DataPDF Available

The Pillars of Hercules as a bathymetric barrier to gene flow promoting isolation in a global deep-sea shark (Centroscymnus coelolepis)

Authors:
The Pillars of Hercules as a bathymetric barrier to gene
flow promoting isolation in a global deep-sea shark
(Centroscymnus coelolepis)
DIANA CATARINO,*HALVOR KNUTSEN,§¶ ANA VER
ISSIMO,**†† ESBEN MOLAND
OLSEN,§¶ PER ERIK JORDE,GUI MENEZES,*HANNE SANNÆ S,DAVID STANKOVI
C,‡‡
JOAN BAPTISTA COMPANY,§§ FRANCIS NEAT,¶¶ ROBERTO DANOVARO,***††† ANTONIO
DELL’ANNO,*** BASTIEN ROCHOWSKI‡‡‡ and SE R GIO STE F ANN I §§§
*MAREMarine and Environmental Sciences Centre, Department of Oceanography and Fisheries, University of the Azores,
Rua Prof. Dr. Frederico Machado, Horta, Azores, Portugal, IMAR-Institute of Marine Research, Department of Oceanography
and Fisheries, University of the Azores, Rua Prof. Dr. Frederico Machado, Horta, Azores, Portugal, Institute of Marine
Research (IMR), Flødevigen N-4817 His, Norway, §University of Agder, Kristiansand N-4604, Norway, Department of
Biosciences, University of Oslo, Centre for Ecological and Evolutionary Synthesis (CEES), PO Box 1066 Blindern, Oslo N-0316,
Norway, **CIBIO-U.P., Centro de Investigacß
~
ao em Biodiversidade e Recursos Gen
eticos, Campus Agr
ario de Vair~
ao, Rua Padre
Armando Quintas, Vair~
ao 4485-661, Portugal, ††College of William and Mary, Virginia Institute of Marine Science, Route
1208, Greate Road, Gloucester Point 23062, VA, USA, ‡‡Department of Life Sciences, University of Trieste, Via Licio Giorgieri
5, Trieste 34127, Italy, §§Institut de Ci
encies del Mar, CSIC, Passeig Mar
ıtim de la Barceloneta 37-49, Barcelona 08003, Spain,
¶¶Marine Laboratory, Marine Scotland-Science, PO Box 101, 375 Victoria Road, Aberdeen AB11 9DB, UK, ***Department of
Life and Environmental Sciences, Universit
a Politecnica delle Marche, Ancona 60131, Italy, †††Stazione Zoologica Anton
Dohrn, Villa Comunale, Naples, Italy, ‡‡‡School of BioSciences, The University of Melbourne, Parkville, Vic. 3010, Australia,
§§§CNR-ISSIA, Via de Marini 6, Genova 16149, Italy
Abstract
Knowledge of the mechanisms limiting connectivity and gene flow in deep-sea ecosys-
tems is scarce, especially for deep-sea sharks. The Portuguese dogfish (Centroscymnus
coelolepis) is a globally distributed and near threatened deep-sea shark. C. coelolepis pop-
ulation structure was studied using 11 nuclear microsatellite markers and a 497-bp frag-
ment from the mtDNA control region. High levels of genetic homogeneity across the
Atlantic (Φ
ST
=0.0091, F
ST
=0.0024, P>0.05) were found suggesting one large popula-
tion unit at this basin. The low levels of genetic divergence between Atlantic and Aus-
tralia (Φ
ST
=0.0744, P<0.01; F
ST
=0.0015, P>0.05) further suggested that this species
may be able to maintain some degree of genetic connectivity even across ocean basins. In
contrast, sharks from the Mediterranean Sea exhibited marked genetic differentiation
from all other localities studied (Φ
ST
=0.3808, F
ST
=0.1149, P<0.001). This finding sug-
gests that the shallow depth of the Strait of Gibraltar acts as a barrier to dispersal and that
isolation and genetic drift may have had an important role shaping the Mediterranean
shark population over time. Analyses of life history traits allowed the direct comparison
among regions providing a complete characterization of this sharks populations. Sharks
from the Mediterranean had markedly smaller adult body size and size at maturity com-
pared to Atlantic and Pacific individuals. Together, these results suggest the existence of
an isolated and unique population of C. coelolepis inhabiting the Mediterranean that
most likely became separated from the Atlantic in the late Pleistocene.
Keywords: barriers to dispersal, deep-sea shark, isolation, Mediterranean, population structure
Received 10 April 2014; revision received 29 October 2015; accepted 2 November 2015
Correspondence: Diana Catarino, Fax: +351 292200411;
E-mail: dianacatarino@uac.pt
©2015 John Wiley & Sons Ltd
Molecular Ecology (2015) 24, 6061–6079 doi: 10.1111/mec.13453
Introduction
The initial idea of large marine populations approximat-
ing panmixia (Caley et al. 1996), showing low genetic
differentiation (Palumbi 1992; Ward et al. 1994; Waples
1998), has recently been challenged (reviewed by Hau-
ser & Carvalho 2008; Salmenkova 2011). Ocean currents,
bathymetry, habitat fragmentation, limited dispersal
and migration have been shown to limit gene flow and
connectivity for many species across their distribution,
promoting genetic divergence. Furthermore, past cli-
matic events such as ice ages have led to the strength-
ening of barriers and promoted vicariance (e.g.
Chevolot et al. 2006; Robalo et al. 2012). Despite some
progress made towards understanding the mechanisms
behind genetic structure in marine organisms, our
knowledge remains scarce regarding deep-sea species
(those living below 200 m depth). The logistic difficul-
ties in sample collection, the limited number of speci-
mens available and taxonomic uncertainty, have
hampered population genetic studies in deep-sea taxa.
Spatial genetic divergence in deep-sea fishes is
thought to be generally low or nearly absent (e.g. Ball
et al. 2000; White et al. 2009; Longmore et al. 2014).
However, there are a few notable exceptions of genetic
differentiation, as in the roundnose grenadier (Knutsen
et al. 2012), the bluemouth rockfish (Aboim et al. 2005)
and the blackspot sea bream (Stockley et al. 2005). These
species differ from each other in their depth distribu-
tion and general life history, and thus, the barriers to
connectivity and mechanisms of differentiation are
likely to be different.
While the number of genetic studies performed in
coastal and pelagic sharks has increased (reviewed by
Dudgeon et al. 2012), few studies have focused on pat-
terns of genetic structure in deep-sea sharks (Straube
et al. 2011; Ver
ıssimo et al. 2011a, 2012; Cunha et al.
2012). These studies indicate that deep-sea sharks main-
tain high levels of gene flow within ocean and some-
times across oceans basins, despite lacking a dispersive
pelagic early life stage (Conrath & Musick 2012). As dis-
tance seems less important in reducing genetic connec-
tivity, deep-sea sharks are an interesting group to test
hypotheses on other factors limiting connectivity.
Just as bathymetric features, such as troughs and
deep basins, can act as barriers to shallow water species
(Knutsen et al. 2009), shallow water passages, sills or
plateaus, are potential factors that may act as barriers
for deep-sea fishes. For instance, the Mediterranean Sea
(hereafter referred to as Mediterranean) is the youngest
deep-sea basin in the world (Emig & Geistdoerfer 2004)
and is connected to the Atlantic Ocean only by the nar-
row and shallow Strait of Gibraltar (12.9 km wide,
284 m deep). The Strait of Gibraltar was also known in
the classical antiquity as the ‘Pillars of Hercules’, due to
the promontories that flank its entrance. The average
depth of the Mediterranean is 1500 m, with the western
basin reaching 3400 m and the eastern reaching 5267 m.
The Mediterranean has been shaped by turbulent geo-
logical events, such as the Messinian salinity crisis (i.e.
the reduction in the Atlantic water inflow resulted in
Mediterranean desiccation c. 5.6 million years ago; Hs
u
et al. 1973; Krijgsman et al. 1999), which led to massive
extinctions. After this event, the Atlantic marine biota
recolonized the Mediterranean arriving in different and
separated demographic waves depending on their bio-
geographic affinities (warmer or colder) mostly during
the drastic climatic fluctuations occurring in the Pleis-
tocene (reviewed by Emig & Geistdoerfer 2004). The
Strait of Gibraltar acts as a physical barrier isolating the
deep-sea Mediterranean from the adjacent Atlantic
Ocean (Smith & Gale 2009), resulting in impoverished
deep-sea communities in the Mediterranean compared
to the Atlantic (Emig & Geistdoerfer 2004). The Strait of
Gibraltar has been reported as a genetic break for shal-
low water and pelagic marine taxa, as some species dis-
play marked genetic differences between the eastern
North Atlantic and the Mediterranean. However, this is
still controversial as other species do not exhibit such
pattern (reviewed by Patarnello et al. 2007). Genetic
studies on exclusively deep-sea species occurring on
either side of the Strait of Gibraltar are currently lack-
ing. Therefore, the Atlantic/Mediterranean transition is
a very interesting area to study genetic connectivity in
deep-sea species.
The Portuguese dogfish (Centroscymnus coelolepis Bar-
bosa du Bocage & de Brito Capello 1864) has been eco-
nomically exploited and is one of the most widespread
deep-sea sharks, reported all over the Atlantic Ocean,
in the western Mediterranean, in the Indian and Wes-
tern Pacific Oceans (Compagno et al. 2005). This deep-
water squaloid shark (order Squaliformes) inhabits
continental and island slopes, seamounts and ocean
ridges (Ebert & Stehmann 2013), and its vertical distri-
bution ranges from 128 to 3675 m depth (Ebert &
Stehmann 2013). In the Mediterranean, the deepest
specimen was collected at 2863 m (Grey 1956). The Por-
tuguese dogfish is a yolk-sac viviparous shark that
gives birth to 12 pups per litter on average and has a
long reproductive cycle with a gestation period of two
or more years (Yano & Tanaka 1988; Girard & Du Buit
1999; Ver
ıssimo et al. 2003; Figueiredo et al. 2008). Spe-
cies longevity is still unknown (Irvine 2004; Cotton
2010), but it is likely to be as high as the maximum esti-
mated age for a close relative, Centroselachus crepidater
(54 years old; Irvine et al. 2006).
©2015 John Wiley & Sons Ltd
6062 D. CATARINO ET AL.
The species is of commercial interest in the eastern
North Atlantic and in Australian waters (Daley et al.
2002; ICES 2012), where intense fishing mortality led to
the stock’s decline and resulted in strict conservation
policies. High fishing pressure together with the low
fecundity and expected long longevity of C. coelolepis
led the International Union for Conservation of Nature
(IUCN) to recently classify the species as globally near
threatened (Stevens & Correia 2003). The European
Union has implemented a zero catch policy for several
deep-water sharks since 2010 including C. coelolepis, and
no by-catch landings are allowed since 2012 (Council
Regulation (EU) No 1225/2010, Council Regulation
(EU) No 1359/2008). In the Mediterranean, commercial
fisheries usually operate down to 800 m depth, and
fisheries below 1000 m depth were banned since 2005
(FAO 2006). Thus, the species is presently not targeted
commercially, being classified in the Mediterranean as
Least Concerned (IUCN; Bradai et al. 2012).
A recent genetic study (Ver
ıssimo et al. 2011a) showed
that the species is genetically undifferentiated and proba-
bly consists in a single panmictic unit along the eastern
Atlantic from Ireland to South Africa. Despite the
observed genetic homogeneity over such a large spatial
scale, the species exhibits differences in life history traits
when considering the Mediterranean. In the Atlantic,
C. coelolepis pups are born around 30 cm total length (TL)
and grow to a maximum of 122 cm TL (Girard & Du Buit
1999). Females attain larger sizes than males, with males
maturing around 7086 cm TL and females around 95
102 cm TL in the Atlantic and Pacific Oceans (Yano &
Tanaka 1988; Girard & Du Buit 1999; Clarke et al. 2001;
Ver
ıssimo et al. 2003). In the Mediterranean, body size is
considerably smaller: 22 cm TL at birth (Torchio &
Michelangeli 1971) and reaching a maximum size of
65 cm TL (Cl
oet al. 2002). Size at maturity is different in
the Mediterranean compared to the Atlantic and Pacific
Oceans, and fecundity is also expected to be lower in the
Mediterranean (Yano & Tanaka 1988; Cl
oet al. 2002).
These striking differences in life history traits between
the Mediterranean and other regions of C. coelolepis range
strongly suggest that the former may harbour a distinct
population of Portuguese dogfish; however, this hypoth-
esis has never been tested.
The aim of the present study was to investigate the
population genetic structure of C. coelolepis, with special
emphasis on the Atlantic/Mediterranean transition, and
to assess to what extent the Strait of Gibraltar may have
acted as a barrier to gene flow, considering the species
bathymetric distribution. As far as we know, this study
is the first to analyse the genetic connectivity in the
Atlantic/Mediterranean transition in a bathyal species.
We further investigate large- (Atlantic/Pacific) and
small-scale (within the Mediterranean) connectivity. Life
history traits, such as body size and size at maturity,
were collected from different locations along the species
range to allow for direct comparison among regions
and to provide a complete characterization of this
species’ populations from the studied areas.
Material and Methods
Sampling
Specimens of Centroscymnus coelolepis were collected by
research vessels or commercial fishing vessels (before
fishery closure) from nine main locations (Fig. 1; Tables 1
and S1, Supporting information): Azores (AZ), South of
Azores (SAZ; Irving and Great Meteor seamounts com-
bined), Madeira (MAD), Portugal mainland (PT), Ireland
(IRE), United Kingdom (UK), Mauritania (MAU), west-
ern Mediterranean (MED) and Australia (AUS). Samples
from IRE and MAU and part of the samples from MAD
and PT were previously used in the study by Ver
ıssimo
et al. (2011a; see Table 1 for details). Samples from the
AZ and SAZ were collected using a ‘stone-buoy’ bottom
long-line gear (Menezes et al. 2009). Sharks from the UK
(Campbell et al. 2011) and AUS were collected using
deep-water bottom trawl nets. Sampling in the Mediter-
ranean was performed using baited traps or deep-water
bottom trawl nets (Tosti et al. 2006; Tecchio et al. 2011).
Biological data (length, weight, sex, maturation) were col-
lected on board from fresh specimens or in the laboratory
from specimens frozen on board. Small portions of white
muscle or fin clips were sampled for genetic analyses and
preserved in 95% ethanol or DMSO buffer saturated with
NaCl (Seutin et al. 1991) prior to DNA extraction.
DNA extraction and PCR amplification
Mitochondrial DNA (mtDNA). Genomic DNA was
extracted using the E.Z.N.A.
â
(EaZy Nucleic Acid Isola-
tion) Mag-Bind Tissue DNA kit (OMEGA Bio-Tek,
USA) according to the manufacturer’s instructions with
the KingFisher mL magnetic particle processor
(ThermoElectron Corporation, USA). The mitochon-
drial control region (CR) was partially amplified for 155
individuals (this study) by polymerase chain reaction
(PCR) using the primers Pro-L and CoeCRH (Keeney
et al. 2003; Ver
ıssimo et al. 2011a). PCR amplification fol-
lowed published protocols (Ver
ıssimo et al. 2011a) with
slight modifications (Appendix S1, Supporting informa-
tion). No template controls were included to check for
possible contamination. Electrophoresis of PCR prod-
ucts was performed on a 1% agarose gel to evaluate the
integrity of the products. Finally, all amplified products
were purified using ExoSAP-IT (USB Corporation,
USA) and sequenced commercially at BMR Genomics
©2015 John Wiley & Sons Ltd
POPULATION STRUCTURE OF THE PORTUGUESE DOGFISH 6063
(www.bmr-genomics.it, Padua, Italy). All haplotypes
were confirmed by sequencing both the forward and
reverse strands for at least one individual per haplotype.
Nuclear DNA (microsatellites). Genomic DNA was
extracted using the E.Z.N.A.
â
Tissue DNA Kit (OMEGA
Bio-Tek) according to the manufacturer’s instructions. A
total of 503 sharks (Table 1) were genotyped for 11
nuclear microsatellite loci described by Ver
ıssimo et al.
(2010, 2011b). Primers were combined in three different
sets (Table S2, Supporting information) after optimiza-
tion. Sets 1 and 3 were amplified in multiplexes of four
loci, whereas set 2 was amplified as a duplex for the loci
Ccoe13 and Ccoe55 and separately for locus Ccoe61.
Amplification followed published protocols (Ver
ıssimo
et al. 2011b) with some modifications (Appendix S1,
Supporting information). PCR products were run for
fragment analyses on an ABI Prism 3130xl (Life Tech-
nologies, USA). All genotypes were scored indepen-
dently by three persons using GENEMAPPER software
(version 4.0; Life Technologies), and inconsistencies
were resolved by replicate runs. Genotypes that could
not be resolved after replicate runs were omitted from
further analyses.
Combined samples
To increase sample sizes per geographical region, some
of the original sample collections were combined based
upon geographical proximity and absence of genetic
divergence between them (exact G-test, see Results
section), resulting in nine (eight for microsatellite data)
Fig. 1 Portuguese dogfish sampled col-
lection sites. Stars indicate samples used
only for mtDNA. Sampled sites abbrevia-
tions are given in Table 1.
©2015 John Wiley & Sons Ltd
6064 D. CATARINO ET AL.
composite collections for the analyses of spatial genetic
structure (see Table 1 for details). Thus, samples from
the Azores, collected from 2003 to 2011, in several geo-
graphically adjacent collection sites around the islands
and seamounts, were combined in one collection (AZ or
AZ/MAR: cf. Table 1). Mediterranean samples were
obtained in several years (2001, 2009 and 2012) at three
main locations: northeast of Barcelona (MED1), south-
east of Ibiza (MED2) and southwest of Sardinia (MED3;
see Fig. 1 and Table 1), and were combined into a sin-
gle MED collection. Likewise, IRE and UK samples
were combined (UK/IRE) based upon the same criteria.
Atlantic results (ATL) refer to all sampled localities
within the Atlantic Ocean (excluding MED), including
South Africa (SAF) in the case of mtDNA.
Statistical analysis
mtDNA sequence analyses. All sequences available for
this species at the National Center for Biotechnology
Information (NCBI) database (Accession nos HQ664432
HQ664449, Ver
ıssimo et al. 2011a) were also included in
the mtDNA CR alignment, with the respective haplo-
type frequencies. This added 192 additional samples to
the data set, for a total of 347 (see Table 1 for details).
All sequences were aligned using SEAVIEW (Galtier et al.
1996) and CLUSTALX (version 1.8.3.; Thompson et al.
1997).
Comparisons among sampling localities and relation-
ships among CR haplotypes were analysed using the
median-joining network method (Bandelt et al. 1999)
and the maximum parsimony approach (Polzin &
Daneschmand 2003), estimated with the NETWORK soft-
ware version 4.6.1.0. (fluxus-engineering.com) using
default parameters and a nucleotide mutation weight of
10 for transitions/transversions, and 30 for insertions/
deletions, as suggested in the manual. A transition-
to-transversion ratio of 3:1 was estimated from the data
and used as a parameter (DNASP version 5.1; Rozas et al.
2003).
Table 1 List of sampling sites with number of Centroscymnus coelolepis specimens collected for genetic analyses (N), and genetic
diversity indices for the microsatellites and mtDNA control region (CR) markers
Sampling area Code Sampling date (year)
Nuclear microsatellites Mitochondrial CR
NH
S
F
IS
AR
s
NH
n
H
d
p
AZ and MAR combined AZ/MAR —— 68 11 (2) 0.706 0.0020
Azores AZ 20032011 115 0.821 0.014 16.5 (11) 8.9 28 9 (2) 0.778 0.0024
Mid-Atlantic Ridge MAR 2004 —— 40* 7 0.660 0.0018
South of the Azores SAZ 2007 25 0.795 0.036 10.5 (4) 8.3 10 3 0.600 0.0015
UK and IRE combined UK/IRE 68 0.818 0.018 14.2 (6) 8.7 48 8 (2) 0.597 0.0017
Ireland IRE 2006, 2007 35
0.817 0.030 12.0 (4) 8.6 38
8 (2) 0.636 0.0019
United Kingdom UK 2011 33 0.820 0.007 11.9 (2) 8.6 10 2 0.467 0.0009
Mauritania MAU 2007 46
0.828 0.065 14.2 (9) 9.2 40
7 0.630 0.0017
Mediterranean combined MED 151 0.657 0.041 10.0 (3) 5.6 55 7 (5) 0.579 0.0015
Northeast of Barcelona MED1 2009, 2012 105 0.665 0.024 9.5 (1) 5.7 39 5 (3) 0.566 0.0014
Southeast of Ibiza MED2 2009 15 0.651 0.021 6.0 5.8 ——
Southwest of
Sardinia
MED3 2001, 2009 31 0.628 0.114 6.6 (1) 5.3 16 4 (1) 0.642 0.0018
Portugal mainland PT 2005, 2006, 2011 37
0.812 0.002 12.3 (4) 8.7 52
9 0.614 0.0015
Madeira MAD 2003 22 0.820 0.044 10.1 (1) 8.6 22
7 (3) 0.671 0.0018
Australia AUS 2009, 2010 39 0.809 0.009 12.5 8.5 27 7 (3) 0.712 0.0023
South Africa SAF 2008 —— 25*6 0.693 0.0018
Southeast of Ibiza samples were not sequenced because the tissue was prioritized for microsatellite analyses, as these markers are
more variable.
H
S
, estimated mean gene diversity; F
is
, inbreeding coefficient; A, mean number of alleles (number of unique alleles); R
s
, mean allelic
richness (minimum sample size of 13 diploid individuals); H
n
, number of haplotypes (unique haplotypes); H
d
, haplotype diversity;
p, nucleotide diversity.
*Samples from MAR and SAF were only available for mtDNA and were sequenced by Ver
ıssimo et al. (2011a), all available at the NCBI.
Samples used before by Ver
ıssimo et al. (2011a) and genotyped again for the 11 microsatellite markers used in this study. For PT
sampling, only 25 of the 37 samples used in the microsatellites genotyping were from Ver
ıssimo et al. (2011a), as the DNA was
degraded. The remaining 12 samples came from recent collected specimens at this location.
Samples sequenced by Ver
ıssimo et al. (2011a) and available at NCBI. To this previous data set, 10 new sequences were add for PT
sampling and seven for MAD sampling.
In bold are the main sampled locations analysed.
©2015 John Wiley & Sons Ltd
POPULATION STRUCTURE OF THE PORTUGUESE DOGFISH 6065
Genetic diversity indices, including number of haplo-
types (H
n
), haplotype diversity (H
d
) and nucleotide
diversity (p), were calculated for each of the nine sam-
pling localities. Genetic distance (Φ
ST
) between localities
was estimated based on the mean number of pairwise
differences among sequences. All analyses were per-
formed in ARLEQUIN (version 3.5.1.2.; Excoffier & Lischer
2010), using 10 000 permutations of the data for esti-
mating P-values when applicable. The false discovery
rate approach (FDR; Benjamini & Hochberg 1995) was
used to correct for multiple comparisons.
Microsatellite analyses. Only samples in which at least
seven loci were successfully amplified were considered
for downstream analysis, to minimize biases caused by
poor DNA quality and missing genotypes. Levels of
genetic variability were estimated by gene diversity (H
S
within samples, H
T
in the total: Nei & Chesser 1983),
observed number of alleles and rarefied allelic richness
at each locus and for each sampling locality using FSTAT
(version 2.9.3; Goudet 2001; Tables 1 and S2, Supporting
information). Each locus/sample combination was anal-
ysed for deviations from the HardyWeinberg equilib-
rium (HWE) expectations by estimating the inbreeding
coefficient F
IS
within loci (Weir & Cockerham 1984;
Table S2, Supporting information) and tested for signifi-
cance using two-tailed probability exact tests in the GEN-
EPOP software (version 4.0; Raymond & Rousset 1995;
Rousset 2008). The FDR approach was applied when
interpreting the resulting P-values. MICRO-CHECKER soft-
ware (version 2.2.3; van Oosterhout et al. 2004) was
used to estimate the presence of null alleles, stutter and
large allele dropout in all loci.
Genetic differences among localities were quantified
by F
ST
, using Weir & Cockerham’s (1984) estimator,
over all sampling localities, as well as for pairs of sam-
ples and subsamples. Statistical significance of pairwise
F
ST
tests was assessed by G-tests for allele frequency
differences in GENEPOP using 10 000 dememorizations
and batches, and 10 000 iterations per batch. P-values
in multiple tests situations were evaluated for signifi-
cance using the FDR approach.
Temporal replicates for the Atlantic were represented
by the IRE (2006/2007, n=35) vs.UK (2011, n=33) sam-
ples, and AZ2003 (n=26) vs.AZ2011 (n=50). Within
the MED basin, temporal replicates were collected in
three different years: 2001 (n=30), 2009 (n=71) and
2012 (n=50). Furthermore for MED, juveniles (n=75)
were compared with adults (n=23). The genetic struc-
ture for the temporal replicates was quantified by F
ST
and tested for genetic divergence as described above.
Genetic differentiation patterns among samples were
visualized by applying principal component analysis
(PCA) on allele frequencies, using PCA-GEN (version
1.2.1; Goudet 1999). Significance of total inertia in each
axis was tested using 10 000 data randomizations.
BAYESCAN (Foll & Gaggiotti 2008) and FDIST2 (imple-
mented in the LOSITAN software: Antao et al. 2008) were
used to check for outlier loci putatively under selection
with emphasis on the Atlantic vs.Mediterranean com-
parisons. Runs were based on the default parameters
for BAYESCAN and based on the recommended options
(on for ‘neutral’ mean F
ST
and force mean F
ST
), the step-
wise-mutation model (SMM) and 50 000 simulations for
LOSITAN. The FDIST2 and BAYESCAN were recently recom-
mended for testing divergent selection due to lower
type I and II error (i.e. detecting false outliers and
rejecting true outliers; Narum & Hess 2011). However,
as both methods have benefits and pitfalls (Hansen
et al. 2010; Narum & Hess 2011), they were used to look
for common trends.
The BOTTLENECK software (version 1.2.02; Piry et al.
1999) was applied to investigate population declines,
using the TPM model with 7090% stepwise mutations
and 1030 variance and based on 10 000 interactions.
The Wilcoxon test was used to check for significant
heterozygosity excess. Past reduction in effective popu-
lation size was also investigated using the M-ratio test
(Garza & Williamson 2001) and comparing it with the
critical value (M
c
), following the conservative parame-
ters suggested by the authors (i.e. h=10, 90% one-step
mutation and 3.5 mean size of non-one-step mutation).
Mitochondrial and nuclear markers analyses. In addition to
F
ST
and Φ
ST
(above), Jost’s Dstatistics (Jost 2008) were
also calculated for both marker types. Pairwise D
est
val-
ues were calculated with the ‘diveRsity’ package (Keenan
et al. 2013) in R(version 3.1.1; R Core Team 2014), using
1000 bootstraps. P-values in multiple tests situations
were evaluated for significance using the FDR approach.
Spatial patterns of genetic divergence were visualized
using BARRIER software (version 2.2; Manni et al. 2004)
using sample coordinates and pairwise F
ST
or Φ
ST
estimates as input. The robustness of each barrier
was accessed by using 1000 bootstrapped matrices
calculated with customized Rcodes followed by the
POPGENOME package (Pfeifer et al. 2014) for the mtDNA
and using the ‘diveRsity’ package for nuclear genes,
both in R.
Approximate Bayesian computation (ABC, Beaumont
et al. 2002), implemented in DIYABC software (version 2.0;
Cornuet et al. 2014), was used to estimate the time since
divergence (t) between Atlantic and Mediterranean and
long-term effective population size (N
e
) of both popula-
tions. For estimation of the demographic parameters,
we tested a total of four different scenarios (see Fig. S1,
Supporting information) to choose among those that
better fits the present data set: Scenario 1constant effec-
©2015 John Wiley & Sons Ltd
6066 D. CATARINO ET AL.
tive population size for ATL and MED after divergence;
Scenario 2MED population size is allowed to change
after divergence while ATL population size is constant
over time; Scenario 3ATL population size is allowed to
change after divergence while MED population size is
constant over time; and Scenario 4both ATL and MED
population sizes are allowed to change after divergence.
The reference table (the base of the parameter estimation)
consisted of 1 000 000 simulated data sets for each sce-
nario. Uniform priors were used when building the refer-
ence table and a range of values were allowed for the
historical parameters under consideration: Atlantic
N1 =1002 000 000; Mediterranean N2 =10500 000;
t=120 000; t1 =115 000, t1 <t; Atlantic N1b =10
80 000; Mediterranean N2b =1020 000. Several trial
runs were performed to choose the range of values that
best fit the observed data, as suggested in Cornuet et al.
(2010). Sharks have low mutation rates (Martin et al.
1992; Martin 1995) and therefore the range for the mean
mutation rate and the individual locus mutation rate was
set to 1.0 910
6
to 3.0 910
4
(Karl et al. 2011) for the 11
microsatellites markers, leaving the other parameters as
default. Regarding the mtDNA CR, the most appropriate
nucleotide substitution model was selected from the hier-
archical series of likelihood ratio test, implemented in
MEGA (version 5.0; Tamura et al. 2011). From the models
available at the DIYABC, the HasegawaKishinoYano
(HKY 1985) model (I=0 and G=0.5) had the lowest
Bayesian information criterion (Schwarz 1978) value and
was used. The range for the mtDNA mutation rate was
set for 6.5 910
9
to 1.0 910
7
(Karl et al. 2011), and the
rest of the parameters were left as default. Gaps were
removed from the sequence data previous to the analy-
ses, because DIYABC does not include them in the current
version. The summary statistics used in the analyses
(Table S3, Supporting information) were the ones recom-
mended by Cornuet et al. (2008, 2010), because they are
sensitive to demographic events and divergence time (i.e.
mean genetic diversity and mean GarzaWilliamson’s M
are sensitive to demographic changes and F
ST
to diver-
gence time). The posterior probability of each scenario
was estimated using the logistic regression (Cornuet et al.
2008; Cornuet et al. 2010) on the 1% of simulated data sets
closest to the observed data set and using 10 intermedi-
ated values. Type error I and Type error II were esti-
mated for the chosen scenario computing 500 data sets
under each competing scenarios. The posterior distribu-
tion of the parameters was estimated by the logit transfor-
mation of the parameters and using 10 000 selected data.
A ‘model check’ was performed at the end of the analy-
ses using all the other summary statistics not used when
building the reference table (Cornuet et al. 2010; Table S3,
Supporting information). The discrepancy between the
model and the real data was visualized by PCA, as
implemented in DIYABC using 10 000 simulated data sets.
Time of divergence in years was assessed assuming a
generation time of 17 years, an average estimate based
on data from FISHBASE (Froese & Pauly 2013) and for a clo-
sely related species (Irvine et al. 2006).
In order to determine the amount of gene flow
between Atlantic and Mediterranean populations and to
obtain an independent estimate of tand N
e
, the isola-
tion-with-migration model (IM, as implemented in the
software IMa2; Hey & Nielsen 2007) was used. To mini-
mize the computation time, only a subset of samples
was analysed. The subset consisted in 53 samples
drawn randomly from both Atlantic and Mediterranean
(altogether 106 samples) and only samples with avail-
able microsatellite genotypes and mtDNA sequences
were considered. Mitochondrial CR sequences and
seven of the eleven microsatellite loci were included in
the model. Ccoe9, Saca3853 and SacaGA11 were
excluded, as they contain short interrupted motifs
(Ver
ıssimo et al. 2011b) and Ccoe75 was excluded due
to alleles with single bp repeats. As for the DIYABC anal-
ysis, the HKY model of sequence evolution was applied
to mitochondrial sequences, and a SMM was assumed
for microsatellite loci. Altogether five parameters were
estimated: current and ancestral population sizes
(hATL, hMED and hANC, respectively), time since
divergence (t) and a single migration parameter (m).
Upper bounds for parameter priors were estimated
from the previous DIYABC analysis and during consecu-
tive preliminary runs of the program, based on initial
estimates of theta as advised in the manual of IMa2.
The final prior values used for population size, migra-
tion rate and divergence time were as follows: q 100,
m1,t 30. We ran 150 Markov chains in parallel
under a geometric heating scheme. Several run time set-
tings with different heating schemes and durations
were explored to assess the mixing of the MCMC simu-
lation and chosen accordingly (ha =0.999, hb =0.4).
For the final simulation, the manual’s suggestion for
large data sets was followed. First, two independent
jobs were run until a suitable burn-in was reached
(1 051 048 and 1 528 526 steps). Next, a new set of runs
were started by reloading the Markov chain state file
with an additional short burn-in period of 100 000 and
afterwards 20 000 genealogies were sampled every 25
steps from a total 500 000 steps. Both Markov chain
state files generated in burn-in runs were used three
times. Finally, all six replicates were combined in L-
mode run with identical parameter settings.
Life history analyses. As not all individuals used in the
genetic analyses had maturity data, a subsample of those
was analysed and summarized for average size, weight,
sex and capture depth for the three main regions anal-
©2015 John Wiley & Sons Ltd
POPULATION STRUCTURE OF THE PORTUGUESE DOGFISH 6067
ysed: Mediterranean, Atlantic and Australia (Pacific). To
increase the sample size and the analyses’ power, a set of
samples collected at the end of 2012 and 2013 (not
screened for genetic data) was added to the Atlantic
(n=32) and Mediterranean (n=43) areas. The biological
data available from our Australian samples were very
limited; thus, a data set previously published (Daley et al.
2002) was included in the analyses (courtesy of K. Gra-
ham). The maturity data (immature vs. mature) were
analysed in detail using generalized linear models
(McCullagh & Nelder 1989) with maturity state as a bin-
ary response variable (juvenile or mature) and individual
body length or body weight as explanatory variables in
two separate models. In both models, effects of area
(Mediterranean vs.Atlantic vs.Pacific) and sex on the
probability of being mature were included in the model.
Akaike’s information criterion (AIC) was used to com-
pare the performance of alternative candidate models
(Burnham & Anderson 1998). The model with the small-
est AIC value is expected to represent the best compro-
mise between lack of precision (including too many
parameters) and bias (too few parameters). Maturity
ogives were constructed based on the selected models,
showing the probability of being mature at a given length
and weight, for each area and sex independently. The
size and weight at which 50% of the sharks were mature
(L
p50
and W
p50
) were extracted from the maturity ogives.
Results
Genetic variability
mtDNA. The CR primers amplified a 497-bp fragment,
and the 347 successfully obtained sequences contained 19
polymorphic sites (13 transitions, 4 transversions and 3
indels). Two indels occurred only in MED sequences,
where 23 of 55 individuals shared one or more indels.
The 19 polymorphic sites resulted in 27 haplotypes, 15 of
which were exclusive to a single locality (AZ: n=2, IRE:
n=2, MED: n=5, MAD: n=3 and AUS: n=3; Table 1),
while 12 were shared by two or more localities. The esti-
mated haplotype diversity varied between 0.467 and
0.778, whereas the nucleotide diversity ranged from
0.0009 to 0.0024, for UK and AZ, respectively. The lower
estimates refer to the very small UK sample (n=10) and
may represent a sampling artefact: for larger, regional
collections (n=4868), the lowest haplotype and nucleo-
tide diversities were observed in the MED (Table 1)
although differences among regions were small.
Microsatellites. In total, 503 sharks were genotyped.
Genetic diversity in the total sample (H
T
) varied among
loci from 0.639 for locus Ccoe2 to 0.954 for Ccoe55
(mean over 11 loci =0.816), with a total number of alle-
les per locus ranging from 5 (Saca3853) to 56 (Ccoe55;
mean =20; Table S2, Supporting information). The
exact G-tests did not find any deviations to HWE in the
final data set (see Appendix S1, Supporting information
for details). Results from MICRO-CHECKER did not report
any consistent pattern of null alleles within loci, stutter
or large allele dropout. The levels of genetic diversity
(H
S
) were similar among Atlantic and AUS, ranging
from 0.795 (SAZ) to 0.828 (MAU), but were significantly
lower (about 19%) in the MED (H
S
=0.657; Mann
Whitney U=23, P<0.05 two-tailed). Allelic richness
(R
s
) displayed a similar pattern, with reduced levels
inside the MED (5.6) compared to elsewhere
(mean =8.7; Table 1; Fig. S2, Supporting information);
however, the differences were not significant (Mann
Whitney U=33, P>0.05 two-tailed).
Spatial and temporal structure
mtDNA. Pairwise Φ
ST
and D
est
estimates were not sig-
nificant within the Atlantic (Table 2, above diagonals).
However, for Φ
ST
comparisons, differences were appar-
ent between Atlantic and Australian samples, where
three comparisons remained significant even after FDR
correction at the 5% level. Regarding D
est
comparisons,
the pairwise values including AUS (D
est
0.04730.2775)
were higher than those among Atlantic collections (D
est
0.00000.0628); however, none of the ATL/AUS compar-
isons were significant after FDR correction. Overall, the
ATL/AUS pairwise comparison was significant and
indicative of genetic structure for Φ
ST
(0.0744, P<0.01),
but not for D
est
(0.1543, P>0.5). In contrast, sharks from
the MED were genetically different from all other locali-
ties, with very high and significant estimates of genetic
differentiation (mean Φ
ST
=0.3808, P<0.001). As men-
tioned above, the gaps present in several of the Mediter-
ranean sequences were not found in any of the ATL/
AUS samples. Deleting these sites and rerunning the
data (as suggested by Pearce 2006) did not change the
main patterns found. Indeed, pairwise Φ
ST
increased
slightly when the gaps were excluded (results not
shown). Pairwise D
est
comparisons including MED also
showed high and significant values (D
est
0.27850.7281;
Table 2). The haplotype network for the mtDNA CR
(Fig. 2) consisted in two central, abundant haplotypes
and several other low-frequency-derived haplotypes.
The two central haplotypes, H1 (n=121) and H2
(n=144), occurred in similar frequencies across locali-
ties except for MED, where H2 was not observed.
Instead, the MED had H1 and a unique haplotype (H5)
occurring in high frequency (55% and 36%, respectively),
differing by one or two indels (position 41 and 53 in the
alignment) from the central ones. In the BARRIER analysis,
the strongest genetic barrier (100% support) was found
©2015 John Wiley & Sons Ltd
6068 D. CATARINO ET AL.
between the MED and all other localities (Fig. 3a). Other
secondary barriers are located around the AZ/MAR,
with statistical support between 68% and 98%.
Microsatellites. Significant genetic differentiation was
found over all sampling localities both across loci (com-
bining temporal samples; F
ST
=0.0581, P<0.001;
Table S2, Supporting information) and for each locus
separately (F
ST
: 0.0107-0.1275, P<0.001; Table S2,
Supporting information). However, no divergence was
found among sampling localities across loci after
removing MED from the analyses (F
ST
=0.0016,
P> 0.05; Table S2, Supporting information). Thus, the over-
all genetic divergence seems to be caused by divergence
between the MED and the other localities (mean
F
ST
= 0.1149, P< 0.001; Table 2). Within the MED, no evi-
dence for genetic structure among the three replicate sites was
found (see Table S4, Supporting information).
No significant genetic structure was found within the
Atlantic (mean F
ST
=0.0024; P>0.05; Table 2) and the
Table 2 Pairwise F-statistics (F
ST
and Φ
ST
) and Jost’s D
est
values between sampling sites for microsatellites (below the diagonals) and
mtDNA control region (above the diagonals)
AZ SAZ UK/IRE MAU MED PT MAD AUS SAF
F-statistics
AZ/MAR 0.0244 0.0267 0.0011 0.3127*** 0.0168 0.0402 0.0140 0.0166
SAZ 0.0096 0.0200 0.0426 0.3802*** 0.0344 0.0413 0.0110 0.0412
UK/IRE 0.0005 0.0035 0.0064 0.4437*** 0.0145 0.0128 0.0997* 0.0118
MAU 0.0007 0.0096 0.0014 0.3970*** 0.0160 0.0024 0.0413 0.0170
MED 0.1098*** 0.1138*** 0.1149*** 0.1154*** 0.4330*** 0.4495*** 0.2722*** 0.3578***
PT 0.0004 0.0064 0.0006 0.0004 0.1143*** 0.0122 0.0825* 0.0003
MAD 0.0013 0.0078 0.0004 0.0007 0.1051*** 0.0009 0.0944* 0.0165
AUS 0.0020 0.0108* 0.0005 0.0018 0.1311*** 0.0002 0.0029 0.0089
Jost D
est
AZ/MAR 0.0183 0.0569 0.0028 0.4914*** 0.0182 0.0362 0.0785 0.0000
SAZ 0.0123 0.0000 0.0000 0.6142** 0.0000 0.0000 0.1788 0.0089
UK/IRE 0.0012 0.0000 0.0141 0.7281*** 0.0006 0.0000 0.2775 0.0628
MAU 0.0000 0.0040 0.0000 0.5594*** 0.0000 0.0039 0.1284 0.0000
MED 0.2681*** 0.2632*** 0.2738*** 0.2681*** 0.6154*** 0.6764*** 0.2785*0.4747***
PT 0.0008 0.0013 0.0000 0.0000 0.2769*** 0.0000 0.1786 0.0143
MAD 0.0008 0.0044 0.0000 0.0000 0.2610*** 0.0001 0.2230 0.0327
AUS 0.0008 0.0071 0.0003 0.0000 0.3293*** 0.0002 0.0088 0.0473
Sampled sites abbreviations are given in Table 1.
Significant at alpha =*0.05, **0.01, ***0.001 after false discovery rate approach.
Bold indicates significant values.
H5
H4
H6
H8 H13
H12
H15
H23
H24
H25
H26
H27
H21
H19
H20
H22
H16
H7
H3
H17
H18
H14
H10
H11
H9
H1
H2
*14%
86% *
86%
86%
86%
71%
AZ/MA
R
SAZ
UK/IRE
MAU
MED
PT
MAD
AUS
SAF
*
Fig. 2 Median-joining network from 497-
bp mtDNA control region sequences.
Size of circles (haplotypes) is propor-
tional to the relative frequencies in the
sample and colour coded according to
sampling locality; abbreviations are given
in Table 1. Each branch indicates single
nucleotide substitutions, except when
noted. Bootstrapping support of each
branch is 100%, except when noted.
*indicates an indel.
©2015 John Wiley & Sons Ltd
POPULATION STRUCTURE OF THE PORTUGUESE DOGFISH 6069
overall ATL/AUS comparison showed also no signifi-
cant differences (F
ST
=0.0015; P>0.05), despite the
existence of one ATL/AUS significant pairwise compar-
isons (Table 2). D
est
values showed the same patterns as
found with F
ST
values, with no genetic structure
detected within Atlantic, or between ATL/AUS locali-
ties. Regarding comparisons including MED, high and
significant pairwise F
ST
and D
est
values were found
(P<0.001 in all cases; Table 2).
Allelic frequencies (Fig. S3, Supporting information)
were very similar between ATL and AUS across all loci;
however, MED exhibited very different allelic frequen-
cies. A shift in the most common alleles was observed
between Atlantic/Australia localities and Mediterranean
ones, where less frequent alleles at the former locations
became more common in the Mediterranean. This
pattern was found also for mtDNA haplotypes. No
temporal heterogeneity among years was detected either
within the Atlantic (8 years period) or the Mediterranean
(11 years period; Table S4, Supporting information).
The BARRIER analysis revealed that the strongest
genetic break was between the MED and all other sam-
ples, with 100% support (Fig. 3b). Other weaker barriers
separated SAZ from the other samples with a support
above 80%. A similar pattern is suggested by the PCA
analyses (Fig. 4), where the MED was clearly separated
from all other samples (PC1: F
ST
=0.0476, 87% of total
inertia, P=0.0001). The PCA further suggested some
separation between SAZ and other samples by PC2;
however, this axis was not significant (PC2:
F
ST
=0.0019, 4% of total inertia, P=1.00). Results from
64
84
92
90
100
100
UK/IRE
AZ PT MED
MAD
SAZ
MAU
68
93
100
100
88
100
98
100
UK/IRE
MED
AZ/MAR
SAZ
MAU
SAF
MAD
PT
mtDNA microsatellites
(a) (b)
Fig. 3 Barrier-inferred gene flow patterns based on (a) mtDNA and (b) 11 microsatellite loci. Thickness of the barriers (red lines)
indicates their rank of importance and numbers indicate the percentage of support. Only barriers with support >60% are shown.
Sampled sites abbreviations are given in Table 1.
–0.6
–0.4
–0.2
0
0.2
0.4
0.6
1 –0.6 –0.2 0.2 0.6 1
SAZ
MED
UK/IRE
PT
MAD
AUS
MAU
AZ
PC2 (4%inertia)
PC1 (87% inertia)
Fig. 4 Principal component analysis (PCA) based on 11
microsatellite loci. Circles around the sampling localities high-
light the separation according to PCA 1 (P<0.001). Sampled
sites abbreviations are given in Table 1.
©2015 John Wiley & Sons Ltd
6070 D. CATARINO ET AL.
the outlier loci tests showed no candidate loci under
directional selection when comparing ATL/MED.
Demographic analyses
For microsatellite data, no bottlenecks were detected
using both the Wilcoxon tests (P>0.05) and the M-ratio
tests (mean observed M0.749, M
c
0.675 for seven
loci and M
c
0.725 for 15 loci).
In the DIYABC analysis, scenario 4 was the best sup-
ported with relatively stronger posterior probability
(mean P=0.99, 95% CI =0.981.00; Table 3; Fig. S4,
Supporting information). All other scenarios received
much lower support (between P=0.00 and 0.01; Table 3;
Fig. S4, Supporting information). Based on the logistic
estimate, the type error I (probability of being the true
model, but not selected) estimated for scenario 4 was 0.2
and the type error II (probability of not being the true
model, but it is selected) was 0.06. Under scenario 4, the
effective population size estimation for ATL (N1) did not
produce a distinct mode regardless the large priors
given. All the other estimated parameters showed dis-
tinct modes (Fig. S5, Supporting information). The MED
N2 was estimated in 90 900 (95% CI: 36 500479 900;
Table 3). The simulations showed that both populations’
sizes after divergence (N1b and N2b) were not very small
(Table 3; Fig. S5, Supporting information) with modal
values of 40 500 (95% CI: 16 30077 400) and 2050 (408
15 300), respectively. Time since divergence was esti-
mated in 2750 generations (95% CI: 145018 100) for sce-
nario 4. Therefore, under this scenario and according to
the generation time used for the species (17 years), the
divergence time between ATL and MED population can
be put around 47 000 years before present (yBP; 95% CI:
24 650307 700). The posterior distributions of the muta-
tion rates suggested that the prior range selected was
appropriate for the analysed data set (Fig. S6, Supporting
information). Model check showed that the observed
and the simulated data were similar (Fig. S6, Supporting
information), suggesting a good model fit.
A summary of the estimated parameters using IMa2
analyses can be found in Table S5 (Supporting informa-
tion). Replicate runs in IMa2 analyses revealed low
levels of migration between both populations (0.756; a
single sharp peak was observed). Clear peaks were also
produced when estimating divergence time (0.495) and
the two population size parameters (Θ
ATL
=15.55 and
Θ
MED
=4.05), while the estimated size of the ancestral
population was on the upper border of the priors
(slightly higher than 100). When converting migration
parameter into per-generation population migration
rates (M=Θ9m/2), peak locations corresponded to
1.530 (M
ATL?MED
) and 5.878 (M
MED?ATL
) migration
events per generation, which means 0.090 and 0.346
Table 3 Posterior parameters values for each scenario estimated with DIYABC
Scenario Parameter Mean Median Mode
95% CI
P(95% CI)Lower Upper
S1 N1 3.77 910
5
3.34 910
5
2.81 910
5
1.12 910
5
8.86 910
5
0.00 (0.000.00)
N2 4.61 910
4
2.62 910
4
9.45 910
3
4.26 910
3
2.34 910
5
t8.91 910
3
8.18 910
3
5.69 910
3
1.78 910
3
1.87 910
4
S2 N1 3.65 910
5
3.30 910
5
2.57 910
5
1.26 910
5
8.09 910
5
0.00 (0.000.00)
N2 3.18 910
5
3.32 910
5
4.80 910
5
7.78 910
4
4.93 910
5
t8.79 910
3
8.04 910
3
6.05 910
3
2.01 910
3
1.84 910
4
t1 4.66 910
3
3.68 910
3
1.92 910
3
3.30 910
2
1.32 910
4
N2b 7.97 910
3
7.14 910
3
4.77 910
3
9.33 910
2
1.85 910
4
S3 N1 1.34 910
6
1.40 910
6
1.92 910
6
4.37 910
5
1.98 910
6
0.01 (0.000.02)
N2 4.08 910
4
2.69 910
4
1.76 910
4
7.60 910
3
1.93 910
5
t7.99 910
3
6.95 910
3
2.93 910
3
1.63 910
3
1.85 910
4
t1 6.43 910
3
6.04 910
3
4.15 910
3
1.2 910
3
1.37 910
4
N1b 4.32 910
4
4.36 910
4
4.46 910
4
7.93 910
3
7.62 910
4
S4 N1 1.49 910
6
1.57 910
6
1.97 910
6
6.25 910
5
1.99 910
6
0.99 (0.981.00)
N2 2.17 910
5
1.92 910
5
9.09 910
4
3.65 910
4
4.79 910
5
t7.10 910
3
5.85 910
3
2.75 910
3
1.45 910
3
1.81 910
4
t1 4.69 910
3
3.93 910
3
2.43 910
3
9.42 910
2
1.23 910
4
N1b 4.78 910
4
4.77 910
4
4.05 910
4
1.63 910
4
7.74 910
4
N2b 4.52 910
3
3.36 910
3
2.05 910
3
4.08 910
2
1.53 910
4
For each estimated parameter are presented the mean, median and modal value, the lower and upper 95% credible intervals and the
probability P(95% CI) of each scenario under consideration. In bold is shown the best supported scenario.
©2015 John Wiley & Sons Ltd
POPULATION STRUCTURE OF THE PORTUGUESE DOGFISH 6071
migrations events per year taking into account the
C. coelolepis generation time (17 years). In order to date
the divergence time in years and to estimate the effec-
tive population size, we applied the geometric mean of
the mutation rates estimated by the DIYABC approach
considering Scenario 1 with constant Ne, namely
5.3 910
6
(2.75 910
6
6.98 910
5
) and 1.81 910
8
(9.2 910
9
8.71 910
8
) for microsatellites and
mtDNA, respectively (Fig. S6, Supporting information).
Applying these rates, IM indicates that the divergence
between Atlantic and Mediterranean populations
occurred at 92 063 yBP (95% CI: 36 267209 233), while
N
e
per generation was estimated as 42 530 (95% CI:
27 48774 257) and 11 077 (95% CI: 642717 641) for
Atlantic and Mediterranean, respectively.
Life history analyses
For both sexes, the largest sharks were found in the
Atlantic. Here, the average body length exceeded the
maximum body length seen in the Mediterranean,
where much smaller sharks dominated the samples
(Table 4). The data supported an interaction effect
between body length and sex, and also an additive effect
of area, on the probability of being mature (Table S6,
Supporting information). In a separate model, the data
supported an interaction effect between body weight
and sex, as well as an additive effect of area, on the
probability of being mature (Table S6, Supporting infor-
mation). These models could not be simplified
(removing terms) without increasing the AIC value. The
smallest mature sharks were found in the Mediterranean
(Table 4; Fig. 5), where females reached L
p50
at 56 cm
and W
p50
at 2.2 kg, while males reached L
p50
at 46 cm
and W
p50
at 0.5 kg. In the Atlantic and Pacific Oceans,
maturity ogives were similar and shifted towards larger
sizes and weights for both females (>100 cm and
>10 kg) and males (>80 cm and >3 kg; Table 4; Fig. 5).
Discussion
Genetic barrier across Gibraltar
The main finding of this study is the existence of a
major genetic discontinuity between the Mediterranean
population of Portuguese dogfish and those inhabiting
the Atlantic/Pacific Oceans. This conclusion was sup-
ported by all genetic markers (11 nuclear and 1 mito-
chondrial). Significant high levels of pairwise F
ST
(>0.1),
Φ
ST
(>0.3) and D
est
(>0.2) between Mediterranean and
all the other localities suggest that gene flow across the
shallow waters of Gibraltar are nearly absent for this
deep-sea species. The low estimated migration rates (1.5
and 5.9 migration events per generation) corroborate
this hypothesis.
Table 4 Biological data sampled from three main regions, showing the sample size (N) of females (F) and males (M), mean and
(range) of capture depth, body length (total length, TL) and body weight (total weight, TW), and the size (L
p50
) and weight (W
p50
)at
first maturation
Sampled region Sex NCapture depth (m) L
p50
(cm) W
p50
(cm)TL (cm) TW (kg)
Mediterranean F 106 2136 (15002853) 41 (2065) 0.6 (0.041.9) 56 2.2
M 78 2081 (15002853) 42 (2158) 0.5 (0.051.4) 46 0.5
Atlantic F 73 1454 (10752475) 105 (64119) 9.2 (0.716.2) 109 10.2
M 134 1333 (10252475) 92 (69104) 5.7 (1.98.3) 83 3.7
Australia F 22 1062 (9001070) 80 (55110) 4.7 (1.013.0) 111* 11.2*
M 91 1009 (3901070) 79 (4897) 3.9 (0.78.0) 84 4.1
Part of the Australian data is from the literature (Daley et al. 2002), courtesy of K. Graham.
*Only two of the 22 females were mature.
20 40 60 80 100 120
0.04.0 8.
0
(a)
0 2 4 6 8 10121416
0.0 4.08.0
(b)
)
e
ru
t
a
m
(P
Body length (cm)
)erutam
(P
Body weight (kg)
Fig. 5 Maturity ogives for Portuguese dogfish, showing the
probability of being mature within the observed ranges of
body length (a) and body weight (b) for females (solid lines)
and males (broken lines) collected from the Mediterranean
(black), Atlantic (red) and Pacific (blue).
©2015 John Wiley & Sons Ltd
6072 D. CATARINO ET AL.
Although the reported bathymetric distribution of
this species (between 128 and 3675 m; Ebert & Steh-
mann 2013) would not rule out transition across the
shallows of Gibraltar, it is highly unlikely as most stud-
ies reported a considerably deeper minimum depth of
occurrence, depending on the region, that is below
500 m off Ireland (Clarke et al. 2001) and below 900 m
in the Azores (Menezes et al. 2009; Table S1, Supporting
information). The average capture depth in this study
was 2109 m within the Mediterranean, with no individ-
uals being caught shallower than 1500 m depth. Such
findings are in accordance with previous studies in the
Mediterranean (Carrass
on et al. 1992; Tecchio et al.
2011), where the existence of a deeper-dwelling commu-
nity of this species seems to prevail, possibly to avoid
interspecific competition with shallower-dwelling spe-
cies of similar trophic level (Carrass
on et al. 1992).
Therefore, the very deep habits of the Portuguese dog-
fish together with the shallow bathymetry of the Strait
of Gibraltar (284 m) may act as a physical barrier to the
movement of individuals, leading to genetic isolation of
the Mediterranean population. In other species where
an Atlantic/Mediterranean genetic barrier has been
found (all epipelagic species with planktonic larvae),
the Almeria-Oran oceanographic density front has been
identified as the barrier to gene flow, rather than the
Strait of Gibraltar (Patarnello et al. 2007 and references
therein). The Almeria-Oran front (Tintore et al. 1988) is
located at the eastern end of the Alboran Sea and can
extend to 300 m deep, limiting extensively the mixing
of the epipelagic eggs and larvae from the Atlantic with
the rest of the Mediterranean and vice versa. The exis-
tence of pelagic larval stages in many species promotes
passive dispersal and gene flow, while the absence of
such larval stages in sharks makes gene flow dependent
on active individual dispersal (Conrath & Musick 2012).
Therefore, in the case of the viviparous deep-sea shark
Centroscymnus coelolepis, the shallow depth of the Strait
of Gibraltar, instead of the Almeria-Oran front, is a bio-
logically more realistic barrier to gene flow. The
hypothesis of bathymetry as a barrier to genetic connec-
tivity is supported by studies in other deep-sea taxa,
such as the roundnose grenadier (Knutsen et al. 2012)
and tusk (Knutsen et al. 2009). A physical barrier at the
Strait of Gibraltar seems also relevant for several
deep-sea species and especially sharks, because only a
fraction of the Atlantic deep-sea species occur inside
the Mediterranean (Emig & Geistdoerfer 2004; Ebert &
Stehmann 2013).
Genetic pattern across the Atlantic and Pacific
In contrast to the marked genetic discontinuity
across the Strait of Gibraltar, the apparent genetic
homogeneity across the Atlantic Ocean suggests a single
Atlantic panmictic population, consistent with the find-
ing of Ver
ıssimo et al. (2011a). The genetic homogeneity
across distant locations (Ver
ıssimo et al. 2011a; this
study) is probably due to the high swimming capacity
and long lifespan of the species (Irvine et al. 2006; Maia
et al. 2012). Therefore, active dispersal is likely to be a
major source of connectivity among areas without major
physical barriers, as has also been implied for other
deep-sea sharks (Straube et al. 2011; Cunha et al. 2012;
Ver
ıssimo et al. 2012).
Active dispersal over a long generation time may also
explain the low level of differentiation between Atlantic
and Australia. Furthermore, females are known to store
sperm after mating (up to several months or years;
Moura et al. 2011) and may mate whenever they get the
opportunity, saving the sperm until suitable conditions
arise to fertilize and produce the young. This strategy
may also contribute to population mixing during long
distance foraging movements.
Demographic events and population divergence
The low levels of genetic variability, as those found
within the Mediterranean, are sometimes associated
with past bottlenecks. However, no evidence of a bottle-
neck event was detected by the tests employed, perhaps
due to limitations of such tests (see Peery et al. 2012 for
details). The best supported scenario chosen by DIYABC,
where both populations N
e
were allowed to change
after divergence, showed high modal values of N
e
after
divergence (N1b and N2b, Fig. S5, Supporting informa-
tion, Table 3), suggesting that no severe bottleneck has
occurred. However, in the Mediterranean, a possible
stronger reduction in the population cannot be entirely
ruled out because the lower credible interval for N2b is
as low as 500. Nevertheless, the reduction in heterozy-
gosity in the Mediterranean was not very large (~19%),
so it is possible that the divergence between the Atlan-
tic and Mediterranean populations may have occurred
simply by postisolation genetic drift over time. Results
from the ABC analyses also suggest that historical N
e
after divergence in the Mediterranean was at least one
order of magnitude smaller than in the Atlantic, which
may had facilitated the action of drift in the Mediter-
ranean. Results from IM approach also support that N
e
is significantly smaller (no overlapping CI) within the
Mediterranean compared with the Atlantic. The shift in
the common allele frequencies across the Strait of
Gibraltar (Fig. S3, Supporting information) and the loss
of the high-frequency H2 Atlantic haplotype in the
Mediterranean are in agreement with what to expect in
an isolated population subjected to the action of genetic
drift (Hedrick 2011) and are consistent with previous
©2015 John Wiley & Sons Ltd
POPULATION STRUCTURE OF THE PORTUGUESE DOGFISH 6073
studies for the area (i.e. P
erez-Losada et al. 2002;
Lemaire et al. 2005). This pattern suggests that demo-
graphic events such as isolation followed by genetic
drift have been acting over time and may be an impor-
tant force shaping the gene pool of this Mediterranean
shark population. Mediterranean populations of some
large pelagic predators such as tunas, swordfish and
dolphinfish have been shown to have a similar genetic
pattern (i.e. genetic differentiation and low genetic
diversity comparing with their Atlantic counterparts),
explained as the result of vicariance and demographic
events (Vi~
nas et al. 2004; Alvarado-Bremer et al. 2005;
Nakadate et al. 2005; D
ıaz-Jaimes et al. 2010).
The ABC results from scenario 4 suggested the sepa-
ration between Atlantic and Mediterranean at about
47 000 yBP (95% CI: 24 650307 700 yBP), at the late
Pleistocene, during the last glacial period (110 000
12 000 yBP; Kukla 2005). However, the relatively wide
credible intervals suggest that an earlier separation may
also be possible. A similar estimate was also provided
by IM approach suggesting that divergence occurred at
92 063 yBP (95% CI: 36 267209 233 yBP). The DIYABC
approach has the advantage of statistically choosing the
scenario that best explains the studied data set and esti-
mating the different parameters under that scenario. On
the other hand, the IM approach has the advantage of
considering and estimating migration rates which DIYABC
has not. As DIYABC does not assume migration after
divergence, the estimates of divergence time and effec-
tive population sizes may be biased. However, in the
present study, the estimated migration rates are very
low suggesting that this pitfall may have little effect on
the DIYABC estimates. Regardless of the exact timing of
the Atlantic/Mediterranean split, the models suggest
that this separation has probably occurred during the
last glacial period, which reached its maximum at the
last glacial maximum (LGM; 26 00020 000 yBP; Clark
et al. 2009). According to the credible intervals, the sep-
aration occurred no later than the LGM, suggesting that
this climatic event had an important role in the separa-
tion of the Atlantic and Mediterranean shark popula-
tions or at least contributed to keep them apart in case
of an earlier split. This finding is in accordance with
Emig & Geistdoerfer’s (2004) suggesting that the condi-
tions for the bathyal and abyssal species during the last
glacial period were very harsh and therefore entrance
in the Mediterranean by Atlantic deep-sea species was
thought to be limited. The geological history of the
Mediterranean and the several periods of drastic cli-
matic variability are thought to play a major role in the
separation of populations for several species, including
elasmobranchs, resulting in genetic heterogeneity
between the isolated groups (Hewitt 1996; Chevolot
et al. 2006; Griffiths et al. 2011). Different thermal
regimes in the water column during different climatic
events may have contributed to the shift in the species
vertical range and thereby promote or prevent the
entrance in the Mediterranean by Atlantic specimens
(Emig & Geistdoerfer 2004).
Life history traits
Portuguese dogfish from the Mediterranean were smal-
ler and reached sexual maturity at smaller sizes com-
pared to those from the Atlantic and Australia. Females
in the Mediterranean reach only about half the maxi-
mum size (65 cm) of that expected for an Atlantic
female shark (122 cm). Furthermore, size at birth seems
to be smaller in the Mediterranean (about 1722 cm;
Table 4; Torchio & Michelangeli 1971; Carrass
on et al.
1992) compared to the Atlantic (about 30 cm; Girard &
Du Buit 1999). The capture of four recently born speci-
mens (2022 cm TL, yolk-sac scar was still present)
within Mediterranean during this study and one (32 cm
TL, D. Catarino, personal observation) more recently
within Atlantic seems to confirm this pattern. Smaller
adult sizes in Mediterranean populations compared to
their Atlantic conspecifics are a common trend for sev-
eral species. This trend may result from a combination
of several factors, in which limited food resources and
higher temperature in the Mediterranean are likely to
play a major role (Stefanescu et al. 1992; Massut
ıet al.
2004). The Mediterranean deep biological communities
live in warmer and saltier waters (~13 °C, >38&;
Tecchio et al. 2011) than those in Atlantic and Pacific
Oceans (Table S1, Supporting information). Also food
availability (downward particle fluxes; biomass and
abundance indices) is lower within the Mediterranean
(Danovaro et al. 1999; Massut
ıet al. 2004) and thereby
deep-sea communities are faced with an oligotrophic
environment. Nevertheless, hotspots of higher produc-
tivity are also present, such as submarine canyons
(Danovaro et al. 2010 and references therein). Therefore,
deep-sea species crossing Gibraltar have not only to
overcome the shallow physical barrier but also to adjust
to the unique environmental conditions of the deep
Mediterranean basin.
As explained above, the genetic break observed at
molecular markers across the Strait of Gibraltar in
C. coelolepis is probably the result of demographic
events (isolation +genetic drift) rather than natural
selection because no signal of selection was detected.
Ambient temperature plays a major role in the meta-
bolic rates of ectothermic organisms, including sharks
(Bernal et al. 2012), and this condition coupled to a
low-food environment as seen in the Mediterranean
deep-sea may set up strong selection pressures on traits
related to feeding activity/strategy and growth.
©2015 John Wiley & Sons Ltd
6074 D. CATARINO ET AL.
Therefore, we suggest that the selection signals should
be further investigated, specifically regarding genes
controlling metabolic activity and growth.
In conclusion, the Mediterranean Portuguese dogfish
population is genetically distinct and isolated from its
Atlantic and Australian counterparts, and the striking
differences in their respective life history traits suggest
they are following different evolutionary trajectories.
Acknowledgements
The authors are very thankful to the several persons and insti-
tutions that provided data, samples and assistance at different
stages of this work: R. Medeiros, A. Rosa, E. Giacomello
(IMAR/DOP, Portugal); H. Ramos (SeaExpert, Portugal); K.
Enersen (IMR, Norway); U. Fernandez, A. Mecho, S. Tecchio
and E. Ramirez-Llodra (ICM-CSIC, Spain); C. Mytilineou
(HCMR, Greece); K. Graham (CFRC, Australia); S. Tanaka
(Tokai University, Japan); M. Mea, F. Mairano (Italy); to the
crews and scientific teams of the R/V Arquip
elago, F/V Santo
Onofre, R/V Garcia del Cid, R/V Sarmiento de Gamboa and MRV
Scotia. This study was performed under the framework of
ReDEco (FP6 ERA-NET, MARIN-ERA/MAR/0003/2008) and
Hermione (FP7-ENV-2008-1, nr 226354) projects. Samples were
collected under other scientific projects: DEECON (ESF, 06-
EuroDEEP-FP-008 & SFRH-EuroDEEP/0002/2007); CONDOR
(EEA Grants, PT0040/2008); SMaRT (M.2.1.2/029/2011) and
DEMERSAIS (Azorean Government); OASIS (FP5, EVK3-CT-
2002-00073-OASIS); PescProf I (INTERREG III B, MAC/4.2/
M12); BIOFUN (ESF, CTM2007-28739-E); DOS MARES
(CTM2010-21810-C03-03); ADIOS (EVK3-CT-2000-00035). DC is
a PhD student funded by FCT (SFRH/BD/65730/2009); HK is
cofunded by the University of Agder and the University of
Oslo (CEES); AV is a postdoctoral fellow funded by FCT
(POPH/FSE; SFRH/BPD/77487/2011); DS is a postdoctoral fel-
low funded by the DIANET programme (FP1527385002); and
SS is a research fellow supported by the Marie Curie grant
cofunded by the EU under the FP7-People-2012-COFUND; Co-
funding of Regional, National and International Programmes,
GA n.600407 and the Bandiera Project RITMARE. MARE is
funded by FCT through the strategic project UID/MAR/
04292/2013. Finally, we would like to thank to the editor and
to the anonymous referees whose comments and suggestions
helped improve the manuscript.
References
Aboim MA, Menezes GM, Schlitt T, Rogers AD (2005) Genetic
structure and history of populations of the deep-sea fish
Helicolenus dactylopterus (Delaroche, 1809) inferred from
mtDNA sequence analysis. Molecular Ecology,14, 13431354.
Alvarado-Bremer JR, Vi~
nas J, Mejuto J, Ely B, Pla C (2005)
Comparative phylogeography of Atlantic bluefin tuna and
swordfish; the combined effects of vicariance, secondary con-
tact, introgression and population expansion on the regional
phylogenies of two highly migratory pelagic fishes. Molecular
Phylogenetics and Evolution,36, 169187.
Antao T, Lopes A, Lopes RJ, Beja-Pereira A, Luikart G (2008)
LOSITAN: a workbench to detect molecular adaptation
based on a F
ST
-outlier method. BMC Bioinformatics,9, 323
327.
Ball AO, Sedberry GR, Zatcoff MS, Chapman RW, Carlin JL
(2000) Population structure of the wreckfish Polyprion ameri-
canus determined with microsatellite genetic markers. Marine
Biology,137, 10771090.
Bandelt HJ, Forster P, R
ohl A (1999) Median-joining networks
for inferring intraspecific phylogenies. Molecular Biology and
Evolution,16,3748.
Beaumont MA, Zhang W, Balding DJ (2002) Approximate
Bayesian Computation in population genetics. Genetics,162,
20252035.
Benjamini Y, Hochberg Y (1995) Controlling the false discovery
rate: a practical and powerful approach to multiple testing.
ournal of the Royal Statistical Society. Series B (Methodological),
57, 289300.
Bernal D, Carlson JK, Goldman KJ, Lowe CG (2012) Energetics,
metabolism and endothermy in sharks and rays. In: Biology of
Sharks and their Relatives, 2nd edn (eds Carrier JC, Musick JA,
Heithaus MR), pp. 211237. CRC Press, Boca Raton, FL, USA.
Bradai MN, Saidi B, Enajjar S (2012) Elasmobranchs of the
Mediterranean and Black Sea: Status, Ecology and Biology. Stud-
ies and Reviews. General Fisheries Commission for the
Mediterranean. FAO, No. 91. Rome.
Burnham KP, Anderson DR (1998) Model Selection and Inference:
A Practical Information-Theoretic Approach. Springer, New
York, New York, USA.
Caley MJ, Carr MH, Hixon MA, Hughes TP, Jones GP, Menge BA
(1996) Recruitment and the local dynamics of open marine popu-
lations.AnnualReview ofEcologyand Systematics,27, 477500.
Campbell N, Neat F, Burns F, Kunzlik P (2011) Species rich-
ness, taxonomic diversity, and taxonomic distinctness of the
deep-water demersal fish community on the Northeast
Atlantic continental slope (ICES Subdivision VIa). ICES Jour-
nal of Marine Science,68, 365376.
Carrass
on M, Stefanescu C, Cartes JE (1992) Diets and bathy-
metric distributions of two bathyal sharks of the Catalan
deep sea (western Mediterranean). Marine Ecology Progress
Series,82,2130.
Chevolot M, Hoarau G, Rijnsdorp AD et al. (2006) Phylogeog-
raphy and population structure of thornback rays (Raja cla-
vata L., Rajidae). Molecular Ecology,15, 36933705.
Clark PU, Dyke AS, Shakun JD et al. (2009) The last glacial
maximum. Science,325, 710.
Clarke MW, Conolly PL, Bracken JJ (2001) Aspects of the
reproduction of the deep water sharks Centroscymnus coelole-
pis and Centrophorus squamosus from west of Ireland and
Scotland. Journal of the Marine Biological Association of the Uni-
ted Kingdom,81, 10191029.
Cl
o S, Dal
u M, Danovaro R, Vacchi M (2002) Segregation of
the Mediterranean Population of Centroscymnus coelolepis
(Chondrichthyes: Squalidae): A Description and Survey.
NAFO SCR Doc. 02/83.
Compagno LJV, Dando M, Fowler S (2005) Sharks of the World.
Princeton University Press, Princeton, New Jersey, USA.
Conrath CL, Musick JA (2012) Reproductive biology of Elasmo-
branchs. In: Biology of Sharks and their Relatives, 2nd edn (eds
Carrier JC, Musick JA, Heithaus MR), pp. 291311. CRC
Press, Boca Raton, FL, USA.
Cornuet JM, Santos F, Beaumont MA et al. (2008) Inferring
population history with DIY ABC: a user-friendly approach
©2015 John Wiley & Sons Ltd
POPULATION STRUCTURE OF THE PORTUGUESE DOGFISH 6075
to approximate Bayesian computation. Bioinformatics,24,
27132719.
Cornuet JM, Ravign
e V, Estoup A (2010) Inference on popula-
tion history and model checking using DNA sequence and
microsatellite data with the software DIYABC (v1.0). BMC
Bioinformatics,11, 401.
Cornuet J-M, Pudlo P, Veyssier J et al. (2014) DIYABC v2.0: a
software to make Approximate Bayesian Computation infer-
ences about population history using Single Nucleotide Poly-
morphism, DNA sequence and microsatellite data.
Bioinformatics,30, 11871189.
Cotton C (2010) Age, growth, and reproductive biology of deep-
water Chondrichthyans. PhD thesis, Faculty School of Marine
Science, VIMS, Gloucester Point, Virginia, USA.
Cunha RL, Coscia I, Madeira C et al. (2012) Ancient divergence
in the trans-oceanic deep-sea shark Centroscymnus crepidater.
PLoS One,7, e49196.
Daley RK, Stevens JD, Graham K (2002) Catch analysis and pro-
ductivity of the deepwater dogfish resource in southern Australia.
CSIRO Marine Research, Fisheries Research and Develop-
ment Corporation (FRDC), and NSW Fisheries, Australia,
FRDC Project No. 1998/108.
Danovaro R, Dinet A, Duineveld G, Tselepides A (1999) Ben-
thic response to particulate fluxes in different trophic envi-
ronments: a comparison between the Gulf of Lions-Catalan
Sea (western-Mediterranean) and the Cretan Sea (eastern-
Mediterranean). Progress in Oceanography,44, 287312.
Danovaro R, Company JB, Corinaldesi C et al. (2010) Deep-sea
biodiversity in the Mediterranean Sea: the known, the
unknown, and the unknowable. PLoS One,5, e11832.
D
ıaz-Jaimes P, Urlbe-Alcocer M, Rocha-Olivares A, Garcia-
de-Leon FJ, Nortmoon P, Durand JD (2010) Global phylo-
geography of the dolphinfish (Coryphaena hippurus): the
influence of large effective population size and recent disper-
sal on the divergence of a marine pelagic cosmopolitan spe-
cies. Molecular Phylogenetics and Evolution,57, 12091218.
Dudgeon CL, Blower DC, Broderick D et al. (2012) A review of
the application of molecular genetics for fisheries manage-
ment and conservation in sharks and rays. Journal of Fish
Biology,80, 17891843.
Ebert DA, Stehmann MFW (2013) Sharks, batoids, and chimaeras
of the North Atlantic. FAO Species Catalogue for Fishery Pur-
poses. No. 7. Rome.
Emig CC, Geistdoerfer P (2004) The Mediterranean deep-sea
fauna: historical evolution, bathymetric variations and geographi-
cal changes. Carnets de G
eologie/Notebooks on Geology.
Article 2004/01 (CG2004_A01_CCE-PG), 10 pp.
Excoffier L, Lischer HEL (2010) Arlequin suite ver 3.5: a new
series of programs to perform population genetics analyses
under Linux and Windows. Molecular Ecology Resources,10,
564567.
FAO (2006) General Fisheries Commission for the Mediter-
ranean. Report of the ninth session of the Scientific Advisory
Committee. FAO Fisheries Report No. 814. Rome.
Figueiredo I, Moura T, Neves A, Gordo LS (2008) Reproductive
strategy of leafscale gulper shark Centrophorus squamosus and
the Portuguese dogfish Centroscymnus coelolepis on the Por-
tuguese continental slope. Journal of Fish Biology,73, 206225.
Foll M, Gaggiotti O (2008) A genome scan method to identify
selected loci appropriate for both dominant and codominant
markers: a Bayesian perspective. Genetics,180, 977993.
Froese R, Pauly D (2013) FishBase (version 12/2013). Available
from www.fishbase.org (accessed 7 February 2014].
Galtier N, Gouy M, Gautier C (1996) SEAVIEW and PHYLO_-
WIN: two graphic tools for sequence alignment and molecu-
lar phylogeny. Computer Applications in the Biosciences,12,
543548.
Garza JC, Williamson EG (2001) Detection of reduction in pop-
ulation size using data from microsatellite loci. Molecular
Ecology,10, 305318.
Girard M, Du Buit MH (1999) Reproductive biology of two
deep-water sharks from the British Isles, Centroscymnus coelo-
lepis and Centrophorus squamosus (Chondrichthyes: Squali-
dae). Journal of the Marine Biological Association of the UK,79,
923931.
Goudet J (1999) PCA-GEN (version 1.2.1). Available from http://
www2.unil.ch/popgen/softwares/pcagen.htm (accessed 5
November 2013).
Goudet J (2001) FSTAT, a program to estimate and test gene
diversities and fixation indices version 2.9.3.1. Available
from www2.unil.ch/popgen/softwares/fstat.htm.
Grey M (1956) The distribution of fishes found below a depth
of 2000 m. Fieldiana (Zoology),36,75183.
Griffiths AM, Sims DW, Johnson A et al. (2011) Levels of con-
nectivity between longnose skate (Dipturus oxyrinchus) in the
Mediterranean Sea and the north-eastern Atlantic Ocean.
Conservation Genetics,12, 577582.
Hansen MM, Meier K, Mensberg KLD (2010) Identifying foot-
prints of selection in stocked brown trout populations: a spa-
tio-temporal approach. Molecular Ecology,19, 17871800.
Hasegawa M, Kishino H, Yano T (1985) Dating the human-ape
splitting by a molecular clock of 30 mitochondrial DNA.
Journal of Molecular Evolution,22, 160174.
Hauser L, Carvalho GR (2008) Paradigm shifts in marine fish-
eries genetics: ugly hypotheses slain by beautiful facts. Fish
and Fisheries,9, 333362.
Hedrick PW (2011) Genetics of Populations, 4th edn. Jones and
Bartlett, Burlington, Massachusetts, USA.
Hewitt GM (1996) Some genetic consequences of ice ages, and
their role in divergence and speciation. Biological Journal of
the Linnean Society,58, 247276.
Hey J, Nielsen R (2007) Integration within the Felsenstein equa-
tion for improved Markov chain Monte Carlo methods in
population genetics. PNAS,104, 27852790.
Hs
u KJ, Ryan WBF, Cita MB (1973) Late Miocene desiccation
of the Mediterranean. Nature,242, 240244.
ICES (2012) Report of the Working Group on Elasmobranch Fishes
(WGEF),1926 June 2012, Lisbon, Portugal. ICES CM 2012/
ACOM:19. 547 pp.
Irvine SB (2004) Age, growth and reproduction of deepwater dog-
fishes from southeastern Australia. PhD thesis. Deakin Univer-
sity, Warrnambool, Victoria, Australia.
Irvine SB, Stevens JD, Laurenson LBJ (2006) Surface bands on
deepwater squalid spines: an alternative method of ageing
Centroselachus crepidater.Canadian Journal of Fisheries and
Aquatic Sciences,63, 617627.
Jost L (2008) G
ST
and its relatives do not measure differentia-
tion. Molecular Ecology,17, 4015.
Karl S, Castro A, Lopez J, Charvet P, Burgess G (2011) Phylo-
geography and conservation of the bull shark (Carcharhinus
leucas) inferred from mitochondrial and microsatellite DNA.
Conservation Genetics,12, 371382.
©2015 John Wiley & Sons Ltd
6076 D. CATARINO ET AL.
Keenan K, McGinnity P, Cross TF, Crozier WW, Prod
ohl PA
(2013) diveRsity: an R package for the estimation of popula-
tion genetics parameters and their associated errors. Methods
in Ecology and Evolution,4, 782788.
Keeney DB, Heupel MR, Hueter RE, Heist EJ (2003) Genetic
heterogeneity among blacktip shark, Carcharhinus limbatus,
continental nurseries along the US Atlantic and Gulf of Mex-
ico. Marine Biology,143, 10391046.
Knutsen H, Jorde PE, Sannaes H et al. (2009) Bathymetric barri-
ers promoting genetic structure in the deepwater demersal
fish tusk (Brosme brosme). Molecular Ecology,18, 31513162.
Knutsen H, Jorde PE, Bergstad OA, Skogen M (2012) Popula-
tion genetic structure in a deepwater fish Coryphaenoides
rupestris: patterns and processes. Marine Ecology Progress
Series,460, 233246.
Krijgsman W, Hilgen FJ, Raffi I, Sierro FJ, Wilson DS (1999)
Chronology, causes and progression of the Messinian salinity
crisis. Nature,400, 652655.
Kukla G (2005) Saalian supercycle, Mindel/Riss interglacial
and Milankovitch’s dating. Quaternary Science Reviews,24,
15731583.
Lemaire C, Versini JJ, Bonhomme F (2005) Maintenance of
genetic differentiation across a transition zone in the sea: dis-
cordance between nuclear and cytoplasmic markers. Journal
of Evolutionary Biology,18,7080.
Longmore C, Trueman CN, Neat F et al. (2014) Oceanic scale
connectivity and life cycle reconstruction in a deep-sea fish.
Canadian Journal of Fisheries and Aquatic Sciences,71, 13121323.
Maia AMR, Wilga CAD, Lauder GV (2012) Biomechanics of
locomotion in sharks, rays and chimaeras. In: Biology of
Sharks and their Relatives, 2nd edn (eds Carrier JC, Musick JA,
Heithaus MR), pp. 125152. CRC Press, Boca Raton, FL, USA.
Manni F, Gu
erard E, Heyer E (2004) Geographic patterns of
(genetic, morphologic, linguistic) variation: how barriers can
be detected by “Monmonier’s algorithm”. Human Biology,76,
173190.
Martin AP (1995) Mitochondrial DNA sequence evolution in
sharks: rates, patterns and phylogenetic inferences. Molecular
Biology and Evolution,12, 11141123.
Martin AP, Naylor GJP, Palumbi SR (1992) Rates of mitochon-
drial DNA evolution in sharks are slow compared with
mammals. Nature,357, 153155.
Massut
ı M, Gordon JDM, Moranta J et al. (2004) Mediterranean
and Atlantic deep-sea fish assemblages: differences in bio-
mass composition and size-related structure. Scientia Marina,
68, 101115.
McCullagh P, Nelder JA (1989) Generalized Linear Models. Chap-
man and Hall/CRC, Boca Raton, Florida, USA.
Menezes GM, Rosa A, Melo O, Pinho MR (2009) Demersal fish
assemblages off the Seine and Sedlo seamounts (northeast
Atlantic). Deep-Sea Research II,56, 26832704.
Moura T, Serra-Pereira B, Gordo LS, Figueiredo I (2011) Sperm
storage in males and females of the deepwater shark Por-
tuguese dogfish with notes on oviducal gland microscopic
organization. Journal of Zoology,283, 210219.
Nakadate M, Vi~
nas J, Corriero A, Clarke S, Suzuki N, Chow S
(2005) Genetic isolation between Atlantic and Mediterranean
albacore populations inferred from mitochondrial and
nuclear DNA markers. Journal of Fish Biology,66, 15451557.
Narum SR, Hess JE (2011) Comparison of FST outlier tests for SNP
loci under selection. Molecular Ecology Resources,11, 184194.
Nei M, Chesser RK (1983) Estimation of fixation indices and
gene diversities. Annals of Human Genetics,47, 253259.
van Oosterhout C, Hutchinson WF, Wills DPM, Shipley P
(2004) MICRO-CHECKER: software for identifying and cor-
recting genotyping errors in microsatellite data. Molecular
Ecology Notes,4, 535538.
Palumbi SR (1992) Marine speciation on a small planet. Trends
in Ecology and Evolution,7, 114118.
Patarnello T, Volckaert FAMJ, Castilho R (2007) Pillars of Her-
cules: is the Atlantic-Mediterranean transition a phylogeo-
graphical break? Molecular Ecology,16, 44264444.
Pearce JM (2006) Minding the gap: frequency of indels in
mtDNA control region sequence data and influence on pop-
ulation genetic analyses. Molecular Ecology,15, 333341.
Peery MZ, Kirby R, Reid BN et al. (2012) Reliability of genetic
bottleneck tests for detecting recent population declines.
Molecular Ecology,21, 34033418.
P
erez-Losada M, Guerra A, Carvalho GR et al. (2002) Extensive
population subdivision of the cuttlefish Sepia officinalis (Mol-
lusca: Cephalopoda) around the Iberian Peninsula indicated
by microsatellite DNA variation. Heredity,89, 417424.
Pfeifer B, Wittelsb
urger U, Ramos-Onsins SE, Lercher MJ
(2014) PopGenome: an efficient swiss army knife for popula-
tion genomic analyses in R. Molecular Biology Evolution,31,
19291936.
Piry S, Lukart G, Cornuet JM (1999) BOTTLENECK: a com-
puter program for detecting recent reductions in the effective
population size using allele frequency data. Journal of Hered-
ity,90, 502503.
Polzin T, Daneschmand SV (2003) On Steiner trees and mini-
mum spanning trees in hypergraphs. Operations Research Let-
ters,31,1220.
R Core Team (2014) R: A Language and Environment for Statisti-
cal Computing. R Foundation for Statistical Computing,
Vienna, Austria. Available from http://www.R-project.org.
(accessed 30 July 2015).
Raymond M, Rousset F (1995) GENEPOP (version 1.2): popula-
tion genetics software for exact tests and ecumenicism. Jour-
nal of Heredity,86, 248249.
Robalo JI, Castilho R, Francisco SM et al. (2012) Northern refu-
gia and recent expansion in the North Sea: the case of the
wrasse Symphodus melops (Linnaeus, 1758). Ecology and Evolu-
tion,2, 153164.
Rousset F (2008) Genepop’007: a complete re-implementation
of the Genepop software for Windows and Linux. Molecular
Ecology Resources,8, 103106.
Rozas J, S
anchez-DelBarrio JC, Messeguer X, Rozas R (2003)
DnaSP, DNA polymorphism analyses by the coalescent and
other methods. Bioinformatics,19, 24962497.
Salmenkova EA (2011) New view on the population genetic
structure of marine fish. Russian Journal of Genetics,47,1279
1287.
Schwarz G (1978) Estimating the dimension of a model. Annals
of Statistics,6, 461464.
Seutin G, White BN, Boag PT (1991) Preservation of avian
blood and tissue samples for DNA analyses. Canadian Journal
of Zoology,69,8290.
Smith AB, Gale A (2009) The pre-Messinian deep-sea Neogene
echinoid fauna of the Mediterranean: surface productivity
controls and biogeographical relationships. Palaeogeography,
Palaeoclimatology, Palaeoecology,281, 115125.
©2015 John Wiley & Sons Ltd
POPULATION STRUCTURE OF THE PORTUGUESE DOGFISH 6077
Stefanescu C, Rucabado J, Lloris D (1992) Depth-size trends in
western Mediterranean demersal deep-sea fishes. Marine
Ecology Progress Series,81, 205213.
Stevens J, Correia JPS (2003) SSG Australia & Oceania Regional
Workshop, Centroscymnus coelolepis. In: IUCN 2013. IUCN
Red List of Threatened Species (Version 2013.2). Available
from www.iucnredlist.org (accessed 14 January 2014).
Stockley B, Menezes GM, Pinho MR, Rogers AD (2005) Genetic
population structure and evidence of an extreme bottleneck
in the black-spot sea bream (Pagellus bogaraveo) from the NE
Atlantic. Marine Biology,146, 793804.
Straube N, Kriwet J, Schliewen UK (2011) Cryptic diversity
and species assignment of large lantern sharks of the
Etmopterus spinax clade from the Southern Hemisphere
(Squaliformes, Etmopteridae). Zoologica Scripta,40,6175.
Tamura K, Peterson D, Peterson N, Stecher G, Nei M, Kumar S
(2011) MEGA5: molecular evolutionary genetics analysis
using maximum likelihood, evolutionary distance, and maxi-
mum parsimony methods. Molecular Biology and Evolution,
28, 27312739.
Tecchio S, Ram
ırez-Llodra E, Sard
aFet al. (2011) Drivers of
deep Mediterranean megabenthos communities along longi-
tudinal and bathymetric gradients. Marine Ecology Progress
Series,439, 181192.
Thompson JD, Gibson TJ, Plewniak F, Jeanmougin F, Higgins
DG (1997) The ClustalX windows interface: flexible strategies
for multiple sequence alignment aided by quality analysis
tools. Nucleic Acids Research,25, 48764882.
Tintore J, Violette PE, Blade I, Cruzado A (1988) A study of an
intense density front in the eastern alboran sea: the Almeria-
Oran front. Journal of Physical Oceanography,18, 13841397.
Torchio M, Michelangeli M (1971) Prima segnalazione in acque
italiane di uno squalide del genere Centroscymnus.Natura,
62, 241245.
Tosti L, Danovaro R, Dell’Anno A et al. (2006) Vitellogenesis in
the deep-sea shark Centroscymnus coelolepis.Chemistry and
Ecology,22, 335345.
Ver
ıssimo A, Gordo L, Figueiredo I (2003) Reproductive biol-
ogy and embryonic development of Centroscymnus coelolepis
in Portuguese mainland waters. ICES Journal of Marine
Science,60, 13351341.
Ver
ıssimo A, McDowell JR, Graves JE (2010) Global population
structure of the spiny dogfish Squalus acanthias, a temperate
shark with an antitropical distribution. Molecular Ecology,19,
16511662.
Ver
ıssimo A, McDowell JR, Graves JE (2011a) Population struc-
ture of a deep-water squaloid shark, the Portuguese dogfish
(Centroscymnus coelolepis). ICES Journal of Marine Science,68,
555563.
Ver
ıssimo A, Moura T, McDowell J, Graves J, Gordo L, Hoelzel
R (2011b) Isolation and characterization of ten nuclear
microsatellite loci for the Portuguese dogfish Centroscymnus
coelolepis.Conservation Genetics Resources,3, 299301.
Ver
ıssimo A, McDowell JR, Graves JE (2012) Genetic population
structure and connectivity in a commercially exploited and
wide-ranging deepwater shark, the leafscale gulper (Cen-
trophorus squamosus). Marine and Freshwater Research,63,505
512.
Vi~
nas J, Alvarado-Bremer J, Pla C (2004) Phylogeography of
the Atlantic bonito (Sarda sarda) in the northern Mediter-
ranean: the combined effects of historical vicariance, popula-
tion expansion, secondary invasion, and isolation by
distance. Molecular Phylogenetics and Evolution,33,3242.
Waples R (1998) Separating the wheat from the chaff: patterns
of genetic differentiation in high gene flow species. Journal of
Heredity,89, 438450.
Ward RD, Woodwark M, Skibinski DOF (1994) A comparison
of genetic diversity levels in marine, freshwater, and anadro-
mous fishes. Journal of Fish Biology,44, 213232.
Weir BS, Cockerham CC (1984) Estimating F-statistics for the
analysis of population structure. Evolution,38, 13581370.
White TA, Stefanni S, Stamford J, Hoelzel AR (2009) Unex-
pected panmixia in a long-lived, deep-sea fish with well
defined spawning habitat and relatively low fecundity.
Molecular Ecology,18, 25632573.
Yano K, Tanaka S (1988) Size at maturity, reproductive cycle,
fecundity, and depth segregation of the deep sea squaloid
sharks Centroscymnus owstoni and C. coelolepis in Suruga Bay,
Japan. Nippon Suisan Gakkaishi,54, 167174.
S.S., H.K. and G.M. supervised and coordinated the
work. D.C., S.S. and H.K. conceived the study. D.C.,
S.S., H.K., R.D. and A.D. initiated the work. D.C., S.S.,
A.V., G.M., J.B.C., F.N., R.D., A.D. and B.R. organized
and collected field samples. D.C. performed laboratory
work with major contributions of S.S. and H.S. D.C.,
H.S. and H.K. scored the microsatellites genotypes.
D.C., H.K. and E.O. analysed the data and performed
the statistical analyses with major contributions of
P.E.J., A.V., D.S. and S.S. D.C. and H.K. led the writing
of the text to which all authors contributed.
Data accessibility
The mtDNA CR sequences can be accessed in GenBank
by the Accession nos HQ664432HQ664449 and
KT320896KT321040 (new sequences from this study).
The file with all used and aligned sequences, the list of
haplotypes with haplotype frequencies, the microsatel-
lite genotypes, the biological data used in the life his-
tory analyses and the command lines used in IMa2 are
available at DRYAD: doi:10.5061/dryad.ss368.
Supporting information
Additional supporting information may be found in the online ver-
sion of this article.
Appendix S1 Material and methods.
Table S1 Portuguese dogfish sampled collection sites with
averaged coordinates in decimal degrees within sampling area.
Table S2 Nuclear microsatellite loci combined in multiplexes
(Set) and annealing temperature (T
a
).
©2015 John Wiley & Sons Ltd
6078 D. CATARINO ET AL.
Table S3 Summary statistics used in the computations in the
DIYABC software.
Table S4 Pairwise F
ST
results for temporal replicates and com-
bined sampled locations using the 11 nuclear loci.
Table S5 IM model estimates: time since divergence (t), effec-
tive population sizes of ancestral and descended populations
(ΘANC, ΘMED, ΘATL) and migration rate (mATL<>MED)
with the lower and upper boundaries of the 95% highest poste-
rior density (HPD) interval.
Table S6 Parameter estimates, with standard error (SE) and P-
values, of the generalized linear models used for describing
Portuguese dogfish maturity ogives.
Fig. S1 DIYABC schematic representation of the four alternative
scenarios used for Approximate Bayesian computation
simulations.
Fig. S2 Genetic diversity patterns across all localities, based on
11 microsatellite loci.
Fig. S3 Graphical representation of the allele frequencies and
haplotype frequencies for the 11 microsatellites markers and
mtDNA CR for the three main regions analysed.
Fig. S4 Relative posterior probability of the demographic sce-
narios estimated using the logistic regression approach.
Fig. S5 DIYABC posterior probability distributions (green lines)
for the six parameters estimated: Atlantic N
e
(N1), Mediter-
ranean N
e
(N2), Atlantic N
e
after divergence (N1b), Mediter-
ranean N
e
after divergence (N2b), time since divergence (t),
and t1 under the best supported scenarios (S4).
Fig. S6 DIYABC analyses with estimated distributions for the
genetic markers and control quality of each scenario.
©2015 John Wiley & Sons Ltd
POPULATION STRUCTURE OF THE PORTUGUESE DOGFISH 6079
... Massutí and Moranta, 2003;Sion et al., 2004). Centroscymnus coelolepis is the deepest-dwelling elasmobranch recorded (>1,417 m), which is consistent with its bathymetric distribution in different areas of the Mediterranean Sea (1,301-2,863 m; Massutí and Moranta, 2003;D'Onghia et al., 2004b;Sion et al., 2004;Tecchio et al., 2011;Catarino et al., 2015). The Norwegian skate D. nidarosiensis, the only recorded batoid species, found at a maximum depth of 1,688 m, allowed us to update its distribution information for the Mediterranean Sea, as it was hitherto found at a maximum depth of 1180 m (e.g. ...
Article
This study compiles information on the size distribution and reproductive patterns of five chondrichthyans and 10 bony species dwelling on the continental slope of the Central-Western Mediterranean (south-eastern Sardinian waters, depth 720-1890 m). The abundance of the relatively less numerous chondrichthyans was found to decrease with depth. Almost all chondrichthyan species had a long reproductive period throughout the year at all depths. In bony fishes, the upper bathymetric stratum (<1100 m) was dominated by species with continuous and seasonal (autumn-winter) spawning cycles. At intermediate and lower depths (>1100 m), continuous reproduction was the most common pattern. Both chondrichthyans and bony fishes exhibited large sizes at maturity and at first maturity (>75% and >60% of their maximum body length, respectively), supporting the hypothesis that most deep-sea fish populations are characterized by a delayed maturity, which makes them highly vulnerable to fishing pressure and over-exploitation. In some sharks, population partitioning was observed, according to life stage as a function of depth: adult/mature females occupied deeper habitats than males and juveniles. Conversely, Dipturus nidarosiensis female and male juveniles (>1100 m) were segregated from the adult/mature population, dwelling between 1000 and 1100 m. In four of the 10 bony fishes examined, we found a clear and significant predominance of females, particularly pronounced in Alepocephalus rostratus, which is clearly related to their K-reproductive strategy. The four key-selected species (Galeus melastomus, A. rostratus, Bathypterois mediterraneus and Trachyrincus scabrus), characterized by different reproductive strategies/modal-ities and feeding behaviours, were differently affected by depth and fluxes of organic matter (Particulate Organic Carbon concentration) from the photic zone towards the sea-bottom as a function of sex and life-stage.
Article
Full-text available
The American eel (Anguilla rostrata) has long been regarded as a panmictic fish and has been confirmed as such in the northern part of its range. In this paper, we tested for the first time whether panmixia extends to the tropical range of the species. To do so, we first assembled a reference genome (975 Mbp, 19 chromosomes) combining long (PacBio and Nanopore and short (Illumina paired‐end) reads technologies to support both this study and future research. To test for population structure, we estimated genotype likelihoods from low‐coverage whole‐genome sequencing of 460 American eels, collected at 21 sampling sites (in seven geographic regions) ranging from Canada to Trinidad and Tobago. We estimated genetic distance between regions, performed ADMIXTURE‐like clustering analysis and multivariate analysis, and found no evidence of population structure, thus confirming that panmixia extends to the tropical range of the species. In addition, two genomic regions with putative inversions were observed, both geographically widespread and present at similar frequencies in all regions. We discuss the implications of lack of genetic population structure for the species. Our results are key for the future genomic research in the American eel and the implementation of conservation measures throughout its geographic range. Additionally, our results can be applied to fisheries management and aquaculture of the species.
Article
Full-text available
Deep-water sharks are highly diverse, vulnerable, and understudied as a group, despite the increasing pressures on their populations. Twenty-five species of deep-water sharks have been recorded in the Azores, an oceanic archipelago in the mid-North Atlantic, that are regularly caught as bycatch in hook-and-line fisheries. Avoiding the bycatch of deep-water sharks presents multiple challenges due to their high catchability, difficulties in correctly identifying species, and the general lack of data on these species. This review summarizes the findings of recent studies from the region, providing an up-to-date science-based framework for mitigating bycatch effects of Azorean hook-and-line fisheries. Several depth-based, area-based, and gear-based measures have been studied that demonstrate the potential to either avoid or increase the survival of deep-water shark bycatch. However, these measures may have limited efficacy for some species (e.g. highly mobile species) and thus, limited widespread applicability. Convincing fishers to avoid deep-water shark bycatch is also a challenge given the antagonistic interactions with sharks damaging the catch and fishing gear, while simultaneously a market incentive for shark liver oil remains. It highlights the need to proactively engage fishers and incentivize the mitigation of bycatch of deep-water sharks in Azorean waters.
Article
Full-text available
Identifying how physical and biotic factors shape genetic connectivity among populations in time and space is essential to our understanding of the evolutionary trajectory as well as the management of marine species. Atlantic cod is a widespread and commercially important marine species displaying several ecotypes with different life history strategies. Using three sets of SNPs: neutral, informative, and genome‐inversion linked, we studied population genetic structure of ~2500 coastal Atlantic cod (CC) from 40 locations along Norway´s 2500km coastline, including nine fjords. We observed: 1) a genetic cline, suggesting a mechanism of isolation by distance, characterized by a declining FST between CC and North East Arctic Cod (NEAC – genetically distinct migratory ecotype) with increasing latitude, 2) that in the north, samples of CC from outer‐fjord areas were genetically more similar to NEAC than were samples of CC from their corresponding inner‐fjord areas, 3) greater population genetic differentiation among CC sampled from outer‐fjord areas along the coast, than among CC sampled from their corresponding inner‐fjord areas, 4) genetic differentiation among samples of CC from both within and among fjords. Collectively, these results permit us to draw two main conclusions. First, that differences in the relative presence of the genetically highly distinct, migratory ecotype NEAC, declining from north to south and from outer to inner fjord, plays the major role in driving population genetic structure of the Norwegian CC. Second, that there is limited connectivity between CC from different fjords. These results suggest that the current management units implemented for this species in Norway should be divided into smaller entities. Furthermore, the situation where introgression from one ecotype drives population genetic structure of another, as is the case here, may exist in other species and geographical regions, thus creating additional challenges for sustainable fisheries management.
Article
Full-text available
Background: The interplay of animal dispersal and environmental heterogeneity is fundamental for the distribution of biodiversity on earth. In the ocean, the interaction of physical barriers and dispersal has primarily been examined for organisms with planktonic larvae. Animals that lack a planktonic life stage and depend on active dispersal are however likely to produce distinctive patterns. Methods: We used available literature on population genetics and phylogeography of elasmobranchs (sharks, rays and skates) to examine how marine barriers and dispersal ecology shape genetic connectivity in animals with active dispersal. We provide a global geographical overview of barriers extracted from the literature and synthesize the geographical and hydrological factors, spatial and temporal scales to characterize different types of barriers. The three most studied barriers were used to analyse the effect of elasmobranch dispersal potential and barrier type on genetic connectivity. Results: We characterized nine broad types of marine barriers, with the three most common barriers being related to ocean bathymetry. The maximum depth of occurrence, maximum body size and habitat of each species were used as proxies for dispersal potential, and were important predictors of genetic connectivity with varying effect depending on barrier type. Environmental tolerance and reproductive behaviour may also play a crucial role in population connectivity in animals with active dispersal. However, we find that studies commonly lack appropriate study designs based on a priori hypotheses to test the effect of physical barriers while accounting for animal behaviour. Main conclusions: Our synthesis highlights the relative contribution of different barrier types in shaping elasmobranch populations. We provide a new perspective on how barriers and dispersal ecology interact to rearrange genetic variation of marine animals with active dispersal. We illustrate methodological sources that can bias the detection of barriers and provide potential solutions for future research in the field.
Article
Full-text available
Skates are characterised by conservative body morphology which hampers identification and leads to frequent taxonomic confusion and market mislabelling. Accurate specimen classification is crucial for reliable stock assessments and effective conservation plans, otherwise the risk of extinction could be unnoticed. The misclassification issue is evident for the genus Dipturus, distributed worldwide, from the continental shelf and slope to the deep sea. In this study, barcode cytochrome oxidase I gene (COI) sequences were used along with species delimitation and specimen assignment methods to improve taxonomy and zoogeography of species of conservation interest inhabiting the Atlantic Ocean and Mediterranean Sea. In this study, we provided new evidence of the occurence of D. nidarosiensis in the Central‐Western Mediterranean Sea and the lack of Atlantic‐Mediterranean genetic divergence. The Atlantic endangered species D. laevis and D. batis clustered together under the same molecular operational taxonomic unit (MOTU) with any delimitation methods used, while the assignment approach correctly discriminated specimens into the two species. These results provided evidence that the presence of the barcode gap is not an essential predictor of identification success, but the use of different approaches is crucially needed for specimen classification, especially when threshold‐ or tree‐based methods result less powerful. The analyses also showed how different putative, vulnerable, species dwelling across South‐Western Atlantic and South‐Eastern Pacific are frequently misidentified in public sequence repositories. Our study emphasised the limits associated to public databases, highlighting the urgency to verify and implement the information deposited therein in order to guarantee accurate species identification and thus effective conservation measures for deep‐sea skates.
Article
Full-text available
The Portuguese dogfish Centroscymnus coelolepis is a wide‐ranging deep‐water shark and a common by‐catch component of the catches of several mid‐ to deep‐water fisheries. In the present study, two new records from the south‐western Atlantic Ocean are reported based on specimens caught by bottom‐longline fishing vessels operating in the Argentinean–Uruguayan Common Fishing Zone. Species identification based on morphology and detailed morphometrics, as well as molecular data are presented for one of the specimens. The distribution of the species over the south‐western Atlantic is discussed on the basis of the available bibliography and a thorough revision of museum collections. The records presented here expand the species' previously acknowledged distribution southwards, from around 21° S to at least 38° S, suggesting it occurs continuously along the shelf break of eastern South America. However, given the limited access to specimens of deep‐water sharks in the region, the abundance and real extent of C. coelolepis distribution in the south‐western Atlantic as well as its interaction with deep‐water fisheries remain to be fully assessed.
Article
Full-text available
Data obtained over a period of twenty years from 214 bottom trawls, towed on a single warp at depths between 402 and 1993 m in the western Mediterranean (Algerian and Balearic basins) and eastern North Atlantic (Rockall Trough and Porcupine Seabight), allowed a standardised comparison of density, biomass composition and size-related structure for both the whole fish fauna and for the most common species found within the deep-sea fish assemblages. All four areas are characterised by distinctly different and well-documented oceanographic conditions, biogeographical affinities and fishing exploitation. The results showed clear differences between the Atlantic and the Mediterranean deep-sea fish fauna, not only in density, species richness and composition, but also in the structure of the biomass that constitutes these assemblages. These differences are discussed in relation to environmental conditions and fishing pattern, which have determined the adaptive responses of both individual species and the whole ecosystem.