Origins and genetic diversity among Atlantic salmon recolonizing upstream areas of a large South...

15
RESEARCH ARTICLE Origins and genetic diversity among Atlantic salmon recolonizing upstream areas of a large South European river following restoration of connectivity and stocking Charles Perrier Je ´ro ˆme Le Gentil Virginie Ravigne Philippe Gaudin Jean-Claude Salvado Received: 8 December 2013 / Accepted: 25 March 2014 Ó Springer Science+Business Media Dordrecht 2014 Abstract The restoration and maintenance of habitat connectivity are major challenges in conservation biology. These aims are especially critical for migratory species using corridors that can be obstructed by anthropogenic barriers. Here, we explored the origins and genetic diver- sity of Atlantic salmon (Salmo salar) recolonizing upstream areas of the largest South European Atlantic salmon population (Adour drainage, France) following restoration of connectivity and stocking. We genotyped 1,009 juvenile individuals, sampled either in continuously inhabited downstream sites or in recently reconnected and recolonized upstream locations, at 12 microsatellite loci. We found significant fine scale genetic structure, with three main genetic clusters corresponding to the Nive, Nivelle and Gaves rivers. Within each of these clusters, samples collected in continuously inhabited and recently recolonized sites had comparable allelic richness and effective population sizes and were only weakly differen- tiated. Genetic structure among basins was also similar among continuously inhabited and recently recolonized sites. The majority of the individuals sampled from recently recolonized sites were assigned to neighboring continuously inhabited downstream sites, but noticeable proportions of fish were assigned to samples collected in more distant sites or identified as putative hybrids. Overall, this study suggests that the restoration of accessibility to upstream areas can allow for the recolonization and effective reproduction of Atlantic salmon from proximate downstream refugia, which does not decrease local diver- sity or disrupt existing genetic structure. Keywords Recolonization Á Genetic diversity Á Dam Á Connectivity Á Assignment Á Salmo salar Introduction The fragmentation of habitat is one of the major human threats to wild populations (Vitousek et al. 1997; Ewers and Didham 2006; Fahrig 2003) and is often due to the construction of artificial barriers, which can result in patch size reduction and patch isolation (Fahrig 2003). For a wide range of species, habitat fragmentation can also modify dispersal and gene flow (Coulon et al. 2010; Van Oort et al. 2011; Pe ´pino et al. 2012). Moreover, population size and effective population sizes as well as genetic diversity can be affected (Couvet 2002; Blanchet et al. 2010; Dixo et al. 2009; Tsuboi et al. 2013; Whiteley et al. 2013). Eventually, fragmentation may affect the evolutionary trajectories of populations and lead up to local extinction. Therefore, to mitigate habitat fragmentation and its impacts on local Charles Perrier and Je ´ro ˆme Le Gentil have contributed equally to this study. C. Perrier De ´partement de Biologie, Universite ´ Laval, Quebec G1V 0A6, Canada C. Perrier (&) Á J. Le Gentil UMR 0985 ESE, INRA, 35042 Rennes, France e-mail: [email protected] J. Le Gentil Á P. Gaudin Á J.-C. Salvado UMR 1224 Ecobiop, INRA, 64310 St Pe ´e sur Nivelle, France V. Ravigne UMR BGPI, CIRAD, TA A 54/K, Campus International de Baillarguet, 34398 Montpellier Cedex 05, France J.-C. Salvado UMR 1224 Ecobiop, Universite ´ de Pau Et des Pays de l’Adour, Campus de Montaury, 64600 Anglet, France 123 Conserv Genet DOI 10.1007/s10592-014-0602-3

Transcript of Origins and genetic diversity among Atlantic salmon recolonizing upstream areas of a large South...

RESEARCH ARTICLE

Origins and genetic diversity among Atlantic salmon recolonizingupstream areas of a large South European river followingrestoration of connectivity and stocking

Charles Perrier • Jerome Le Gentil •

Virginie Ravigne • Philippe Gaudin •

Jean-Claude Salvado

Received: 8 December 2013 / Accepted: 25 March 2014! Springer Science+Business Media Dordrecht 2014

Abstract The restoration and maintenance of habitatconnectivity are major challenges in conservation biology.

These aims are especially critical for migratory species

using corridors that can be obstructed by anthropogenicbarriers. Here, we explored the origins and genetic diver-

sity of Atlantic salmon (Salmo salar) recolonizing

upstream areas of the largest South European Atlanticsalmon population (Adour drainage, France) following

restoration of connectivity and stocking. We genotyped

1,009 juvenile individuals, sampled either in continuouslyinhabited downstream sites or in recently reconnected and

recolonized upstream locations, at 12 microsatellite loci.

We found significant fine scale genetic structure, with threemain genetic clusters corresponding to the Nive, Nivelle

and Gaves rivers. Within each of these clusters, samples

collected in continuously inhabited and recently

recolonized sites had comparable allelic richness andeffective population sizes and were only weakly differen-

tiated. Genetic structure among basins was also similar

among continuously inhabited and recently recolonizedsites. The majority of the individuals sampled from

recently recolonized sites were assigned to neighboring

continuously inhabited downstream sites, but noticeableproportions of fish were assigned to samples collected in

more distant sites or identified as putative hybrids. Overall,

this study suggests that the restoration of accessibility toupstream areas can allow for the recolonization and

effective reproduction of Atlantic salmon from proximate

downstream refugia, which does not decrease local diver-sity or disrupt existing genetic structure.

Keywords Recolonization ! Genetic diversity ! Dam !Connectivity ! Assignment ! Salmo salar

Introduction

The fragmentation of habitat is one of the major humanthreats to wild populations (Vitousek et al. 1997; Ewers

and Didham 2006; Fahrig 2003) and is often due to the

construction of artificial barriers, which can result in patchsize reduction and patch isolation (Fahrig 2003). For a wide

range of species, habitat fragmentation can also modify

dispersal and gene flow (Coulon et al. 2010; Van Oort et al.2011; Pepino et al. 2012). Moreover, population size and

effective population sizes as well as genetic diversity can

be affected (Couvet 2002; Blanchet et al. 2010; Dixo et al.2009; Tsuboi et al. 2013; Whiteley et al. 2013). Eventually,

fragmentation may affect the evolutionary trajectories of

populations and lead up to local extinction. Therefore, tomitigate habitat fragmentation and its impacts on local

Charles Perrier and Jerome Le Gentil have contributed equally to thisstudy.

C. PerrierDepartement de Biologie, Universite Laval, Quebec G1V 0A6,Canada

C. Perrier (&) ! J. Le GentilUMR 0985 ESE, INRA, 35042 Rennes, Francee-mail: [email protected]

J. Le Gentil ! P. Gaudin ! J.-C. SalvadoUMR 1224 Ecobiop, INRA, 64310 St Pee sur Nivelle, France

V. RavigneUMR BGPI, CIRAD, TA A 54/K, Campus International deBaillarguet, 34398 Montpellier Cedex 05, France

J.-C. SalvadoUMR 1224 Ecobiop, Universite de Pau Et des Pays de l’Adour,Campus de Montaury, 64600 Anglet, France

123

Conserv Genet

DOI 10.1007/s10592-014-0602-3

populations, numerous conservation actions have been

undertaken to conserve, restore (Clewell and Aronson2006; Aronson 2011; De Groot et al. 2013), and increase

habitat connectivity (Brown et al. 2013). Therefore,

effective conservation and restoration of habitat connec-tivity requires knowledge of how fragmentation and

reconnection impacts local species.

Freshwater ecosystems are particularly subject to frag-mentation due to anthropogenic activities (Nilsson et al.

2005; Hall et al. 2011). One of the main causes of frag-mentation of freshwater habitats has been the widespread

construction of dams for irrigation purposes, drinking

water retention, watermills, hydroelectric power plants, andrecreational activities. Their impacts on freshwater eco-

systems are well recognized, and range from modifications

to population genetic diversity (Horreo et al. 2011; Neraasand Spruell 2001; Wofford et al. 2005) to changes in

species assemblages (Poulet 2007; Boet et al. 1999; Brown

et al. 2013; Grenouillet et al. 2008). Barriers sizes andpermeability are important parameters affecting the

movement of fish species (Raeymaekers et al. 2009) and

thus their demographic and genetic characteristics. Inaddition, swimming ability of the fish and their capacity to

disperse through barriers are often species specific (Haro

et al. 2004). Accordingly, recent studies demonstrated thatthere are species-specific modifications of gene flow and

genetic structure due to weirs (Blanchet et al. 2010) and

waterfalls (Gomez-Uchida et al. 2009). In particular,anadromous species like Atlantic salmon, which migrate

between spawning grounds located in rivers and feeding

zones at sea (Jonsson and Jonsson 2011), can be highlyimpacted by reduced connectivity as a result of dams

(Brown et al. 2013; Hall et al. 2012). Therefore, many

conservation studies and programs aimed at documentingthe impact of artificial barriers on the demography and

evolutionary trajectories of anadromous fish populations

have highlighted the importance of the restoration of hab-itat connectivity.

In many regions of the world, there are regulations to

re-establish connectivity among freshwater habitats thathave been disconnected by weirs and dams to restore fish

movement within watersheds. Migratory fish species have

been the main targets of these restoration regulations sincethey are highly impacted by habitat fragmentation and

because they are of high biological, economic and societal

interest. Partial restoration of river connectivity can beprovided by the addition of fishways (Coutant and Whitney

2000; Brown et al. 2013; Larinier and Boyer-Bernard

1991). Even though fishways have proven their effective-ness, dam removal remains the best option to effectively

improve fish movement (Brown et al. 2013; Oldani and

Baigun 2002; Mallen-Cooper and Brand 2007). In additionto restoring fish movement, dam removal can also allow

for the restoration of spawning grounds that were previ-

ously buried under sediment (Bednarek 2001). Geneticanalyses of these recolonizing individuals together with

baseline information from samples collected in geograph-

ically close and continuously inhabited sites may help toidentify the origin and genetic diversity of these colonists

and help us to better understand the recolonization process

(Kiffney et al. 2009; Perrier et al. 2010; Griffiths et al.2011; Winans et al. 2010; Ikediashi et al. 2012). Several

recent studies have found that the Atlantic salmon recol-onizing rivers mainly came from nearby rivers but that

some fish also came from more distant source populations

(Perrier et al. 2010; Griffiths et al. 2011; Ikediashi et al.2012). However, there have not yet been any studies

aiming at identifying the origin, among productive down-

stream areas, of salmon recolonizing recently reconnectedupstream parts of rivers. In addition to determining the

origin of the immigrants, such studies within a single river

system will provide insight into the diversity and effectivepopulation size in recently recolonized river sections, two

population genetic parameters predicted to influence the

success of recolonization (Naish et al. 2013; Oakley 2013;Fraser et al. 2007).

As an alternative approach to sustain declining or re-

establish extinct populations of migratory salmonids, thestocking of wild or captive hatchery-reared individuals is

often used. However, although stocking may help rees-

tablish populations, it has been highly criticized because ofits potential negative effect on the fitness of wild popula-

tions in the long term (Aprahamian et al. 2003; Araki and

Schmid 2010; Fraser 2008). Indeed, the fitness of stockedfish may be reduced due to the effects of unintentional

selection and domestication during early life in hatcheries

that affects survival, migration and reproductive success ofstocked fish (Araki et al. 2008; Milot et al. 2013; Theriault

et al. 2011). Moreover, stocking with individuals origi-

nating from highly differentiated stocks has resulted in theloss of neutral genetic integrity in salmonids (Marie et al.

2010; Bourret et al. 2011; Perrier et al. 2013a, b; Hansen

et al. 2009). Since fine scale genetic structure may existwithin large basins (Dionne et al. 2009; Ensing et al. 2011;

Vaha et al. 2007; Primmer et al. 2006), even the use of

local wild fish to produce stocked fry may ultimately resultin the disruption of local genetic structure and local

adaptation if subpopulations are not carefully considered

(Eldridge et al. 2009; Pearse et al. 2011). Therefore,genetic analyses can be used not only to document the

origin of fish recolonizing recently reconnected sites but

also to estimate the relative contributions of ‘‘natural’’recolonization versus stocking (Beaudou et al. 1994;

Griffiths et al. 2011).

In this study, we explore the distribution of geneticdiversity within the largest Atlantic salmon catchment in

Conserv Genet

123

Southern Europe (Adour, France), to determine the origin

of Atlantic salmon recolonizing upstream areas followingrecent restoration of connectivity and stocking. The Adour

hydro-geographic basin has a surface of 16,880 km2 and

harbours a large wild Atlantic salmon population, which isan important target of both recreational and commercial

fishing (Vauclin 2007). On the basis of relatively few

samples, Perrier et al. (2011b) revealed that Atlantic sal-mon samples collected in Nive, Nivelle and Adour rivers

(see the map Fig. 1) clustered together and were geneti-cally differentiated from other French stocks. Nevertheless,

given the size of the Adour basin and the distance among

spawning grounds where Atlantic salmon have beenobserved, a fine scale population genetic structure may

exist. Of principal interests the connectivity of several

upstream areas previously inaccessible to migratory fishdue to the presence of impassable dams have been restored

since 1986 (see Fig. 1; Table 1 and methods to locate

impassable and passable dams). In parallel, salmon fry andparr, mainly the offspring of wild caught parents, have

been released since the 1970s into various tributaries of the

Adour catchment in order to sustain populations and to aidin the reestablishment of new populations in recently

reconnected sites (Marty 1984; Beall et al. 1995; Davaine

et al. 1996). How this history of human intervention hasimpacted the Salmon populations in this area is unknown.

Therefore, the specific aims of the present study were to:

(1) investigate the fine scale genetic population diversityand structure within the Adour catchment and identify

conservation units, (2) determine whether Atlantic salmon

recolonizing recently restored habitats originated from

proximate downstream sites in the basin or from the other

basins, (3) test whether the genetic diversity, effectivepopulation size, and genetic structure within recently

recolonized areas are reduced compared to continuously

inhabited sites, (4) test whether stocking impacted thedistribution of genetic diversity within the basin, and (5)

discuss the relevance of our results for the sustainable

management and conservation of this important Southern-European Atlantic salmon stock and generally discuss the

effects of the restoration of connectivity in freshwaterfishes.

Materials and methods

Study site and Atlantic salmon population

The Adour catchment is a 16,880 km2 river-system.

Atlantic salmon (Salmo salar) spawning sites can be foundin several major sub-drainages including: the Nive

(993 km2), the Saison (631 km2), the Gave d’Oloron

(2,000 km2), the Gave d’Aspe (598 km2), the Gave d’Os-sau (493 km2) and the Gave de Pau (2,600 km2). The

Nivelle River (244 km2) estuary is twenty kilometres away

from the Adour estuary and harbours a small Atlanticsalmon population. On average over the last decade, 6,500

adults annually return to the Adour Drainage, and 300,000

parr 0? are produced per year (Barracou 2008; Le Gentilet al. in prep). Thus, the Adour River is currently the most

productive drainage in France as well as in Southern Eur-

ope. The Adour population is large when compared to mostNorthern-European populations, although much larger

populations do exist (Tonteri et al. 2009).

Numerous dams have been built on the Adour River sincethe beginning of the eighteenth century, and habitat frag-

mentation reached its maximum during the 1940’s due to

hydro-electric power plant construction. Dam constructiondecreased or prevented the migratory fish from reaching

upstream spawning sites (Fig. 1; Table 1). As a result, the

abundance and catch of Atlantic salmon dramaticallydeclined in the 1960’s in the Adour basin (Barracou 2008).

For example, more than 10,000 fish were caught annually at

the beginning of the twentieth century but \500 fish werecaught in 1976 (Marty and Bousquet 2001). The intensifi-

cation of agriculture, urbanization, industrialization and

overexploitation at sea may have also contributed to thisdecline in Atlantic salmon. Since 1986, in response to

changes in regulations, much work has been done to

improve the accessibility of upstream spawning sites tomigratory fish, principally Atlantic salmon (Barracou 2008).

Connectivity restoration mainly consisted of building fish-

ways on dams isolating sectors 2, 3, 5, 7, 8, 11, 12, 13 and 16(Fig. 1). The dam that isolated sector 6 was destructed. The

N

1

23

4

56

7

8

9

10

11

12

13

15

14 16

Nive

Nivelle

Lurgorieta

Béhérobie Arnéguy

Gaves Adour

Pau Ossau

Aspe

Lourdios Vert

Saison

Oloron

Fig. 1 Location of the sampling sites that were continuouslyinhabited (blue circles), or which have been recently recolonized(green circles). Dams which were passable (green stars) or impass-able (red stars) at the time of sampling are also noted. For site numbercorrespondence, see Table 1. (Color figure online)

Conserv Genet

123

Tab

le1

Ch

arac

teri

stic

so

fsa

mp

lin

glo

cati

on

san

dg

enet

icd

iver

sity

ind

ices

of

thei

rco

rres

po

nd

ing

sam

ple

sin

clu

din

gF

IS,

exp

ecte

dh

eter

ozy

go

sity

(He)

,al

leli

cri

chn

ess

(Ar)

,p

riv

ate

alle

lic

rich

nes

s(P

Ar)

,ef

fect

ive

size

(Ne)

,an

dce

nsu

ssi

zep

erb

asin

(Nc

corr

esp

on

ds

toth

eav

erag

en

um

ber

of

anad

rom

ou

sfi

shen

teri

ng

the

bas

ins,

Lan

ge

etal

.2

01

1)

Bas

inR

iver

Tri

bu

tary

Co

ord

inat

esL

oca

tio

nD

ista

nce

tori

ver

mo

uth

(km

)

Yea

ro

fre

sto

rati

on

of

acce

ssib

ilit

y

N gen

oty

ped

Nfu

llsi

bs

%o

ffu

llsi

bs

Naf

ter

rem

ov

ing

full

sib

s

FIS

He

Ar

PA

rN

eN

cp

erb

asin

Niv

elle

Niv

elle

Niv

elle

43

"200

5200 N

/0

1"3

30 0

600 W

11

7–

30

82

72

20

.08

0.8

48

.59

0.3

43

3.8

10

5

Niv

elle

43

"180

3300 N

/0

1"3

10 4

900 W

2*

22

19

92

45

51

14

00

.01

0.8

48

.61

0.3

11

24

.0

Lu

rgo

riet

a4

3"1

802

800 N

/0

1"3

30 5

800 W

3*

22

19

92

19

94

71

00

.01

0.8

58

.67

0.4

13

02

.6

Ad

ou

rN

ive

Niv

e4

3"2

004

400 N

/0

1"2

60 5

300 W

45

9–

25

00

25

0.0

40

.80

7.2

40

.08

59

5.8

56

0

Lau

rhib

ar/

Beh

ero

bie

43

"150

4000 N

/0

1"2

40 3

100 W

5*

67

20

01

11

63

63

18

00

.03

0.8

27

.84

0.2

71

07

.6

Arn

egu

y4

3"1

208

900 N

/0

1"2

60 7

400 W

6*

70

19

92

17

45

02

91

24

0.0

50

.80

7.2

80

.18

59

5.5

Gav

esS

aiso

n4

3"3

102

700 N

/0

0"8

70 1

900 W

7*

88

19

86

27

00

27

0.0

50

.85

8.4

30

.28

14

2.6

9,2

00

Sai

son

43

"190

5100 N

/0

0"9

10 3

300 W

8*

10

51

99

59

87

79

10

.02

0.8

48

.45

0.3

0.0

40

0.5

Gav

ed

’Olo

ron

43

"260

8500 N

/0

0"6

80 4

100 W

91

06

–6

53

56

20

.02

0.8

38

.21

0.1

92

38

.4

Gav

ed

’Olo

ron

43

"230

1100 N

/0

0"6

30 7

000 W

10

11

5–

47

71

54

00

.06

0.8

28

.07

0.2

64

34

.8

Ver

t4

3"1

603

100 N

/0

0"6

80 1

500 W

11

*1

23

20

01

42

13

31

29

0.0

70

.83

8.3

30

.26

78

.6

Gav

ed

’Asp

e4

3"0

308

800 N

/0

0"6

00 4

200 W

12

*1

41

19

92

47

71

54

00

.03

0.8

38

.04

0.2

68

7.9

Lo

urd

ios

43

"070

6900 N

/0

0"6

60 9

80W

13

*1

34

19

92

30

41

32

60

.00

0.8

78

.93

0.2

38

9.0

Gav

ed

’Oss

au4

3"1

109

800 N

/0

0"4

70 7

60W

14

13

8–

56

19

34

37

0.0

70

.85

8.3

40

.28

10

4.7

Gav

ed

’Oss

au4

3"0

806

000 N

/0

0"4

20 1

00W

15

*1

49

19

93

54

61

14

80

.08

0.8

38

.12

0.3

13

58

3.8

Gav

ed

eP

au4

3"0

906

000 N

/0

0"1

70 1

00W

16

*1

57

19

92

49

13

27

36

0.0

30

.82

7.7

80

.15

84

.0

To

tal

16

92

41

87

19

73

70

.04

0.8

38

.18

0.2

6

Conserv Genet

123

Dam that isolated sector 15 was equipped with a fish lift.

Thanks to these improvements in habitat connectivity sec-tors 7, 2–3–6–12–13–16, 15, 8 and 5–11 became accessible

in 1986, 1992, 1993, 1995 and 2001, respectively.

To sustain the Atlantic salmon populations inhabitingthe Adour drainage as well as facilitate the recolonization

of upstream habitats, stocking programs were implemented

in the 1970s (Marty 1984; Beall et al. 1995; Davaine et al.1996; Marty and Bousquet 2001). Eggs, young of the year

and smolts were stocked in several tributaries. Until 1990,eggs, fry and smolts produced from local and non-local

parents (mainly originating from Scotland) were stocked in

the Gaves, Nive and Nivelle. Since 1990, local parentswere used in the Nive, Nivelle and Gaves Rivers (Marty

and Bousquet 2001). At present, stocking is sustained only

in the Gave de Pau. As presented in (Perrier et al. 2013b),little introgression by non-local strains occurred in the

Adour populations and this introgression has remained

stable over the last decades.

Sampling

During September and October 2005, 0? and 1? year old

juvenile Atlantic salmon were sampled by cutting off a

small part of their pectoral fins (non-lethal sampling). Finswere stored in 95 % ethanol. A total of 1,009 individuals

were sampled in 16 locations distributed in the Adour

drainage (Fig. 1; Table 1). To increase the number ofindividuals analysed per location, but to limit biases

resulting from sampling individuals from the same family,

each sampling site consisted in a section of 10–30 km long.Individuals sampled were wild fish for all but site #14,

where a restocking operation occurred in 2004.

DNA extraction and genotyping

DNA extraction, amplification and genotyping were com-pleted according to the procedures described in (Horreo

et al. 2008). A total of twelve microsatellites were selected

on the basis of consistency of amplification, ease of scoringand variability: Ssa197, Ssa202, Ssa171 (O’Reilly et al.

1996), SSsp1605, SSsp2210, SSspG7, SSsp2201 (Paterson

et al. 2004), Ssosl417, Ssosl85 (Slettan et al. 1995),SsaD144b, SSa157a (King et al. 2005), Ssa289 (McConnell

et al. 1995). A thirteenth microsatellite was used to dis-

tinguish S. salar from Salmo trutta and detect possiblehybrids between these species (Perrier et al. 2011a).

Data quality

ML-relate (Kalinowski et al. 2006) was used to detect full-

sibs. Full-sibs were subsequently removed from the dataset

to avoid bias caused by the overrepresentation of individ-

uals from the same family (Hansen et al. 1997). We usedthe software Micro-Checker to test for the presence of null

alleles (Van Oosterhout et al. 2004). Linkage disequilib-

rium was estimated using Genepop 3.4 (Raymond andRousset 1995b), with sequential Bonferroni correction

(Raymond and Rousset 1995a). FIS and tests for Hardy–

Weinberg disequilibrium were conducted withFSTAT2.9.3.2 based on 1,000 permutations.

Analysis of genetic diversity within samples

Observed (Ho) and expected heterozygoties (He) wereestimated using Genepop 3.4. The number of alleles per

locus and population, allelic richness (Ar) and private

allelic richness (PAr) were estimated using Hp-rare 1.0(Kalinowski 2005). Effective population size (Ne) was

estimated for each location using the LDNe method (Wa-

ples and Do 2008) implemented in NeEstimator V2.0 (Doet al. 2013). We used an allele frequency threshold of 0.05.

We tested for data normality using a Shapiro test, and then

compared average Ar and Ne among recently recolonizedand continuously inhabited sites using student’s t tests for

normally distributed data and Mann–Withney tests for non-

normal distributed data. Census size (Nc) were obtainedfrom Lange et al.’s (2011) and corresponded to the average

number of anadromous fish entering in the Nive, Nivelle

and Gaves rivers.

Analysis of genetic structure among sites

We used Genepop 3.4 to estimate FST between sites. Sig-

nificance of FST was estimated using 1,000 permutations.

We compared average FST among: (A) recently recolon-ized sites and sites that were continuously inhabited within

each basins, (B) recently recolonized sites from different

basins and (C) sites from different basins that remainedinhabited by Atlantic salmon. We tested for sstatistical

significance using student’s t tests. We performed three

AMOVAs using Arlequin v3.5 (Excoffier and Lischer2010). The first AMOVA was conducted using Nivelle,

Nive and Gaves as groups to test for hierarchical structure.

The second AMOVA was conducted using the samegrouping but on continuously inhabited sites only. The last

AMOVA was conducted using the same grouping but on

recently recolonized sites only. These last two AMOVAswere conducted to investigate whether genetic differenti-

ation was lower among recently recolonized sites com-

pared to continuously inhabited ones at both hierarchicallevels. A neighbor-joining dendrogram based on pairwise

Nei (Da) genetic distances (Nei et al. 1983) was con-

structed with Populations 1.2.30 (http://bioinformatics.org/*tryphon/populations/). Confidence estimates of tree

Conserv Genet

123

topology were calculated by 1,000 bootstraps of loci.

Dendrograms were visualized using TreeView (Page1996).

Bayesian clustering and assignment of individuals

We examined the clustering of populations and individuals

using Baps v2.0 (Corander et al. 2004) and Structure(Pritchard et al. 2000). While Baps rapidly and accurately

finds main clusters, the estimation of individual admixturemay be less accurate. Alternatively, Structure accurately

estimates population and individual admixture but is rela-

tively slow and may have difficulty finding main clusterswhen some populations are under or over-represented

(Kalinowski 2010). First, we used the population clustering

option implemented in BAPS to delineate main geneticclusters in the dataset, with a maximum number of

potential clusters set to 10. Population and individual

admixture was subsequently tested on the basis of thesepopulation-clustering results. Second, the Bayesian clus-

tering method implemented in the software Structure was

used to delineate K genetic clusters. The best K valueswere defined according to the DK procedure as described

by (Evanno et al. 2005), using STRUCTURE HAR-

VESTER (Earl and vonHoldt 2012). A total of 10 runswere computed for each value of K tested, from 1 to 10.

Each run started with a burn-in period of 50,000 steps

followed by 300,000 Markov Chain Monte Carlo (MCMC)replicates. We used an admixture model, without prior

information regarding population clustering.

We used the software Geneclass2 (Piry et al. 2004)following the methods of (Baudouin and Lebrun 2000) to

assign individuals sampled in either recently recolonized

sites or continuously inhabited ones by Atlantic salmon to abaseline constituted of samples from continuously inhab-

ited sites (sites 1, 4, 9, 10 and 14). This was done to

identify the source populations for the newly foundedpopulations. Individuals that had scores \70 % were con-

sidered as potential hybrids or migrants from un-sampled

populations.

Results

Data quality

Of the 1,009 sampled individuals, 960 individuals were

successfully genotyped with a minimum of 66 % of indi-

vidual amplification success. The total amplification suc-cess was 99.68 %. Using the SSAD486 marker to identify

species, we found that 5 individuals were S. trutta, 26

individuals were hybrids between S. trutta and S. salar, and5 could not be identified (amplification failed at this locus).

We discarded these 36 individuals and conducted the

subsequent analyses on the 924 remaining Atlantic salmon.Following the results from ML-relate, we removed a total

of 187 (20 %) individuals that were found to have full-sibs

in our dataset. These 187 removed individuals represented0–47 % of the total number of individuals at each site

(median value of 15 %). We thus conducted all the sub-

sequent analyses on a total of 737 individuals (Table 1).Micro-Checker detected no null alleles. No linkage dis-

equilibrium was detected by Genepop among loci(p [ 0.05 for all them after Bonferroni corrections), thus,

all loci were considered to be genetically independent.

Only eight out of 192 FIS computed in Fstat were signif-icant. No locus presented significant deviation from Hardy

to Weinberg Equilibrium over all populations. At the

population level, locations 6, 11 and 15 yielded signifi-cantly smaller observed than expected heterozygosities

(Table 1).

Analysis of genetic diversity within samples

Loci had 6 (Ssa289) to 49 (Ssa157a) alleles over the entiredataset, with a median value of 24 and a total of 322 alleles

over all loci. Average He (Expected heterozygosity) over

all loci per population varied from 0.80 to 0.87 (Table 1),0.84 on average in the Nivelle basin, 0.81 in the Nive basin

and 0.84 in the Gaves basin. He was on average of 0.83 in

continuously inhabited locations and of 0.84 in recentlyrecolonized sites. Average Ar (Allelic richness) over all

loci varied from 7.28 to 8.93 depending on the location. Ar

was on average 8.62 in the Nivelle basin, 7.45 in the Nivebasin and 8.27 in the Gaves basin. Ar was on average 8.09

in continuously inhabited locations and 8.23 in recently

recolonized sites (Fig. 2). Ar was not significantly differentbetween continuously inhabited locations and recently

recolonized ones (Table 2, t test, t = -0.50, df = 7.26,

p value = 0.63). Average PAr (Private allelic richness)varied from 0.15 to 0.41 depending on the location. PAr

was on average 0.35 in the Nivelle basin, 0.18 in the Nive

basin and 0.25 in the Gaves basin. PAr was on average 0.23in continuously inhabited locations and 0.27 in recently

recolonized sites. Effective size (Ne) varied from 33.8 to

3,583.8 depending on the location (Table 1). Overall, themedian value of Ne was 238.4 in continuously inhabited

locations and 124.0 in recently recolonized sites (Fig. 2).

These two medians of Ne estimates were not significantlydifferent (Table 2, Mann–Whitney test, W = 30,

p value = 0.83).

Analysis of genetic structure among sites

FST among sites ranged from 0.000 to 0.054 (Table 2).Table 2 and Fig. 3 show small FST among continuously

Conserv Genet

123

inhabited sites and recently recolonized ones within each

basin (FSTA = 0.012 on average), but relatively high FST

among sites located in different basins, either recently

recolonized ones (FSTB 0.035 on average) or continuously

inhabited sites (FSTC 0.042 on average). These three FST

were significantly different (FSTB vs FSTC t = -2.54,

df = 10.99, p value = 0.03; FSTA vs FSTB t test, t =

-10.30, df = 27.75, p value = 5.6e -11; FSTA vs FSTCt test, t = -10.36, df = 12.87, p value = 1.3e-07).

Considering all the sites, AMOVAs revealed that2.53 % of the genetic variance was found among basins

and 1.08 % among populations within basins (Table 3).

When considering only continuously inhabited sites,AMOVAs revealed that 3.75 % of the variance was found

among basins and 0.47 % among populations within

basins. When considering only recently recolonized sites,AMOVAs revealed that 2.15 % of the variance was found

among basins and 1.22 % among populations within

basins. These two last AMOVAs showed that geneticstructure was higher among continuously inhabited sites

than among recently recolonized ones, according to the

5–95 % confidence intervals.A neighbor-joining dendrogram of Nei’s genetic dis-

tance revealed three genetically distinct populations cor-

responding to the Nivelle, Nive and Gaves basins (Fig. 4).This dendrogram also reveals relatively little differentia-

tion among sites within each basin.

Bayesian clustering and assignment of individuals

While the best k value found by BAPS was k = 3 (Fig. 5),the first delta k pick found in STRUCTURE result corre-

sponded to k = 4, followed by a smaller pick at k = 8.

However, the existence of several clusters in Gaves fork = 4 and in Gaves and Nive for k = 8 did not clearly

corresponded to any geographic grouping and individuals

were highly admixed. We therefore proposed that the mostrealistic number of genetic clusters was three, corre-

sponding to the three main basins.

According to the assignment conducted using Geneclass,an average of 86 % of the individuals sampled in continu-

ously inhabited sites were assigned to their site of origin, 95,

100 and 78 % on average in the Nivelle, Nive and Gavesrivers, respectively (Table 4), suggesting a relatively high

power to assign fish to the three different basins. In contrast,

an average of 54 % of putatively local individuals was foundin recently recolonized sites (48, 71 and 50 % in the Nivelle,

Nive and Gaves rivers, respectively). Within the continu-

ously inhabited sites, 0 % of the individuals that could not beassigned to a basin were putative migrants from other basins

and 12 % of individuals were hybrids or migrants from other

unsampled populations. Within the recently recolonizedsites, the individuals non-assigned to the basin were 4 %

putative migrants from other basins and 38 % hybrids or

migrants from other unsampled populations.

Table 2 Table of FST among sites. (Color table online)

Significant values are indicated in bold and non-significant ones in italic. A color gradient help to visualize the hierarchical differentiation amongpopulations

Conserv Genet

123

Discussion

Fine scale genetic diversity within continuously

inhabited sites

The significant genetic differentiation (FST) observed

among continuously inhabited sites within the Adourcatchment was comparable to the FST observed among

other Atlantic salmon populations at the within river scale

using markers with similar level of polymorphism (Elliset al. 2011; Dionne et al. 2009; Vaha et al. 2007; Primmer

et al. 2006). This significant differentiation among popu-

lations from these three rivers contrasted with the results ofPerrier et al. (2011b), who only found one major genetic

cluster within the Adour River. However, Perrier et al.

(2011b) analyzed a much smaller number of individuals.The clear clustering of individuals in three genetically and

geographically distinct groups suggests limited dispersal

and gene flow among the Nive, Nivelle and Gaves rivers.

This is in line with the relatively strict homing behaviour of

Atlantic salmon (Stabell 1984) that has been suggested tobe accurate to the tributary level (Vaha et al. 2008; Dillane

et al. 2008). This existing fine scale genetic structure may

also be linked to fine scale local adaptation (Vaha et al.2008, 2007) and should be taken into account for man-

agement (Fraser and Bernatchez 2001). Hence, these

results support the local management strategy that has been

Fig. 2 Boxplots of allelic richness and effective population size incontinuously inhabited and recently recolonized sites

Fig. 3 Boxplots of FST: A between continuously inhabited andrecently recolonized sites in each basin; B between recentlyrecolonized sites among each basins; C between continuouslyinhabited sites among each basins

Table 3 Analysis of molecular variance partitioning genetic struc-ture among and within the 3 main drainages (Gaves, Nivelle, Nive)

Populationconsidered

Source ofvariation

% ofvariation

U-Statistic mean(5–95 %)

All, n = 16 Among groups 2.534 0.031 (0.021–0.041)

Amongpopulationswithin groups

1.076 0.011 (0.009–0.013)

Continuouslyinhabited,n = 5

Among groups 3.749 0.037 (0.027–0.049)

Amongpopulationswithin groups

0.469 0.005 (0.001–0.010)

Recentlyrecolonized,n = 11

Among groups 2.145 0.021 (0.017–0.027)

Amongpopulationswithin groups

1.222 0.012 (0.011–0.014)

Conserv Genet

123

applied since the 1990’s, which considers the Nive, Nivelleand Gaves rivers as three distinct conservation units.

Effective population sizes estimated for the continu-

ously inhabited sites in these three rivers were relativelyhigh compared to other rivers located in Southern Europe

(Perrier et al. 2013b; Nikolic et al. 2009). In particular, the

populations located at the edge of this species’ range (likethe study population) have dramatically declined (Boylan

and Adams 2006; Parrish et al. 1998; Dumas and Prouzet

2003; Prouzet 1990). While the size of Atlantic salmonpopulations inhabiting the Adour basin has declined during

the past decades, the effective population sizes we detected

appear may be high enough for maintaining geneticdiversity on the short term. However, these effective pop-

ulation sizes are relatively low compared to what is needed

for the long-term conservation of wild populations(Frankham 2002, 2005; Traill et al. 2010). Nevertheless,

the role of migration among populations within the Adour

basin or even from distant rivers should not be neglected insustaining long-term effective population sizes and genetic

diversity (Gomez-Uchida et al. 2013; Palstra and Ruzzante2011; Kuparinen et al. 2010). Moreover, high proportions

of mature male parr may also increase effective population

size in these southern populations (Saura et al. 2008;Garcia-Vazquez et al. 2000; Martinez et al. 2000; Moran

et al. 1996).

The low admixture and low proportions of putativemigrants or hybrids within continuously inhabited sites

suggests a relatively low impact of stocking. In particular,

the use of geographically and genetically distant popula-tions to stock the river from the Adour catchment may have

resulted in noticeable admixture (Perrier et al. 2013b;

Finnengan and Stevens 2008; Hansen et al. 2009) and in alack of differentiation among locations due to a local

homogenisation of the genetic diversity (Marie et al. 2010;

Eldridge and Naish 2007). Similarly, since stocked fish canhave a higher dispersal than their wild counterparts (Pe-

dersen et al. 2007; Quinn 1993), even stocking local but

hatchery-reared individuals may have led to admixtureamong local clusters. However, admixture appeared low

within continuously inhabited sites and these locations

were significantly differentiated. This may be due to arelatively low return rate (Perrier et al. 2013a) and fitness

0.05

1

3*

50

2*

50

4

6*

985*

97

77

7*

8*

4713*

9

10

75

14

34

15*

14

16*

3

12*

2

11*Gaves

Nive

Nivelle

Fig. 4 Neighbor joining tree based on Nei 1983 genetic distances.Bootstrap values are indicated upside the node. Recently recolonizedsites are indicated with an asterisk

Fig. 5 Bayesian individual clustering obtained using the softwareBAPS for k = 3 the software STRUCTURE for k = 8 and k = 4.Vertical bars represent proportions of membership of each individual

to each cluster represented, which are represented by different colors.Recently recolonized sites are indicated with an asterisk. (Color figureonline)

Conserv Genet

123

of non-native salmon stocked before the 90’s (Milot et al.

2013; Perrier et al. 2013b; Theriault et al. 2011). Relativelylow survival and/or low straying of local fish stocked since

the 90’s using Nive, Nivelle and the Gaves basins as

conservation units may also explain low admixture amongthese populations. Overall, while we cannot rule out the

impacts of historical stocking on present neutral genetic

structure among populations or on their adaptive potential,our results suggests that stocking did not significantly

affect genetic structure within continuously inhabited sites

from the Adour basin.

The recolonization of reconnected sites by Atlantic

salmon

A relatively high production of fry was found in recently

reconnected spawning grounds (Barracou 2008), illustrat-ing that the construction of fishways on dams may aid in

the effective recolonization of Atlantic salmon populations

(i.e. recolonization followed by successful reproduction).Even though noticeable proportions of fish sampled in

recently recolonized sites were assigned to distant sites or

were identified as putative hybrids, the majority of theindividuals were assigned to neighboring downstream sites,

suggesting a higher colonization success of local fish.

Accordingly, the genetic structure among samples from asingle river was small, suggesting a relatively high con-

tribution of fish having local origin, either wild or stocked,

to the establishment of new populations. This was also

supported by our finding that there is a similar genetic

structure among basins for both continuously inhabited andrecently recolonized sites. These results agree with (Perrier

et al. 2010; Ikediashi et al. 2012) data, which found that

large proportions of fish recolonizing depopulated riversoriginated from nearby rivers. However, in contrast with

the previous studies documenting recolonization processes

in Atlantic salmon by inferring the origin of adults’ indi-viduals recolonizing rivers (Griffiths et al. 2011; Perrier

et al. 2010; Ikediashi et al. 2012), here we genotypedjuveniles caught in recolonized parts. Thus, we character-

ized both a recolonization and a successful reestablishment

of Atlantic salmon populations. Nevertheless, it did notallow us to compare reproductive success among adult fish

returning to upstream areas. Hence, the fact that large

proportions of fry caught in recently recolonized habitatswere assigned to downstream sites of the same rivers

(Nive, Nivelle, Gaves, respectively) does not necessarily

indicate that stocked or wild adult fish had high homingrates but instead may suggest that adults with local genetic

characteristics had high reproductive success relative to

potential migrants. This illustrates that even though dis-persal of both wild and stocked fish might have contributed

to the reproductive effort within recently reconnected sites,

the parents with the highest reproductive success, overall,originated from close downstream sites. This result agrees

with studies showing that immigrants often come from

nearby sites (Perrier et al. 2010; Ikediashi et al. 2012) andwith those finding locally adapted fish may have higher

reproductive success (Dionne et al. 2008; McGinnity et al.

2007; Hendry 2004). However, it was not possible to dis-entangle the relative contributions of stocking and coloni-

zation of wild fish since recent stocking has used local

parents for each of the three rivers. Pedigree reconstructionwould have allowed us to address this question (Milot et al.

2013; Araki et al. 2007; Theriault et al. 2011), but this was

not possible due to the prohibitively extensive and costlygenotyping effort needed to address this issue in such a

large population. Given that local fish might be locally

adapted, as widely suggested for Salmonids (Bourret et al.2013; Primmer 2011; Garcia de Leaniz et al. 2007; Taylor

1991), the fact that local fish may have had a higher

reproductive contribution than non-local fish has importantimplications for the long term recolonization of the Adour

basin. In particular, local fish might be more prone to

establish a viable population than exogenous individuals.We expected a relatively low effective population sizes

in recently recolonized locations compared to continuously

inhabited sites, as a result of founder effects. However,effective population sizes in these two groups were rela-

tively similar. Such comparable effective sizes among

recently recolonized and continuously inhabited sites couldbe explained by relatively high contemporary gene flow

Table 4 Assignation of Atlantic salmon among rivers and tributaries(sampling sites grouped by basins)

River Sites 1 (%) 4 (%) 9, 10, 14(%)

Genetically admixedindividuals (%)

Nivelle 1 95 0 5 0

2* 35 5 13 48

3* 60 0 10 30

Nive 4 0 100 0 0

5* 0 65 15 20

6* 1 77 9 13

Gaves 7* 7 7 48 37

8* 4 3 42 51

9 0 0 85 15

10 0 0 73 28

11* 0 14 41 45

12* 0 0 45 55

13* 0 0 65 35

14 0 0 78 22

15* 0 0 60 40

16* 0 0 50 50

* Recently recolonized sites

Conserv Genet

123

(Waples and England 2011). Of particular interest, effec-

tive population size tended to be larger for recentlyrecolonized habitats within the Nivelle River. This result

suggests that the new population that recolonized reopened

areas is even larger than the populations located down-stream in continuously inhabited sites. This may occur

because upstream habitats are, in general, more suitable for

Atlantic salmon than downstream sites (Barracou 2008). Inturn, this observation may not be linked to stocking oper-

ations since hatcheries generally use a reduced numbers ofparents harbouring reduced diversity compared to the

population of origin (Araki and Schmid 2010). Indeed,

these results likely suggest a relatively low impact ofstocked Atlantic salmon, which is in line with recent

studies on the impact of stocking on the breeding system in

this species (Milot et al. 2013; Jonsson and Jonsson 2006).In the case of the Nivelle River the effective population

size appeared larger than the census size. Such a result is

difficult to explain without invoking a potentially largecontribution of precocious parr (Johnstone et al. 2013;

Saura et al. 2008; Jones and Hutchings 2001) but could also

be linked to gene flow from the Nive and Gave rivers.Indeed, census size has been estimated through an

exhaustive monitoring of adult anadromous salmon but did

not included precocious parr (Lange et al. 2011). Withinthe Nive and the Gaves rivers, the effective population

sizes tended to be smaller than the census sizes, which is

more in lines with the expectations of a typical Atlanticsalmon population (Palstra and Ruzzante 2011) in which a

large variance among breeders exists (Richard et al. 2013;

Fleming 1996). Overall, the relatively large effectivepopulation sizes estimated for recently recolonized sites

might allow for the conservation of a relatively high level

of genetic diversity in these new populations, which maylimit short-term extinction risks (Traill et al. 2010;

Frankham 2005). Accordingly, within each of the three

rivers, samples collected in continuously inhabited and inrecently recolonized sites had comparable allelic richness,

suggesting no or only a weak loss of genetic diversity

during the recolonization process. This is a critical obser-vation because effective population size and genetic

diversity is positively linked to the effectiveness of selec-

tion relative to drift (Charlesworth 2009; Olson-Manninget al. 2012).

Conclusion

Overall, our results suggest that restoring accessibility toupstream areas can allow for the recolonization of Atlantic

salmon. This recolonization mainly comes from individuals

from proximate downstream sites, with neither a decreaseof local diversity nor disruption of existing genetic

structure. Along with previous studies (Perrier et al. 2010;

Ikediashi et al. 2012; Kiffney et al. 2009; Griffiths et al.2011; Schreiber and Diefenbach 2005), this study shows

connectivity restoration as an effective way to support

recolonization of rivers from which salmonids have beenpreviously extirpated. Nevertheless, the ecological resto-

ration policy should also aim to reconnect several other

upstream tributaries and improve water quality. Indeed, theactual census size of the Atlantic salmon population in the

Adour basin is much smaller than the estimated potentialcapacity (Barracou 2008). While the implementation of

fishways in the Adour drainage allowed an effective

recolonization of Atlantic salmon within several upstreamareas, little is known about the impacts of these fish pas-

sages on the recolonization dynamic of other migratory

species. While Atlantic salmon can cross over relativelychallenging fishways, several other migratory fish having

lower swimming capacities may need more specific and

less challenging fishways to effectively recolonize depop-ulated areas. More broadly, even though the recolonization

of some fish migratory species can be enabled by fishways,

such devices may not compensate the overall ecosystem-wide dramatic impacts of dams (Brown et al. 2013).

Acknowledgments We acknowledge all participants to the collec-tion of samples and of various historical and environmental data, witha special attention to A. Manicki, J. Chat, D. Barracou. We also thankP. Regnacq, JB Torterotot and Anne Dalziel for their help whileanalyzing data and writing the paper. We thank two anonymousreviewers and the associate editor C. Primmer for their very con-structive comments. Authors also thank all French organizations thatprovided their technical assistance for electric fishing: the NationalInstitute for Agricultural Research (INRA), the National Office ofWater and Aquatic Media (ONEMA) and Migradour. This work wasfunded by the European Union INTERREG IIIB program [AtlanticSalmon Arc Project (ASAP)] and the European Union INTERREGIVB program [Atlantic Arc Resource Conservation (AARC)].

References

Aprahamian MW, Smith KM, McGinnity P, McKelvey S, Taylor J(2003) Restocking of salmonids—opportunities and limitations.Fish Res 62:211–227

Araki H, Schmid C (2010) Is hatchery stocking a help or harm?Evidence, limitations and future directions in ecological andgenetic surveys. Aquaculture 308:S2–S11. doi:10.1016/j.aquaculture.2010.05.036

Araki H, Ardren WR, Olsen E, Cooper B, Blouin MS (2007)Reproductive success of captive-bred steelhead trout in the wild:evaluation of three hatchery programs in the Hood River. ConservBiol 21(1):181–190. doi:10.1111/j.1523-1739.2006.00564.x

Araki H, Berejikian BA, Ford MJ, Blouin MS (2008) Fitness ofhatchery-reared salmonids in the wild. Evol Appl 1(2):342–355.doi:10.1111/j.1752-4571.2008.00026.x

Aronson J (2011) Ecological restoration in 2010: insight from fourcontinents. Conserv Biol 25(1):206–208. doi:10.1111/j.1523-1739.2010.01621.x

Conserv Genet

123

Barracou D (2008) Synthese de la qualite des frayeres a saumons surle bassin de l’Adour: accessibilite des geniteurs et survie desjuveniles. Atlantic Salmon Arc Project, Migradour

Baudouin L, Lebrun P (2000) An operational bayesian approach forthe identification of sexually reproduced cross-fertilized popu-lations using molecular markers. In: Dore C, Dosba F, Baril C(eds) International symposium on molecular markers for char-acterizing genotypes and identifying cultivars in horticulture.Montpellier, France, Mar 06–08 2000, pp 81–93

Beall E, Davaine P, Bazin D, Bousquet B (1995) Repeuplement duGave de Pau. Utilisation de souches etrangeres et localesdomestiquees pour le repeuplement des rivieres a Salmonides.Rapport DIREN Midi Pyrenees code INRA 2713A, Stationd’Hydrobiologie INRA, St Pee sur Nivelle

Beaudou D, Baril D, Roche B, LeBaron M, CattaneoBerrebi G,Berrebi P (1994) Recolonization in a devastated Corsican river:respective contribution of wild and domestic brown trout. In:International symposium on fish and their habitat. Lyon, France,Dec 06–08 1994, pp 259–266

Bednarek AT (2001) Undamming rivers: a review of the ecologicalimpacts of dam removal. Environ Manag 27(6):803–814. doi:10.1007/s002670010189

Blanchet S, Rey O, Etienne R, Lek S, Loot G (2010) Species-specificresponses to landscape fragmentation: implications for manage-ment strategies. Evol Appl 3(3):291–304. doi:10.1111/j.1752-4571.2009.00110.x

Boet P, Belliard J, Berrebi-dit-Thomas R, Tales E (1999) Multiplehuman impacts by the City of Paris on fish communities in theSeine river basin, France. Hydrobiologia 410:59–68

Bourret V, O’Reilly PT, Carr JW, Berg PR, Bernatchez L (2011)Temporal change in genetic integrity suggests loss of localadaptation in a wild Atlantic salmon (Salmo salar) populationfollowing introgression by farmed escapees. Heredity 106:500–510

Bourret V, Dionne M, Kent MP, Lien S, Bernatchez L (2013)Landscape Genomics in Atlantic Salmon (Salmo salar): search-ing for gene-environment interactions driving local adaptation.Evolution 67(12):3469–3487. doi:10.1111/evo.12139

Boylan P, Adams CE (2006) The influence of broad scale climaticphenomena on long term trends in Atlantic salmon populationsize: an example from the River Foyle, Ireland. J Fish Biol68(1):276–283. doi:10.1111/j.0022-1112.2006.00893.x

Brown JJ, Limburg KE, Waldman JR, Stephenson K, Glenn EP,Juanes F, Jordaan A (2013) Fish and hydropower on the U.S.Atlantic coast: failed fisheries policies from half-way technol-ogies. Conserv Lett. doi:10.1111/conl.12000

Charlesworth B (2009) Fundamental concepts in genetics: effectivepopulation size and patterns of molecular evolution and varia-tion. Nat Rev Genet 10(3):195–205. doi:10.1038/nrg2526

Clewell AF, Aronson J (2006) Motivations for the restoration ofecosystems. Conserv Biol 20(2):420–428. doi:10.1111/j.1523-1739.2006.00340.x

Corander J, Waldmann P, Marttinen P, Sillanpaa MJ (2004) BAPS 2:enhanced possibilities for the analysis of genetic populationstructure. Bioinformatics 20(15):2363–2369. doi:10.1093/bioinformatics/bth250

Coulon A, Fitzpatrick JW, Bowman R, Lovette IJ (2010) Effects ofhabitat fragmentation on effective dispersal of Florida scrub-jays. Conserv Biol 24(4):1080–1088. doi:10.1111/j.1523-1739.2009.01438.x

Coutant CC, Whitney RR (2000) Fish behavior in relation to passagethrough hydropower turbines: a review. Trans Am Fish Soc129(2):351–380

Couvet D (2002) Deleterious effects of restricted gene flow infragmented populations. Conserv Biol 16(2):369–376. doi:10.1046/j.1523-1739.2002.99518.x

Davaine P, Beall E, Glise S (1996) Repeuplement du Gave de Pau :utilisation de souches domestiquees de saumon atlantique.INRA, St Pee sur Nivelle

De Groot RS, Blignaut J, Van Der Ploeg S, Aronson J, Elmqvist T,Farley J (2013) Benefits of investing in ecosystem restoration.Conserv Biol. doi:10.1111/cobi.12158

Dillane E, McGinnity P, Coughlan JP, Cross MC, de Eyto E,Kenchington E, Prodohl P, Cross TF (2008) Demographics andlandscape features determine intrariver population structure inAtlantic salmon (Salmo salar L.): the case of the River Moy inIreland. Mol Ecol 17:4786–4800

Dionne M, Caron F, Dodson J, Bernatchez L (2008) Landscapegenetics and hierarchical genetic structure in Atlantic salmon:the interaction of gene flow and local adaptation. Mol Ecol17(10):2382–2396

Dionne M, Caron F, Dodson JJ, Bernatchez L (2009) Comparativesurvey of within-river genetic structure in Atlantic salmon;relevance for management and conservation. Conserv Genet10(4):869–879. doi:10.1007/s10592-008-9647-5

Dixo M, Metzger JP, Morgante JS, Zamudio KR (2009) Habitatfragmentation reduces genetic diversity and connectivity amongtoad populations in the Brazilian Atlantic Coastal Forest. BiolConserv 142(8):1560–1569. doi:10.1016/j.biocon.2008.11.016

Do C, Waples RS, Peel D, Macbeth GM, Tillett BJ, Ovenden JR(2013) NeEstimator v2: re-implementation of software for theestimation of contemporary effective population size (Ne) fromgenetic data. Mol Ecol Resour. doi:10.1111/1755-0998.12157

Dumas J, Prouzet P (2003) Variability of demographic parametersand population dynamics of Atlantic salmon (Salmo salar L.) ina southwest French river. ICES J Mar Sci 60(2):356–370

Earl D, vonHoldt B (2012) STRUCTURE HARVESTER: a websiteand program for visualizing STRUCTURE output and imple-menting the Evanno method. Conserv Genet Resour4(2):359–361. doi:10.1007/s12686-011-9548-7

Eldridge WH, Naish KA (2007) Long-term effects of translocationand release numbers on fine-scale population structure amongcoho salmon (Oncorhynchus kisutch). Mol Ecol16(12):2407–2421. doi:10.1111/j.1365-294X.2007.03271.x

Eldridge WH, Myers JM, Naish KA (2009) Long-term changes in thefine-scale population structure of coho salmon populations(Oncorhynchus kisutch) subject to extensive supportive breed-ing. Heredity 103(4):299–309. doi:10.1038/hdy.2009.69

Ellis JS, Sumner KJ, Griffiths AM, Bright DI, Stevens JR (2011)Population genetic structure of Atlantic salmon, Salmo salar L.,in the River Tamar, southwest England. Fish Manag Ecol18(3):233–245. doi:10.1111/j.1365-2400.2010.00776.x

Ensing D, Prodohl PA, McGinnity P, Boylan P, O’Maoileidigh N,Crozier WW (2011) Complex pattern of genetic structuring inthe Atlantic salmon (Salmo salar L.) of the River Foyle systemin northwest Ireland: disentangling the evolutionary signal frompopulation stochasticity. Ecol Evol 1(3):359–372. doi:10.1002/ece3.32

Evanno G, Regnault S, Goudet J (2005) Detecting the number ofclusters of individuals using the software STRUCTURE: asimulation. Mol Ecol 14:2611–2620

Ewers RM, Didham RK (2006) Confounding factors in the detectionof species responses to habitat fragmentation. Biol Rev81(1):117–142. doi:10.1017/s1464793105006949

Excoffier L, Lischer HEL (2010) Arlequin suite ver 3.5: a new seriesof programs to perform population genetics analyses underLinux and Windows. Mol Ecol Resour 10(3):564–567. doi:10.1111/j.1755-0998.2010.02847.x

Fahrig L (2003) Effects of Habitat fragmentation on biodiversity.Annu Rev Ecol Evol Syst 34: 487–515. doi:10.2307/30033784

Finnengan AK, Stevens JR (2008) Assessing the long-term geneticimpact of historical stocking events on contemporary

Conserv Genet

123

populations of Atlantic salmon, Salmo salar. Fish Manag Ecol15(4):315–326. doi:10.1111/j.1365-2400.2008.00616.x

Fleming IA (1996) Reproductive strategies of Atlantic salmon:ecology and evolution. Rev Fish Biol Fish 6(4):379–416. doi:10.1007/Bf00164323

Frankham R (2002) Introduction to conservation genetics. CambridgeUniversity Press, Cambridge 617 pp

Frankham R (2005) Genetics and extinction. Biol Conserv126(2):131–140. doi:10.1016/j.biocon.2005.05.002

Fraser DJ (2008) How well can captive breeding programs conservebiodiversity? A review of salmonids. Evol Appl 1(4):535–586.doi:10.1111/j.1752-4571.2008.00036.x

Fraser DJ, Bernatchez L (2001) Adaptive evolutionary conservation:towards a unified concept for defining conservation units. MolEcol 10(12):2741–2752. doi:10.1046/j.0962-1083.2001.01411.x

Fraser DJ, Jones MW, McParland TL, Hutchings JA (2007) Loss ofhistorical immigration and the unsuccessful rehabilitation ofextirpated salmon populations. Conserv Genet 8(3):527–546.doi:10.1007/s10592-006-9188-8

Garcia de Leaniz C, Fleming IA, Einum S, Verspoor E, Jordan WC,Consuegra S, Aubin-Horth N, Lajus D, Letcher BH, Youngson AF,Webb JH, Vollestad LA, Villanueva B, Ferguson A, Quinn TP(2007) A critical review of adaptive genetic variation in Atlanticsalmon: implications for conservation. Biol Rev 82(2):173–211

Garcia-Vazquez E, Moran P, Martinez JL, Perez J, de Gaudemar B,Beall E (2000) Alternative mating strategies in Atlantic salmonand brown trout. In: Symposium on DNA-based profiling ofmating systems and reproductive behaviors in poikilothermicvertebrates. New Haven, Ct, Jun 17–20 2000, pp 146–149

Gomez-Uchida D, Knight TW, Ruzzante DE (2009) Interaction oflandscape and life history attributes on genetic diversity, neutraldivergence and gene flow in a pristine community of salmonids.Mol Ecol 18(23):4854–4869. doi:10.1111/j.1365-294X.2009.04409.x

Gomez-Uchida D, Palstra FP, Knight TW, Ruzzante DE (2013)Contemporary effective population and metapopulation size (Neand meta-Ne): comparison among three salmonids inhabiting afragmented system and differing in gene flow and its asymme-tries. Ecol Evol 3(3):569–580. doi:10.1002/ece3.485

Grenouillet G, Brosse S, Tudesque L, Lek S, Baraille Y, Loot G(2008) Concordance among stream assemblages and spatialautocorrelation along a fragmented gradient. Divers Distrib14(4):592–603. doi:10.1111/j.1472-4642.2007.00443.x

Griffiths AM, Ellis JS, Clifton-Dey D, Machado-Schiaffino G, BrightD, Garcia-Vazquez E, Stevens JR (2011) Restoration versusrecolonisation: the origin of Atlantic salmon (Salmo salar L.)currently in the River Thames. Biol Conserv 144(11):2733–2738. doi:10.1016/j.biocon.2011.07.017

Hall C, Jordaan A, Frisk M (2011) The historic influence of dams ondiadromous fish habitat with a focus on river herring andhydrologic longitudinal connectivity. Landsc Ecol 26(1):95–107.doi:10.1007/s10980-010-9539-1

Hall CJ, Jordaan A, Frisk MG (2012) Centuries of anadromous foragefish loss: consequences for ecosystem connectivity and produc-tivity. Bioscience 62(8):723–731. doi:10.1525/bio.2012.62.8.5

Hansen MM, Nielsen EE, Mensberg KLD (1997) The problem ofsampling families rather than populations: relatedness amongindividuals in samples of juvenile brown trout Salmo trutta L.Mol Ecol 6(5):469–474. doi:10.1046/j.1365-294X.1997.t01-1-00202.x

Hansen MM, Fraser DJ, Meier K, Mensberg KLD (2009) Sixty yearsof anthropogenic pressure: a spatio-temporal genetic analysis ofbrown trout populations subject to stocking and populationdeclines. Mol Ecol 18:2549–2562

Haro A, Castro-Santos T, Noreika J, Odeh M (2004) Swimmingperformance of upstream migrant fishes in open-channel flow: a

new approach to predicting passage through velocity barriers.Can J Fish Aquat Sci 61(9):1590–1601

Hendry AP (2004) Selection against migrants contributes to the rapidevolution of ecologically dependent reproductive isolation. EvolEcol Res 6(8):1219–1236

Horreo JL, Machado-Schiaffino G, Griffiths A, Bright D, Stevens J,Garcia-Vazquez E (2008) Identification of differential brood-stock contribution affecting genetic variability in hatchery stocksof Atlantic salmon (Salmo salar). Aquaculture 280(1–4):89–93.doi:10.1016/j.aquaculture.2008.05.004

Horreo JL, Martinez JL, Ayllon F, Pola IG, Monteoliva JA, HelandM, Garcia-Vazquez E (2011) Impact of habitat fragmentation onthe genetics of populations in dendritic landscapes. Freshw Biol56(12):2567–2579. doi:10.1111/j.1365-2427.2011.02682.x

Ikediashi C, Billington S, Stevens JR (2012) The origins of Atlanticsalmon (Salmo salar L.) recolonizing the River Mersey innorthwest England. Ecol Evol 2(10):2537–2548. doi:10.1002/ece3.353

Johnstone DL, O’Connell MF, Palstra FP, Ruzzante DE (2013)Mature male parr contribution to the effective size of ananadromous Atlantic salmon (Salmo salar) population over30 years. Mol Ecol 22(9):2394–2407. doi:10.1111/mec.12186

Jones MW, Hutchings JA (2001) The influence of male parr body sizeand mate competition on fertilization success and effectivepopulation size in Atlantic salmon. Heredity (Edinb) 86(Pt6):675–684

Jonsson B, Jonsson N (2006) Cultured Atlantic salmon in nature: areview of their ecology and interaction with wild fish. In: ICES/NASCO symposium on interactions between aquaculture andwild stocks of atlantic salmon and other diadromous fish species.Bergen, Norway, Oct 18–21 2006, pp 1162–1181. doi:10.1016/j.icesjms.2006.03.004

Jonsson B, Jonsson N (2011) Migrations ecology of Atlantic salmonand brown trout. In: vol 33 fish & fisheries series. Springer,Netherlands, pp 247–325. doi:10.1007/978-94-007-1189-1_6

Kalinowski ST (2005) Hp-rare 1.0: a computer program forperforming rarefaction on measures of allelic richness. MolEcol Notes 5(1):187–189

Kalinowski ST (2010) The computer program STRUCTURE does notreliably identify the main genetic clusters within species:simulations and implications for human population structure.Heredity 106(4):625–632

Kalinowski ST, Wagner AP, Taper ML (2006) ml-relate: a computerprogram for maximum likelihood estimation of relatedness andrelationship. Mol Ecol Notes 6(2):576–579. doi:10.1111/j.1471-8286.2006.01256.x

Kiffney PM, Pess GR, Anderson JH, Faulds P, Burton K, Riley SC(2009) Changes in fish communities following recolonization ofthe Cedar River, Wa, USA by pacific salmon after 103 years oflocal extirpation. River Res Appl 25:438–452

King TL, Eackless MS, Letcher BH (2005) Primer note microsatelliteDNA markers for the study of Atlantic salmon (Salmo salar)kinship, population structure, and mixed-fishery analyses. MolEcol Notes 5:130–132

Kuparinen A, Tufto J, Consuegra S, Hindar K, Merila J, de Leaniz CG(2010) Effective size of an Atlantic salmon (Salmo salar L.)metapopulation in Northern Spain. Conserv Genet 11(4):1559–1565

Lange F, Prevost E, Brun M (2011) Les populations de saumons,truites de mer et grandes aloses de la Nivelle en 2010. INRA, StPee sur Nivelle

Larinier M, Boyer-Bernard S (1991) Devalaison des smolts etefficacite d’un exutoire de devalaison a l’usine hydroelectriqued’Halsou sur la Nive. Bull Fr Peche Piscic 321:72–92

Mallen-Cooper M, Brand DA (2007) Non-salmonids in a salmonidfishway: what do 50 years of data tell us about past and future

Conserv Genet

123

fish passage? Fish Manag Ecol 14(5):319–332. doi:10.1111/j.1365-2400.2007.00557.x

Marie AD, Bernatchez L, Garant D (2010) Loss of genetic integritycorrelates with stocking intensity in brook charr (Salvelinusfontinalis). Mol Ecol 19:2025–2037

Martinez JL, Moran P, Perez J, De Gaudemar B, Beall E, Garcia-Vazquez E (2000) Multiple paternity increases effective size ofsouthern Atlantic salmon populations. Mol Ecol 9(3):293–298

Marty A (1984) Le saumon dans les basins de l’Adour et de laNivelle. Synthesis report ONEMA, Pau

Marty A, Bousquet B (2001) SITUATION DES POISSONS MIG-RATEURS AMPHIHALINS SUR LEBASSIN DE L’ADOUR.Bulletin Francais de La Peche et de la Piciculture.(357–360):345–356. doi:10.1051/kmae/2001054

McConnell SKJ, O’Reilly P, Hamilton L, Wright JM, Bentzen P(1995) Polymorphic microsatellite loci from Atlantic Salmon(Salmo salar): genetic differentiation of North American andEuropean populations. Can J Fish Aquat Sci 52:1863–1872

McGinnity P, de Eyto E, Cross TF, Coughlan J, Whelan K, FergusonA (2007) Population specific smolt development, migration andmaturity schedules in Atlantic salmon in a natural riverenvironment. Aquaculture 273(2–3):257–268. doi:10.1016/j.aquaculture.2007.10.008

Milot E, Perrier C, Papillon L, Dodson JJ, Bernatchez L (2013)Reduced fitness of Atlantic salmon released in the wild after onegeneration of captive breeding. Evol Appl 6(3):472–485. doi:10.1111/eva.12028

Moran P, Pendas AM, Beall E, GarciaVazquez E (1996) Geneticassessment of the reproductive success of Atlantic salmonprecocious parr by means of VNTR loci. Heredity77(6):655–660. doi:10.1038/Hdy.1996.193

Naish KA, Seamons TR, Dauer MB, Hauser L, Quinn TP (2013)Relationship between effective population size, inbreeding andadult fitness-related traits in a steelhead (Oncorhynchus mykiss)population released in the wild. Mol Ecol. doi:10.1111/mec.12185

Nei M, Tajima F, Tateno Y (1983) Accuracy of estimated phyloge-netic trees from molecular-data.2. Gene-frequency data. J MolEvol 19(2):153–170

Neraas LP, Spruell P (2001) Fragmentation of riverine systems: thegenetic effects of dams on bull trout (Salvelinus confluentus) inthe Clark Fork River system. Mol Ecol 10:1153–1164

Nikolic N, Butler JRA, Bagliniere JL, Laughton R, McMyn IAG,Chevalet C (2009) An examination of genetic diversity andeffective population size in Atlantic salmon populations. GenetRes 91(6):395–412. doi:10.1017/s0016672309990346

Nilsson C, Reidy CA, Dynesius M, Revenga C (2005) Fragmentationand flow regulation of the world’s large river systems. Science308(5720):405–408. doi:10.1126/science.1107887

O’Reilly PT, Hamilton L, McConnell SKJ, Wright JM (1996) Rapidanalysis of genetic variation in Atlantic salmon (Salmo salar) byPCR multiplexing of dinucleotide and tetranucleotide microsat-ellites. Can J Fish Aquat Sci 53:2292–2298

Oakley CG (2013) Small effective size limits performance in a novelenvironment. Evol Appl 6(5):823–831. doi:10.1111/Eva.12068

Oldani NO, Baigun CRM (2002) Performance of a fishway system ina major South American dam on the Parana River (Argentina–Paraguay). River Res Appl 18(2):171–183. doi:10.1002/rra.640

Olson-Manning CF, Wagner MR, Mitchell-Olds T (2012) Adaptiveevolution: evaluating empirical support for theoretical predic-tions. Nat Rev Genet 13(12):867–877. doi:10.1038/nrg3322

Van Oosterhout C, Hutchinson WF, Wills DPM, Shipley P (2004)Micro-checker: software for identifying and correcting genotyp-ing errors in microsatellite data. Mol Ecol Notes 4:535–538

Page RDM (1996) TreeView: an application to display phylogenetictrees on personal computers. Comput Appl Biosci 12(4):357–358

Palstra FP, Ruzzante DE (2011) Demographic and genetic factorsshaping contemporary metapopulation effective size and itsempirical estimation in salmonid fish. Heredity 107(5):444–455.http://www.nature.com/hdy/journal/v107/n5/suppinfo/hdy201131s1.html

Parrish DL, Behnke RJ, Gephard SR, McCormick SD, Reeves GH(1998) Why aren’t there more Atlantic salmon (Salmo salar)?Can J Fish Aquat Sci 55(S1):281–287. doi:10.1139/d98-012

Paterson S, Piertney SB, Knox D, Gilbey J, Verspoor E (2004)Characterization and PCR multiplexing of novel highly variabletetranucleotide Atlantic salmon (Salmo salar L.) microsatellites.Mol Ecol Notes 4:160–162

Pearse DE, Martinez E, Garza JC (2011) Disruption of historicalpatterns of isolation by distance in coastal steelhead. ConservGenet 12(3):691–700. doi:10.1007/s10592-010-0175-8

Pedersen S, Rasmussen G, Nielsen EE, Karlsson L, Nyberg P (2007)Straying of Atlantic salmon, Salmo salar, from delayed andcoastal releases in the Baltic Sea, with special focus on theSwedish west coast. Fish Manag Ecol 14(1):21–32

Pepino M, Rodrıguez MA, Magnan P (2012) Fish dispersal infragmented landscapes: a modeling framework for quantifyingthe permeability of structural barriers. Ecol Appl22(5):1435–1445. doi:10.1890/11-1866.1

Perrier C, Evanno G, Belliard J, Guyomard R, Bagliniere JL (2010)Natural recolonization of the Seine River by Atlantic salmon(Salmo salar) of multiple origins. Can J Fish Aquat Sci67(1):1–4. doi:10.1139/F09-190

Perrier C, Grandjean F, Le Gentil J, Cherbonnel C, Evanno G (2011a)A species-specific microsatellite marker to discriminate Euro-pean Atglantic salmon, brown trout, and their hybrids. ConservGenet Resour 3(1):131–133

Perrier C, Guyomard R, Bagliniere J-L, Evanno G (2011b) Determi-nants of hierarchical genetic structure in Atlantic salmonpopulations: environmental factors vs. anthropogenic influences.Mol Ecol 20(20):4231–4245. doi:10.1111/j.1365-294X.2011.05266.x

Perrier C, Bagliniere J-L, Evanno G (2013a) Understanding admix-ture patterns in supplemented populations: a case study com-bining molecular analyses and temporally explicit simulations inAtlantic salmon. Evol Appl 6(2):218–230. doi:10.1111/j.1752-4571.2012.00280.x

Perrier C, Guyomard R, Bagliniere J-L, Nikolic N, Evanno G (2013b)Changes in the genetic structure of Atlantic salmon populations overfour decades reveal substantial impacts of stocking and potentialresiliency. Ecol Evol 3(7):2334–2349. doi:10.1002/ece3.629

Piry S, Alapetite A, Cornuet JM, Paetkau D, Baudouin L, Estoup A(2004) GENECLASS2: a software for genetic assignment andfirst-generation migrant detection. J Hered 95(6):536–539.doi:10.1093/jhered/esh074

Poulet N (2007) Impact of weirs on fish communities in a piedmontstream. River Res Appl 23(9):1038–1047. doi:10.1002/rra.1040

Primmer CR (2011) Genetics of local adaptation in salmonid fishes.Heredity 106(3):401–403

Primmer CR, Veselov AJ, Zubchenko A, Poututkin A, Bakhmet I,Koskinen MT (2006) Isolation by distance within a river system:genetic population structuring of Atlantic salmon, Salmo salar,in tributaries of the Varzuga River in northwest Russia. Mol Ecol15:653–666

Pritchard JK, Stephens P, Donelly P (2000) Inference of populationstructure using multilocus genotype data. Genetics 155:945–959

Prouzet P (1990) Stock characteristics of Atlantic salmon Salmo salarin france a review. Aquat Living Resour 3(2):85–98

Quinn TP (1993) a review of homing and straying of wild andhatchery-produced salmon. Fish Res 18:29–44

Raeymaekers JAM, Raeymaekers D, Koizumi I, Geldof S, VolckaertFAM (2009) Guidelines for restoring connectivity around water

Conserv Genet

123

mills: a population genetic approach to the management ofriverine fish. J Appl Ecol 46(3):562–571. doi:10.1111/j.1365-2664.2009.01652.x

Raymond M, Rousset F (1995a) An exact test for populationdifferentiation. Evolution 49:1280–1283

Raymond M, Rousset F (1995b) GENEPOP (version 1.2): populationgenetics software for exact tests and ecumenicism. J Hered86:248–249

Richard A, Dionne M, Wang J, Bernatchez L (2013) Does catch andrelease affect the mating system and individual reproductivesuccess of wild Atlantic salmon (Salmo salar L.)? Mol Ecol22(1):187–200. doi:10.1111/mec.12102

Saura M, Caballero A, Caballero P, Moran P (2008) Impact ofprecocious male parr on the effective size of a wild population ofAtlantic salmon. Freshw Biol 53(12):2375–2384. doi:10.1111/j.1365-2427.2008.02062.x

Schreiber A, Diefenbach G (2005) Population genetics of theEuropean trout (Salmo trutta L.) migration system in the riverRhine: recolonisation by sea trout. Ecol Freshw Fish 14(1):1–13.doi:10.1111/j.1600-0633.2004.00072.x

Slettan A, Olsaker I, Lie O (1995) Atlantic Salmon microsatellites atthe SSOSL25, SSOSL85 SSOSL311 and SSOSL417 loci. AnimGenet 26:281

Stabell OB (1984) Homing and olfaction in Salmonids—a criticalreview with special reference to the Atlantic salmon. Biol RevCamb Philos Soc 59:333–388

Taylor EB (1991) A review of local adaptation in salmonidae, withparticular reference to pacific and Atlantic salmon. Aquaculture98(1–3):185–207

Theriault V, Moyer GR, Jackson LS, Blouin MS, Banks MA (2011)Reduced reproductive success of hatchery coho salmon in thewild: insights into most likely mechanisms. Mol Ecol20(9):1860–1869. doi:10.1111/j.1365-294X.2011.05058.x

Tonteri A, Veselov AJ, Zubchenko AV, Lumme J, Primmer CR(2009) Microsatellites reveal clear genetic boundaries amongAtlantic salmon (Salmo salar) populations from the Barents andWhite seas, northwest Russia. Can J Fish Aquat Sci66(5):717–735. doi:10.1139/f09-010

Traill LW, Brook BW, Frankham RR, Bradshaw CJA (2010)Pragmatic population viability targets in a rapidly changingworld. Biol Conserv 143(1):28–34. doi:10.1016/j.biocon.2009.09.001

Tsuboi J-I, Iwata T, Morita K, Endou S, Oohama H, Kaji K (2013)Strategies for the conservation and management of isolatedsalmonid populations: lessons from Japanese streams. FreshwBiol 58(5):908–917. doi:10.1111/fwb.12096

Vaha JP, Erkinaro J, Niemela E, Primmer CR (2007) Life-history andhabitat features influence the within-river genetic structure ofAtlantic salmon. Mol Ecol 16(13):2638–2654. doi:10.1111/j.1365-294X.2007.03329.x

Vaha JP, Erkinaro J, Niemela E, Primmer CR (2008) Temporallystable genetic structure and low migration in an Atlantic salmonpopulation complex: implications for conservation and manage-ment. Evol Appl 1(1):137–154. doi:10.1111/j.1752-4571.2007.00007.x

Van Oort H, McLellan BN, Serrouya R (2011) Fragmentation,dispersal and metapopulation function in remnant populations ofendangered mountain caribou. Anim Conserv 14(3):215–224.doi:10.1111/j.1469-1795.2010.00423.x

Van Oosterhout C, Hutchinson WF, Wills DPM, Shipley P (2004)MICRO-CHECKER: software for identifying and correctinggenotyping errors in microsatellite data. Mol Ecol Notes4(3):535–538. doi:10.1111/j.1471-8286.2004.00684.x

Vauclin V (2007) French Implementation plan of NASCO’s resolu-tions and agreements with regard to the protection, the manage-ment and the exploitation of the Atlantic salmon and its Habitats.Report ONEMA, Orleans

Vitousek PM, Mooney HA, Lubchenco J, Melillo JM (1997) Humandomination of earth’s ecosystems. Science 277(5325):494–499.doi:10.1126/science.277.5325.494

Waples RS, Do C (2008) ldne: a program for estimating effectivepopulation size from data on linkage disequilibrium. Mol EcolResour 8(4):753–756. doi:10.1111/j.1755-0998.2007.02061.x

Waples RS, England PR (2011) Estimating contemporary effectivepopulation size on the basis of linkage disequilibrium in the faceof migration. Genetics 189(2):633–644. doi:10.1534/genetics.111.132233

Whiteley AR, Coombs JA, Hudy M, Robinson Z, Colton AR, NislowKH, Letcher BH (2013) Fragmentation and patch size shapegenetic structure of brook trout populations. Can J Fish AquatSci 70(5):678–688. doi:10.1139/cjfas-2012-0493

Winans GA, Baird MC, Baker J (2010) A genetic and pheneticbaseline before the recolonization of steelhead above HowardHanson Dam, Green River, Washington. North Am J Fish Manag30(3):742–756

Wofford JEB, Gresswell RE, Banks MA (2005) Influence of barriersto movement on within-watershed genetic variation of coastalcutthroat trout. Ecol Appl 15(2):628–637. doi:10.1890/04-0095

Conserv Genet

123