You are on page 1of 12

Aquaculture 544 (2021) 737043

Contents lists available at ScienceDirect

Aquaculture
journal homepage: www.elsevier.com/locate/aquaculture

Genetic diversity of domestic brown trout stocks in Europe


Patrick Berrebi a, *, Ákos Horvath b, Andrea Splendiani c, Stefan Palm d, Rafał Bernaś e
a
Genome - Recherche & Diagnostic, 697 avenue de Lunel, 34400 Saint-Just, France
b
Hungarian University of Agriculture and Life Sciences, Institute of Aquaculture and Environmental Safety, H-2100 Gödöllő, Páter K. u. 1., Hungary
c
Dipartimento di Scienze della Vita e dell'Ambiente (DiSVA), Università Politecnica delle Marche, Via Brecce Bianche, 60131 Ancona, Italy
d
Swedish University of Agricultural Sciences, Department of Aquatic Resources, Institute of Freshwater Research, Stångholmsvägen 2, SE-178 93 Drottningholm, Sweden
e
Inland Fisheries Institute in Olsztyn, Department of Migratory Fishes, Rutki, 83-330 Żukowo, Poland

A R T I C L E I N F O A B S T R A C T

Statement of relevance: Genetic composition of Brown trout (Salmo trutta) is composed of numerous geographical forms in the wild and a multitude of stocks
hatchery strains all over Europe. reared in hatcheries. Practices impacting this species are domestication and fish farming. Thousands of hatch­
eries are producing domestic trout which are frequently released in the wild without real knowledge of the origin
Keywords: and genetic composition of the strains and of the receiving populations. The present study contains an analysis of
Brown trout (Salmo trutta)
the genetic structure (using twelve microsatellites) of 26 hatchery strains from Europe (Sweden, Poland, Czech
Hatchery stocks
Republic, Slovakia, Hungary, France, Italy and Spain) and one from North America (Minnesota). Several new
Microsatellites
Domestic strains observations improved our knowledge on this domestic part of S. trutta. First, a cross-border commercial strain of
Conservation the North Atlantic lineage occupies most of European hatcheries and the American one, as a probable conse­
quence of intensive exchanges of broodstock materials over time. In addition to the common Europe-wide
commercial Atlantic strain, local strains stemming from domesticated regional wild fish also occur. Second,
the level of genetic polymorphism of most hatchery strains is high to very high, likely reflecting genetic
admixture counteracting expected losses of diversity through random genetic drift and domestication. This study
emphasizes the value of identifying the genetic composition of hatchery stocks used for releases. It further
stresses the need for caution when stocking a common stock across the whole geographical range of a species,
with risks for reducing the intraspecific genetic diversity and local adaptation.

1. Introduction the last century. According to the D-loop marker of the mitochondrial
DNA, five main geographic lineages exist: Atlantic (AT), Mediterranean
Brown trout (Salmo trutta L.) is one of the most commonly manipu­ (ME), Adriatic (AD), marble trout (MA, north of Adriatic Sea) and
lated fish in the world, together with other salmonids as well as cichlid, Danube (DA) (Bernatchez et al. 1992). Several secondary lineages,
silurid and cyprinid species. In Europe, manipulation consists mainly of phylogenetically at the base of the main ones, have also been described,
domestication and production of eggs, fry, sub-adults and adults. e.g. Duero (DU), Tigris (TI) and North Africa (NA) (Suarez et al. 2001;
Stocking practices to enhance natural populations by release of Bardakci et al. 2006; Tougard et al. 2018, respectively).
hatchery-produced trout have been common in Europe for over a cen­ The taxonomic organization of brown trout has not reached a general
tury (Hansen and Loeschcke 1994). The main purpose of stocking with consensus. Among taxonomy-oriented ichthyologists, focusing mainly
hatchery-reared brown trout is to support angling and, in case of on morphology, many species have been described; Kottelat and Freyhof
anadromous forms, coastal fisheries. Started in the 18th century, arti­ (2007) reviewed the published nomenclature and reached a total
ficial production has exploded since the 1950s (Naish et al. 2008; ICES number of 28 nominal species. This number has been augmented by
2019). more than fifteen additional species during the last 15 years (Delling and
The natural range of brown trout covers Europe, Western Asia and Doadrio 2005; Turan et al. 2009, 2010, 2011, 2012, 2014a, 2014b;
North Africa. However, the real present range of this species is world­ Delling 2010; Doadrio et al. 2015). Among more molecular oriented
wide because of active introduction to all continents, mainly in the scientists, this part of the Salmo genus diversity is most often simply
beginning (Elliott, 1989) but also in the middle (Berrebi et al. 2020) of referred to as brown trout, Salmo trutta or the “brown trout complex”

* Corresponding author.
E-mail address: patrick.berrebi@laposte.net (P. Berrebi).

https://doi.org/10.1016/j.aquaculture.2021.737043
Received 10 November 2020; Received in revised form 4 June 2021; Accepted 10 June 2021
Available online 17 June 2021
0044-8486/© 2021 Elsevier B.V. All rights reserved.
P. Berrebi et al. Aquaculture 544 (2021) 737043

(Sanz 2018). understanding of genetic structure across European brown trout hatch­
At the same time as numerous studies on the genetic diversity in wild ery strains can provide a better understanding of effects from decades of
European brown trout and other salmonids have been carried out, in traditional domestic management. Such knowledge also allows infer­
step with the rapid development in molecular methodology (e.g. ence on the history of used hatchery strains, biodiversity protection,
Antunes et al. 1999; Sanz 2018), natural populations are constantly population management on particular river basins and monitoring of
being admixed with domestic strains, resulting in progressive genetic stocking activities. A description of European domestic stocks may
homogenization (i.e. reduced differentiation between populations, often further be of value for research on the evolution of introduced trout
paralleled by increased diversity at the local level). As an example, based outside Europe, especially in order to detect the geographic origin of
on genotyping of 19 microsatellite loci in adult lake trout (Salvelinus historic introductions (Dieterman et al. 2012; Hoxmeier and Dieterman
namaycush), a halving of genetic differentiation and an increase in ge­ 2013; French et al. 2016; Berrebi et al. 2020).
netic diversity was observed among stocked populations compared to in
unstocked ones (Valiquette et al. 2014). Examples of mixed genetic 2. Material and methods
lineages after stocking can also be found for brown trout in almost all
European countries (e.g. Largiadèr and Scholl 1995; Apostolidis et al. 2.1. Sampling
1996; Lerceteau-Köhler et al. 2013; Almodóvar et al., 2006; Fruciano
et al. 2014; Berrebi et al. 2019; Bernaś and Wąs-Barcz 2020). Clear The present analysis includes 779 trout sampled from 26 European
conflicts of interests often exist between groups of anglers and ichthy­ hatcheries (see supplementary document), with a concentration on its
ologists advocating large-scaled releases of farm-raised trout to support eastern part. Northern and western samples have been added in order to
fisheries on one side, and nature conservationists (including other an­ develop a comparative perspective, together with a North American one
glers and most evolutionary biologists) trying to preserve local native (stemming from 19th century introductions, when acclimatization so­
trout diversity on the other side. cieties encouraged the introduction of non-native species at the world
While intra-specific diversity and genetic structure in wild brown scale with the hope of their adaptation).
trout often follows a logic related to hydrology, geography and post­ Tissue samples (fin-clips in ethanol) were collected between years
glacial recolonization, the genetic structure of domestic strains is un­ 2000 and 2017, and sent to Montpellier (France) where the molecular
predictable, as it is entirely driven by human actions such as strain analyses were done (sub-contracted by Labofarm, private company at
creations, long-distance translocations and strain admixtures. Therefore, Loudéac, France). Eleven collectors assured sampling in 9 countries.
knowledge about the diversity of domestic strains across Europe is Samples and collectors are described in Table 1, whereas geographic
crucial to understand effects of stocking: the more similar captive stocks positions are shown in Fig. 1.
are to each other and distinct from local wild populations, the greater
risks are associated with stocking and supplementation as these may 2.2. Molecular techniques
result in an overall reduction in genetic diversity (Valiquette et al. 2014;
Bohling et al. 2016). Fin clips were cleared of alcohol and a small piece of fin was used for
On the basis of molecular information, Bohling et al. (2016) pro­ DNA extraction, following the improved Chelex extraction protocol of
posed a classification for French hatcheries which can be easily extended Estoup et al. (1996). A set of twelve nuclear microsatellite loci were
to stocks in other parts of Europe: ComATL refers to genetically similar, selected for analysis (Mst543, Mst85, Omm1105, Omy21Dias, Oneμ9,
internationally-distributed, Atlantic trout strains (the only one category Sfo1, Ssa197, SsoSL311, SsoSL438, SsoSL417, Str591 and StrBS131) ac­
of strain dispatched everywhere in France), whereas LocATL refers to cording to high levels polymorphism observed in previous studies of
strains derived from local rivers within the Atlantic watershed. Simi­ brown trout. The characteristics of the markers are given in Table 2. The
larly, LocMED denotes stocks derived from local rivers within the twelve loci were amplified within three multiplex PCR reactions
Mediterranean watershed. Finally, the MIX category refers to assem­ (Table 2). PCR amplifications were carried out using the Qiagen multi­
blages stemming from mixing between ComATL and local-based plex PCR kit in a 10 μL volume containing 3 μL of genomic DNA diluted
(Atlantic or Mediterranean) strains. at 10 ng/μL, 5 μL of Qiagen PCR Master Mix, 1 μL of Qiagen Q-solution,
The production of hatchery fish is typically associated with various 1 μL of primer mix of variable concentration (Table 2) with forward
constraints, resulting in losses of genetic diversity and low effective primers labelled at 5′ end using different fluorescent dyes (FAM, HEX, or
population sizes (e.g. Wąs and Bernaś 2016), founder effects (e.g. Skaala NED). Amplifications were conducted using a GeneAmp PCR System
et al. 2004) and random genetic drift (e.g. Klütsch et al. 2019). 2700 thermal cycler (Applied Biosystems), according to the supplier's
Furthermore, unnatural selection pressures in hatcheries, differing instructions (Qiagen multiplex PCR kit): initial denaturation step at
markedly from those in nature, result in domestication affecting multi­ 95 ◦ C for 15 min; followed by 35 cycles of denaturation at 94 ◦ C (30 s),
ple traits like warm water adaptation, growth rate and stress adaptation annealing (59 ◦ C, 90 s) and extension (72 ◦ C, 59 s); with a final extension
(e.g. Gjedrem 2000). In France, for example, experimental within-group step at 59 ◦ C during 30 min. Amplified PCR fragments were then diluted
selection to improve growth provided significant improvement in body and separated on a capillary ABIPRISM 3130xl sequencer (Applied
weight in captive-reared brown trout (Chevassus et al. 1992, 2004) and Biosystems) with the use of GeneScan500Rox dye as standards size.
an unknown number of farms today rear stocks derived from this Fragment lengths were assessed using the GENEMAPPER v4.1 software
experimentally selected population. Additionally, many hatcheries are system (Life Technologies TM).
working with closed reproductive cycles using only their own “gene
pool” without a proper supplement of fish from the wild. Consequently, 2.3. Population parameters and equilibriums
genetic origin, levels of introgression and inbreeding of stocks varies,
and the success and risk associated with release of hatchery fish in Levels of genetic diversity were estimated using the GENETIX software
natural environments is highly unpredictable (e.g. Aparicio et al. 2005; 4.05.2 program (Belkhir et al. 2004) for the following four basic pa­
Berrebi et al. 2000; Delling et al. 2000). rameters: Ho (observed heterozygosity based on the relative proportions
The main objective of the present work is to assess the genetic of homozygote and heterozygote genotypes); He (expected heterozy­
structure and levels of diversity within and among brown trout from gosity in a theoretical population in Hardy-Weinberg-Equilibrium
hatcheries distributed around Europe, from central Sweden to western (HWE)); Hnb (unbiased heterozygosity compensating for sample size
Spain, with a particular focus on Eastern European countries. So far, few disparity according to Nei 1978); A (mean number of alleles per locus).
genetic studies have been devoted to hatchery brown trout strains (but Additionally, estimates of allelic richness (which allows comparison
see, e.g., Bohling et al. 2016), and a broader survey is necessary. A better of allele numbers without the bias associated with different sample

2
P. Berrebi et al. Aquaculture 544 (2021) 737043

Table 1
Listing and characteristics of the 26 samples analyzed, totalizing 779 trouts. See Fig. 1 for map. Expected lineages are the types of strain deduced from local infor­
mation, generally without genetic analysis (see supplementary document).
N◦ Map Hatchery name Expected lineage Year COUNTRY N Collector

St. 01 Dalälven (VBF river strain) Local Atlantic 2002 Sweden 18 Stefan Palm
St. 02 Dalälven (HBF river strain) Local Atlantic 2003 Sweden 16 Stefan Palm
St. 03 Gullspang (Lake Vänern strain) Local Atlantic 2010 Sweden 33 Stefan Palm
St. 04 Bartush Local Atlantic 2017 Poland 23 Rafał Bernaś
St. 05 Rutki Local Atlantic 2016 Poland 40 Rafał Bernaś
St. 06 Dabie Domestic Atlantic 2017 Poland 29 Rafał Bernaś
St. 07 Aquamar Local Atlantic 2016 Poland 30 Rafał Bernaś
St. 08 Czarci Jar Local Atlantic 2017 Poland 24 Rafał Bernaś
St. 09 Folutsz Local Atlantic 2017 Poland 23 Rafał Bernaś
St. 10 Turnov (Austrian strain) Unknown 2009 Czech Rep. 26 Jan Kohout (1)
St. 11 Turnov (Italian strain) Unknown 2009 Czech Rep. 28 Jan Kohout (1)
St. 12 Borová Lada Unknown 2007 Czech Rep. 28 Jan Kohout (1)
St. 13 Východná Unknown 2007 Slovakia 29 Jan Kohout (1)
St. 14 Slovryb Unknown 2017 Slovakia 24 Rafał Bernaś
St. 15 Lillafüred domestic Atlantic 2015 Hungary 30 Akos Horvath
St. 16 Cantiano Domestic Atlantic 2006 Italy 43 Vincenzo Caputo (2)
St. 17 Visso Domestic Atlantic 2000 Italy 40 Vincenzo Caputo (2)
St. 18 Roquebillière Local Mediterranean 2008 France 29 Genesalm (3)
St. 19 Murgat Domestic Atlantic 2008 France 30 Genesalm (3)
St. 20 Brantôme (Dronne strain) Local Atlantic 2007 France 56 Genesalm (3)
St. 21 Babeau (Gravezon strain) Local Mediterranean 2014 France 40 Eric Ravel (4)
St. 22 Soueich Domestic Atlantic 2011 France 29 François Dauba (5)
St. 23 Cauterets Domestic Atlantic 2014 France 32 Eric Ravel (4)
St. 24 Baga Unknown 2014 Spain 25 Jose-Luis García-Marín (6)
St. 25 O Veral-Lugo Unknown 2004 Spain 27 Carmen Bouza (7)
St. 26 Lanesboro Domestic Atlantic 2016 USA 27 Doug Dieterman (8)

Collectors: (1) Institute of Zoology, Bratislava, Slovakia; (2) Università Politecnica delle Marche, Ancona, Italy; (3) French national project (2009–2013) for brown
trout genetic structure, Université Montpellier II, Montpellier, France; (4) Fédération de pêche de l'Hérault, Octon, France; (5) INTP-ENSAT, Castanet-Tolosan, France;
(6) Universidad de Girona, Girona, Spain; (7) Universidad de Santiago de Compostela, Lugo, Spain; (8) Minnesota Department of Natural Resources, Lake City,
Minnesota, USA.

sizes) and private allelic richness (alleles limited in a single population) Markov Chain Monte Carlo simulations to partition individuals into K
were calculated using the rarefaction method in the HP-RARE 1.1. soft­ clusters (here set between 1 and 10). Assignment criteria are the mini­
ware (Kalinowski 2005). The frequency of private alleles provide an mization of deviations from HWE and linkage disequilibria (LD) within
indication of the extent of gene flow between populations (Slatkin 1985; clusters. Different run lengths were used at each step (from 50,000 to
Barton and Slatkin 1986) and is a useful criterion in genetic diversity 100,000 burn-in and 100,000 to 200,000 total lengths, repeated 5 to 10
preservation, by allowing evaluation of the uniqueness of each popula­ times for each K) depending on convergence (proportion of identical
tion in terms of allele composition (Foulley and Ollivier, 2006). structure among the 5 repeated runs at each step). The admixture
Deviations from panmixia were estimated through the FIS parameter, ancestry model and correlated allele frequency options were chosen.
using the estimator f of Weir and Cockerham (1984). Panmixia (FIS = 0) Sampling location was not used as prior information. The ΔK method
characterizes a population in which reproduction between all adult (Evanno et al. 2005) was applied as a decision help, as implemented in
members is at random. Deviations in terms of a deficiency (FIS > 0) or STRUCTURE HARVESTER (Earl and von Holdt 2012). In order to detect
excess (FIS < 0) of heterozygotes may be observed in numerous bio­ differentiated subgroups of hatchery populations, a ‘hierarchical STRUC­
logical cases, the most frequent being related to sampling of population TURE analysis’ (cf. Vähä et al. 2007; Marić et al. 2017; Berrebi et al. 2019)
mixtures, migration, or a low population size (few parents, few families). was performed. After the first analysis and determination of the most
Deviations of FIS from zero were tested statistically using 5000 permu­ probable K according to the ΔK method of Evanno et al. (2005), which
tations of alleles within each locus, using the GENETIX software. Likewise, define the first hierarchical level, each sub-group of the first level was
genetic differentiation between samples was quantified using the FST analyzed separately, allowing a grouping of individuals without elimi­
parameter through the estimator Ɵ (Weir and Cockerham 1984), using nating mixed individuals. We repeated the same procedure until no
5000 permutations of individuals among samples to assess the signifi­ more substructure was observed (indicated by red crosses in Fig. 5). This
cance of departures of Ɵ from zero. hierarchical approach was preferred because a global run tended to
detect only major groups.

2.4. Research of hatchery genetic clustering


2.5. Parentage investigation
In order to depict the overall genetic structure of the involved
hatchery samples in a single diagram, a Factorial Correspondence In order to describe each strain in terms of family structure and
Analysis (FCA, Benzécri 1973) was first performed, as implemented abundance of genitors, parentage structure was explored using the
within the GENETIX program. This method is well adapted to genotypic COLONY software (Jones and Wang 2010). The main objective was full-sib
data (simplified mathematical details are given in She et al. 1987). Each and half-sib dyads detection, together with the number of families
trout is positioned in a hyperspace according to its 24 alleles. Clusters within each analyzed sample. A genotyping error rate of 0.001 was used,
(clouds) detected on the diagram correspond to homogeneous groups following Palm et al. (2008). Sibs identified were not removed because,
gathering individuals with similar multilocus genotypes, regardless of with the exception of a few samples, there was no excess of sibs.
their geographical origin. Effective population size was estimated, using the sibship assignment
An alternate clustering method is the assignment analysis imple­ method implemented in COLONY (Wang 2009). This method makes
mented in the program STRUCTURE 2.3.2 (Pritchard et al. 2000). It runs several assumptions that could be violated in reality. The most critical

3
P. Berrebi et al. Aquaculture 544 (2021) 737043

Fig. 1. Geographic position of the 25 European samples of domestic stocks of brown trout (Salmo trutta) (the 26th one, from Minnesota, USA, is not shown). The
numbers refer to the first column, of Table 1.

Nb/Ne ratios if necessary (Waples et al. 2013). However, from the point
Table 2
of view of breeding stocks, Nb values seem more informative because the
Twelve microsatellite loci used to characterize domestic stocks of brown trout
number of fish in stocks may change over time depending on production
(Salmo trutta) in Europe and North America.
needs.
Locus Multiplex Primers concentration Reference

Mst543 C 0.15 μM Presa et al. 1994


Mst85 B 0.15 μM Presa and Guyomard 1996 2.6. Phylogenetic analysis
Omm1105 A 0.8 μM Rexroad et al. 2002
Omy21Dias B 0.1 μM Holm and Bendixen 2000
As a complement to the clustering analyses described above, a
Oneμ9 C 0.2 μM Scribner et al. 1996
Sfo1 A 0.1 μM Angers et al. 1995 neighbor-joining (NJ) tree was constructed based on pairwise Nei's
Ssa197 A 0.2 μM O'Reilly et al. 1996 standard genetic distance without sample size correction (DST, Nei
SsoSL311 C 0.6 μM Slettan et al. 1995 1972) with 10,000 bootstrap replications, using the POPTREE2 software
SsoSL417 A 0.1 μM Slettan et al. 1995
(Takezaki et al. 2010).
SsoSL438 C 0.1 μM Slettan et al. 1996
Str591 B 0.2 μM Presa and Guyomard 1996
Because most samples are of Atlantic lineage, the two Mediterranean
StrBS131 C 0.4 μM Charles et al. 2005 ones (St. 18 and 21) were treated as outgroups. The objective of this kind
of phylogeny is to link apparently independent hatchery stocks implic­
itly suggesting a history of exchanges, which has not been recorded. This
assumption is that the sample has been taken at random (with respect to information allows understanding how hatchery exchanges can shape
kinship) from a single cohort of the population. If there are several co­ the constitution of trout strains.
horts in the sample, it is possible that some individuals are actually
parents to other sampled individuals. As a consequence, some parental-
3. Results
offsprings relationships may be mistakenly inferred to be full-sibs (FS),
increasing the estimate of FS frequency and decreasing the estimate of
3.1. Population parameters
effective size. In the present study, this potential bias was neglected
since most samples contained single cohorts. Thus, the effective size
The results obtained on diversity, panmixia (Table 3) and differen­
estimates produced mostly relate to the effective number of annual
tiation (Table 4) are based on predefined samples from different
breeders (Nb), rather than to the effective population size per generation
hatchery strains. About 70% of the hatchery stocks were in HWE. Few
(Ne) (cf. Waples et al. 2013). Knowing that the values of Nb and Ne are
hatchery strains showed low polymorphism (St. 5, 6 and 21), most
correlated (Ferchaud et al. 2016), values of Ne can be estimated from
exhibited high polymorphism (yellow and orange cells in Table 3) and

4
P. Berrebi et al. Aquaculture 544 (2021) 737043

Table 3
Different parameters of population genetic diversity, highlighted by heat colors: grey, white, yellow, orange and red cells indicate increasing levels of polymorphism.
Details on the meaning of each parameter are given in the text (Material and methods). (For interpretation of the references to colour in this figure legend, the reader
is referred to the web version of this article.)

one, the Východná sample, was exceptionally polymorphic compared to French ones, the former leaving the latter ones at the next step. This
French hatchery strains considered as highly polymorphic (Bohling et al. second step also clustered east European hatchery strains (Poland, Czech
2016). Allelic richness (30 individuals) followed more or less the level of Republic, Slovakia and Hungary) with the North America (Minnesota)
heterozygosity, being greatest in the St. 13 and St. 22 stocks (8.22 and captive population (St. 26).
8.03). High allelic richness values were also obtained for stocks St. 14
and St. 17. By contrast, the lowest values were observed in the popu­ 3.3. Phylogenetic relationships among strains
lation St. 21 and St. 11 (Table 4). Private alleles are described in Table 3;
their richness was significantly higher in stocks St. 18, St. 20 and St. 25 A bootstrapped NJ phylogenetic tree including the examined 26
than in the remaining sampling. No private alleles were observed in brown trout stocks revealed that they were divided in two main sub­
stocks St. 4 and St. 6. groups of uneven size (Fig. 6). The first subgroup comprised most
samples, probably of Atlantic origin, subdivided into at least three
3.2. Clustering analyses smaller clusters: Swedish, Eastern European and French/Italian stocks.
The second subgroup represented two Mediterranean stocks (St. 18, St.
The FCA diagrams depicted in Figs. 2 to 4 clearly indicated a com­ 21) clearly separated with high bootstrap support from the presumed
mon origin of most European hatchery stocks (i.e. the “black ellipse” of Atlantic populations, and with two Southern European stocks (St. 20,
Figs. 2 to 4), constituting a homogenous group including all countries, France and St. 25, Spain) located in-between.
also USA (Fig. 5). Outside this common group, several local strains
emerged of presumed Mediterranean and Atlantic origin (Figs. 2 and 3). 3.4. Familial structure and effective population size
The hierarchical STRUCTURE analysis presented in Fig. 5 required seven
main steps, based on different subsets of the material. Without neces­ In Table 5 are indicated, for each strain sample, the number of
sarily describing all these steps in detail (see Fig. 5), some important detected fullsib families and the proportion (%) of unrelated vs fullsib
results can be highlighted. The first step separated Mediterranean strains and halfsib dyads. Estimates of current effective population sizes are also
from Atlantic ones, with the exception of the Brantôme hatchery (St. 20) shown. Proportions of unrelated individuals within samples were high,
and the O-Veral hatchery (St. 25), breeding distinct local southern from 68 to 97.4% (mean 93.5%), whereas full-sib proportions varied
Atlantic populations that clustered with Mediterranean strains. The from 0 to 15.4% (1%) and half-sib proportions from 2.6 to 16.6 (5.5%).
second step clustered Swedish hatcheries (St. 1 to 3) with Italian and Finally, the effective size estimates were very heterogeneous, ranging

5
P. Berrebi et al. Aquaculture 544 (2021) 737043

Table 4
Heat-map of pairwise FST comparing each fish farm. All estimates were significantly different from zero at the p < 0.05 level (after Bonferroni correction), with the
only exception of the two Swedish strains from the same river (samples 1 and 2, green cell). Mediterranean strains (St.18 and 21) are especially differentiated from
the rest. (For interpretation of the references to colour in this figure legend, the reader is referred to the web version of this article.)
N° Map Hatchery name 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26
St. 01 Dalälven (VBF river strain) 0 -0,01 0,12 0,15 0,16 0,15 0,08 0,19 0,10 0,14 0,20 0,15 0,13 0,09 0,09 0,07 0,06 0,23 0,13 0,14 0,40 0,07 0,09 0,19 0,18 0,14
St. 02 Dalälven (HBF river strain) 0 0,10 0,13 0,14 0,14 0,07 0,17 0,08 0,12 0,19 0,12 0,12 0,08 0,08 0,06 0,04 0,20 0,11 0,13 0,38 0,06 0,07 0,17 0,15 0,11
St. 03 Gullspang (Lake Vänern strain) 0 0,19 0,18 0,18 0,14 0,23 0,15 0,18 0,25 0,17 0,18 0,16 0,13 0,13 0,11 0,23 0,15 0,17 0,43 0,11 0,15 0,24 0,21 0,16
St. 04 Bartush 0 0,15 0,13 0,10 0,05 0,11 0,14 0,18 0,10 0,09 0,09 0,10 0,13 0,12 0,21 0,17 0,18 0,38 0,11 0,12 0,13 0,21 0,14
St. 05 Rutki 0 0,03 0,09 0,17 0,09 0,16 0,24 0,14 0,18 0,09 0,13 0,14 0,12 0,28 0,16 0,20 0,45 0,15 0,14 0,23 0,24 0,13
St. 06 Dabie 0 0,10 0,16 0,08 0,14 0,23 0,14 0,17 0,08 0,10 0,13 0,11 0,27 0,14 0,20 0,44 0,14 0,14 0,21 0,23 0,12
St. 07 Aquamar 0 0,12 0,04 0,09 0,17 0,11 0,10 0,05 0,06 0,06 0,06 0,21 0,12 0,12 0,37 0,07 0,06 0,14 0,14 0,09
St. 08 Czarci Jar 0 0,13 0,18 0,21 0,13 0,12 0,13 0,14 0,15 0,16 0,21 0,19 0,20 0,39 0,13 0,14 0,18 0,24 0,18
St. 09 Folutsz 0 0,08 0,18 0,11 0,10 0,03 0,05 0,08 0,05 0,20 0,11 0,13 0,38 0,07 0,07 0,13 0,15 0,05
St. 10 Turnov (Austrian strain) 0 0,19 0,15 0,13 0,09 0,07 0,08 0,07 0,23 0,13 0,13 0,40 0,07 0,10 0,14 0,14 0,08
St. 11 Turnov (Italian strain) 0 0,14 0,14 0,17 0,16 0,16 0,15 0,26 0,19 0,21 0,40 0,17 0,18 0,19 0,25 0,20
St. 12 Borová Lada 0 0,05 0,11 0,11 0,13 0,11 0,18 0,16 0,17 0,37 0,12 0,12 0,13 0,19 0,13
St. 13 Východná 0 0,08 0,09 0,11 0,09 0,15 0,15 0,14 0,32 0,09 0,09 0,11 0,17 0,13
St. 14 Slovryb 0 0,04 0,07 0,06 0,20 0,10 0,14 0,37 0,09 0,08 0,14 0,16 0,08
St. 15 Lillafüred 0 0,07 0,06 0,20 0,10 0,12 0,38 0,06 0,08 0,13 0,14 0,07
St. 16 Cantiano 0 0,02 0,19 0,08 0,10 0,36 0,02 0,06 0,14 0,15 0,11
St. 17 Visso 0 0,19 0,07 0,09 0,34 0,01 0,04 0,12 0,13 0,08
St. 18 Roquebillière 0 0,22 0,21 0,35 0,18 0,17 0,22 0,23 0,21
St. 19 Murgat 0 0,15 0,39 0,08 0,10 0,18 0,18 0,13
St. 20 Brantôme (Dronne strain) 0 0,35 0,08 0,12 0,18 0,15 0,15
St. 21 Babeau (Gravezon strain) 0 0,34 0,34 0,38 0,38 0,40
St. 22 Soueich 0 0,04 0,12 0,13 0,10
St. 23 Cauterets 0 0,14 0,14 0,11
St. 24 Baga 0 0,20 0,15
St. 25 O Veral-Lugo 0 0,16
St. 26 Lanesboro 0

Fig. 2. Factorial Correspondence Analysis (FCA) di­


agram presenting the total material (779 brown trout
individuals sampled in 26 locations in Europe and
North America) with most samples concentrated in
the black ellipse (top right), probably representing
strains belonging to the Atlantic lineage, and three
samples outside (likely representing strains from the
Mediterranean lineage). The numbers between pa­
rentheses are those of the first column of Table 1. (For
interpretation of the references to colour in this
figure legend, the reader is referred to the web
version of this article.)

from only 4 to 74 (mean 37.7). This variation may reflect different 4. Discussion
breeding practices (e.g. number of spawners used, proportion of males/
females, etc.). A few hatchery samples deviated from the average values: Population genetic surveys of brown trout frequently describe
the Turnov hatchery (Italian strain; St. 11; Czech Republic) had the admixture between native wild populations and introduced domestic
lowest effective size and proportion of unrelated genotypes (4 and 68% stocks. Such induced hybridization is generally considered as a negative
respectively) and the highest full-sib and half-sib proportions (15.4 and consequence of stocking, reducing the original natural diversity and its
16.6% respectively). At the opposite, the Soueich hatchery sample (St. adaptive value (e.g. Hansen et al. 2000; Poteaux et al. 2000; Kohout
22; France) had the highest effective size and unrelated proportion (74 et al. 2012; Valiquette et al. 2014; Berrebi et al. 2017; Leitwein et al.
and 97.4%) and the lowest fullsib and halfsib proportions (0 and 2.6%). 2018). While analyzing trout from very different natural contexts and
Most other samples showed intermediate values, except Babeau hatch­ disparate geographic regions, most of these surveys refer to a domestic
ery (St. 21, France) close to the Turnov excess, and Isère (St. 19, France) lineage characterized by four main mtDNA haplotypes named 1, 2, 3 and
hatchery close to the Soueich one. 4 or Atcs1 to 4 (Cortey et al. 2004). However, a more complete genetic
description of European domestic stocks based on nuclear markers has
remained missing. By analyzing twelve microsatellites in 26 European

6
P. Berrebi et al. Aquaculture 544 (2021) 737043

Fig. 3. Factorial Correspondence Analysis (FCA) diagram of 683 brown trout individuals sampled in 23 locations in Europe and North America following removal of
three strains of presumed Mediterranean lineage origin (cf. Fig. 2); the “black ellipse” is still dominant, but four samples (11, 12 and 13, and especially 20) are clearly
different, while included in the black ellipse of the Fig. 2. (For interpretation of the references to colour in this figure legend, the reader is referred to the web version
of this article.)

Fig. 4. Factorial Correspondence Analysis (FCA) diagram of 542 brown trout individuals sampled in 19 locations in Europe and North America following removal of
three strains of presumed Mediterranean lineage origin as well as 3 + 4 deviating samples identified in Figs. 2 and 3; the “black ellipse” has been further subdivided,
with most samples in the middle and few of them at the border. (For interpretation of the references to colour in this figure legend, the reader is referred to the web
version of this article.)

(with one American) domestic stocks, the present survey confirmed the cases (e.g. the Atlantic St. 20 pooled with Mediterranean strains ac­
domination of an “international strain” of north Atlantic origin, and has cording to the Bayesian assignment analysis). More fundamental dif­
revealed high diversity within hatcheries. ferences emerged between multidimensional analysis and the other two
To understand the history of hatchery establishments and the ex­ methods. In Figs. 2 to 4, the mass of Atlantic samples are not structured
change of genes between hatcheries, we used three main statistical and, in particular, an east-west structure does not appear. Such differ­
methods for clustering our samples (i.e. multidimensional FCA, Bayesian ences can be explained by a less discriminating power of the FCA (east-
assignment and a phylogenetic NJ-tree). Common main structures west structure) or by strongly different methods of discrimination
among these analyses were (1) a clear subdivision between Atlantic and (search for subgroups approaching populations in equilibrium HW and
Mediterranean lineage strains, and (2) further subdivision of the Atlantic LD for the Bayesian assignment vs. genetic distances calculated pairwise
lineage into three groups including the well-separated Swedish samples, between samples and used for the NJ tree).
hatcheries in eastern Europe (Poland, Slovakia, Czech Republic), less One main result is that all-around Europe, in countries with different
strongly separated from those in the west (France and Italy). traditions and histories (north-south for climate, east-west for the recent
However, although results generally agreed between methods, some political regimes), genetically similar commercial strains are bred (i.e.
discrepancies also appeared. Most of these differences concerned special the black ellipse in Figs. 2 to 4, the main central composite cluster in

7
P. Berrebi et al. Aquaculture 544 (2021) 737043

trout as previously proposed by Tortonese (1970) and Giuffra et al.


(1996).
On the other hand, there are also signs of differentiation, even among
the larger group of samples with presumed common Atlantic origin.
Pairwise FST estimates are highly significant in all cases (Table 4), which
indicates genetic drift coupled with no recent exchanges of genetic
materials among the hatcheries. Local stocks may have been “enriched”
by local wild fish or by exchanges between hatcheries. For example, in
Poland, most hatcheries breed their own stock, without or with only
occasional inclusion of wild fish. However, it has been observed that
some river managers stock their own waters with imported fish, and
after a few seasons they may capture adult specimens for broodstock
renovation. In Poland this process occurs on a small scale (Bernaś and
Wąs-Barcz 2020) and a clear genetic separation between Polish northern
and southern strains has been detected (Figs. 5 and 6). The situation in
Poland illustrates how some mixing of imported domestic strains with
local wild populations may proceed, and how hatchery stocks from
different geographic areas may still be genetically similar. Another
example is given by Kohout et al. (2012) showing that the Východná and
Slovryb hatcheries (here St. 13 and 14), settled in the Danube basin,
were composed of only 1/4 to 1/3 of Danubian lineage, admixed with
the Atlantic one. The exchange of breeding material between Czech,
Slovak and Polish breeding centers has and had a long tradition, strongly
related to the history of this region. Geographically, this area is drained
by the upper catchment areas of the Elbe, Vistula and Danube rivers.
Therefore, the presence of mixed lines in this area is largely the result of
human activity (Kohout et al. 2012; Bernaś and Wąs-Barcz 2020).
The second main result concerns levels of polymorphism within
strains. High genetic variability at the intra-population level and low
differentiation at the inter-population level are quite common within
domestic trout stocks (Bohling et al. 2016). While it is frequently
claimed that with time, due to management leading to consanguinity (i.
e. low effective population sizes at hatcheries), captive populations
should lose genetic diversity (Li et al. 2004), the present hatchery strains
show generally high levels of genetic variability. The level of poly­
Fig. 5. Representation of the sequential assignment analysis of domestic stocks morphism is similar to that of French hatcheries investigated earlier,
of brown trout (Salmo trutta) using STRUCTURE software. The sequential considered to be highly polymorphic in comparison with wild pop­
analysis on separate parts of the total material was continued until no more ulations and displaying high values at all measured parameters (0.67
subdivision could be found (indicated by a red cross). The first column repre­ \Hnb\0.78 and 5.7\A\8.3, Berrebi et al. 2000; Bohling et al. 2016). In
sents the first partition (ΔK method) for K = 2: mainly Atlantic and Mediter­ the present survey, 85% of the strains showed high (0.61 < Hnb < 0.70)
ranean lineages. The second column shows results for the partition of the
to very high (0.71 < Hnb < 0.81) heterozygosity (Table 3). This high
Atlantic lineage (for K = 2) into two major groups (mainly west and north vs.
level of genetic diversity is potentially driven by past admixture (gene
east Europe) and the independent partition of the Mediterranean lineage into
four separate population clusters (for K = 4). The colors within each column
flow) mediated by imports of fish from other hatcheries (Chevassus et al.
have no sense and were chosen at random. The sample numbers are those of the 1992). As discussed above, fish exchanges between trout farms is a
first column of Table 1. (For interpretation of the references to colour in this frequent practice in Europe (van Houdt et al. 2005). Therefore, this
figure legend, the reader is referred to the web version of this article.) practice has most likely created a high genetic variation in the main
domestic strain as a whole, and at the same time reduced regional and
Fig. 6), suggesting a common origin. Most likely this dominating cluster local differentiation among local hatchery derivatives of this common
corresponds to what Bohling et al. (2016) have referred to as ComATL strain.
(Common Atlantic strains). One possible reason for this widespread Low estimates of effective size (<30) were observed in eight strains,
genetic similarity is inter-hatchery trade, mainly of domestic commer­ indicating that the level of inbreeding is probably high. The associated
cial Atlantic strains whose success is primarily due to high hatching rates parentage analysis also confirms this interpretation, since a clear family
(close to 90–100%), even using standard facilities (Chevassus et al. structure could be observed in more than half the strains (Table 5).
1992, 2004; Czerniawski et al. 2010). The third main result showed that, despite an apparent common
The scale of these exchanges seems to be Europe-wide. For example, nature of domestic trout all around Europe, what may be called “local
the phylogenetic position of the Spanish Baga strain (St. 24) which strains” (i.e. domesticated from local wild fish) are not rare. In Sweden,
clustered with samples from the Czech Republic (Fig. 6) is symptomatic. the three samples included (St. 1 to 3) are all from hatchery strains
Another demonstration of exchanges is the Turnov hatchery strains (in known to be recently derived from local natural populations (Palm et al.
Czech Republic) originating from Italy and Austria. Furthermore, a ge­ 2003, 2012). In France, the St. 18 (Doubs River flowing to the Medi­
netic similitude was observed among French and Italian stocks (Fig. 5 terranean Rhône River), St. 21 (from Gravezon small river flowing to
and partly Fig. 6). This closeness has been already suggested by Caputo Mediterranean Orb River, near Montpellier, strain created with only 15
et al. (2004) who detected in Italian domestic strains several mtDNA genitors 15 years ago) and St. 20 (Dronne Atlantic River flowing to the
haplotypes identical to Danish haplotypes described by Hansen and Dordogne-Garonne) are also local strains. In Spain, the O Veral strain
Loeschcke (1996). We concluded that this observation was in accor­ (St. 25) is clearly of local origin, according to the present results, so of
dance with the Danish (or Scandinavian) origin of domestic Atlantic Atlantic lineage. All of these strains were found to have high numbers of
private alleles.

8
P. Berrebi et al. Aquaculture 544 (2021) 737043

Fig. 6. Bootstrapped neighbor joining (NJ) construction revealing two uneven independent subgroups. The first one, composed of several imbricated subgroups,
concentrates 24 samples belonging to the Atlantic lineage. The second includes only two Mediterranean stocks (St. 18, St. 21, in purple) clearly separated with high
bootstrap support from Atlantic samples. (For interpretation of the references to colour in this figure legend, the reader is referred to the web version of this article.)

Considering the results of this study, several collective recommen­ (Oncorhynchus kisutch) has been shown to have this type of difference
dations for local authorities managing breeding programs could be between the wild form and a genetically identical domestic line. The
made. In closed breeding stocks, it is essential to keep number of differently methylated regions affect biological functions linked to the
effective breeders (Nb) high. This is the most important because it capacity of the young stages to migrate towards the ocean (Le Luyer
prevents the effects of inbreeding and genetic drift (Tave 1999; Bentsen et al. 2017). Epigenetic modifications induced by the captive environ­
and Olesen 2002). In the case of open farms, engaged in local breeding ment are therefore a potential explanation for reductions in fitness of
programs, renewing the stock, they should be based on local fish, pref­ stocked populations.
erably from the same river. This will keep the genetic distance between Another potential impact of domestic trout releases into the natural
breeding and wild fish at a low level. Above all, fishing users should stick environment relates to deleterious mutations. In a study on conse­
to the principle of not mixing the population. In addition, some breeding quences of stocking based on nearly 5000 SNPs, supplemented lake trout
farms also conduct a directional selection of the desired traits, e.g. populations were found to have fewer deleterious alleles than unstocked
phenotypic or related to rapid growth (e.g. Sanchez et al. 2001; Mam­ ones, despite an overall increase in neutral genetic diversity in the
brini et al. 2004). However, we must be aware of the consequences of former group (Ferchaud et al. 2018). An earlier study had also revealed a
releasing such fish, especially given the possibility of non-directional decrease in genetic integrity of the same stocked lake trout populations
selection with unknown effects on the population (O'Sullivan et al. as they regrouped in assignment analyzes according to their common
2019). stocking source (Valiquette et al. 2014). These studies suggest that
This study highlights the need for caution before stocking. As elevated gene flow due to stocking may assist with the purging of
observed herein (Table 3) and by Valiquette et al. (2014) in their study deleterious mutations in small local populations, where genetic drift
of lake trout, stocking may induce locally increased levels of genetic may otherwise overrule natural selection. At the same time, however, it
polymorphism that is simultaneously associated with loss of inter- is important to consider the source of the stocked fish in order to pre­
population diversity and of local adaptation. The main criticism of serve local adaptations and genetic integrity of recipient populations.
stocking with domestic trout is the loss of fitness in the recipient pop­ From a biogeographic and conservation point of view, stocking
ulation (Feulner et al. 2013; Miller et al. 2004; Wollebaek et al. 2012; without obvious necessity can lead to a reduction in the overall genetic
Hansen et al. 2009). Artificial selection increasing survival in fish cul­ diversity of brown trout on the European scale. Even when a local strain
ture but not in the wild, has often been cited as the most likely expla­ is created with broodstock from nearby rivers, it is advisable to release
nation for lower fitness. An alternative hypothesis, however, concerns domestic trout only in the area inhabited by genetically similar pop­
DNA methylation (epigenetic modifications). Coho salmon ulations. Note that such locally derived stocks may still become quickly

9
P. Berrebi et al. Aquaculture 544 (2021) 737043

Table 5
Results from the sibship assignment method adopted by COLONY (Wang 2004) at one typing error rate (0.001). Non default COLONY job settings as follows: typing
error rate 0.001, mating system I: male and female polygamy, mating system II: with inbreeding, length of run: medium, analysis method: full-likelihood, rest of setting
as default. Fullsib and halfsib dyad are displayed both as absolute value and relative frequency (between brackets).
N◦ Map Hatchery name Fullsib families Unrelated dyads Fullsib dyads Halfsib dyads Ne (CI95 L - U)

St. 01 Dalälven (VBF river strain) 0 303 (93.5%) 0 21 (6.5%) 29 (17–61)


St. 02 Dalälven (HBF river strain) 0 246 (96.1%) 0 10 (3.9%) 48 (24–143)
St. 03 Gullspang (Lake Vänern strain) 2 1033 (94.9%) 1 (0.1%) 55 (5.1%) 37 (23–62)
St. 04 Bartush 1 495 (93.6%) 1 (0.2%) 33 (6.2%) 29 (17–55)
St. 05 Rutki 0 1501 (93.8%) 0 99 (6.2%) 32 (19–55)
St. 06 Dabie 1 764 (90.8%) 1 (0.1%) 76 (9.0%) 21 (11–39)
St. 07 Aquamar 1 862 (95.7%) 1 (0.1%) 38 (4.2%) 44 (27–83)
St. 08 Czarci Jar 0 584 (93.4%) 0 41 (6.6%) 29 (16–52)
St. 09 Folutsz 1 502 (94.9%) 1 (0.2%) 26 (4.9%) 36 (22–65)
St. 10 Turnov (Austrian strain) 1 694 (95.2%) 1 (0.1%) 34 (4.7%) 39 (22–70)
St. 11 Turnov (Italian strain) 3 496 (68.0%) 112 (15.4%) 121 (16.6%) 4 (2− 12)
St. 12 Borová Lada 0 750 (95.7%) 0 34 (4.3%) 44 (26–76)
St. 13 Východná 3 749 (95.5%) 5 (0.6%) 30 (3.8%) 38 (24–68)
St. 14 Slovryb 0 556 (96.5%) 0 20 (3.5%) 55 (33–107)
St. 15 Lillafüred 5 857 (95.2%) 7 (0.8%) 36 (4.0%) 35 (21–84)
St. 16 Cantiano 3 1734 (93.8%) 32 (1.7%) 83 (4.5%) 25 (15–44)
St. 17 Visso 0 1549 (96.8%) 0 51 (3.2%) 38 (21–79)
St. 18 Roquebillière 3 790 (93.9%) 3 (0.3%) 48 (5.7%) 30 (18–54)
St. 19 Isère 1 847 (94.1%) 1 (0.1%) 52 (5.8%) 61 (39–95)
St. 20 Brantôme (Dronne strain) 1 3021 (96.3%) 3 (0.1%) 112 (3.6%) 52 (34–79)
St. 21 Babeau (Gravezon strain) 3 1352 (84.5%) 53 (3.3%) 195 (12.2%) 10 (5–24)
St. 22 Soueich 0 819 (97.4%) 0 22 (2.6%) 74 (44–147)
St. 23 Cauterets 0 986 (96.3%) 0 38 (3.7%) 52 (33–89)
St. 24 Baga 3 579 (92.6%) 12 (1.9%) 34 (5.4%) 21 (11–42)
St. 25 O Veral-Lugo 0 702 (96.3%) 0 27 (3.7%) 50 (30–94)
St. 26 Lanesboro 0 930 (96.8%) 0 31 (3.2%) 45 (26–86)

differentiated from wild recipient populations through drift and derived from “cosmopolitan” fish of presumed Atlantic lineage origin
domestication. In general, stocking should thus be avoided if there are (ComATL) in most investigated European countries (possibly except
not good reasons, e.g. to assist recovery of local stocks after their main Sweden, although analysis of additional samples from Northern Euro­
problems have been remedied. Stocking into areas with naturally pean hatcheries appears necessary), and in North America where the
reproducing populations is generally questionable (Fernández-Cebrián species has been introduced from Europe. This huge nebula of geneti­
et al. 2014), due to the biological risks associated with such stocking cally similar strains, used for restocking at the continental level,
aimed at supporting “put and take” fisheries (with bycatches of wild necessarily reduces the natural diversity of the species. However, we
trout, genetic introgression, etc.). In Britain, where self-reproducing have also observed several hatcheries breeding local strains from
salmonids population persist, only stocking with triploid female in­ restricted areas with seemingly local characteristics, capable of limiting
dividuals is permitted, on the other hand, supportive breeding with local this loss of overall diversity if locally used for stocking.
strains is allowed, but rarely done (Andrew Ferguson personal
communication). Still, in the Mediterranean countries (mainly Italy or Declaration of Competing Interest
France), stocking with local strains is often interpreted as a simple
replacement of the commercial brown Atlantic trout by the so-called The following authors declare no conflict of interest.
“Mediterranean trout” without worrying about the numerous diffi­
culties affecting the management of wild biodiversity (Laikre et al. Acknowledgements
2020). There is a risk of simply substituting wild populations by
domesticated ones, without any advantage in conservation and biodi­ The authors are grateful to numerous colleagues who have provided
versity of freshwater fish. hatchery fin clips for analyses: Jan Kohout, Liběchov, Czech Republic;
The results obtained in this work should be developed in further Vincenzo Caputo Barucchi, Ancona, Italy; Eric Ravel, Octon, France;
studies using new technologies and allowing for a deeper analysis of the Francis Dauba, Castanet-Tolosan, France; José-Luis García-Marín, Ger­
observed processes. Particularly promising and interesting would be the ona, Spain; Carmen Bouza, Lugo, Spain; Douglas Dieterman, Lake City,
analysis of selection signals occurring in breeding stocks and by Minnesota, USA. Sampling in Hungary was supported by the Ministry of
checking their generality. Recent studies in genomics using medium and Innovation and Technology within the framework of the Thematic
high-dense SNP microarray methods have shown the ability to detect Excellence Programme 2020, Institutional Excellence Subprogramme
such phenomena (e.g. Gutierrez et al. 2016; Linløkken et al. 2017) and (TKP2020-IKA-12). In Poland work was supported by project 2016/21/
indicate that other selection processes may take place in breeding stocks D/NZ9/00405 of the Ministry of Science and Higher Education and
and in wild populations (Bernaś et al. 2020). In addition to SNP statutory topic S030/005 in the IFI Olsztyn. The authors also thank José-
microarrays, RAD-Seq technology should be applied as a tool for rapid Luis García-Marín (Universitat de Girona) for reviewing an earlier
and efficient generation of QTL-targeted and genome-wide marker data version of the text.
(Houston et al. 2012). The high efficiency of methods based on RAD
sequencing in population differentiation may reveal further patterns of Appendix A. Supplementary data
genetic differentiation, in addition to those identified herein, between
brown trout breeding populations from different parts of the species' Supplementary data to this article can be found online at https://doi.
range (cf. Bradbury et al. 2015; Lemopoulos et al. 2019). org/10.1016/j.aquaculture.2021.737043.
In conclusion, the most important finding is that we observed a
common hatchery strain (slightly diversified due to local practices)

10
P. Berrebi et al. Aquaculture 544 (2021) 737043

References Dieterman, D.J., Hoxmeier, R.J.H., Staples, D.F., 2012. Factors influencing growth of
individual brown trout in three streams of the upper Midwestern United States. Ecol.
Freshw. Fish 21, 483–493.
Almódovar, A., Nicola, G.G., Elvira, B., Garcia-Marin, J.-L., 2006. Introgression
Doadrio, I., Perea, S., Yahyaoui, A., 2015. Two new species of Atlantic trout
variability among Iberian brown trout evolutionary significant units: the influence of
(Actinopterygii, Salmonidae) from Morocco. Graellsia 71, e031.
local management and environmental features. Freshw. Biol. 51, 1175–1187.
Earl, D.A., von Holdt, B.M., 2012. STRUCTURE HARVESTER: a website and program for
Angers, B., Bernatchez, L., Angers, A., Desgroseillers, L., 1995. Specific microsatellite loci
visualizing STRUCTURE output and implementing the Evanno method. Conserv.
for brook charr reveal strong population subdivision on a microgeographic scale.
Genet. Resour. 4, 359–361.
J. Fish Biol. 47, 177–185.
Elliott, J.M., 1989. Wild brown trout Salmo trutta: an important national and
Antunes, A., Apostolidis, A., Berrebi, P., Duguid, A., Ferguson, A., García-Marín, J.-L.,
international resource. Freshw. Biol. 21, 1–5.
Guyomard, R., Hansen, M.M., Hindar, K., Koljonen, M.L., Laikre, L., Largiadèr, C.,
Estoup, A., Largiader, C.R., Perrot, E., Chourrout, D., 1996. Rapid one-tube DNA
Martínez, P., Nielsen, E.E., Palm, S., Ruzzante, D., Ryman, N., Triantaphyllidis, C.,
extraction for reliable PCR detection of fish polymorphic markers and transgenes.
1999. Conservation genetic management of brown trout (Salmo trutta) in Europe.
Mol. Mar. Biol. Biotechnol. 5, 295–298.
Concerted Action on Identification, Management and Exploitation of Genetic
Evanno, G., Regnaut, S., Goudet, J., 2005. Detecting the number of clusters of individuals
Resources in the Brown Trout (Salmo trutta) (“TROUTCONCERT”; EU FAIR CT97-
using the software STRUCTURE: a simulation study. Mol. Ecol. 14, 2611–2620.
3882). Laikre, L. ed. 91 pp.
Ferchaud, A.-L., Perrier, C., April, J., Hernandez, C., Dionne, M., Bernatchez, L., 2016.
Aparicio, E., Garcia-Berthou, E., Araguas, R.M., Martinez, P., Garcia-Marin, J.-L., 2005.
Making sense of the relationships between Ne, Nb and Nc towards defining
Body pigmentation pattern to assess introgression by hatchery stocks in native Salmo
conservation thresholds in Atlantic salmon (Salmo salar). Heredity 117, 268–278.
trutta Mediterranean streams. J. Fish Biol. 67, 931–949.
Ferchaud, A.-L., Laporte, M., Perrier, C., Bernatchez, L., 2018. Impact of supplementation
Apostolidis, A., Karakousis, Y., Triantaphyllidis, C., 1996. Genetic divergence and
on deleterious mutation distribution in an exploited salmonid. Evol. Appl. 11,
phylogenetic relationships among Salmo trutta L. (brown trout) populations from
1053–1065.
Greece and other European countries. Heredity 76, 551–560.
Fernández-Cebrián, R., Araguas, R.M., García-Marín, J.-L., 2014. Genetic risks of
Bardakci, F., Degerli, N., Ozdemir, O., Basibuyuk, H.H., 2006. Phylogeography of the
supplementing trout populations with native stocks: a simulation case study from
Turkish brown trout Salmo trutta L.: mitochondrial DNA PCR-RFLP variation. J. Fish
current Pyrenean populations. Can. J. Fish. Aquat. Sci. 71, 1243–1255.
Biol. 68, 36–55.
Feulner, P.G.D., Gratten, J., Kijas, J.W., Visscher, P.M., Pemberton, J.M., Slate, J., 2013.
Barton, N.H., Slatkin, M., 1986. A quasi-equilibrium theory of the distribution of rare
Introgression and the fate of domesticated genes in a wild mammal population. Mol.
alleles in a subdivided population. Heredity. 56, 409–415.
Ecol. 22, 4210–4221.
Belkhir, K., Borsa, P., Goudet, J., Bonhomme, F., 2004. GENETIX 4.05: logiciel sous
Foulley, J.L., Ollivier, L., 2006. Estimating allelic richness and its diversity. Livest. Sci.
Windows pour la génétique des populations. Laboratoire Génome et Population,
101, 150–158.
CNRS-UPR, Université de Montpellier II, Montpellier, France.
French, W.E., Vondracek, B., Ferrington Jr., J.C., Finlay, J.C., Dieterman, D.J., 2016.
Bentsen, H.B., Olesen, I., 2002. Designing aquaculture mass selection programs to avoid
Winter diet of brown trout Salmo trutta in groundwater-dominated streams: influence
high inbreeding rates. Aquaculture 204, 349–359.
of environmental factors on spatial and temporal variation. J. Fish Biol. 89,
Benzécri, J.P., 1973. L'analyse des données. Dunod, Paris.
2449–2464.
Bernaś, R., Wąs-Barcz, A., 2020. Genetic structure of important resident brown trout
Fruciano, C., Pappalardo, A.M., Tigano, C., Ferrito, V., 2014. Phylogeographical
breeding lines in Poland. J. Appl. Genet. 61, 239–247.
relationships of Sicilian brown trout and the effects of genetic introgression on
Bernaś, R., Poćwierz-Kotus, A., Árnyasi, M., Kent, M.P., Lien, S., Wenne, R., 2020.
morphospace occupation. Biol. J. Linn. Soc. 112, 387–398.
Genetic differentiation in hatchery and stocked populations of sea trout in the
Giuffra, E., Forneris, G., Guyomard, R., 1996. Diversita genetica e filogenesi dei
southern Baltic: selection evidence at SNP loci. Genes 11, 184.
salmonidi del bacino del Po. In: Proceedings of the Fourth AIIAD National Meeting,
Bernatchez, L., Guyomard, R., Bonhomme, F., 1992. DNA sequence variation of the
Riva del Garda. TEMI, Trento, pp. 21–32.
mitochondrial control region among geographically and morphologically remote
Gjedrem, T., 2000. Genetic improvement of cold-water fish species. Aquac. Res. 31,
European brown trout Salmo trutta populations. Mol. Ecol. 1, 161–173.
25–33.
Berrebi, P., Poteaux, C., Fissier, M., Cattaneo-Berrebi, G., 2000. Stocking impact and
Gutierrez, A.P., Yáñez, J.M., Davidson, W.S., 2016. Evidence of recent signatures of
allozyme diversity in brown trout from Mediterranean southern France. J. Fish Biol.
selection during domestication in an Atlantic salmon population. Mar. Genomics 26,
56, 949–960.
41–50.
Berrebi, P., Jesenšek, D., Crivelli, A.J., 2017. Natural and domestic introgressions in the
Hansen, M.M., Loeschcke, V., 1994. Effect of releasing hatchery-reared brown trout to
marble trout population of Soca River (Slovenia). Hydrobiologia 785, 277–291.
wild trout populations. Conserv. Genet. 68, 273–289.
Berrebi, P., Caputo Barrucchi, V., Splendiani, A., Muracciole, S., Sabatini, S., Palmas, F.,
Hansen, M.M., Loeschcke, V., 1996. Genetic differentiation among Danish brown trout
Tougard, C., Arculeo, M., Marić, S., 2019. Brown trout (Salmo trutta L.) high genetic
populations, as detected by RFLP analysis of PCR amplified mitochondrial DNA
diversity around Tyrrhenian Sea as revealed by nuclear and mitochondrial markers.
segments. J. Fish Biol. 48, 422–436.
Hydrobiologia 826, 209–231.
Hansen, M.M., Ruzzante, D.E., Nielsen, E.E., Mensberg, K.-L.D., 2000. Microsatellite and
Berrebi, P., Marić, S., Snoj, A., Hasegawa, K., 2020. Brown trout in Japan - introduction
mitochondrial DNA polymorphism reveals life-history dependent interbreeding
history, distribution and genetic structure. Knowl. Manag. Aquat. Ec. 421, 1–16.
between hatchery and wild brown trout (Salmo trutta L.). Mol. Ecol. 9, 583–594.
Bohling, J., Haffray, P., Berrebi, P., 2016. Genetic diversity and population structure of
Hansen, M.M., Fraser, D.J., Meier, K., Mensberg, K.-L.D., 2009. Sixty years of
domestic brown trout (Salmo trutta) in France. Aquaculture 462, 1–9.
anthropogenic pressure: a spatio-temporal genetic analysis of brown trout
Bradbury, I.R., Hamilton, L.C., Dempson, B., Robertson, M.J., Bourret, V., Bernatchez, L.,
populations subject to stocking and population declines. Mol. Ecol. 18, 2549–2562.
Verspoor, E., 2015. Transatlantic secondary contact in Atlantic salmon, comparing
Holm, L.E., Bendixen, C., 2000. Oncorhynchus mykiss Clone TAA72-13, Sequence Tagged
microsatellites, a single nucleotide polymorphism array and restriction-site
Site [available on internet at http://www.ncbi.nlm.nih.gov/genbank/]. Accession
associated DNA sequencing for the resolution of complex spatial structure. Mol. Ecol.
number AF239038.
24, 5130–5144.
Houston, R.D., Davey, J.W., Bishop, S.C., Lowe, N.R., Mota-Velasco, J.C., Hamilton, A.,
Caputo, V., Giovannotti, M., Nisi Cerioni, P., Caniglia, M.L., Splendiani, A., 2004. Genetic
Guy, D.R., Tinch, A.E., Thomson, M.L., Blaxter, M.L., Gharbi, K., Bron, J.E.,
diversity of brown trout in central Italy. J. Fish Biol. 65, 403–418.
Taggart, J.B., 2012. Characterisation of QTL-linked and genome-wide restriction
Charles, K., Guyomard, R., Hoyheim, B., Ombredane, D., Baglinière, J.-L., 2005. Lack of
site-associated DNA (RAD) markers in farmed Atlantic salmon. BMC Genomics 13,
genetic differentiation between anadromous and resident sympatric brown trout
244.
(Salmo trutta) in a Normandy population. Aquat. Living Resour. 18, 65–69.
Hoxmeier, R.J.H., Dieterman, D.J., 2013. Seasonal movement, growth and survival of
Chevassus, B., Krieg, F., Guyomard, R., Blanc, J., Quillet, E., 1992. The genetics of the
brook trout in sympatry with brown trout in Midwestern US streams. Ecol. Freshw.
brown trout: twenty years of French research. Icel. Agric. Sci. 6, 109–124.
Fish 22, 530–542.
Chevassus, B., Quillet, E., Krieg, F., Hollebecq, M.-G., Mambrini, M., Fauré, A., Labbé, L.,
ICES, 2019. Baltic salmon and trout assessment working group (WGBAST). ICES Sci. Rep.
Hiseux, J.-P., Vandeputte, M., 2004. Enhanced individual selection for selecting fast
1 (23) https://doi.org/10.17895/ices.pub.4979, 312 pp.
growing fish: the “PROSPER” method, with application on brown trout (Salmo trutta
Jones, O.R., Wang, J.I., 2010. COLONY: a program for parentage and sibship inference
fario). Genet. Sel. Evol. 36, 643–661.
from multilocus genotype data. Mol. Ecol. Resour. 10, 551–555.
Cortey, M., Pla, C., Garcia-Marin, J.-L., 2004. Historical biogeography of Mediterannean
Kalinowski, S.T., 2005. HP-rare: a computer program for performing rarefaction on
Trout. The role of allopatry and dispersal events. Mol. Phylogenet. Evol. 33,
measures of allelic diversity. Mol. Ecol. Notes 5, 187–189.
831–844.
Klütsch, C.F.C., Maduna, S.N., Polikarpova, N., Forfang, K., Aspholm, P.E., Nyman, T.,
Czerniawski, R., Pilecka-Rapacz, M., Domagała, J., 2010. Growth and survival of brown
Eiken, H.G., Amundsen, P.A., Hagen, S.B., 2019. Genetic changes caused by
trout fry (Salmo trutta M. fario L.) in the wild, reared in the hatchery on different
restocking and hydroelectric dams in demographically bottlenecked brown trout in a
feed. Electron. J. Pol. Agricult. Univ. (EJPAU) 13 (2). #04 (Available Online: http
transnational subarctic riverine system. Ecol. Evol. 9 (10), 6068–6081.
://www.ejpau.media.pl/volume13/issue2/art-04.html.
Kohout, J., Jaskova, I., Papousek, I., Sediva, A., Slechta, V., 2012. Effects of stocking on
Delling, B., 2010. Diversity of western and southern Balkan trouts, with the description
the genetic structure of brown trout, Salmo trutta, in Central Europe inferred from
of a new species from the Louros River, Greece (Teleostei: Salmonidae). Ichthyol.
mitochondrial and nuclear DNA markers. Fish. Manag. Ecol. 19, 252–263.
Explor. Fres. 21, 331–344.
Kottelat, M., Freyhof, J., 2007. Handbook of European Freshwater Fishes. Publications
Delling, B., Doadrio, I., 2005. Systematics of the trouts endemic to Moroccan lakes, with
Kottelat, Cornol, Switzerland.
description of a new species (Teleostei: Salmonidae). Ichthyol. Explor. Fres. 16,
Laikre, L., Hoban, S., Bruford, M.W., Segelbacher, G., Allendorf, F.W., Gajardo, G.,
49–64.
González Rodríguez, A., Hedrick, P.W., Heuertz, M., Hohenlohe, P.A., Jaffé, R.,
Delling, B., Crivelli, A., Rubin, J.-F., Berrebi, P., 2000. Morphological variation in hybrids
Johannesson, K., Liggins, L., MacDonald, A.J., OrozcoterWengel, P., Reusch, T.B.H.,
between Salmo marmoratus and alien Salmo species in the Volarja stream, Soca River
Rodríguez-Correa, H., Russo, I.-R.M., Ryman, N., Vernesi, C., 2020. Post-2020 goals
basin, Slovenia. J. Fish Biol. 57, 1199–1212.
overlook genetic diversity. Science 367, 1083–1085.

11
P. Berrebi et al. Aquaculture 544 (2021) 737043

Largiadèr, C.R., Scholl, A., 1995. Effects of stocking on the genetic diversity of brown Sanchez, M., Chevassus, B., Labbé, L., Quillet, E., Mambrini, M., 2001. Selection for
trout populations of the Adriatic and Danubian drainages in Switzerland. J. Fish Biol. growth of brown trout (Salmo trutta) affects feed intake but not feed efficiency.
47, 209–225. Aquat. Living Resour. 14, 41–48.
Le Luyer, J., Laporte, M., Beacham, T., Kaukinen, K., Whithler, R., Leong, J.S., Sanz, N., 2018. Phylogeographic history of brown trout: a review. In: Lobon-Cervia, J.,
Rondeau, E.B., Koop, B., Bernatchez, L., 2017. Parallel epigenetic modifications Sanz, N. (Eds.), Brown Trout: Biology, Ecology and Management. John Wiley & Sons
induced by hatchery rearing in a Pacific Salmon. P. Natl. Acad. Sci. Biol. 114, Ltd., pp. 17–63
12964–12969. Scribner, K.T., Gust, J.R., Fields, R.L., 1996. Isolation and characterization of novel
Leitwein, M., Gagnaire, P.-A., Desmarais, D., Berrebi, P., Guinand, B., 2018. Genomic salmon microsatellite loci: cross-species amplification and population genetic
consequences of a recent three-way admixture in supplemented wild brown trout applications. Can. J. Fish. Aquat. Sci. 53, 833–841.
populations revealed by ancestry tracts. Mol. Ecol. 27, 3466–3483. She, J.X., Autem, M., Kotoulas, G., Pasteur, N., Bonhomme, F., 1987. Multivariate
Lemopoulos, A., Prokkola, J.M., Uusi-Heikkilä, S., Vasemägi, A., Huusko, A., analysis of genetic exchanges between Solea aegyptiaca and Solea senegalensis
Hyvärinen, P., Koljonen, M.-L., Koskiniemi, J., Vainikka, A., 2019. Comparing (Teleosts, Soleidae). Biol. J. Linn. Soc. 32, 357–371.
RADseq and microsatellites for estimating genetic diversity and relatedness Skaala, Ø., Høyheim, B., Glover, K., Dahle, G., 2004. Microsatellite analysis in
–implications for brown trout conservation. Ecol. Evol. 9 (4), 2106–2120. domesticated and wild Atlantic salmon (Salmo salar L.): allelic diversity and
Lerceteau-Köhler, E., Schliewen, U., Kopun, T., Weiss, S., 2013. Genetic variation in identification of individuals. Aquaculture 240, 131–143.
brown trout Salmo trutta across the Danube, Rhine, and Elbe headwaters: a failure of Slatkin, M., 1985. Rare alleles as indicators of gene flow. Evolution. 39, 53–65.
the phylogeographic paradigm? BMC Evol. Biol. 13, 176. Slettan, A., Olsaker, I., Lie, Ø., 1995. Atlantic salmon, Salmo salar, microsatellites at the
Li, Q., Park, C., Endo, T., Kijima, A., 2004. Loss of genetic variation at microsatellite loci SSOSL25, SSOSL85, SSOSL311, SSOSL417 loci. Anim. Genet. 26, 277–285.
in hatchery strains of the Pacific abalone (Haliotis discus hannai). Aquaculture 235, Slettan, A., Olsaker, I., Lie, Ø., 1996. Polymorphic Atlantic salmon, Salmo salar L.,
207–222. microsatellites at the SSOSL438, SSOSL439 and SSOSL444 loci. Anim. Genet. 27,
Linløkken, A.N., Haugen, T.O., Kent, M.P., Lien, S., 2017. Genetic differences between 57–64.
wild and hatchery-bred brown trout (Salmo trutta L.) in single nucleotide Suarez, J., Bautista, J.M., Almodovar, A., Machordom, A., 2001. Evolution of the
polymorphisms linked to selective traits. Ecol. Evol. 7, 4963–4972. mitochondrial control region in Palaearctic brown trout (Salmo trutta) populations:
Mambrini, M., Medale, F., Sanchez, M.P., Recalde, B., Chevassus, B., Labbe, L., the biogeographical role of the Iberian Peninsula. Heredity 87, 198–206.
Quillet, E., Boujard, T., 2004. Selection for growth in brown trout increases feed Takezaki, N., Nei, M., Tamura, K., 2010. POPTREE2: software for constructing
intake capacity without affecting maintenance and growth requirements. J. Anim. population trees from allele frequency data and computing other population
Sci. 82, 2865–2875. statistics with windows interface. Mol. Biol. Evol. 27, 747–752.
Marić, S., Sušnik Bajec, S., Schöffmann, J., Kostov, V., Snoj, A., 2017. Phylogeography of Tave, D., 1999. Inbreeding and brood stock management. In: FAO Fisheries Technical
stream-dwelling trout in the Republic of Macedonia and a molecular genetic basis for Paper 392. FAO, Rome, 122 pp.
revision of the taxonomy proposed by S. Karaman. Hydrobiologia 785, 249–260. Tortonese, E., 1970. Osteichthyes, parte I. Fauna d’Italia, Vol. X. Calderini, Bologna.
Miller, L.M., Close, T., Kapuscinski, R., 2004. Lower fitness of hatchery and hybrid Tougard, C., Justy, F., Guinand, B., Douzery, E.J.P., Berrebi, P., 2018. Salmo macrostigma
rainbow trout compared to naturalized populations in Lake Superior tributaries. Mol. (Teleostei, Salmonidae): nothing more than a brown trout (S. trutta) lineage? J. Fish
Ecol. 13, 3379–3388. Biol. 93, 302–310.
Naish, K.A., Taylor, J.E., Levin, P.S., Quinn, T.P., Winton, J.R., Huppert, D., Hilborn, R., Turan, D., Kottelat, M., Engin, S., 2009. Two new species of trouts, resident and
2008. An evaluation of the effects of conservation and fishery enhancement migratory, sympatric in streams of northern Anatolia (Salmoniformes: Salmonidae).
hatcheries on wild populations of salmon. Adv. Mar. Biol. 53, 61–194. Ichthyol. Explor. Fres. 20, 333–364.
Nei, M., 1972. Genetic distance between populations. Am. Nat. 106, 283–292. Turan, D., Kottelat, M., Engin, S., 2010. Two new species of trouts, resident and
Nei, M., 1978. Estimation of average heterozygosity and genetic distance from a small migratory, sympatric in streams of northern Anatolia (Salmoniformes: Salmonidae).
number of individuals. Genetics 89, 583–590. Ichthyol. Explor. Fres. 20, 289–384.
O'Reilly, P.T., Hamilton, L.C., McConnell, S.K., Wright, J.M., 1996. Rapid analysis of Turan, D., Kottelat, M., Bektas, Y., 2011. Salmo tigridis, a new species of trout from the
genetic variation in Atlantic salmon (Salmo salar) by PCR multiplexing of Tigris River, Turkey (Teleostei: Salmonidae). Zootaxa 2993, 23–33.
dinucleotids and tetranucleotids microsatellites. Can. J. Fish. Aquat. Sci. 53, Turan, D., Kottelat, M., Engin, S., 2012. The trouts of the Mediterranean drainages of
2292–2298. southern Anatolia, Turkey, with description of three new species (Teleostei:
O’Sullivan, R.J., Aykanat, T., Johnston, S.E., Kane, A., Poole, R., Rogan, G., Prodöhl, P. Salmonidae). Ichthyol. Explor. Fres. 23, 219–236.
A., Primmer, C.R., McGinnity, P., Reed, T.E., 2019. Evolutionary stasis of a heritable Turan, D., Doğan, E., Kaya, C., Kanyılmaz, M., 2014a. Salmo kottelati, a new species of
morphological trait in a wild fish population despite apparent directional selection. trout from Alakır stream, draining to the Mediterranean in southern Anatolia,
Ecol. Evol. 9, 7096–7111. Turkey (Teleostei, Salmonidae). ZooKeys 462, 135–151.
Palm, S., Dannewitz, J., Järvi, T., Petersson, E., Prestegaard, T., Ryman, N., 2003. Lack of Turan, D., Kottelat, M., Engin, S., 2014b. Two new species of trouts from the Euphrates
molecular genetic divergence between sea-ranched and wild sea trout (Salmo trutta). drainage, Turkey (Teleostei: Salmonidae). Ichthyol. Explor. Fres. 24, 275–287.
Mol. Ecol. 12, 2057–2071. Vähä, J.P., Erkinaro, J., Niemelä, E., Primmer, C.R., 2007. Life-history and habitat
Palm, S., Dannewitz, J., Järvi, T., Koljonen, M.-L., Prestegaard, T., Olsén, K.H., 2008. No features influence the within-river genetic structure of Atlantic salmon. Mol. Ecol.
indications of Atlantic salmon (Salmo salar) shoaling with kin in the Baltic Sea. Can. 16, 2638–2654.
J. Fish. Aquat. Sci. 65, 1738–1748. Valiquette, E., Perrier, C., Thibault, I., Bernatchez, L., 2014. Loss of genetic integrity in
Palm, S., Dannewitz, J., Johansson, D., Laursen, F., Norrgård, J., Prestegaard, T., wild lake trout populations following stocking: insights from an exhaustive study of
Sandström, A., 2012. Populationsgenetisk kartläggning av Vänerlax. Aqua reports 72 lakes from Québec, Canada. Evol. Appl. 7, 625–644.
2012, 4. Institutionen för akvatiska resurser, Sveriges lantbruksuniversitet, van Houdt, J.K.J., Pinceel, J., Flamand, M.-C., Briquet, M., Dupont, E., Volckaert, F.A.M.,
Drottningholm Lysekil Öregrund, 64 pp. (In Swedish with English Abstract). Bare, P.V., 2005. Migration barriers protect indigenous brown trout (Salmo trutta)
Poteaux, C., Guyomard, R., Berrebi, P., 2000. Single and joint gene segregation in populations from introgression with stocked hatchery fish. Conserv. Genet. 6,
intraspecific hybrids of brown trout (Salmo trutta L.) lineages. Aquaculture 186, 175–191.
1–12. Wang, J., 2009. A new method for estimating effective population sizes from a single
Presa, P., Guyomard, R., 1996. Conservation of microsatellites in three species of sample of multilocus genotypes. Mol. Ecol. 18, 2148–2164.
salmonids. J. Fish Biol. 49, 1326–1329. Waples, R.S., Luikart, G., Faulkner, J.R., Tallmon, D.A., 2013. Simple life history traits
Presa, P., Krieg, F., Estoup, A., Guyomard, R., 1994. Diversité et gestion génétique de la explain key effective population size ratios across diverse taxa. Proc. R. Soc. B 280,
truite commune: apport de l’étude du polymorphisme des locus protéiques et 20131339.
microsatellites. Genet. Sel. Evol. 26 (Suppl. 1), 183–202. Wąs, A., Bernaś, R., 2016. Long-term and seasonal genetic differentiation in wild and
Pritchard, J.K., Stephens, M., Donnelly, P., 2000. Inference of population structure using enhanced stocks of sea trout (Salmo trutta m. trutta L.) from the Vistula River, in the
multilocus genotype data. Genetics 155, 945–959. southern Baltic - management implications. Fish. Res. 175, 57–65.
Rexroad, C.E., Coleman, R.L., Hershberger, W.K., Killefer, J., 2002. Rapid Weir, B., Cockerham, C., 1984. Estimating F-statistics for the analysis of population
communication: thirty-eight polymorphic microsatellite markers for mapping in structure. Evolution 38, 1358–1370.
rainbow trout. J. Anim. Sci. 80, 541–542. Wollebaek, J., Knut, H.R., Brabrand, A., Heggenes, J., 2012. Interbreeding of genetically
distinct native brown trout (Salmo trutta) populations designates offspring fitness.
Aquaculture 356, 158–168.

12

You might also like