlantic Cod Feeding on Piscine Reovirus (PRV) Infected Atlantic Salmon Originating from a Commercial Fish Farm. PLoS ONE 2013, 8, e60924. [Google Scholar] [CrossRef]
  346. Brown, C.; Miltiadou, D.; Tsigenopoulos, C.S. Prevalence and survival of escaped European seabass Dicentrarchus labrax in Cyprus identified using genetic markers. Aquac. Environ. Interact. 2013, 7, 49–59. [Google Scholar] [CrossRef]
  347. Noble, T.H.; Smith-Keune, C.; Jerry, D.R. Genetic investigation of the large-scale escape of a tropical fish, barramundi Lates calcarifer, from a sea-cage facility in northern Australia. Aquac. Environ. Interact. 2014, 5, 173–183. [Google Scholar] [CrossRef]
  348. An, H.S.; Kim, E.M.; Lee, J.W.; Dong, C.M.; Lee, B.I.; Kim, Y.C. Novel Polymorphic Microsatellite Loci for the Korean Black Scraper (Thamnaconus modestus), and Their Application to the Genetic Characterization of Wild and Farmed Populations. Int. J. Mol. Sci. 2011, 12, 4104–4119. [Google Scholar] [CrossRef]
  349. An, H.S.; Kim, E.M.; Lee, J.W.; Kim, D.J.; Kim, Y.C. New polymorphic microsatellite markers in the Korean mi-iuy croaker, Miichthys miiuy, and their application to the genetic characterization of wild and farmed populations. Anim. Cells Syst. 2012, 16, 41–49. [Google Scholar] [CrossRef]
  350. An, H.S.; Lee, J.W.; Kim, H.Y.; Kim, J.B.; Chang, D.S.; Park, J.Y.; Myeong, J.I.; An, C.M. Genetic differences between wild and hatchery populations of Korean spotted sea bass (Lateolabrax maculatus) inferred from microsatellite markers. Genes Genom. 2013, 35, 671–680. [Google Scholar] [CrossRef]
  351. Bahri-Sfar, L.; Lemaire, C.; Chatain, B.; Divanach, P.; Ben Hassine, O.K.; Bonhomme, F. Impact of aquaculture on the genetic structure of Mediterranean populations of Dicentrarchus labrax. Aquat. Living Resour. 2005, 18, 71–76. [Google Scholar] [CrossRef]
  352. Karahan, B.; Gokcek, E.O.; Gamsiz, K.; Magoulas, A. Microsatellite polymorphism of seabass (Dicentrarchus labrax) breeders: Natural and cultured stocks. Fresenius Environ. Bull. 2014, 23, 1289–1294. [Google Scholar]
  353. Karaiskou, N.; Triantafyllidis, A.; Katsares, V.; Abatzopoulos, T.J.; Triantaphyllidis, C. Microsatellite variability of wild and farmed populations of Sparus aurata. J. Fish Biol. 2009, 74, 1816–1825. [Google Scholar] [CrossRef]
  354. Loukovitis, D.; Sarropoulou, E.; Vogiatzi, E.; Tsigenopoulos, C.S.; Kotoulas, G.; Magoulas, A.; Chatziplis, D. Genetic variation in farmed populations of the gilthead sea bream Sparus aurata in Greece using microsatellite DNA markers. Aquac. Res. 2012, 43, 239–246. [Google Scholar] [CrossRef]
  355. Coughlan, J.P.; Imsland, A.K.; Galvin, P.T.; Fitzgerald, R.D.; Naevdal, G.; Cross, T.F. Microsatellite DNA variation in wild populations and farmed strains of turbot from Ireland and Norway: A preliminary study. J. Fish Biol. 1998, 52, 916–922. [Google Scholar] [CrossRef]
  356. Kim, W.J.; Kim, K.K.; Han, H.S.; Nam, B.H.; Kim, Y.O.; Kong, H.J.; Noh, J.K.; Yoon, M. Population structure of the olive flounder (Paralichthys olivaceus) in Korea inferred from microsatellite marker analysis. J. Fish Biol. 2010, 76, 1958–1971. [Google Scholar] [CrossRef]
  357. Sekino, M.; Hara, M.; Taniguchi, N. Loss of microsatellite and mitochondrial DNA variation in hatchery strains of Japanese flounder Paralichthys olivaceus. Aquaculture 2002, 213, 101–122. [Google Scholar] [CrossRef]
  358. Kohout, J.; Jaskova, I.; Papousek, I.; Sediva, A.; Slechta, V. Effects of stocking on the genetic structure of brown trout, Salmo trutta, in Central Europe inferred from mitochondrial and nuclear DNA markers. Fish. Manag. Ecol. 2012, 19, 252–263. [Google Scholar] [CrossRef]
  359. Was, A.; Wenne, R. Genetic differentiation in hatchery and wild sea trout (Salmo trutta) in the Southern Baltic at microsatellite loci. Aquaculture 2002, 204, 493–506. [Google Scholar] [CrossRef]
  360. Palm, S.; Dannewitz, J.; Jarvi, T.; Petersson, E.; Prestegaard, T.; Ryman, N. Lack of molecular genetic divergence between sea-ranched and wild sea trout (Salmo trutta). Mol. Ecol. 2003, 12, 2057–2071. [Google Scholar] [CrossRef]
  361. Koljonen, M.L.; Tahtinen, J.; Saisa, M.; Koskiniemi, J. Maintenance of genetic diversity of Atlantic salmon (Salmo salar) by captive breeding programmes and the geographic distribution of microsatellite variation. Aquaculture 2002, 212, 69–92. [Google Scholar] [CrossRef]
  362. Skaala, O.; Hoyheim, B.; Glover, K.; Dahle, G. Microsatellite analysis in domesticated and wild Atlantic salmon (Salmo salar L.): Allelic diversity and identification of individuals. Aquaculture 2004, 240, 131–143. [Google Scholar] [CrossRef]
  363. Withler, R.E.; Supernault, K.J.; Miller, K.M. Genetic variation within and among domesticated Atlantic salmon broodstocks in British Columbia, Canada. Anim. Genet. 2005, 36, 43–50. [Google Scholar] [CrossRef]
  364. Blanchet, S.; Paez, D.J.; Bernatchez, L.; Dodson, J.J. An integrated comparison of captive-bred and wild Atlantic salmon (Salmo salar): Implications for supportive breeding programs. Biol. Conserv. 2008, 141, 1989–1999. [Google Scholar] [CrossRef]
  365. Bingham, D.M.; Kennedy, B.M.; Hanson, K.C.; Smith, C.T. Loss of Genetic Integrity in Hatchery Steelhead Produced by Juvenile-Based Broodstock and Wild Integration: Conflicts in Production and Conservation Goals. N. Am. J. Fish. Manag. 2014, 34, 609–620. [Google Scholar] [CrossRef]
  366. Ha, H.P.; Nguyen, T.T.T.; Poompuang, S.; Na-Nakorn, U. Microsatellites revealed no genetic differentiation between hatchery and contemporary wild populations of striped catfish, Pangasianodon hypophthalmus (Sauvage 1878) in Vietnam. Aquaculture 2009, 291, 154–160. [Google Scholar] [CrossRef]
  367. Na-Nakorn, U.; Moeikum, T. Genetic diversity of domesticated stocks of striped catfish, Pangasianodon hypophthalmus (Sauvage 1878), in Thailand: Relevance to broodstock management regimes. Aquaculture 2009, 297, 70–77. [Google Scholar] [CrossRef]
  368. McKinna, E.M.; Nandlal, S.; Mather, P.B.; Hurwood, D.A. An investigation of the possible causes for the loss of productivity in genetically improved farmed tilapia strain in Fiji: Inbreeding versus wild stock introgression. Aquac. Res. 2010, 41, e730–e742. [Google Scholar] [CrossRef]
  369. Diyie, R.L.; Agyarkwa, S.K.; Armah, E.; Amonoo, N.A.; Owusu-Frimpong, I.; Osei-Atweneboana, M.Y. Genetic variations among different generations and cultured populations of Nile Tilapia (Oreochromis niloticus) in Ghana: Application of microsatellite markers. Aquaculture 2021, 544, 737070. [Google Scholar] [CrossRef]
  370. Panagiotopoulou, H.; Popovic, D.; Zalewska, K.; Weglenski, P.; Stankovic, A. Microsatellite multiplex assay for the analysis of Atlantic sturgeon populations. J. Appl. Genet. 2014, 55, 505–510. [Google Scholar] [CrossRef]
  371. Santos, C.H.; Santana, G.X.; Sa Leitao, C.S.; Paula-Silva, M.N.; Almeida-Val, V.M. Loss of genetic diversity in farmed populations of Colossoma macropomum estimated by microsatellites. Anim. Genet. 2016, 47, 373–376. [Google Scholar] [CrossRef]
  372. Aziz, D.; Siraj, S.S.; Daud, S.K.; Panandam, J.M.; Othman, M.F. Genetic Diversity of Wild and Cultured Populations of Penaeus monodon using Microsatellite Markers. J. Fish. Aquat. Sci. 2011, 6, 614–623. [Google Scholar] [CrossRef]
  373. Nahavandi, R.; Hafezamini, P.; Shamsudin, M.N. Genetic diversity of intensive cultured and wild tiger shrimp Penaeus monodon (Fabricius) in Malaysia using microsatellite markers. Afr. J. Biotechnol. 2011, 10, 15501–15508. [Google Scholar] [CrossRef]
  374. Maggioni, R.; Moura Coimbra, M.R.; da Costa, R.B.; Diniz, F.M.; Molina, W.F.; de Oliveira, D.M.; Puchnick-Legat, A. Genetic variability of marine shrimp in the Brazilian industry. Pesqui. Agropecu. Bras. 2013, 48, 968–974. [Google Scholar] [CrossRef]
  375. Schneider, K.J.; Tidwell, J.H.; Gomelsky, B.; Pomper, K.W.; Waldbieser, G.C.; Saillant, E.; Mather, P.B. Genetic diversity of cultured and wild populations of the giant freshwater prawn Macrobrachium rosenbergii (de Man, 1879) based on microsatellite analysis. Aquac. Res. 2013, 44, 1425–1437. [Google Scholar] [CrossRef]
  376. Geng, H.-J.; Zhou, Z.-C.; Dong, Y.; He, C.-B.; Zou, L.-L. Genetic structure analysis of wild and cultured populations of the Japanese sea urchin (Strongylocentrotus intermedius) using microsatellites. J. Fish. China 2009, 33, 549–556. [Google Scholar]
  377. Carlsson, J.; Morrison, C.L.; Reece, K.S. Wild and aquaculture populations of the eastern oyster compared using microsatellites. J. Hered. 2006, 97, 595–598. [Google Scholar] [CrossRef]
  378. Miller, P.A.; Elliott, N.G.; Koutoulis, A.; Kube, P.D.; Vaillancourt, R.E. Genetic Diversity of Cultured, Naturalized, and Native Pacific Oysters, Crassostrea gigas, Determined from Multiplexed Microsatellite Markers. J. Shellfish Res. 2012, 31, 611–617. [Google Scholar] [CrossRef]
  379. Lin, G.; Lo, L.C.; Yue, G.H. Genetic Variations in Populations from Farms and Natural Habitats of Asian Green Mussel, Perna viridis, in Singapore Inferred from Nine Microsatellite Markers. J. World Aquac. Soc. 2012, 43, 270–277. [Google Scholar] [CrossRef]
  380. Rhode, C.; Maduna, S.N.; Roodt-Wilding, R.; Bester-van der Merwe, A.E. Comparison of population genetic estimates amongst wild, F1 and F2 cultured abalone (Haliotis midae). Anim. Genet. 2014, 45, 456–459. [Google Scholar] [CrossRef] [PubMed]
  381. Evans, B.; Bartlett, J.; Sweijd, N.; Cook, P.; Elliott, N.G. Loss of genetic variation at microsatellite loci in hatchery produced abalone in Australia (Haliotis rubra) and South Africa (Haliotis midae). Aquaculture 2004, 233, 109–127. [Google Scholar] [CrossRef]
  382. Herbinger, C.M.; Doyle, R.W.; Pitman, E.R.; Paquet, D.; Mesa, K.A.; Morris, D.B.; Wright, J.M.; Cook, D. DNA fingerprint based analysis of paternal and maternal effects on offspring growth and survival in communally reared rainbow trout. Aquaculture 1995, 137, 245–256. [Google Scholar] [CrossRef]
  383. Aho, T.; Ronn, J.; Piironen, J.; Bjorklund, M. Impacts of effective population size on genetic diversity in hatchery reared Brown trout (Salmo trutta L.) populations. Aquaculture 2006, 253, 244–248. [Google Scholar] [CrossRef]
  384. Horvath, A.; Hoitsy, G.; Kovacs, B.; Sipos, D.K.; Osz, A.; Bogataj, K.; Urbanyi, B. The effect of domestication on a brown trout (Salmo trutta m fario) broodstock in Hungary. Aquac. Int. 2014, 22, 5–11. [Google Scholar] [CrossRef]
  385. Ward, R.D.; Jorstad, K.E.; Maguire, G.B. Microsatellite diversity in rainbow trout (Oncorhynchus mykiss) introduced to Western Australia. Aquaculture 2003, 219, 169–179. [Google Scholar] [CrossRef]
  386. Elliott, N.G.; Reilly, A. Likelihood of bottleneck event in the history of the Australian population of Atlantic salmon (Salmo salar L.). Aquaculture 2003, 215, 31–44. [Google Scholar] [CrossRef]
  387. Reilly, A.; Elliott, N.G.; Grewe, P.M.; Clabby, C.; Powell, R.; Ward, R.D. Genetic differentiation between Tasmanian cultured Atlantic salmon (Salmo salar L.) and their ancestral Canadian population: Comparison of microsatellite DNA and allozyme and mitochondrial DNA variation. Aquaculture 1999, 173, 459–469. [Google Scholar] [CrossRef]
  388. Wachirachaikarn, A.; Prakoon, W.; Nguyen, T.T.T.; Prompakdee, W.; Na-Nakorn, U. Loss of genetic variation of Phalacronotus bleekeri (Gunther, 1864) in the hatchery stocks revealed by newly developed microsatellites. Aquaculture 2011, 321, 298–302. [Google Scholar] [CrossRef]
  389. Wang, L.; Shi, X.; Su, Y.; Meng, Z.; Lin, H. Loss of genetic diversity in the cultured stocks of the large yellow croaker, Larimichthys crocea, revealed by microsatellites. Int. J. Mol. Sci. 2012, 13, 5584–5597. [Google Scholar] [CrossRef] [PubMed]
  390. An, H.S.; Shin, E.H.; Lee, J.W.; Nam, M.M.; Myeong, J.I.; An, C.M. Comparative genetic variability between broodstock and offspring populations of Korean starry flounder used for stock enhancement in a hatchery by using microsatellite DNA analyses. Genet. Mol. Res. 2013, 12, 6319–6330. [Google Scholar] [CrossRef] [PubMed]
  391. Austin, J.D.; Johnson, A.; Matthews, M.; Tringali, M.D.; Porak, W.F.; Allen, M.S. An assessment of hatchery effects on Florida bass (Micropterus salmoides floridanus) microsatellite genetic diversity and sib-ship reconstruction. Aquac. Res. 2012, 43, 628–638. [Google Scholar] [CrossRef]
  392. De La Rosa-Reyna, X.F.; Sifuentes-Rincon, A.M.; Parra-Bracamonte, G.M.; Arellano-Vera, W. Identification of Two Channel Catfish Stocks, Ictalurus punctatus, Cultivated in Northeast Mexico. J. World Aquac. Soc. 2014, 45, 104–114. [Google Scholar] [CrossRef]
  393. Fopp-Bayat, D. Microsatellite DNA variation in the Siberian sturgeon, Acipenser baeri (Actinopterygii, Acipenseriformes, Acipenseridae), cultured in a Polish fish farm. Acta Ichthyol. Piscat. 2010, 40, 21–25. [Google Scholar] [CrossRef]
  394. Fopp-Bayat, D.; Furgala-Selezniow, G. Application of Microsatellite DNA Variation in Russian Sturgeon (Acipenser gueldenstaedtii) and Sterlet Acipenser Ruthenus Cultured in a Polish Fish Farm. Pol. J. Nat. Sci. 2010, 25, 173–181. [Google Scholar] [CrossRef]
  395. Kaczmarczyk, D.; Luczynski, M.; Brzuzan, P. Genetic variation in three paddlefish (Polyodon spathula Walbaum) stocks based on microsatellite DNA analysis. Czech J. Anim. Sci. 2012, 57, 345–352. [Google Scholar] [CrossRef]
  396. Frost, L.A.; Evans, B.S.; Jerry, D.R. Loss of genetic diversity due to hatchery culture practices in barramundi (Lates calcarifer). Aquaculture 2006, 261, 1056–1064. [Google Scholar] [CrossRef]
  397. Loughnan, S.R.; Domingos, J.A.; Smith-Keune, C.; Forrester, J.P.; Jerry, D.R.; Beheregaray, L.B.; Robinson, N.A. Broodstock contribution after mass spawning and size grading in barramundi (Lates calcarifer, Bloch). Aquaculture 2013, 404, 139–149. [Google Scholar] [CrossRef]
  398. He, L.; Xie, Y.-N.; Lu, W.; Wang, Y.; Chen, L.-L.; Mather, P.B.; Zhao, Y.-L.; Wang, Y.-P.; Wang, Q. Genetic diversity in three redclaw crayfish (Cherax quadricarinatus, von Martens) lines developed in culture in China. Aquac. Res. 2012, 43, 75–83. [Google Scholar] [CrossRef]
  399. Hulak, M.; Kaspar, V.; Kohlmann, K.; Coward, K.; Tesitel, J.; Rodina, M.; Gela, D.; Kocour, M.; Linhart, O. Microsatellite-based genetic diversity and differentiation of foreign common carp (Cyprinus carpio) strains farmed in the Czech Republic. Aquaculture 2010, 298, 194–201. [Google Scholar] [CrossRef]
  400. Loukovitis, D.; Ioannidi, B.; Chatziplis, D.; Kotoulas, G.; Magoulas, A.; Tsigenopoulos, C.S. Loss of genetic variation in Greek hatchery populations of the European sea bass (Dicentrarchus labrax L.) as revealed by microsatellite DNA analysis. Mediterr. Mar. Sci. 2015, 16, 197–200. [Google Scholar] [CrossRef]
  401. Lizeth Cuevas-Rodriguez, B.; Parra-Bracamonte, M.; Garcia-Ulloa, M.; Maria Sifuentes-Rincon, A.; Rodriguez-Gonzalez, H. Genetic diversity of commercial species of the tilapia genus Oreochromis in Mexico. Int. J. Aquat. Sci. 2014, 5, 58–66. [Google Scholar]
  402. Li, Q.; Park, C.; Endo, T.; Kijima, A. Loss of genetic variation at microsatellite loci in hatchery strains of the Pacific abalone (Haliotis discus hannai). Aquaculture 2004, 235, 207–222. [Google Scholar] [CrossRef]
  403. Straus, K.M.; Vadopalas, B.; Davis, J.P.; Friedman, C.S. Reduced Genetic Variation and Decreased Effective Number of Breeders in Five Year-Classes of Cultured Geoducks (Panopea generosa). J. Shellfish Res. 2015, 34, 163–169. [Google Scholar] [CrossRef]
  404. Povh, J.A.; Ribeiro, R.P.; Lopera-Barrero, N.M.; Jacometo, C.B.; Vargas, L.; Gomes, P.C.; Lopes, T.d.S. Microsatellite analysis of pacu broodstocks used in the stocking program of Paranapanema River, Brazil. Sci. Agric. 2011, 68, 308–313. [Google Scholar] [CrossRef]
  405. Sriphairoj, K.; Kamonrat, W.; Na-Nakom, U. Genetic aspect in broodstock management of the critically endangered Mekong giant catfish, Pangasianodon gigas in Thailand. Aquaculture 2007, 264, 36–46. [Google Scholar] [CrossRef]
  406. Knibb, W.; Whatmore, P.; Lamont, R.; Quinn, J.; Powell, D.; Elizur, A.; Anderson, T.; Remilton, C.; Nguyen Hong, N. Can genetic diversity be maintained in long term mass selected populations without pedigree information?—A case study using banana shrimp Fenneropenaeus merguiensis. Aquaculture 2014, 428, 71–78. [Google Scholar] [CrossRef]
  407. Souza de Lima, A.P.; Bezerra Cabral da Silva, S.M.; Cavalcanti Oliveira, K.K.; Maggioni, R.; Moura Coimbra, M.R. Genetics of two marine shrimp hatcheries of the Pacific white shrimp Litopenaeus vannamei (Boone, 1931) in Pernambuco, Brazil. Cienc. Rural 2010, 40, 325–331. [Google Scholar]
  408. Na-Nakorn, U.; Yashiro, R.; Wachirachaikarn, A.; Prakoon, W.; Pansaen, N. Novel microsatellites for multiplex PCRs in the Humpback grouper, Cromileptes altivelis (Valenciennes, 1828), and applications for broodstock management. Aquaculture 2010, 306, 57–62. [Google Scholar] [CrossRef]
  409. Faber-Hammond, J.; Phillips, R.B.; Park, L.K. The Sockeye Salmon Neo-Y Chromosome Is a Fusion between Linkage Groups Orthologous to the Coho Y Chromosome and the Long Arm of Rainbow Trout Chromosome. Cytogenet. Genome Res. 2012, 136, 69–74. [Google Scholar] [CrossRef] [PubMed]
  410. Rengmark, A.H.; Slettan, A.; Skaala, Ø.; Lie, Ø.; Lingaas, F. Genetic variability in wild and farmed Atlantic salmon (Salmo salar) strains estimated by SNP and microsatellites. Aquaculture 2006, 253, 229–237. [Google Scholar] [CrossRef]
  411. Kong, L.; Bai, J.; Li, Q. Comparative assessment of genomic SSR, EST–SSR and EST–SNP markers for evaluation of the genetic diversity of wild and cultured Pacific oyster, Crassostrea gigas Thunberg. Aquaculture 2014, 420–421, S85–S91. [Google Scholar] [CrossRef]
  412. Sánchez-Velásquez, J.J.; Pinedo-Bernal, P.N.; Reyes-Flores, L.E.; Yzásiga-Barrera, C.; Zelada-Mázmela, E. Genetic diversity and relatedness inferred from microsatellite loci as a tool for broodstock management of fine flounder Paralichthys adspersus. Aquac. Fish. 2022, 7, 664–674. [Google Scholar] [CrossRef]
  413. Berrebi, P.; Horvath, Á.; Splendiani, A.; Palm, S.; Bernaś, R. Genetic diversity of domestic brown trout stocks in Europe. Aquaculture 2021, 544, 737043. [Google Scholar] [CrossRef]
  414. McDonald, G.J.; Danzmann, R.G.; Ferguson, M.M. Relatedness determination in the absence of pedigree information in three cultured strains of rainbow trout (Oncorhynchus mykiss). Aquaculture 2006, 233, 65–78. [Google Scholar] [CrossRef]
  415. De Mestral, L.G.; O’Reilly, P.T.; Jones, R.; Flanagan, J.; Herbinger, C.M. Preliminary assessment of the environmental and selective effects of a captive breeding and rearing programme for endangered Atlantic salmon, Salmo salar. Fish. Manag. Ecol. 2013, 20, 75–89. [Google Scholar] [CrossRef]
  416. Garant, D.; Dodson, J.J.; Bernatchez, L. A genetic evaluation of mating system and determinants of individual reproductive success in Atlantic salmon (Salmo salar L.). J. Hered. 2001, 92, 137–145. [Google Scholar] [CrossRef]
  417. Gruenthal, K.M.; Drawbridge, M.A. Toward responsible stock enhancement: Broadcast spawning dynamics and adaptive genetic management in white seabass aquaculture. Evol. Appl. 2012, 5, 405–417. [Google Scholar] [CrossRef] [PubMed]
  418. Sekino, M.; Saitoh, K.; Yamada, T.; Kumagai, A.; Hara, M.; Yamashita, Y. Microsatellite-based pedigree tracing in a Japanese flounder Paralichthys olivaceus hatchery strain: Implications for hatchery management related to stock enhancement program. Aquaculture 2003, 221, 255–263. [Google Scholar] [CrossRef]
  419. Borrell, Y.J.; Alvarez, J.; Vazquez, E.; Pato, C.F.; Tapia, C.M.; Sanchez, J.A.; Blanco, G. Applying microsatellites to the management of farmed turbot stocks (Scophthalmus maximus L.) in hatcheries. Aquaculture 2004, 241, 133–150. [Google Scholar] [CrossRef]
  420. Castro, J.; Bouza, C.; Presa, P.; Pino-Querido, A.; Riaza, A.; Ferreiro, I.; Sanchez, L.; Martinez, P. Potential sources of error in parentage assessment of turbot (Scophthalmus maximus) using microsatellite loci. Aquaculture 2004, 242, 119–135. [Google Scholar] [CrossRef]
  421. Estoup, A.; Gharbi, K.; SanCristobal, M.; Chevalet, C.; Haffray, P.; Guyomard, R. Parentage assignment using microsatellites in turbot (Scophthalmus maximus) and rainbow trout (Oncorhynchus mykiss) hatchery populations. Can. J. Fish. Aquat. Sci. 1998, 55, 715–725. [Google Scholar] [CrossRef]
  422. Norris, A.T.; Bradley, D.G.; Cunningham, E.P. Parentage and relatedness determination in farmed Atlantic salmon (Salmo salar) using microsatellite markers. Aquaculture 2000, 182, 73–83. [Google Scholar] [CrossRef]
  423. Bekkevold, D.; Hansen, M.M.; Loeschcke, V. Male reproductive competition in spawning aggregations of cod (Gadus morhua, L.). Mol. Ecol. 2002, 11, 91–102. [Google Scholar] [CrossRef]
  424. Garber, A.F.; Tosh, J.J.; Fordham, S.E.; Hubert, S.; Simpson, G.; Symonds, J.E.; Robinson, J.A.B.; Bowman, S.; Trippel, E.A. Survival and growth traits at harvest of communally reared families of Atlantic cod (Gadus morhua). Aquaculture 2010, 307, 12–19. [Google Scholar] [CrossRef]
  425. Senanan, W.; Pechsiri, J.; Sonkaew, S.; Na-Nakorn, U.; Sean-In, N.; Yashiro, R. Genetic relatedness and differentiation of hatchery populations of Asian seabass (Lates calcarifer) (Bloch, 1790) broodstock in Thailand inferred from microsatellite genetic markers. Aquac. Res. 2015, 46, 2897–2912. [Google Scholar] [CrossRef]
  426. Gold, J.R.; Renshaw, M.A.; Saillant, E.; Vega, R.R. Spawning frequency of brood dams and sires in a marine fish stock-enhancement hatchery. J. Fish Biol. 2010, 77, 1030–1040. [Google Scholar] [CrossRef] [PubMed]
  427. Borrell, Y.J.; Alvarez, J.; Blanco, G.; Martinez de Murguia, A.; Lee, D.; Fernandez, C.; Martinez, C.; Cotano, U.; Alvarez, P.; Sanchez Prado, J.A. A parentage study using microsatellite loci in a pilot project for aquaculture of the European anchovy Engraulis encrasicolus L. Aquaculture 2011, 310, 305–311. [Google Scholar] [CrossRef]
  428. Porta, J.; Porta, J.M.; Martinez-Rodriguez, G.; Alvarez, M.C. Genetic structure and genetic relatedness of a hatchery stock of Senegal sole (Solea senegalensis) inferred by microsatellites. Aquaculture 2006, 251, 46–55. [Google Scholar] [CrossRef]
  429. Kuo, H.-C.; Hsu, H.-H.; Chua, C.S.; Wang, T.-Y.; Chen, Y.-M.; Chen, T.-Y. Development of Pedigree Classification Using Microsatellite and Mitochondrial Markers for Giant Grouper Broodstock (Epinephelus lanceolatus) Management in Taiwan. Mar. Drugs 2014, 12, 2397–2407. [Google Scholar] [CrossRef]
  430. Camara, M.D.; Evans, S.; Langdon, C.J. Parental relatedness and survival of Pacific oysters from a naturalized population. J. Shellfish Res. 2008, 27, 323–336. [Google Scholar] [CrossRef]
  431. Liu, P.; Xia, J.H.; Lin, G.; Sun, F.; Liu, F.; Lim, H.S.; Pang, H.Y.; Yue, G.H. Molecular Parentage Analysis is Essential in Breeding Asian Seabass. PLoS ONE 2012, 7, e51142. [Google Scholar] [CrossRef]
  432. Luo, W.; Zeng, C.; Yi, S.; Robinson, N.; Wang, W.; Gao, Z. Heterosis and combining ability evaluation for growth traits of blunt snout bream (Megalobrama amblycephala) when crossbreeding three strains. Chin. Sci. Bull. 2012, 59, 857–864. [Google Scholar] [CrossRef]
  433. Rodriguez-Barreto, D.; Consuegra, S.; Jerez, S.; Cejas, J.R.; Martin, V.; Lorenzo, A. Using molecular markers for pedigree reconstruction of the greater amberjack (Seriola dumerili) in the absence of parental information. Anim. Genet. 2013, 44, 596–600. [Google Scholar] [CrossRef]
  434. Rodzen, J.A.; Famula, T.R.; May, B. Estimation of parentage and relatedness in the polyploid white sturgeon (Acipenser transmontanus) using a dominant marker approach for duplicated microsatellite loci. Aquaculture 2004, 232, 165–182. [Google Scholar] [CrossRef]
  435. Saillant, E.; Renshaw, M.A.; Gatlin, D.M.; Neill, W.H., III; Vega, R.R.; Gold, J.R. An experimental assessment of genetic tagging and founder representation in hatchery-reared red drum (Sciaenops ocellatus) used in stock enhancement. J. Appl. Ichthyol. 2009, 25, 108–113. [Google Scholar] [CrossRef]
  436. Vallecillos, A.; María-Dolores, E.; Villa, J.; Rueda, F.M.; Carrillo, J.; Ramis, G.; Soula, M.; Afonso, J.M.; Armero, E. Development of the First Microsatellite Multiplex PCR Panel for Meagre (Argyrosomus regius), a Commercial Aquaculture Species. Fishes 2022, 7, 117. [Google Scholar] [CrossRef]
  437. Wang, X.; Liu, X.-L.; Zhang, J.-Q.; Zhang, C.-S.; Huang, H.; Xiang, J.-H. Kinship analysis and genetic variation monitoring in Litopenaeus vannamei breeding program using microsatellite DNA markers. J. Fish. China 2009, 33, 832–839. [Google Scholar]
  438. Wang, H.; Cui, Z.; Wu, D.; Guo, E.; Liu, Y.; Wang, C.; Su, X.; Li, T. Application of microsatellite DNA parentage markers in the swimming crab Portunus trituberculatus. Aquac. Int. 2012, 20, 649–656. [Google Scholar] [CrossRef]
  439. Hara, M.; Sekino, M. Parentage testing for hatchery-produced abalone Haliotis discus hannai based on microsatellite markers: Preliminary evaluation of early growth of selected strains in mixed family farming. Fish. Sci. 2007, 73, 831–836. [Google Scholar] [CrossRef]
  440. Slabbert, R.; Bester, A.E.; D’Amato, M.E. Analyses of Genetic Diversity and Parentage Within a South African Hatchery of the Abalone Haliotis midae Linnaeus Using Microsatellite Markers. J. Shellfish Res. 2009, 28, 369–375. [Google Scholar] [CrossRef]
  441. Selvamani, M.J.P.; Degnan, S.M.; Degnan, B.N. Microsatellite genotyping of individual abalone larvae: Parentage assignment in aquaculture. Mar. Biotechnol. 2001, 3, 478–485. [Google Scholar] [CrossRef]
  442. Lu, X.; Wang, H.; Liu, B.; Lin, Z. Microsatellite-based genetic and growth analysis for a diallel mating design of two stocks of the clam, Meretrix meretrix. Aquac. Res. 2012, 43, 260–270. [Google Scholar] [CrossRef]
  443. Lu, X.; Wang, H.; Liu, B.; Xiang, J. An effective method for parentage determination of the clam (Meretrix meretrix) based on SSR and COI markers. Aquaculture 2011, 318, 223–228. [Google Scholar] [CrossRef]
  444. Boudry, P.; Collet, B.; Cornette, F.; Hervouet, V.; Bonhomme, F. High variance in reproductive success of the Pacific oyster (Crassostrea gigas, Thunberg) revealed by microsatellite-based parentage analysis of multifactorial crosses. Aquaculture 2002, 204, 283–296. [Google Scholar] [CrossRef]
  445. Lallias, D.; Taris, N.; Boudry, P.; Bonhomme, F.; Lapegue, S. Variance in the reproductive success of flat oyster Ostrea edulis L. assessed by parentage analyses in natural and experimental conditions. Genet. Res. 2010, 92, 175–187. [Google Scholar] [CrossRef]
  446. Cheng, P.; Yang, A.G.; Wu, B.; Zhou, L.Q.; Li, X. The applicability analysis on microsatellite markers for parentage determination of different shell color lines of japanese scallop Patinopecten yessoensis. Acta Hydrobiol. Sin. 2011, 35, 768–775. [Google Scholar]
  447. Novel, P.; Porta, J.; Fernandez, J.; Mendez, T.; Braulio Gallardo-Galvez, J.; Bejar, J.; Carmen Alvarez, M. Critical points for the maintenance of genetic variability over a production cycle in the European sea bass, Dicentrarchus labrax. Aquaculture 2013, 416, 8–14. [Google Scholar] [CrossRef]
  448. Puritz, J.B.; Renshaw, M.A.; Abrego, D.; Vega, R.R.; Gold, J.R. Reproductive Variance of Brood Dams and Sires used in Restoration Enhancement of Spotted Seatrout in Texas Bays and Estuaries. N. Am. J. Aquac. 2014, 76, 407–414. [Google Scholar] [CrossRef]
  449. Villanueva, B.; Verspoor, E.; Visscher, P.M. Parental assignment in fish using microsatellite genetic markers with finite numbers of parents and offspring. Anim. Genet. 2002, 33, 33–41. [Google Scholar] [CrossRef] [PubMed]
  450. Doyle, R.W.; Perez-Enriquez, R.; Takagi, M.; Taniguchi, N. Selective recovery of founder genetic diversity in aquacultural broodstocks and captive, endangered fish populations. Genetica 2001, 111, 291–304. [Google Scholar] [CrossRef] [PubMed]
  451. Kaczmarczyk, D. Genassemblage software, a tool for management of genetic diversity in human-dependent populations. Conserv. Genet. Resour. 2015, 7, 49–51. [Google Scholar] [CrossRef]
  452. Kaczmarczyk, D. Techniques Based on the Polymorphism of Microsatellite DNA as Tools for Conservation of Endangered Populations. Appl. Ecol. Environ. Res. 2019, 17, 1599–1615. [Google Scholar] [CrossRef]
  453. Ford, M.J.; Williamson, K.S. The Aunt and Uncle Effect Revisited-The Effect of Biased Parentage Assignment on Fitness Estimation in a Supplemented Salmon Population. J. Hered. 2010, 101, 33–41. [Google Scholar] [CrossRef]
  454. Prystupa, S.; McCracken, G.R.; Perry, R.; Ruzzante, D.E. Population abundance in arctic grayling using genetics and close-kin mark-recapture. Ecol. Evol. 2021, 11, 4763–4773. [Google Scholar] [CrossRef]
  455. Steele, C.A.; Anderson, E.C.; Ackerman, M.W.; Hess, M.A.; Campbell, N.R.; Narum, S.R.; Campbell, M.R. A validation of parentage-based tagging using hatchery steelhead in the Snake River basin. Can. J. Fish. Aquat. Sci. 2013, 70, 1046–1054. [Google Scholar] [CrossRef]
  456. Vandeputte, M.; Haffray, P. Parentage assignment with genomic markers: A major advance for understanding and exploiting genetic variation of quantitative traits in farmed aquatic animals. Front. Genet. 2014, 5, 432. [Google Scholar] [CrossRef]
  457. Wacker, S.; Skaug, H.J.; Forseth, T.; Solem, Ø.; Ulvan, E.M.; Fiske, P.; Karlsson, S. Considering sampling bias in close-kin mark–recapture abundance estimates of Atlantic salmon. Ecol. Evol. 2021, 11, 3917–3932. [Google Scholar] [CrossRef]
  458. Trenkel, V.M.; Charrier, G.; Lorance, P.; Bravington, M.V. Close-kin mark–recapture abundance estimation: Practical insights and lessons learned. ICES J. Mar. Sci. 2022, 79, 413–422. [Google Scholar] [CrossRef]
  459. Silva, N.M.L.; Ianella, P.; Yamagishi, M.E.B.; Rocha, J.L.; Teixeira, A.K.; Farias, F.G.; Guerrelhas, A.C.; Caetano, A.R. Development and validation of a low-density SNP panel for paternity and kinship analysis and evaluation of genetic variability and structure of commercial Pacific white shrimp (Litopenaeus vannamei) populations from Brazil. Aquaculture 2022, 560, 738540. [Google Scholar] [CrossRef]
  460. Kho, J.; McCracken, G.; McDermid, J.; Ruzzante, D. Life history implications of kinship structure in an Atlantic herring schooling aggregation. Authorea Prepr. 2022. [Google Scholar] [CrossRef]
  461. Yu, D.; Gao, X.; Shen, Z.; Fujiwara, M.; Yang, P.; Chang, T.; Zhang, F.; Wu, X.; Duan, Z.; Liu, H. Novel insights into the reproductive strategies of wild Chinese sturgeon (Acipenser sinensis) populations based on the kinship analysis. Water Biol. Secur. 2023, 2023, 100134. [Google Scholar] [CrossRef]
  462. Wąs-Barcz, A.; Bernaś, R. Parentage-based tagging and parentage analyses of stocked sea trout in Vistula River commercial catches. J. Appl. Genet. 2023. [Google Scholar] [CrossRef]
  463. Anderson, J.H.; Faulds, P.L.; Atlas, W.I.; Pess, G.R.; Quinn, T.P. Selection on breeding date and body size in colonizing coho salmon, Oncorhynchus kisutch. Mol. Ecol. 2010, 19, 2562–2573. [Google Scholar] [CrossRef]
  464. Dionne, M.; Miller, K.M.; Dodson, J.J.; Bernatchez, L. MHC standing genetic variation and pathogen resistance in wild Atlantic salmon. Philos. Trans. R. Soc. B-Biol. Sci. 2009, 364, 1555–1565. [Google Scholar] [CrossRef]
  465. de Eyto, E.; McGinnity, P.; Huisman, J.; Coughlan, J.; Consuegra, S.; Farrell, K.; O’Toole, C.; Tufto, J.; Megens, H.-J.; Jordan, W.; et al. Varying disease-mediated selection at different life-history stages of Atlantic salmon in fresh water. Evol. Appl. 2011, 4, 749–762. [Google Scholar] [CrossRef]
  466. Mallik, A.; Chakrabarty, U.; Dutta, S.; Mondal, D.; Mandal, N. Study on the Distribution of Disease-Resistant Shrimp Identified by DNA Markers in Respect to WSSV Infection in Different Seasons Along the Entire East Coast of India Aiming to Prevent White Spot Disease in Penaeus monodon. Transbound. Emerg. Dis. 2016, 63, e48–e57. [Google Scholar] [CrossRef]
  467. Mukherjee, K.; Mandal, N. A Microsatellite DNA Marker Developed for Identifying Disease-resistant Population of Giant Black Tiger Shrimp, Penaeus monodon. J. World Aquac. Soc. 2009, 40, 274–280. [Google Scholar] [CrossRef]
  468. Domingos, J.A.; Smith-Keune, C.; Harrison, P.; Jerry, D.R. Early prediction of long-term family growth performance based on cellular processes—A tool to expedite the establishment of superior foundation broodstock in breeding programs. Aquaculture 2014, 428, 88–96. [Google Scholar] [CrossRef]
  469. Vandeputte, M.; Dupont-Nivet, M.; Haffray, P.; Chavanne, H.; Cenadelli, S.; Parati, K.; Vidal, M.-O.; Vergnet, A.; Chatain, B. Response to domestication and selection for growth in the European sea bass (Dicentrarchus labrax) in separate and mixed tanks. Aquaculture 2009, 286, 20–27. [Google Scholar] [CrossRef]
  470. Overturf, K.; Casten, M.T.; LaPatra, S.L.; Rexroad, C.; Hardy, R.W. Comparison of growth performance, immunological response and genetic diversity of five strains of rainbow trout (Oncorhynchus mykiss). Aquaculture 2003, 217, 93–106. [Google Scholar] [CrossRef]
  471. Haffray, P.; Vandeputte, M.; Petit, V.; Pincent, C.; Chatain, B.; Chapuis, H.; Meriaux, J.C.; Coudurier, B.; Quillet, E.; Dupont-Nivet, M. Minimizing maternal effect in salmonid families mixed since eyed stages and a posteriori DNA-pedigreed. Livest. Sci. 2012, 150, 170–178. [Google Scholar] [CrossRef]
  472. Gislason, H.; Karstensen, H.; Christiansen, D.; Hjelde, K.; Helland, S.; Baeverfjord, G. Rib and vertebral deformities in rainbow trout (Oncorhynchus mykiss) explained by a dominant-mutation mechanism. Aquaculture 2010, 309, 86–95. [Google Scholar] [CrossRef]
  473. Douxfils, J.; Mathieu, C.; Mandiki, S.N.M.; Milla, S.; Henrotte, E.; Wang, N.; Vandecan, M.; Dieu, M.; Dauchot, N.; Pigneur, L.M.; et al. Physiological and proteomic evidences that domestication process differentially modulates the immune status of juvenile Eurasian perch (Perca fluviatilis) under chronic confinement stress. Fish. Shellfish Immunol. 2011, 31, 1113–1121. [Google Scholar] [CrossRef]
  474. Garber, A.; Fordham, S.; Tosh, J.; Glebe, B.; Forward, B.; Manning, A.; MacPhee, D.; Robertson, W. Atlantic Salmon Performance Selection and Broodstock Development Program for use in commercial saltwater aquaculture production on the east coast of Canada. In Aquaculture Canada 2012—Frontiers: Bridging Technology and Economic Growth; Wade, J., Reid, G.K., Eds.; Aquaculture Association of Canada: St. Andrews, NB, Canada, 2012; pp. 24–25. [Google Scholar]
  475. Portnoy, D.S.; Hollenbeck, C.M.; Vidal, R.R.; Gold, J.R. A Comparison of Neutral and Immune Genetic Variation in Atlantic Salmon, Salmo salar L. in Chilean Aquaculture Facilities. PLoS ONE 2014, 9, e99358. [Google Scholar] [CrossRef]
  476. Mas-Munoz, J.; Blonk, R.; Schrama, J.W.; van Arendonk, J.; Komen, H. Genotype by environment interaction for growth of sole (Solea solea) reared in an intensive aquaculture system and in a semi-natural environment. Aquaculture 2012, 410, 230–235. [Google Scholar] [CrossRef]
  477. Le Boucher, R.; Vandeputte, M.; Dupont-Nivet, M.; Quillet, E.; Ruelle, F.; Vergnet, A.; Kaushik, S.; Allamellou, J.M.; Medale, F.; Chatain, B. Genotype by diet interactions in European sea bass (Dicentrarchus labrax L.): Nutritional challenge with totally plant-based diets. J. Anim. Sci. 2013, 91, 44–56. [Google Scholar] [CrossRef]
  478. Ma, L.; Saillant, E.; Gatlin, D.M.; Neill, W.H., III; Vega, R.R.; Gold, J.R. Heritability of cold tolerance in red drum. N. Am. J. Aquac. 2007, 69, 381–387. [Google Scholar] [CrossRef]
  479. Saillant, E.; Wang, X.; Ma, L.; Gatlin, D.M.; Vega, R.R.; Gold, J.R. Genetic effects on tolerance to acute cold stress in red drum, Sciaenops ocellatus L. Aquac. Res. 2008, 39, 1393–1398. [Google Scholar] [CrossRef]
  480. Wolters, W.R.; Burr, G.S.; Palti, Y.; Vallejo, R.L. Phenotypic and Genetic Variation in Two North American Arctic Charr, Salvelinus alpinus, Stocks Cultured in a Recirculating Aquaculture System. J. World Aquac. Soc. 2013, 44, 473–485. [Google Scholar] [CrossRef]
  481. Norris, A.T.; Cunningham, E.P. Estimates of phenotypic and genetic parameters for flesh colour traits in fanned Atlantic salmon based on multiple trait animal model. Livest. Prod. Sci. 2004, 89, 209–222. [Google Scholar] [CrossRef]
  482. Castro, J.; Pino-Querido, A.; Hermida, M.; Chavarrias, D.; Romero, R.; Garcia-Cortes, L.A.; Toro, M.A.; Martinez, P. Heritability of skeleton abnormalities (lordosis, lack of operculum) in gilthead seabream (Spares aurata) supported by microsatellite family data. Aquaculture 2008, 279, 18–22. [Google Scholar] [CrossRef]
  483. Chen, G.; Liu, H.; Yu, X.; Luo, W.; Tong, J. Estimation of heritabilities and quantitative trait loci for growth traits of bighead carp (Hypophthalmichthys nobilis). Aquaculture 2023, 566, 739213. [Google Scholar] [CrossRef]
  484. Wan, S.; Li, Q.; Yu, H.; Liu, S.; Kong, L. Estimating heritability for meat composition traits in the golden shell strain of Pacific oyster (Crassostrea gigas). Aquaculture 2020, 516, 734532. [Google Scholar] [CrossRef]
  485. Huang, X.; Zhang, C.; Hu, J.; Tao, S.; Tang, J.; Huang, L.; Zheng, C.; Xu, S.; Wang, Y. Analyses of Growth Performance and Realized Heritability of the Second Generation of Indo-Pacific Pampus argenteus. J. Fish Biol. 2023, 102, 596–604. [Google Scholar] [CrossRef]
  486. Hansen, T.J.; Penman, D.; Glover, K.A.; Fraser, T.W.K.; Vågseth, T.; Thorsen, A.; Sørvik, A.G.E.; Fjelldal, P.G. Production and verification of the first Atlantic salmon (Salmo salar L.) clonal lines. BMC Genet. 2020, 21, 71. [Google Scholar] [CrossRef]
  487. Manan, H.; Noor Hidayati, A.B.; Lyana, N.A.; Amin-Safwan, A.; Ma, H.; Kasan, N.A.; Ikhwanuddin, M. A review of gynogenesis manipulation in aquatic animals. Aquac. Fish. 2022, 7, 1–6. [Google Scholar] [CrossRef]
  488. Galbusera, P.; Volckaert, F.A.M.; Ollevier, F. Gynogenesis in the African catfish Clarias gariepinus (Burchell, 1822) III. Induction of endomitosis and the presence of residual genetic variation. Aquaculture 2000, 185, 25–42. [Google Scholar] [CrossRef]
  489. Poompuang, S.; Sukkorntong, C. Microsatellite-centromere mapping in walking catfish Clarias macrocephalus (Gunther, 1864) using gynogenetic diploids. Aquac. Res. 2011, 42, 210–220. [Google Scholar] [CrossRef]
  490. Miao, L.; Li, M.-Y.; Chen, Y.-Y.; Guo, X.-F.; Xu, Y.-M.; Li, X.-M.; Chen, J. Development of Microsatellite Markers for Miiuy Croaker, Miichthys miiuy (Sciaenidae), and Their Application in Verifying Gynogenesis of Large Yellow Croaker, Pseudosciaena crocea, Induced with M. miiuy Sperm. J. World Aquac. Soc. 2015, 46, 83–91. [Google Scholar] [CrossRef]
  491. Castro, J.; Bouza, C.; Sanchez, L.; Cal, R.M.; Piferrer, F.; Martinez, P. Gynogenesis assessment using microsatellite genetic markers in turbot (Scophthalmus maximus). Mar. Biotechnol. 2003, 5, 584–592. [Google Scholar] [CrossRef]
  492. Chen, S.-L.; Tian, Y.-S.; Yang, J.-F.; Shao, C.-W.; Ji, X.-S.; Zhai, J.-M.; Liao, X.-L.; Zhuang, Z.-M.; Su, P.-Z.; Xu, J.-Y.; et al. Artificial Gynogenesis and Sex Determination in Half-Smooth Tongue Sole (Cynoglossus semilaevis). Mar. Biotechnol. 2009, 11, 243–251. [Google Scholar] [CrossRef]
  493. Flynn, S.R.; Matsuoka, M.; Reith, M.; Martin-Robichaud, D.J.; Benfey, T.J. Gynogenesis and sex determination in shortnose sturgeon, Acipenser brevirostrum Lesuere. Aquaculture 2006, 253, 721–727. [Google Scholar] [CrossRef]
  494. Fopp-Bayat, D. Meiotic gynogenesis revealed not homogametic female sex determination system in Siberian sturgeon (Acipenser baeri Brandt). Aquaculture 2010, 305, 174–177. [Google Scholar] [CrossRef]
  495. Grunina, A.S.; Skoblina, M.N.; Recoubratsky, A.V.; Kovalev, K.V.; Barmintseva, A.E.; Goncharov, B.F. Obtaining gynogenetic progeny of Siberian sturgeon (Acipenser baerii) using eggs matured and ovulated in vitro. J. Appl. Ichthyol. 2011, 27, 701–705. [Google Scholar] [CrossRef]
  496. Fopp-Bayat, D.; Ocalewicz, K. Activation of the Albino Sterlet Acipenser ruthenus Eggs by UV-Irradiated Bester Hybrid Spermatozoa to Provide Gynogenetic Progeny. Reprod. Domest. Anim. 2015, 50, 554–559. [Google Scholar] [CrossRef]
  497. Saber, M.H.; Hallajian, A. Study of sex determination system in ship sturgeon, Acipenser nudiventris using meiotic gynogenesis. Aquac. Int. 2014, 22, 273–279. [Google Scholar] [CrossRef]
  498. Saber, M.H.; Noveiri, S.B.; Pourkazemi, M.; Yarmohammadi, M. Induction of gynogenesis in stellate sturgeon (Acipenser stellatus Pallas, 1771) and its verification using microsatellite markers. Aquac. Res. 2008, 39, 1483–1487. [Google Scholar] [CrossRef]
  499. Fopp-Bayat, D. Induction of Diploid Gynogenesis in Wels Catfish (Silurus glanis) Using UV-Irradiated Grass Carp (Ctenopharyngodon idella) Sperm. J. Exp. Zool. Part A-Ecol. Genet. Physiol. 2010, 313A, 24–27. [Google Scholar] [CrossRef]
  500. Liu, S.; Qin, Q.; Wang, Y.; Zhang, H.; Zhao, R.; Zhang, C.; Wang, J.; Li, W.; Chen, L.; Xiao, J.; et al. Evidence for the Formation of the Male Gynogenetic Fish. Mar. Biotechnol. 2010, 12, 160–172. [Google Scholar] [CrossRef]
  501. Liu, Y.; Wang, G.; Liu, Y.; Hou, J.; Wang, Y.; Si, F.; Sun, Z.; Zhang, X.; Liu, H. Production and verification of heterozygous clones in Japanese flounder, Paralichthys olivaceus by microsatellite marker. Afr. J. Biotechnol. 2011, 10, 17088–17094. [Google Scholar]
  502. Wu, B.; Yang, A.-G.; Wang, Q.-Y.; Liu, Z.-H.; Zhou, L.-Q. Microsatellite analysis of gynogenetic diploids induced by heterologous sperm in Chlamys farreri. J. Fish. China 2009, 33, 542–548. [Google Scholar]
  503. Hou, J.; Saito, T.; Fujimoto, T.; Yamaha, E.; Arai, K. Androgenetic doubled haploids induced without irradiation of eggs in loach (Misgurnus anguillicaudatus). Aquaculture 2014, 420, S57–S63. [Google Scholar] [CrossRef]
  504. Nagoya, H.; Kawamura, K.; Ohta, H. Production of androgenetic amago salmon Oncorhynchus masou ishikawae with dispermy fertilization. Fish. Sci. 2010, 76, 305–313. [Google Scholar] [CrossRef]
  505. Li, Y.; Cai, M.; Wang, Z.; Guo, W.; Liu, X.; Wang, X.; Ning, Y. Microsatellite-centromere mapping in large yellow croaker (Pseudosciaena crocea) using gynogenetic diploid families. Mar. Biotechnol. 2008, 10, 83–90. [Google Scholar] [CrossRef]
  506. David, L.; Blum, S.; Feldman, M.W.; Lavi, U.; Hillel, J. Recent duplication of the common carp (Cyprinus carpio L.) genome as revealed by analyses of microsatellite loci. Mol. Biol. Evol. 2003, 20, 1425–1434. [Google Scholar] [CrossRef]
  507. Havelka, M.; Hulak, M.; Bailie, D.A.; Prodoehl, P.A.; Flajshans, M. Extensive genome duplications in sturgeons: New evidence from microsatellite data. J. Appl. Ichthyol. 2013, 29, 704–708. [Google Scholar] [CrossRef]
  508. Huang, S.; Cao, X.; Tian, X.; Luo, W.; Wang, W. Production of Tetraploid Gynogenetic Loach Using Diploid Eggs of Natural Tetraploid Loach, Misgurnus anguillicaudatus, Fertilized with UV-Irradiated Sperm of Megalobrama amblycephala without Treatments for Chromosome Doubling. Cytogenet. Genome Res. 2016, 147, 260–267. [Google Scholar] [CrossRef]
  509. Tao, M.; Song, Z.-Y.; Xiao, J.; Liu, S.-J.; Luo, K.-K.; Zou, T.-M.; Wang, J.; Liu, W.; Hu, J.; Zhao, R.-R.; et al. Cytogenetic and molecular genetic analysis of gynogenesis in Megalobrama amblycephala using spermatozoa of Erythroculter ilishaeformis. Zool. Res. 2013, 34, 479–486. [Google Scholar]
  510. Morishima, K.; Horie, S.; Yamaha, E.; Arai, K. A cryptic clonal line of the loach Misgurnus anguillicaudatus (Teleostei: Cobitidae) evidenced by induced gynogenesis, interspecific hybridization, microsatellite genotyping and multilocus DNA fingerprinting. Zool. Sci. 2002, 19, 565–575. [Google Scholar] [CrossRef]
  511. Fopp-Bayat, D.; Woznicki, P. Verification of ploidy level in sturgeon larvae. Aquac. Res. 2006, 37, 1671–1675. [Google Scholar] [CrossRef]
  512. Fopp-Bayat, D.; Woznicki, P. Test of Mendelian segregation among 10 microsatellite loci in the fourth generation of bester (Huso huso L. x Acipenser ruthenus L.). Aquac. Res. 2008, 39, 1377–1382. [Google Scholar] [CrossRef]
  513. Imsland, A.K.D.; Haugen, T.; Havardstun, S.; Mangor-Jensen, A. Triploid Induction in Atlantic Cod (Gadus morhua L.) by the Use of Different Pressure Levels. J. Appl. Aquac. 2014, 26, 252–262. [Google Scholar] [CrossRef]
  514. Peruzzi, S.; Chatain, B. Pressure and cold shock induction of meiotic gynogenesis and triploidy in the European sea bass, Dicentrarchus labrax L.: Relative efficiency of methods and parental variability. Aquaculture 2000, 189, 23–37. [Google Scholar] [CrossRef]
  515. Miller, P.A.; Elliott, N.G.; Vaillancourt, R.E.; Koutoulis, A.; Henshall, J.M. Assignment of parentage in triploid species using microsatellite markers with null alleles, an example from Pacific oysters (Crassostrea gigas). Aquac. Res. 2016, 47, 1288–1298. [Google Scholar] [CrossRef]
  516. Li, C.-Y.; Xu, M.-Y.; Zhao, J.-L.; Qian, Y.-Z.; Qian, D.; Wu, C. Microsatellite Analysis of Genetic Characteristics in Spotted Mandarinfish Siniperca schezeri female X Mandarinfish, S. chuatsi male Hybrids. Fish. Sci. 2014, 33, 97–102. [Google Scholar]
  517. Lafarga de la Cruz, F.; Amar-Basulto, G.; Angel del Rio-Portilla, M.; Gallardo-Escarate, C. Genetic Analysis of an Artificially Produced Hybrid Abalone (Haliotis rufescens × Haliotis Discus Hannai) in Chile. J. Shellfish. Res. 2010, 29, 717–724. [Google Scholar] [CrossRef]
  518. Luo, X.; Ke, C.-H.; You, W.-W.; Wang, D.-X.; Chen, F. Molecular identification of interspecific hybrids between Haliotis discus hannai Ino and Haliotis gigantea Gmelin using amplified fragment-length polymorphism and microsatellite markers. Aquac. Res. 2010, 41, 1827–1834. [Google Scholar] [CrossRef]
  519. Gao, B.Q.; Liu, P.; Li, J.; Wang, Q.Y.; Li, X.P. Genetic diversity of different populations and improved growth in the F1 hybrids in the swimming crab (Portunus trituberculatus). Genet. Mol. Res. 2014, 13, 10454–10463. [Google Scholar] [CrossRef]
  520. Vale, L.; Dieguez, R.; Sanchez, L.; Martinez, P.; Vinas, A. A sex-associated sequence identified by RAPD screening in gynogenetic individuals of turbot (Scophthalmus maximus). Mol. Biol. Rep. 2014, 41, 1501–1509. [Google Scholar] [CrossRef] [PubMed]
  521. Li, L.; Lian, Z.; Xiao, W.; Xing, L.; Liu, Y.; Sai, Q.; Yu, Z.; Tian, Y.; Wang, Y.; Liu, Z. Induction of meiotic gynogenesis in Lanzhou catfish (Silurus lanzhouensis) with heterologous sperm and evidence for female homogamety. Aquac. Res. 2022, 53, 3318–3330. [Google Scholar] [CrossRef]
  522. Biagi, C.A.; Leggatt, R.A.; Sakhrani, D.; Wetklo, M.; Vandersteen, W.E.; Christensen, K.A.; Rondeau, E.B.; Watson, B.M.; Wellband, K.W.; Koop, B.F.; et al. Timing of Postfertilization Pressure Shock Treatment for the Production of Mitotic Gynogens in Six Salmonid Species. N. Am. J. Aquac. 2022, 84, 505–515. [Google Scholar] [CrossRef]
  523. Wang, Y.; Liao, A.M.; Geng, C.; Tan, H.; Zhao, R.; Wang, S.; Wen, M.; Luo, K.; Qin, Q.; Zhang, C.; et al. The formation and study of allogynogenesis Hemibarbus maculatus Bleeker. Reprod. Breed. 2023, 3, 1–7. [Google Scholar] [CrossRef]
  524. Quere, N.; Guinand, B.; Kuhl, H.; Reinhardt, R.; Bonhomme, F.; Desmarais, E. Genomic sequences and genetic differentiation at associated tandem repeat markers in growth hormone, somatolactin and insulin-like growth factor-1 genes of the sea bass, Dicentrarchus labrax. Aquat. Living Resour. 2010, 23, 285–296. [Google Scholar] [CrossRef]
  525. Ozaki, A.; Okamoto, H.; Yamada, T.; Matuyama, T.; Sakai, T.; Fuji, K.; Sakamoto, T.; Okamoto, N.; Yoshida, K.; Hatori, K.; et al. Linkage analysis of resistance to Streptococcus iniae infection in Japanese flounder (Paralichthys olivaceus). Aquaculture 2010, 308, S62–S67. [Google Scholar] [CrossRef]
  526. Slettan, A.; Olsaker, I.; Lie, O. Segregation studies and linkage analysis of Atlantic salmon microsatellites using haploid genetics. Heredity 1997, 78, 620–627. [Google Scholar] [CrossRef]
  527. Gilbey, J.; Verspoor, E.; McLay, A.; Houlihan, D. A microsatellite linkage map for Atlantic salmon (Salmo salar). Anim. Genet. 2004, 35, 98–105. [Google Scholar] [CrossRef]
  528. Woram, R.A.; McGowan, C.; Stout, J.A.; Gharbi, K.; Ferguson, M.M.; Hoyheim, B.; Davidson, E.A.; Davidson, W.S.; Rexroad, C.; Danzmann, R.G. A genetic linkage map for Arctic char (Salvelinus alpinus): Evidence for higher recombination rates and segregation distortion in hybrid versus pure strain mapping parents. Genome 2004, 47, 304–315. [Google Scholar] [CrossRef] [PubMed]
  529. Cheng, L.; Liu, L.; Yu, X.; Wang, D.; Tong, J. A linkage map of common carp (Cyprinus carpio) based on AFLP and microsatellite markers. Anim. Genet. 2010, 41, 191–198. [Google Scholar] [CrossRef] [PubMed]
  530. Gu, Y.; Lu, C.; Zhang, X.; Li, C.; Yu, J.; Sun, X. Genetic mapping and QTL analysis for body weight in Jian carp (Cyprinus carpio var. Jian) compared with mirror carp (Cyprinus carpio L.). Chin. J. Oceanol. Limnol. 2015, 33, 636–649. [Google Scholar] [CrossRef]
  531. Chistiakov, D.A.; Tsigenopoulos, C.S.; Lagnel, J.; Guo, Y.M.; Hellemans, B.; Haley, C.S.; Volckaert, F.A.M.; Kotoulas, G. A combined AFLP and microsatellite linkage map and pilot comparative genomic analysis of European sea bass Dicentrarchus labrax L. Anim. Genet. 2008, 39, 623–634. [Google Scholar] [CrossRef]
  532. Tsigenopoulos, C.S.; Louro, B.; Chatziplis, D.; Lagnel, J.; Vogiatzi, E.; Loukovitis, D.; Franch, R.; Sarropoulou, E.; Power, D.M.; Patarnello, T.; et al. Second generation genetic linkage map for the gilthead sea bream Sparus aurata L. Mar. Genom. 2014, 18, 77–82. [Google Scholar] [CrossRef] [PubMed]
  533. Sekino, M.; Hara, M. Linkage maps for the Pacific abalone (genus Haliotis) based on microsatellite DNA markers. Genetics 2007, 175, 945–958. [Google Scholar] [CrossRef]
  534. Massault, C.; Hellemans, B.; Louro, B.; Batargias, C.; Van Houdt, J.K.J.; Canario, A.; Volckaert, F.A.M.; Bovenhuis, H.; Haley, C.; de Koning, D.J. QTL for body weight, morphometric traits and stress response in European sea bass Dicentrarchus labrax. Anim. Genet. 2014, 41, 337–345. [Google Scholar]
  535. Hollenbeck, C.M.; Portnoy, D.S.; Gold, J.R. A genetic linkage map of red drum (Sciaenops ocellatus) and comparison of chromosomal syntenies with four other fish species. Aquaculture 2015, 435, 265–274. [Google Scholar] [CrossRef]
  536. Portnoy, D.S.; Renshaw, M.A.; Hollenbeck, C.M.; Gold, J.R. A genetic linkage map of red drum, Sciaenops ocellatus. Anim. Genet. 2010, 41, 630–641. [Google Scholar] [CrossRef]
  537. Ma, H.; Chen, S.; Yang, J.; Chen, S.; Liu, H. Genetic linkage maps of barfin flounder (Verasper moseri) and spotted halibut (Verasper variegatus) based on AFLP and microsatellite markers. Mol. Biol. Rep. 2011, 38, 4749–4764. [Google Scholar] [CrossRef] [PubMed]
  538. Martin-Robichaud, D.; Reid, D.; Jackson, T.; Reith, M. Use of molecular genetic markers in Atlantic halibut broodstock management. Bull. Aquac. Assoc. Can. 2004, 104, 19–25. [Google Scholar]
  539. Ruan, X.; Wang, W.; Kong, J.; Yu, F.; Huang, X. Genetic linkage mapping of turbot (Scophthalmus maximus L.) using microsatellite markers and its application in QTL analysis. Aquaculture 2010, 308, 89–100. [Google Scholar] [CrossRef]
  540. Sanchez-Molano, E.; Cerna, A.; Toro, M.A.; Bouza, C.; Hermida, M.; Pardo, B.G.; Cabaleiro, S.; Fernandez, J.; Martinez, P. Detection of growth-related QTL in turbot (Scophthalmus maximus). BMC Genom. 2011, 12, 473. [Google Scholar] [CrossRef]
  541. Hermida, M.; Rodriguez-Ramilo, S.T.; Hachero-Cruzado, I.; Herrera, M.; Sciara, A.A.; Bouza, C.; Fernandez, J.; Martinez, P. First genetic linkage map for comparative mapping and QTL screening of brill (Scophthalmus rhombus). Aquaculture 2011, 420, S111–S120. [Google Scholar] [CrossRef]
  542. Agresti, J.J.; Seki, S.; Cnaani, A.; Poompuang, S.; Hallerman, E.M.; Umiel, N.; Hulata, G.; Gall, G.A.E.; May, B. Breeding new strains of tilapia: Development of an artificial center of origin and linkage map based on AFLP and microsatellite loci. Aquaculture 2000, 185, 43–56. [Google Scholar] [CrossRef]
  543. Liu, F.; Sun, F.; Li, J.; Xia, J.H.; Lin, G.; Tu, R.J.; Yue, G.H. A microsatellite-based linkage map of salt tolerant tilapia (Oreochromis mossambicus x Oreochromis spp.) and mapping of sex-determining loci. BMC Genom. 2013, 14, 58. [Google Scholar] [CrossRef]
  544. Araneda, C.; Lam, N.; Diaz, N.F.; Cortez, S.; Perez, C.; Neira, R.; Iturra, P. Identification, development, and characterization of three molecular markers associated to spawning date in Coho salmon (Oncorhynchus kisutch). Aquaculture 2013, 296, 21–26. [Google Scholar] [CrossRef]
  545. Haidle, L.; Janssen, J.E.; Gharbi, K.; Moghadam, H.K.; Ferguson, M.M.; Danzmann, R.G. Determination of quantitative trait loci (QTL) for early maturation in rainbow trout (Oncorhynchus mykiss). Mar. Biotechnol. 2008, 10, 579–592. [Google Scholar] [CrossRef]
  546. Perry, G.M.L.; Danzmann, R.G.; Ferguson, M.M.; Gibson, J.P. Quantitative trait loci for upper thermal tolerance in outbred strains of rainbow trout (Oncorhynchus mykiss). Heredity 2001, 86, 333–341. [Google Scholar] [CrossRef] [PubMed]
  547. Sakamoto, T.; Danzmann, R.G.; Okamoto, N.; Ferguson, M.M.; Ihssen, P.E. Linkage analysis of quantitative trait loci associated with spawning time in rainbow trout (Oncorhynchus mykiss). Aquaculture 1999, 173, 33–43. [Google Scholar] [CrossRef]
  548. Martyniuk, C.J.; Perry, G.M.L.; Mogahadam, H.K.; Ferguson, M.M.; Danzmann, R.G. The genetic architecture of correlations among growth-related traits and male age at maturation in rainbow trout. J. Fish Biol. 2003, 63, 746–764. [Google Scholar] [CrossRef]
  549. Boulton, K.; Massault, C.; Houston, R.D.; de Koning, D.J.; Haley, C.S.; Bovenhuis, H.; Batargias, C.; Canario, A.V.M.; Kotoulas, G.; Tsigenopoulos, C.S. QTL affecting morphometric traits and stress response in the gilthead seabream (Sparus aurata). Aquaculture 2011, 319, 58–66. [Google Scholar] [CrossRef]
  550. Loukovitis, D.; Batargias, C.; Sarropoulou, E.; Apostolidis, A.P.; Kotoulas, G.; Magoulas, A.; Tsigenopoulos, C.S.; Tsigenopoulos, C.S. Quantitative trait loci affecting morphology traits in gilthead seabream (Sparus aurata L.). Anim. Genet. 2013, 44, 480–483. [Google Scholar] [CrossRef]
  551. Loukovitis, D.; Sarropoulou, E.; Batargias, C.; Apostolidis, A.P.; Kotoulas, G.; Tsigenopoulos, C.S.; Chatziplis, D. Quantitative trait loci for body growth and sex determination in the hermaphrodite teleost fish Sparus aurata L. Anim. Genet. 2012, 43, 753–759. [Google Scholar] [CrossRef] [PubMed]
  552. Punhal, L.; Laghari, M.Y.; Narejo, N.T.; Zhang, X.; Xu, P.; Sun, X.; Zhang, Y. QTL for Short-Duration Vigorous Swimming Movements in Common Carp (Cyprinus carpio L.) Based on LDH Activity. Pak. J. Zool. 2014, 46, 383–390. [Google Scholar]
  553. Qiu, G.-F.; Xiong, L.-W.; Liu, Z.-Q.; Yan, Y.-L.; Shen, H. A first generation microsatellite-based linkage map of the Chinese mitten crab Eriocheir sinensis and its application in quantitative trait loci (QTL) detection. Aquaculture 2016, 451, 223–231. [Google Scholar] [CrossRef]
  554. Petersen, J.L.; Baerwald, M.R.; Ibarra, A.M.; May, B. A first-generation linkage map of the Pacific lion-paw scallop (Nodipecten subnodosus): Initial evidence of QTL for size traits and markers linked to orange shell color. Aquaculture 2012, 350, 200–209. [Google Scholar] [CrossRef]
  555. Bai, Z.; Han, X.; Luo, M.; Lin, J.; Wang, G.; Li, J. Constructing a microsatellite-based linkage map and identifying QTL for pearl quality traits in triangle pearl mussel (Hyriopsis cumingii). Aquaculture 2015, 437, 102–110. [Google Scholar] [CrossRef]
  556. Baranski, M.; Rourke, M.; Loughnan, S.; Hayes, B.; Austin, C.; Robinson, N. Detection of QTL for growth rate in the blacklip abalone (Haliotis rubra Leach) using selective DNA pooling. Anim. Genet. 2008, 39, 606–614. [Google Scholar] [CrossRef]
  557. Li, H.; Liu, X.; Zhang, G. Consensus Microsatellite-Based Linkage Map for the Hermaphroditic Bay Scallop (Argopecten irradians) and Its Application in Size-Related QTL Analysis. PLoS ONE 2012, 7, e46926. [Google Scholar] [CrossRef]
  558. Liu, B.; Wang, Q.; Li, J.; Liu, P.; He, Y. A genetic linkage map of marine shrimp Penaeus (Fenneropenaeus) chinensis based on AFLP, SSR, and RAPD markers. Chin. J. Oceanol. Limnol. 2010, 28, 815–825. [Google Scholar] [CrossRef]
  559. Houston, R.D.; Haley, C.S.; Hamilton, A.; Guyt, D.R.; Tinch, A.E.; Taggart, J.B.; McAndrew, B.J.; Bishop, S.C. Major quantitative trait loci affect resistance to infectious pancreatic necrosis in Atlantic salmon (Salmo salar). Genetics 2008, 178, 1109–1115. [Google Scholar] [CrossRef]
  560. Ozaki, A.; Sakamoto, T.; Khoo, S.; Nakamura, K.; Coimbra, M.R.M.; Akutsu, T.; Okamoto, N. Quantitative trait loci (QTLs) associated with resistance/susceptibility to infectious pancreatic necrosis virus (IPNV) in rainbow trout (Oncorhynchus mykiss). Mol. Genet. Genom. 2001, 265, 23–31. [Google Scholar]
  561. Li, J.; Boroevich, K.A.; Koop, B.F.; Davidson, W.S. Comparative Genomics Identifies Candidate Genes for Infectious Salmon Anemia (ISA) Resistance in Atlantic Salmon (Salmo salar). Mar. Biotechnol. 2011, 13, 232–241. [Google Scholar] [CrossRef] [PubMed]
  562. Vallejo, R.L.; Palti, Y.; Liu, S.; Evenhuis, J.P.; Gao, G.; Rexroad, C.E.; Wiens, G.D., III. Detection of QTL in Rainbow Trout Affecting Survival When Challenged with Flavobacterium psychrophilum. Mar. Biotechnol. 2014, 16, 349–360. [Google Scholar] [CrossRef] [PubMed]
  563. Rodriguez, M.F.; LaPatra, S.; Williams, S.; Famula, T.; May, B. Genetic markers associated with resistance to infectious hematopoietic necrosis in rainbow and steelhead trout (Oncorhynchus mykiss) backcrosses. Aquaculture 2004, 241, 93–115. [Google Scholar] [CrossRef]
  564. Liu, P.; Wang, L.; Wan, Z.Y.; Ye, B.Q.; Huang, S.; Wong, S.M.; Yue, G.H. Mapping QTL for Resistance Against Viral Nervous Necrosis Disease in Asian Seabass. Mar. Biotechnol. 2016, 18, 107–116. [Google Scholar] [CrossRef]
  565. Massault, C.; Franch, R.; Haley, C.; de Koning, D.J.; Bovenhuis, H.; Pellizzari, C.; Patarnello, T.; Bargelloni, L. Quantitative trait loci for resistance to fish pasteurellosis in gilthead sea bream (Sparus aurata). Anim. Genet. 2011, 42, 191–203. [Google Scholar] [CrossRef]
  566. Niu, H.; Nie, H.; Zhu, D.; Yang, F.; Yan, X. Identification of EST_SSR markers associated with growth-related traits in the Manila clam Ruditapes philippinarum. Acta Ecol. Sin. 2015, 35, 1910–1916. [Google Scholar]
  567. Plough, L.V.; Hedgecock, D. Quantitative Trait Locus Analysis of Stage-Specific Inbreeding Depression in the Pacific Oyster Crassostrea gigas. Genetics 2011, 189, 1473. [Google Scholar] [CrossRef] [PubMed]
  568. Lallias, D.; Gomez-Raya, L.; Haley, C.S.; Arzul, I.; Heurtebise, S.; Beaumont, A.R.; Boudry, P.; Lapègue, S. Combining Two-Stage Testing and Interval Mapping Strategies to Detect QTL for Resistance to Bonamiosis in the European Flat Oyster Ostrea edulis. Mar. Biotechnol. 2009, 11, 570–584. [Google Scholar] [CrossRef] [PubMed]
  569. Sauvage, C.; Boudry, P.; de Koning, D.J.; Haley, C.S.; Heurtebise, S.; Lapegue, S. QTL for resistance to summer mortality and OsHV-1 load in the Pacific oyster (Crassostrea gigas). Anim. Genet. 2010, 41, 390–399. [Google Scholar] [CrossRef] [PubMed]
  570. Martenot, C.; Fourour, S.; Oden, E.; Jouaux, A.; Travaille, E.; Malas, J.P.; Houssin, M. Detection of the OsHV-1 mu Var in the Pacific oyster Crassostrea gigas before 2008 in France and description of two new microvariants of the Ostreid Herpesvirus 1 (OsHV-1). Aquaculture 2012, 338, 293–296. [Google Scholar] [CrossRef]
  571. Martenot, C.; Lethuillier, O.; Fourour, S.; Oden, E.; Trancart, S.; Travaille, E.; Houssin, M. Detection of undescribed ostreid herpesvirus 1 (OsHV-1) specimens from Pacific oyster, Crassostrea gigas. J. Invertebr. Pathol. 2015, 132, 182–189. [Google Scholar] [CrossRef]
  572. Morrissey, T.; McCleary, S.; Collins, E.; Henshilwood, K.; Cheslett, D. An investigation of ostreid herpes virus microvariants found in Crassostrea gigas oyster producing bays in Ireland. Aquaculture 2015, 442, 86–92. [Google Scholar] [CrossRef]
  573. Qiu, Y.; Lu, H.; Zhu, J.; Chen, X.; Wang, A.; Wang, Y. Characterization of novel EST-SSR markers and their correlations with growth and nacreous secretion traits in the pearl oyster Pinctada martensii (Dunker). Aquaculture 2014, 420–421, S92–S97. [Google Scholar] [CrossRef]
  574. Martinez, V.; Dettleff, P.; Lopez, P.; Fernandez, G.; Jedlicki, A.; Yanez, J.M.; Davidson, W.S. Assessing footprints of selection in commercial Atlantic salmon populations using microsatellite data. Anim. Genet. 2013, 44, 223–226. [Google Scholar] [CrossRef]
  575. Cao, Z.; Ding, W.; Ren, H. Phenotypic Changes in Cyprinus carpio var. Jian Introduced by Sperm-Mediated Transgenesis of Rearranged Homologous DNA Fragments. Appl. Biochem. Biotechnol. 2013, 171, 189–197. [Google Scholar] [CrossRef] [PubMed]
  576. Matsuo, M.Y.; Nonaka, M. Repetitive elements in the major histocompatibility complex (MHC) class I region of a teleost, medaka: Identification of novel transposable elements. Mech. Dev. 2004, 121, 771–777. [Google Scholar] [CrossRef]
  577. Todd, C.D.; Walker, A.M.; Ritchie, M.G.; Graves, J.A.; Walker, A.F. Population genetic differentiation of sea lice (Lepeophtheirus salmonis) parasitic on Atlantic and Pacific salmonids: Analyses of microsatellite DNA variation among wild and farmed hosts. Can. J. Fish. Aquat. Sci. 2004, 61, 1176–1190. [Google Scholar] [CrossRef]
  578. Ai, C.H.; Li, B.J.; Xia, J.H. Mapping QTL for cold-tolerance trait in a GIFT-derived tilapia line by ddRAD-seq. Aquaculture 2022, 556, 738273. [Google Scholar] [CrossRef]
  579. Chen, C.-C.; Huang, C.-W.; Lin, C.-Y.; Ho, C.-H.; Pham, H.N.; Hsu, T.-H.; Lin, T.-T.; Chen, R.-H.; Yang, S.-D.; Chang, C.-I.; et al. Development of Disease-Resistance-Associated Microsatellite DNA Markers for Selective Breeding of Tilapia (Oreochromis spp.) Farmed in Taiwan. Genes 2022, 13, 99. [Google Scholar] [CrossRef] [PubMed]
  580. Kourkouta, C.; Tsipourlianos, A.; Power, D.M.; Moutou, K.A.; Koumoundouros, G. Variability of key-performance-indicators in commercial gilthead seabream hatcheries. Sci. Rep. 2022, 12, 17896. [Google Scholar] [CrossRef] [PubMed]
Figure 1. An integrated concept map showing the main fields of applications of microsatellite DNA markers related to aquatic exploited animal populations.
Figure 1. An integrated concept map showing the main fields of applications of microsatellite DNA markers related to aquatic exploited animal populations.
Genes 14 00808 g001
Table 1. Examples of applications of protein allozymes in studies of aquatic species.
Table 1. Examples of applications of protein allozymes in studies of aquatic species.
TopicSpeciesRegionReferences
Genetic structure of wild populations
European anchovy, Engraulis encrasicolusAdriatic, NE Atlantic[23]
sea trout, Salmo truttaNorway, NE Atlantic[24]
Baltic Sea, NE Atlantic[25]
Atlantic salmon, Salmo salarBritish Isles, NE Atlantic[26]
NE Atlantic[27]
Atlantic herring, Clupea harengusSweden, NE Atlantic[28]
red mullet, Mullus barbatus and M. surmuletusMediterranean Sea, NE Atlantic[29]
mullets, M. cephalus, M. soiuy, Liza ramada, L. aurata, L. abu, L. saliens, L. carinata, Chelon labrosus, C. labrosusMediterranean Sea, NE Atlantic[30,31,32]
tuna, Auxis thazard, A. rochei, Euthynnus affinis, Katsuwonus pelamis, Sarda orientalis, Thunnus tonggol, T. albacaresIndian Ocean[33]
chum salmon, Oncorhynchus ketaAlaska, N. Pacific[34]
chum salmon, Oncorhynchus ketaIslands of Japan, NE Pacific[35]
atlantic salmon, S. salar, brown trout, S. trutta, and their hybrids [36]
sockeye salmon, Oncorhynchus nerkaIslands of Japan, NE Pacific[37]
Brook charr, Salvelinus fontinalisNewfoundland, NW Atlantic[38]
Atlantic cod, Gadus morhuaNorway, NE Atlantic[39]
brown trout, S. truttaEurope, NE Atlantic[40]
horse mackerel, Trachurus trachurusMediterranean, NE Atlantic[41]
Atka mackerel, Pleurogrammus monopterygiusAleuts, N Pacific[42]
turbot, Psetta maxima, brill, Scophthalmus rhombusEurope, NE Atlantic[43]
red king crab, Paralithodes camtschaticusBering Sea, Gulf of Alaska, N. Pacific[44]
Norway lobster, Nephrops norvegicusNorth Sea, Aegean Sea, NE Atlantic[45]
clam, Macoma balthicaPacific and Atlantic coast, N America[46]
coot clam, Mulinia lateralisNW Atlantic[47]
blue mussel, Mytilus edulisNW Atlantic[48]
mussels M. edulis, M. trossulus, M. galloprovincialisNorthern and Southern Hemispreres[49]
Phylogenetic relationships
Salvelinus alpinus, S. malma, S. confluentus, S. leucomaenisNorth America[50]
23 cyprinid species (Alburninae and Leuciscinae) Central Europe[51]
hake Merluccius australis, M. hubbsSW Atlantic[52]
Sepia officinalis, S. orbignyana, S. elegansIberian Penisula, NE Atlantic[53]
Conservation genetics
brown trout, S. truttaIberian Penisula, NE Atlantic[54]
sockeye salmon, O. nerkaKenai River drainage, Alaska, N Pacific[55]
salmonidsE Pacific[56]
Changes in populations associated with fishery exploitation, management units, stocking
management unitssalmon, S. salarN Baltic, NE Atlantic[57]
mixed stock analysischum salmon, O. ketaYukon River, Alaska, NE Pacific[58]
fisheries managementchinook salmon, pink salmon, chum salmonNE Pacific[59]
stocking and escapees from aquaculture native cutthroat trout, Oncorhynchus clarki introgressed with stocked rainbow trout, O. mykissGreat Basin, Oregon, NE Pacific[60]
effects of stockingbrown trout, S. trutta, morpha farioMediterranean[61]
management units—stockingblack bream, Acanthopagrus butcheriS Australia[62]
interbreeding between the farmed escapes strain and the wild populationS. salarGlenarm River, N Ireland, NE Atlantic[63]
impact of the accidental and deliberate introduction of non-native salmonids on the genetic make-up of natural populationsS. salar, S. truttaEurope, N Atlantic[64]
Aquaculture genetics
gynogenesisprucian carps, C. gibelioUkraine, Europe[65]
less genetic diversity in hatchery stocks in comparison with natural populationsturbot, Scophthalmus maximusIberian Penisula, NE Atlantic[66]
hatchery rearing and implications of the use of reared fish in enhancement programmessalmon, S. salarN Ireland, NE Atlantic[67]
selected lines for growth had lower mean heterozygosity, lower percentages of polymorphic loci, and fewer alleles per locus than control lines; alleles correlated with growth ratechannel catfish, Ictalurus punctatusAlabama, USA, N. America[68]
spontaneous gynogens, aneuploids, or possibly incompatible with regulatory loci, hybridization, and triploidsCoho, Oncorhynchus kisutch, chum, O. keta, chinook, O. tshawytscha, Wshington, E Pacific[69]
loss of genetic variability in hatchery stock over 16 years brown trout, S. truttaFinland, Baltic, NE Atlantic[70]
reduction in polymorphism in hatchery stock introduced to ChileCoho salmon, O. kisutchChile, SE Pacific[71]
Genetic improvementPacific oyster, Crassostrea gigasAustralia imported to hatcheries from Japan Tasmania, Australia[72,73]
Table 2. Examples of mitochondrial DNA applications in studies of aquatic species.
Table 2. Examples of mitochondrial DNA applications in studies of aquatic species.
TopicSpeciesRegionReferences
Phylogenetic and phylogeographic analyses
phylogeographyoysters C. gigas, C. angulata, C. sikamea, C. ariakensis, C. hongkongensisChina, Asia[89]
phylogeography and glacial refugiaM. edulis, M. trossulus, M. galloprovincialisEurope[90]
phylogeography and glacial refugia schizothoracine fishes Tibet, Asia[91]
phylogeographyyellowtail amberjack, Seriola lalandiS Atlantic, Pacific[92]
phylogeographybrown trout, S. truttaEurope[93,94]
phylogeographybivalve pustulose ark, Anadara tuberculosaN South America, Pacific[95]
phylogeographyblack scraper, Thamnaconus modestusEast China and Japan Sea[96]
phylogeographyhilsa shad, Tenualosa ilishaN Indian Ocean[97]
Population differentiation studies
breakpoint between Atlantic and Mediterranean populationsgreater amberjack, Seriola dumeriliMediterranean, NE Atlantic[98]
large scale differentiationchum salmon, O. ketaPacific Rim, N Pacific[99,100]
postglacial colonizationbrown trout, S. truttaCentral Europe, NE Atlantic[101,102]
MtDNA of M. edulis in M. trossulusmussel, Mytilus trossulusBaltic, Europe[103,104]
European anchovies, E. encrasicolusMediterranean Sea, NE Atlantic[105]
mussel, Mytilus galloprovincialisCentral East Mediterranean[13]
no population differentiationAmerican eel, Anguilla rostrate, European eel, Anguilla anguillaN Atlantic[106]
differences between northern and southern populations horseshoe crab, Limulus polyphemusNW Atlantic[107]
differentiation between Alaska and CanadaOncorhynchus tshawytschaNE Pacific[108]
cod, G. morhuaE Atlantic[109]
cncordance alloz, mtDNA and nuclear DNAsockeye salmon, O. nerkaAlaska, N Pacific[74]
bay scallop, Argopecten irradiansNW Atlantic[110,111]
MtDNA of herringAtlantic herring C. harengus, Pacific herring C. pallasiiN Europe, NE Atlantic[112]
Chinese mitten crab, Eriocheir sinensisChina, Asia[113]
Conservation genetics
operational taxonomic units for conservation of endemic specieswhite-clawed crayfish, Austropotamobius italicusN Italy, Europe[114]
endemic; restitution of population by stockingAdriatic sturgeon, Acipenser naccariiAdriatic, Mediterranean Sea[115]
delimitation of evolutionarily significant units fish, Piaractus brachypomusAmazon and Orinoco basins, S America[116]
endemic; stock enhancementfish, honmoroko Gnathopogon caerulescensJapan[117]
endemic; stock enhancementcyprinid fish, Platypharodon extremusTibet, Asia[118]
management units huchen, Hucho huchoSlovenia, Europe[119]
wedge clam, Donax trunculusIberian Peninsula, NE Atlantic[120]
Species identification and taxonomy
rougheye rockfish, Sebastes aleutianusAlaska, N Pacific[78]
species and hybridsEuropean flounder, Platichthys flesus, plaice, Pleuronectes platessaBaltic, NE Atlantic[121]
species and hybridsOreochromis niloticus, Tilapia zillii, O. aureus, Sarotherodon galilaeusrivers in S China, Asia[122]
thinlip grey mullet, L. ramadaOder river, Poland, Europe[123]
Introgression from restocking
brown trout, S. truttaIberian Peninsula, Mediterranean, Europe[124]
introgression depends on environmental conditions in streamsbrown trout, S. truttacentral Italy, Europe[125]
low introgressionbrown trout, S. truttaSE Europe[94]
changes in mtDNA haplotype frequencyred sea bream, Pagrus majorKagoshima Bay, Japan, Asia[126]
low impactJapanese Spanish mackerel, Scomberomorus niphoniusSeto Inland Sea, Japan, Asia[127]
Introductions and invasions
two main clusters suggest secondary introduction inside EuropePacific oyster, C. gigasEurope[128]
origin: Francepike-perch, Sander luciopercaTunisia, N Africa[129]
identification of invasive species of tilapia originating from aquaculturetilapiaChina, Asia[122]
identification of invasive species of tilapia originating from aquaculturetilapiaJapan, Asia[130]
Studies related to fisheries
fisheries management and conservation12 speciesBaltic, NE Atlantic[131]
low usefulnessPacific salmonsNW Pacific[59]
potential application in mixed stock analysischum salmon, Oncorhynchus ketaYukon river, NW Pacific[58]
taxa identification in the commercial marine fisheryChondrichthyes (sharks, rays)Indian Economic Zone, N Indian Ocean[132]
elasmobranch species identification36 elasmobranch speciesMalta, Mediterranean, Europe[133]
Aquaculture
polymorphism in breeding stocksrainbow trout, Oncorhynchus mykissFinland, Europe[134]
Hybridization experimentsabalones Haliotis discus discus, H. madaka, H. gigantea.Japan, Asia[135]
lower polymorphism in cultured stocksgiant freshwater prawn, Macrobrachium rosenbergiiChina[136]
assessment of genetic diversity in aquaculture strainscommon carp, Cyprinus carpiocental Europe[17,137]
analysis of interspecific hybrid developmental stagescroakers, Larimichthys crocea x L. polyactisChina, Asia[138]
Seafood testing and forensics
identification of shark species in food productshammerhead shark, Sphyrna lewini, basking shark, Cetorhynus maximus [139]
fish product mislabeling (COI)fish speciesFrance, Europe[140]
barcoding and fish mislabelingfish speciesN America[141]
mislabeling of food productsseafoodEurope[142]
fish product mislabelingfish speciesGermany, Europe[143]
mislabeling of productscommon sole, Solea soleaGermany, Europe[144]
mislabeling of productsfish speciesTasmania, Australia[145]
mislabelingsturgeon caviarAustria[146]
mislabeling of sushi componentstuna, Thunnus sp., eel, Anguilla sp.England, Europe[147]
tracing illegal aquatic wildlife tradefish and mammalsPhilippines[148]
Table 3. Examples of species with population genetic structures found using microsatellites.
Table 3. Examples of species with population genetic structures found using microsatellites.
SpeciesRegionReferences
rainbow trout, Oncorhynchus mykiss; pink salmon, O. gorbuscha; chum salmon, O. keta;
coho salmon, O. kisutch; sockeye salmon, O. nerka; chinook salmon, O. tshawytscha; bull trout, Salvelinus confluentus
Elwha River, Waschington, USA, North America[178]
brook charr, S. fontinalisLa Mauricie lakes, Canada, N America[179]
brown trout, S. truttaNorth Atlantic, Mediterranean, Europe[180]
Atlantic salmon, S. salarIceland, Norway and Ireland, Europe
Nova Scotia, Canada, Noth America
[181,182,183]
broad whitefish, Coregonus nasusMackenzie River, Canada, North America[184]
Arctic charr, S. alpinusLabrador, Canada, Nunavut, Alaska, North America[185]
mulloway, Argyrosomus japonicusSouthern Australia [186]
Curimbatá, Prochilodus lineatusParaná River basin, South America[187]
Salminus franciscanus, Brycon orthotaeniaSão Francisco River system, Brazil, South America[188]
red snapper, Lutjanus campechanusGulf of Mexico, USA, North America[189]
European hake, Merluccius merlucciusNorth Atlantic, Mediterranean, Europe[190]
cod, G. morhuaNorth Atlantic, Europe[39,191,192,193,194]
Pacific herring, C. pallasiBering Sea and Alaskan waters, N Pacific [195]
Atlantic mackerel, Scomber scombrusNorth Atlantic[196]
yellowfin tuna, Thunnus albacaresCentral West Pacific [197]
native cobia, Rachycentron canadumGulf of Thailand and Andaman Sea, Southern Asia[198]
Atlantic sturgeon, Acipenser o. oxyrinchusAtlantic coast of North America[199]
Persian sturgeon, Acipenser persicusCaspian Sea, South Eastern Asia[200]
black carp, Mylopharyngodon piceusYangtze River, China, Asia[201]
European perch, Perca fluviatilisWulungu and Jili lakes, Kalaeerqisi River, North-West China, Asia[176]
Nile tilapia, O. niloticusKenya, Africa[202]
fish hardyhead, Craterocephalus fluviatilis Murray–Darling Basin, south-eastern Australia [203]
catfish, Pseudoplatystoma magdaleniatumMagdalena and Cauca rivers, Colombia, South America[204]
catfishes, Pseudoplatystoma corruscans, P. reticulatumParaguay, Parana, and Uruguay River basins, South America[205]
Hypophthalmus donascimientoiSolimões, Amazon River, Brazil, South America[206]
greater amberjack, Seriola dumeriliAtlantic and Mediterranean [98]
white oci, Seriola rivolianaMediterranean[98]
longnose skates, Zearaja chilensis, Dipturus trachydermaChile, South America[207]
longfin squid, Loligo pealeiiNorth West Atlantic [208]
sea cucumber, Holothuria mammataMediterranean and Atlantic [209]
Abalone, Haliotis asininaHeron Reef, Queensland, Australia[210]
flat oyster, Ostrea edulisEurope[211]
blue mussel, Mytilus chilensisSouthern Chile, South America[212]
Table 4. Comparison of wild populations and hatchery stocks based on the use microsatellites.
Table 4. Comparison of wild populations and hatchery stocks based on the use microsatellites.
SpeciesRegionReferences
black scraperThamnaconus modestusKorea, Asia[348]
mi-iuy croakerMiichthys miiuyKorea, Asia[349]
spotted sea bassLateolabrax maculatusKorea, Asia[350]
sea bassD. labraxMediterranean, Europe[351,352]
gilthead seabreamSparus aurataGreece, Europe[353,354]
turbotScophthalmus maximusIreland and Norway, Europe[355]
olive flounderParalichthys olivaceusKorea, Asia[356]
olive flounderP. olivaceusJapan, Asia[357]
brown troutSalmo truttaCzechs, Slovakia, Europe[358]
sea troutS. truttaPoland, Europe[359]
sea troutS. truttaRiver Dalalven, Sweden, Europe[360]
Atlantic salmonS. salarNorth Europe, Canada, North America[361,362,363,364]
steelheadO. mykissWashington, Pacific, USA, North America[365]
Arctic charrSalvelinus alpinusNorth America[185]
striped catfishPangasianodon hypophthalmusVietnam, Thailand, Asia[366,367]
tilapiaOreochromisSouth-West Pacific, Fiji, Asia[368]
Nile tilapiaO. niloticusGhana, Africa[369]
Atlantic sturgeonA. oxyrinchus, A. sturioUSA, Central Europe[370]
Colossoma macropomumBrazil, South America[371]
prawnPenaeus monodonMalaysia, Asia[372,373]
marine shrimpLitopenaeus vannameiBrazil, South America[374]
giant freshwater prawnMacrobrachium rosenbergiiUSA, Hawaii, Israel, India, Myanmar [375]
Japanese sea urchinStrongylocentrotus intermediusChina, Asia[376]
oysterCrassostrea virginicaChesapeake Bay, USA, Northe America[377]
Pacific oyster C. gigasAustralia, France, Korea, Japan, Asia[378]
green mussel Perna viridisSoutheast Asia[379]
abaloneHaliotis midaeSouth Africa [380]
abaloneH. rubraAustralia[381]
Disclaimer/Publisher’s Note: The statements, opinions and data contained in all publications are solely those of the individual author(s) and contributor(s) and not of MDPI and/or the editor(s). MDPI and/or the editor(s) disclaim responsibility for any injury to people or property resulting from any ideas, methods, instructions or products referred to in the content.

Share and Cite

MDPI and ACS Style

Wenne, R. Microsatellites as Molecular Markers with Applications in Exploitation and Conservation of Aquatic Animal Populations. Genes 2023, 14, 808. https://doi.org/10.3390/genes14040808

AMA Style

Wenne R. Microsatellites as Molecular Markers with Applications in Exploitation and Conservation of Aquatic Animal Populations. Genes. 2023; 14(4):808. https://doi.org/10.3390/genes14040808

Chicago/Turabian Style

Wenne, Roman. 2023. "Microsatellites as Molecular Markers with Applications in Exploitation and Conservation of Aquatic Animal Populations" Genes 14, no. 4: 808. https://doi.org/10.3390/genes14040808

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