ArticlePDF Available

Atlantic bonito (Sarda) genomic analysis reveals population differentiation across Northeast Atlantic and mediterranean locations: Implications for fishery management

Authors:
Marine Environmental Research xxx (xxxx) xxx
Please cite this article as: Judith Ollé-Vilanova et al., Marine Environmental Research, https://doi.org/10.1016/j.marenvres.2024.106408
Available online 13 February 2024
0141-1136/© 2024 Elsevier Ltd. All rights reserved.
Atlantic bonito (Sarda) genomic analysis reveals population differentiation
across Northeast Atlantic and mediterranean locations: Implications for
shery management
Judith Oll´
e-Vilanova
a
, Ghailen Hajjej
b
, David Macias
c
, S´
amar Saber
c
, Pedro G. Lino
d
,
Rub´
en Mu˜
noz-Lechuga
d
,
e
, SidAhmed Baibbat
f
, Fambaye Ngom Sow
g
,
Nguessan Constance Diaha
h
, Rosa M. Araguas
a
, Núria Sanz
a
, Jordi Vinas
a
,
*
a
Laboratori Ictiologia Genetica, Departament de Biologia, Universitat de Girona, 17003, Girona, Catalunya, Spain
b
National Institute of Marine Science and Technology, Tunisia
c
Centro Oceanogr´
aco de M´
alaga. Instituto Espanol de Oceanografía (IEO-CSIC), Spain
d
Instituto Portugues Do Mar e da Atmosfera (IPMA), Avenida 5 de Outubro S/n, 8700-305, Olh˜
ao, Portugal
e
Department of Biology, Faculty of Marine and Environmental Sciences, University of C´
adiz, 11510, Puerto Real, C´
adiz, Spain
f
Laboratoire des Peches (Dakhla), Morocco
g
Centre De Recherches Oceanographiques de Dakar, Senegal
h
Centre National de Recherches Oc´
eanologiques, Abidjan, Cote dIvoire
ARTICLE INFO
Keywords:
Atlantic bonito
ddRadSeq
Population genomics
Atlantic -mediterranean transition
Fishery management
Pelagic
Genetics
Conservation
ABSTRACT
The Atlantic bonito (Sarda, family Scombridae) is a pelagic species and one of the most exploited small tuna
species. Despite its economic importance, biological information is scarce with no associated management di-
rectives. In this study, using a population genomic approach resulted in a lack of panmixia of two genetic pools
with different effective population sizes (east central-tropical Atlantic and northeast Atlantic-Mediterranean) and
an intermixing zone in Atlantic Morocco. The absence of genetic heterogeneity between the locations in Atlantic -
Mediterranean transitional zone adds new evidence that challenges the Strait of Gibraltar as a phylogeographic
barrier for marine pelagic species. These results are proposed to the ICCAT Commission to establish management
areas for this species, although they are not consistent with the recently adopted pelagic ecoregions. Finally, a
panel of highly informative SNPs was developed for efcient and low-cost population assignment and the
analysis of unresolved population structures.
1. Introduction
In the epipelagic realm, assessing the population structure of marine
species is particularly challenging. The apparent lack of biogeographical
barriers and the potential homogenizing effect of marine currents,
together with the biological characteristics of epipelagic organisms,
such as migratory behavior, long lifespan, large population sizes, and
pelagic larval stages, are thought to promote low levels of genetic dif-
ferentiation (Cowen et al., 2007). However, this classical view of the
population structure of these pelagic species has been challenged in
numerous studies. A combination of several biotic and abiotic factors
contributes to the formation of genetically differentiated populations
(Oll´
e et al., 2022; R.S. Waples, 1998). Historical paleogeographic
changes at ocean levels may have contributed to the isolation of pop-
ulations, leaving a genetic ngerprint that can currently be detected
genetically (Allaya et al., 2015; Forde et al., 2023). Furthermore, species
life history traits and strategies, such as the duration of larval stages
combined with phylogeographic discontinuities caused by oceanic
fronts or gyres, could promote local retention and self-recruitment of
larvae, contributing to the genetic differentiation of populations (Pasc-
ual et al., 2017; Van Wyngaarden et al., 2017).
A classic example of a gene ow barrier is the gateway between the
Atlantic Ocean and the Mediterranean Sea that is he Strait of Gibraltar
and its associated Almeria-Oran oceanographic gyre (Patarnello et al.,
2007). Conversely, for some species, especially large pelagic sh, the
Strait of Gibraltar does not seem to act as a barrier to gene ow. For
* Corresponding author.
E-mail address: jordi.vinas@udg.edu (J. Vinas).
Contents lists available at ScienceDirect
Marine Environmental Research
journal homepage: www.elsevier.com/locate/marenvrev
https://doi.org/10.1016/j.marenvres.2024.106408
Received 10 November 2023; Received in revised form 8 February 2024; Accepted 12 February 2024
Marine Environmental Research xxx (xxxx) xxx
2
instance, in pelagic species such as Atlantic bluen and albacore tunas
(Davies et al., 2011; Rodríguez-Ezpeleta et al., 2019), it has been
observed that the genetic pool of Mediterranean populations extends
into the waters of the eastern Atlantic, crossing this elusive biogeo-
graphic barrier. Thus, it is essential to identify these biogeographical
barriers that can lead to the isolation of populations to promote the
conservation of the species. Particularly in the case of species considered
shery resources, knowledge of the population structure is essential to
propose management measures (Nielsen et al., 2023; Reiss et al., 2009).
This is even more crucial when the lack of knowledge of not only the
population structure but also basic biological traits of the species hinders
the proposal of appropriate management measures (Lucena-Fr´
edou
et al., 2021).
A paradigmatic case of a commercially important species is the
Atlantic bonito, whose existing biological information is scarce and
focused on specic areas (Kahraman et al., 2014; Macías et al., 2005).
Atlantic bonito belongs to the family Scombridae, which also includes
other highly prized species such as Bluen tuna (Thunnus thynnus), al-
bacore (T. alalunga), and Little tunny (Euthynnus alletteratus). Although
it is considered a small tuna species, it can reach up to 100 cm in length
and weigh more than 7 kg (Collette and Nauen, 1983). It is mainly
targeted by coastal and artisanal sheries, although signicant catches
are also from bycatch (Juan-Jord´
a et al., 2020), being one most
exploited small tuna sh species with more than 20,000 mt per year
(Lucena-Fr´
edou et al., 2021). Despite Atlantic bonito of this high eco-
nomic, there are currently no management measures for this species. In
this regard, the International Commission for the Conservation of
Atlantic Tunas (ICCAT) initiated a specic research plan in 2017 to
deepen knowledge of biology (focused on reproduction and growth) and
population structure, prioritizing Atlantic bonito, as well as the little
tunny and wahoo as a species objective (SCRS, 2019).
The aim of the present study is to analyze the genetic population
structure of Atlantic bonito in localities in the eastern Atlantic and
western Mediterranean, located on both sides of the Strait of Gibraltar.
In a preliminary study, using a single mitochondrial marker and with a
geographic range similar to that used in the present study, genetic dif-
ferentiation between localities in the eastern Atlantic and Mediterra-
nean areas was detected (Vi˜
nas et al., 2020). Here, we use a genomic
approach, ddRadseq, which has proven to be a highly effective tool for
determining population structure in pelagic species, particularly when a
shallow level of genetic differentiation is expected, and allows the
detection of possible genomic regions susceptible to selection (Narum
et al., 2013). Thus, the importance of this study is, at least, twofold: i) to
evaluate the possible biogeographical barrier of the Strait of Gibraltar in
this pelagic sh species, and ii) to acquire essential knowledge of the
population structure to promote sustainable sheries.
2. Material and methods
As part of the ICCAT SMT program (Fr´
edou et al., 2021), samples of
92 Atlantic bonitos were collected from the years 20182019 in seven
locations in the western Mediterranean and eastern Atlantic species
distribution. Locations were distributed in three ICCATs statistical areas
(see Table 1 and Fig. 1). A small muscle tissue or portion of the n was
excised, preserved in 96% ethanol, and shipped to the Laboratori dIc-
tiologia Gen`
etica of the Universitat de Girona.
Total DNA extraction was performed using the DNeasy Blood &
Tissue Kit (Qiagen), including an RNase A digestion step. DNA quantity
and integrity were assessed using a Qubit dsDNA HS Assay Kit (Life
Technologies) and a 1% agarose gel. For RAD-seq library preparation, a
total amount of 100 ng of DNA per sample was digested with the two
endonucleases, PstI and MspI. Adaptor and padding sequences were
added to each fragment. Paired-end sequencing of 151 bp was per-
formed using Illumina NovaSeq 6000 S
4
(Genomics Center, University of
Minnesota). Once the reads were sequenced, padding sequences and
restriction sites were removed following the sequencing service di-
rections with gbstrim. pl (https://bitbucket.org/jgarbe/gbstrim/src/
master/). After removing the padding sequences and restriction sites,
reads with different lengths were trimmed to 80 bp using Process_radtags
in STACKS 2 ((Rochette et al., 2019). Process_radtags were also used to
lter reads that presented low-quality scores (sliding window 0.25 and
score threshold 20). FastQC and MultiQC (FastQC,2015) were used to
assess read quality during cleansing. Reads were mapped against the
Euthynnus afnis reference genome (BPLY00000000.1 VRT
02-OCT-2021) using BWA-Mem (Li and Durbin, 2009) with default pa-
rameters. From the aligned BAM les, a loci catalog was created under
gstacks in STACKS2. The populations program from STACKS2 was used
to extract information only from biallelic loci and generate the VCF le
with the following parameters: 0.05 as the minimum rate to accept a
polymorphism; maximum observed heterozygosis for a locus 0.6, to
avoid paralogous loci; minimum percentage of individuals in a popu-
lation of 0.75, and two as the minimum populations to retain a locus.
2.1. Filtering radseq
STACKs output data were ltered for variant depths with less than
0.05 missing data for each marker across all samples using the R package
vcfR (Knaus and Grünwald, 2017). Detection loci under selection were
realized using two independent methods, considering only loci under
selection the ones shared by both. The OutFLANK R package (Whitlock
and Lotterhos, 2015) was used to infer the F
ST
distribution of neutral loci
by trimming the extreme values. OutFLANK was used in the ltered data
set using the following parameters left and right trimming fractions of
0.05, minimum heterozygosity of 0.1, and FDR q-value of 0.01. The
second approach was realized using the pcadapt package (Luu et al.,
2017; Priv´
e et al., 2020), which infers the loci under selection based on
principal component analysis (PCA). Pcadpat determines the population
structure based on PCA and computes a vector containing K z scores to
infer the extent to which an SNP is related to the rst K principal
components. Using the command scree plot, the number of PCs to be
retained was selected when the eigenvalues were attened. Outliers
were selected using the FDR Bonferroni correction with a q-value of
0.01.
Table 1
Sample description and basic genetic diversity indices for each location. ICCAT statistical area; N, sample size; Ar, Allele richness; Obs_Het, Observed heterozygosity;
Exp_Het, Expected Heterozygosity; F
IS
, F-statistic inbreeding coefcient; HWE, Hardy-Weinberg Equilibrium.
Location Code N ICCAT area Ar Obs_Het Exp_Het F
IS
Test global HWE
East Atlantic
Abidjan Cˆ
ote DIvoire CIV 14 AT-SE/BIL97 1.788 (1.7131.856) 0.220 0.210 0.039 (0.1070.045) 1.000
Hann- Senegal SEN 14 AT-NE/BIL94B 1.783 (1.7101.842) 0.228 0.210 0.067 (0.1350.070) 1.000
Dahkla- Morocco MOR 14 AT-NE/BIL94B 1.787 (1.6951.839) 0.233 0.214 0.068 (0.1480.067) 1.000
Peniche- Portugal North PRT_N 11 AT-NE/BIL94B 1.791 (1.7331.848) 0.222 0.214 0.036 (0.1040.052) 1.000
Olhao- Portugal South PRT_S 14 AT-NE/BIL94B 1.809 (1.7361.870) 0.223 0.216 0.031 (0.0880.045) 1.000
Mediterranean
Malaga- Spain ESP 14 MD/BIL95 1.792 (1.6781.854) 0.230 0.211 0.066 (0.1350.065) 1.000
Gabes- Tunis TUN 11 MD/BIL95 1.772 (1.6761.836) 0.232 0.214 0.067 (0.1500.077) 1.000
J. Oll´
e-Vilanova et al.
Marine Environmental Research xxx (xxxx) xxx
3
Fig. 1. Map of sampling localities. Sampling codes described in Table 1. Dotted grey lines and text represented the ICCATs BIL statistical areas used as references for
management. Colour of each location is maintained thorough the study.
J. Oll´
e-Vilanova et al.
Marine Environmental Research xxx (xxxx) xxx
4
2.2. Genetic population analysis
Basic genetic population indices, including genetic diversity, het-
erozygosity, allelic richness (Ar), and F
IS
, were estimated using the
command basicStats from the package diveRsity (Keenan et al., 2013).
Discriminant analysis of principal components (DAPC) (Jombart et al.,
2010) was run using the adegenet (T. Jombart and Ahmed, 2011)
package. The number of PCs to retain was inferred by cross-validation
and using the a-score as described in the adegenet package. Then, to
identify the optimal number of clusters (K) it has run sequentially with
increasing values of K. The Bayesian Information Criterion (BIC) of the
different clustering solutions are compared, accepting the most optimal
solution the one with the lowest BIC. Finally, the number of clusters and
the group membership probabilities were estimated using the distribu-
tion of the discriminant functions. Additionally, to the supervised DAPC
method, the genetic structure was also validated by an unsupervised
non-model-based approach using a principal component analysis (PCA)
as implemented in the package adegenet. Pairwise F
ST
among locations
was tested using the command pairwise. neifst using the package hierfstat
(Goudet, 2005). Analysis of molecular variance (AMOVA) (Excofer
et al., 1992) testing different possible population structures was realized
using the command poppr. amova from the package Poppr (Kamvar et al.,
2014).
Several panels with the most signicant SNPs were obtained from the
genetically separated locations using the SambaR package using the
ndstructure function which detects the SNPs with the highest standard
deviation in population minor allele frequencies (de Jong et al., 2021).
Each panel was tested using a the unsupervised PCA analysis in adegenet.
The effective population size (N
e
) of each genetically differentiated
population was inferred using NeEstimatorV2.1 (Do et al., 2014) soft-
ware based on linkage disequilibrium using the nonrandom association
of alleles at different gene loci algorithm. Finally, isolation by distance
was evaluated by testing the correlation by the linearized pairwise F
ST
and geographical distance using a Mantel test with 1000 permutations
using the adegenet (Jombart, 2008) and ade4 (Dray and Dufour, 2007)
packages.
3. Results
The number of demultiplexed sequence reads ranged from 11.4 to
34.1 million per individual, with an average of 17.4 (±4.8). Once reads
were cleansed and padding and restriction enzyme sequences removed,
the remaining reads ranged from 8.4 to 25.9 million, with an average of
12.6 (±3.7) per sample, representing an average of 72.5% (±3.9) of the
reads kept per sample. The RAD loci catalog consisted of 629,186 loci
with a mean of 137.4 sites per loci, and coverage ranged from 25.6x to
68.9x with an average of 35.3x (±7.9x), which resulted in a total of
98.842 variants with 27.96% missing data. Finally, the data set with less
than 2.5% missing data consisted of 8770 SNPs.
Similar levels of genetic diversity were obtained for all locations
(Table 1), with a lack of inbreeding and an excess of heterozygotes, as
noted by signicant negative F
IS
values across locations. Of the 8770
SNPs, only 4 SNPs were detected as outliers by both approximations
(Outank and PCAadapt). BLAST results of these loci are described in
Supplementary Information Table S1. None of the loci resulted in sig-
nicant alignment with a well annotated gene. Since only 4 SNPs
(0.045% of the total) were considered under selection, we decided to
keep a unique data set including all the SNPs for all subsequent analyses.
Fig. 2. Discriminant Analysis of Principal Components (DAPC). A, Number of clusters (K) identied using the Bayesian Information Criterion (BIC). B, Individual
density plot on the rst discriminant function. C, correspondence between locations and inferred clusters. D, DAPC scatterplot where dots represent individuals,
colours denoting sampling origin and inclusion of 95% inertia ellipses. Colour codes for locations as same as Fig. 1.
J. Oll´
e-Vilanova et al.
Marine Environmental Research xxx (xxxx) xxx
5
The BIC criterion of the number of clusters was the lowest in the
aggrupation with K =2 (Fig. 2A). The two groups can be recognized by
plotting the densities of individuals on the membership of a discriminant
function (Fig. 2B). The rst group including the samples from the central
east Atlantic (CIV, SEN and MOR) and a second group of samples from
the northeast Atlantic and Mediterranean (PRT_N, PRT_S, ESP and TUN).
The individuals ‘s location membership to each group (in this case two
groups) can also be summarized plotting the group membership to each
individual (Fig. 2C), and also the scatter plot (Fig. 2D). This population
structure was further conrmed by unsupervised PCA by aggregating
the locations into the same two groups detected previously (Fig. 3).
Finally, the posterior probably membership probability (Fig. 4) de-
picts a situation of limited intermixing of locations. Accordingly, no
shared genotypes were found between CIV and TUN. On the other hand,
MOR is situated in an intermediate situation. Some genotype represen-
tatives from ESP can be found from TUN up to SEN (but not in CIV), with
a reduction in ESP genotypes as geographical distance increases. On the
other hand, the connection between the two localities on each side of the
strait (PRT_S and ESP) can be observed by the high number of shared
genotypes between them (Fig. 4). According to the posterior probability,
these two locations present 32% and 25% of external genotypes,
respectively. In both cases, around 70% of these external genotypes were
from the location on the other side of the strait. The pattern of the two
groups of samples was also observed in the pairwise F
ST
s results (Table 2
and Fig. 5A). As expected, the two locations with the highest F
ST
s values
correspond to the two locations most geographically separated (i.e., CIV
and TUN) (Fig. 5B). Therefore, it is not surprising to nd a slight but
positive correlation between geographical distance and genetic differ-
entiation (Mantel test of isolation by distance p value: 0.029). Further-
more, it can be observed some degree of genetic differentiation within
groups. For instance, the Tunis (TUN) location is signicantly differen-
tiated than the other three locations of the same aggregation (PRT_N,
PRT_S, ESP). In contrast, there is no clear pattern of differentiation in the
other group. The MOR locality is signicantly differentiated from SEN,
but not from CIV location.
The analysis of molecular variance (AMOVA) also supports the ag-
gregation of locations in two groups. Between groups variance (F
CT
and
Va) is maximized when the locations were grouped into the two areas
(central east Atlantic, and northeast Atlantic/Mediterranean) (Supple-
mentary Table S2). Remarkably, when the locations were grouped
following a geographical arrangement (i.e., Atlantic vs Mediterranean
locations), F
CT
and Va were not statistically signicant (Supplementary
Table S2).
Accepting the population structure of K =2, samples were grouped
according to this pattern. Subsequently, the effective population size of
each genetically differentiated unit can be estimated. The group of lo-
cations in the central east Atlantic (CIV, SEN and MOR) resulted in a
relatively low effective population size of Ne =439.0 individuals,
whereas the other group (PRT_N, PRT_S, ESP and TUN) resulted in a
larger estimated effective population size of Ne =1555.7 individuals.
Finally, the validation of the panels of the most signicant SNPs was
based on the population structure of the two groups. PCAs depicted that
using only 100 selected SNPs is adequate to detect the observed genetic
structure (See Supplementary Fig. 1). This panel of SNPs could be used
for a low-cost and efcient methodology of evaluating the population
structure in Atlantic bonito.
4. Discussion
The population structure of Atlantic bonito along the eastern
AtlanticMediterranean region results in a lack of panmixia, with two
differentiated genetic pools (Figs. 2 and 3). One genetic cluster includes
two locations within the Mediterranean (Tunis and Spain) and two lo-
cations along the coast of Portugal in the Northeast Atlantic. The other
genetic pool is composed of the three central-tropical Atlantic locations
of Morocco, Senegal, and Cˆ
ote dIvoire. Although all results indicate the
presence of two distinct populations, the relatively small sample size
may compromise these results. However, previously, in a preliminary
study using a single molecular marker (mtDNA control region), Vi˜
nas
et al., (2020) also supported the genetic structure found here using
ddRADseq. Therefore, the well-known higher resolution of genomic
approaches over conventional genetic methods, even with small sample
sizes (Andrews et al., 2016; Hohenlohe et al., 2021), conrms the pop-
ulation structure of two well-differentiated genetic pools. These results
substantiate the capability of this species to form in genetically differ-
entiated locations at large oceanic levels (between the northeast Atlantic
and northwest Atlantic (Vi˜
nas et al., 2010) and also at a more regional
scale, within the Mediterranean (Turan, 2015; Vi˜
nas et al., 2004).
The genetic structure of the population obtained here highlights the
absence of genetic heterogeneity between localities on both sides of the
Strait of Gibraltar with (PRT_S and ESP) as indicated by the large
number of genotypes shared between these two locations (Fig. 4) lack of
FST signicance (Table 2). This indicates that the west Mediterranean
genetic pool extends into the northeastern Atlantic Ocean, as suggested
by the migratory movement of this species across Strait of Gibraltar (Rey
et al., 1984; Rey and Cort, 1981). The inuence of the Mediterranean
population on the eastern Atlantic in a highly vagile species was also
Fig. 3. Principal component analysis (PCA), using the whole RAD-seq dataset,
showing individuals and 95% inertia ellipses on the coordinates of the rst two
components. Colour codes for locations as same as Fig. 1.
Fig. 4. Posterior probability membership for each individual grouped by lo-
cations. Colour codes for locations as same as Fig. 1.
J. Oll´
e-Vilanova et al.
Marine Environmental Research xxx (xxxx) xxx
6
observed in swordsh (Xiphias gladius) (Smith et al., 2015), with a
strikingly similar pattern to the Atlantic bonito. In swordsh, the Med-
iterranean genetic pool extends to the Atlantic coast of Morocco.
Concordance with this pattern has also been observed in other small
tuna species: Bullet tuna (Auxis rochei) (Oll´
e-Vilanova et al., 2022),
Frigate tuna (Auxis thazard) (Oll´
e et al., 2019) and Little tunny
(Euthynnus alletteratus) (Oll´
e et al., 2022), where no genetic differenti-
ation was detected between localities on either side of the Strait of
Gibraltar. However, in these cases, the geographical location of the gene
ow boundary in the Atlantic could not be determined due to the lack of
appropriate samples in the Atlantic region. All these results, together
with the well-established migratory movement of these species through
the strait, fuel the controversy about whether the Strait of Gibraltar
represents a phylogeographic barrier for marine species (Patarnello
et al., 2007). In most of these cases, the population structure is more
dependent on oceanographic conditions (such as gyres and others) and
philopatric behavior of the species (Oll´
e-Vilanova et al., 2022; Patar-
nello et al., 2007; Smith et al., 2015), rather than the Strait of Gibraltar
acting as a biogeographical barrier.
Additional evidence for the existence of two population units is the
different demography of each genetic differentiated pool. The estima-
tion of the effective population size was approximately threefold higher
for the Mediterranean northeast Atlantic population (Ne =1555.7) than
for the central tropical Atlantic population (Ne =439.0). Assessment of
Ne in marine species can be extremely misleading (Marandel et al.,
2019; R. S. Waples, 2016), although it can be overcome when hundreds
or thousands of markers are used (F. Marandel et al., 2020; R. K. Waples
et al., 2016), as was done in this study. However, considering that in
recent studies on large pelagic species, the Ne/N ratio has been esti-
mated to be as low as 0.25 (R. S. Waples et al., 2018), it would imply that
both populations, but especially the centraol-tropical population, could
be compromised by not having enough genetic diversity to adapt to
environmental and anthropogenic threats.
The result of genetic differentiation in these two units seems to
contradict the positive IBD correlation also detected here. However, the
identication of a correlation between genetic differentiation and
geographic distance does not exclude the presence of genetically
differentiated structures. Previously, a population structure based on
IBD was detected in this species in the Mediterranean (Turan, 2015;
Vi˜
nas et al., 2004). Thus, given all the results, it can be inferred that
population structure based on IBD for Atlantic bonito would encompass
an extensive area from the eastern Mediterranean to the Atlantic African
coast. It remains to be determined whether this IBD pattern extends into
the western Atlantic, as well as what occurs in other unsampled regions
such as the Cantabrian and North Sea, where this species is also present.
Furthermore, in addition to IBD, adaptation could contribute to the
genetic differentiation of populations, as has been seen in other small
pelagic species, such as anchovy (G. Catanese et al., 2017; Montes et al.,
2016) or horse-mackerel (Fuentes-Pardo et al., 2023). This possibility
could not be ruled out; however, it should be noted that only four (out of
800 SNPs) were detected under selection. Thus, and putting all the re-
sults together, Atlantic bonito stands out as a canonical example of
population structure in a marine pelagic species, where the combination
of the assumed absence of barriers is expected to produce a population
structure of isolation by distance. However, it does not exclude the
formation of discrete genetic groups, probably as a consequence of the
combination of elusive biogeographic barriers with philopatric behavior
(Pascual et al., 2017).
Another signicant aspect is the intermediate situation of the
Moroccan location. Although genetically grouped with Ivory Coast and
Senegal, Morocco is the location that shares more genotypes with other
areas (approximately 33.3% of the genotypes come from other locations
(Fig. 4), suggesting that the inuence of the northern genetic pool,
although diluted, could reach up to Morocco. Studies on the migratory
movements of Atlantic bonito are scarce, and although its high migra-
tory capacity of more than 35 km/day has been reported (Rey et al.,
1984), the species seems to show a philopatric behavior toward its
breeding areas, with an evidence that some individuals reside
year-round in the western Mediterranean (Sabat´
es and Recasens, 2001).
In the case of Morocco, this region seems to act as a breeding ground, as
Table 2
Pairwise FSTs values and associated p values among all locations. Below diagonal F
ST
s values; above diagonal, p values. Locations codes as Table 1.
CIV SEN MOR PRT_N PRT_S ESP TUN
CIV 0.00E+00 1.27E-01 0.00E+00 0.00E+00 0.00E+00 0.00E+00
SEN 0.003 0.00E+00 0.00E+00 0.00E+00 0.00E+00 0.00E+00
MOR 0.001 0.003 0.00E+00 0.00E+00 0.00E+00 0.00E+00
PRT_N 0.010 0.008 0.007 5.46E-02 0.00E+00 0.00E+00
PRT_S 0.008 0.007 0.005 0.001 1.00E-04 0.00E+00
ESP 0.010 0.009 0.006 0.004 0.002 0.00E+00
TUN 0.015 0.013 0.012 0.005 0.004 0.006
Fig. 5. A, Heatmap of the pairwise F
ST
s values; colour gradients represent F
ST
s values, from lowest (dark green) to highest (dark red). B, Boxplot of the F
ST
s values for
each location when compared to the rest of locations. Colour codes for locations as same as Fig. 1.
J. Oll´
e-Vilanova et al.
Marine Environmental Research xxx (xxxx) xxx
7
indicated by the presence of larvae on the Moroccan coast (Arkhipov and
Pak, 2019; Rodríguez-Roda and Di Centa, 1981), together with the
capture of actively reproductive individuals in this region (Baibbat et al.,
2017; Petukhova, 2020). Migratory philopatric behavior toward this
breeding ground seems to be conrmed by tagging, where at the
beginning of the reproductive season, sh from the western Mediterra-
nean were tagged in the tuna trap of Ceuta and then recovered in both
the South and North Atlantic areas of the Strait (Rey et al., 1984; Rey
and Cort, 1981). Within the Mediterranean, the Tunis location seem to
be genetically differentiated from rest of locations of the same group.
Adding new evidence of the capability of this species to form genetically
differentiated populations on a relatively small scale. However, for this
location there is little biological information to support this putative
distinct population, but with some evidence ruling out the Tunisian
region as a breeding area (Koched et al., 2016). Therefore, further
analysis is needed to establish the population structure of this species in
the southern Mediterranean.
5. Applications
Fishery management should be based on the identication of
demographically independent genetic units (Hauser and Carvalho,
2008; R. S. Waples et al., 2008). Despite its economic importance, this
species lacks a specic management directive or stock structure. This is
probably a consequence of the deciency of knowledge of life-history
traits and historical catches (Lucena-Fr´
edou et al., 2021). To ll this
gap, in 2017, ICCAT prioritized biological research on the reproduction,
growth, and structure of Atlantic bonito and two other economically
relevant small tuna species (i.e., little tunny tuna and wahoo) (SCRS,
2019).
With no appropriate stock differentiation, assessment and manage-
ment ICCAT uses the BIL (Billsh areas) areas only as a starting point to
collect sheries catch and statistical data on what is called small tuna
species. Remarkably, there is incongruence between the results pre-
sented here and these statistical areas (see Table 1 and Fig. 1 for the
assignation of locations to the ICCATs statistical areas). The genetically
differentiated Mediterranean northeast Atlantic pool comprises two
different ICCAT statistical areas (Stocks): the Mediterranean BIL 95 and
the northeast Atlantic BIL 94B. Similarly, the other genetic pool also
comprises two ICCAT statistical areas, Morocco, and Senegal, in the BIL
94B area, whereas Cˆ
ote dIvoire is in BIL 97.
Recently, ICCAT, along with other tuna shery management com-
missions, committed to adopting an ecosystem approach to sheries
management in which spatial management units should be ecologically
meaningful (Todorovi´
c et al., 2019). In a recent paper, Todorovi´
c et al.
(2019) proposed up to seven pelagic ecoregions in the Atlantic Med-
iterranean area based on the existing knowledge of biogeographic
classications of the pelagic environment, together with the spatial
distribution and dynamics of the main shing eets species. Our results
also challenge the boundaries of these regions. For instance, the
continuous genetic pool between the Mediterranean and the northeast
Atlantic comprises two Todorovi´
cs ecoregions: Mediterranean and
Northern Atlantic Temperate. The other genetic pool is completely
included within the Atlantic tropical ecoregion. It should be noted,
however, that these ecoregions are formulated without considering data
from small tuna species. Thus, the results presented here should serve as
a starting point for shery managers to delimit the boundaries of man-
agement stocks.
In this study, we developed a panel of a few highly informative SNPs
that, with their genotyping, are potentially sufcient to infer the pop-
ulation structure detected by more than 8000 SNPs. This is a streamlined
application for inferring possible population structure and population
assignment of individuals using low-cost genotyping tools (such as
Fluiding, Massarray, TaqMan among others) that have been successfully
applied in other sh species (Catanese et al., 2016; King and Stevens,
2021). As mentioned before, at least two open questions remain
unresolved: the complexity of population structure within the Mediter-
ranean Sea (especially in the south), and the population genetics of the
intermixing zone between the two genetically differentiated pools (the
area between southern Morocco and Mauritania). In addition, this panel
of SNPs could help in inferring the population structure of other
non-sampled areas. For instance, using mitochondrial DNA and nuclear
markers, Vi˜
nas et al. (2010) described a genetically differentiated pop-
ulation in the West Atlantic. Thus, the application of this panel of
markers could help resolve all these questions efciently.
CRediT authorship contribution statement
Judith Oll´
e-Vilanova: Formal analysis, Investigation, Writing
original draft. Ghailen Hajjej: Resources. David Macias: Validation,
Writing original draft. Sama Saber: Validation, Writing original
draft. Pedro G. Lino: Validation, Writing original draft. Rub´
en
Mu˜
noz-Lechuga: Validation, Writing original draft. SidAhmed
Baibbat: Resources. Fambaye Ngom Sow: Resources. Nguessan
Constance Diaha: Resources. Rosa M. Araguas: Formal analysis,
Validation. Núria Sanz: Formal analysis, Validation. Jordi Vinas:
Conceptualization, Formal analysis, Funding acquisition, Project
administration, Supervision, Writing original draft, Writing review &
editing.
Declaration of competing interest
The authors declare the following nancial interests/personal re-
lationships which may be considered as potential competing interest.
Jordi Vinas reports nancial support and administrative support
were provided by International Commission for the Conservation of
Atlantic Tunas. If there are other authors, they declare that they have no
known competing nancial interests or personal relationships that could
have appeared to inuence the work reported in this paper.
Data availability
Data will be made available on request.
Acknowledgements
‘‘This work was carried out under the provision of the ICCAT Small
Tunas Year Program (SMTYP). The contents of this paper do not
necessarily reect the point of view of ICCAT, which has no re-
sponsibility over them, and in no ways anticipate the Commissions
future policy in this area.’’ We would like to thank all the members of
the ICCAT Small Tuna Species Group for their comments during the
meetings that really helped in the construction of this study.
This research was funded by ICCAT Small Tunas Year Program
(SMTYP) and partially by the European Union through the EU Grant
Agreement No. S12.819116Strengthening the scientic.
basis for decision-making in ICCAT.
Appendix A. Supplementary data
Supplementary data to this article can be found online at https://doi.
org/10.1016/j.marenvres.2024.106408.
References
Allaya, H., Faleh, A.B., Hattour, A., Trabelsi, M., Vi˜
nas, J., 2015. Disparate past
demographic histories of three small Scombridae (Actinopterygii) species in
Tunisian waters. Hydrobiologia 758 (1), 1930. https://doi.org/10.1007/s10750-
015-2261-x.
Andrews, K.R., Good, J.M., Miller, M.R., Luikart, G., Hohenlohe, P.A., 2016. Harnessing
the power of RADseq for ecological and evolutionary genomics. Nat. Rev. Genet. 17
(2), 8192. https://doi.org/10.1038/nrg.2015.28.
J. Oll´
e-Vilanova et al.
Marine Environmental Research xxx (xxxx) xxx
8
Arkhipov, A.G., Pak, R.A., 2019. Number dynamics of mass ichthyoplankton species in
the waters of Morocco. J. Ichthyol. 59 (3), 344351. https://doi.org/10.1134/
S0032945219020012.
Baibbat, S.a., Malouli, I., Abid, N., Kell, L., Lucena, F., Benazzouz, B., 2017. Reproduction
of atlantic bonito (Sarda sarda) in south of the Moroccan atlantic waters. ICCAT
Collective Volume of Scientic Papers 74 (1), 95107.
Catanese, G., Montes, I., Iriondo, M., Estonba, A., Iudicone, D., Procaccini, G., 2016. High
resolution SNPs selection in Engraulis encrasicolus through Taqman OpenArray.
Fish. Res. 177, 3138. https://doi.org/10.1016/j.shres.2016.01.014.
Catanese, G., Watteaux, R., Montes, I., Barra, M., Rumolo, P., Borme, D., Procaccini, G.,
2017. Insights on the drivers of genetic divergence in the European anchovy. Sci.
Rep. 7 (1), 4180. https://doi.org/10.1038/s41598-017-03926-z.
Collette, B.B., Nauen, D.A., 1983. FAO species catalogue, Vol. 2. Scombrids of the world:
an annotated and illustrated catalogue of tunas, mackerels, bonitos, and related
species known to date. FAO Fisheries Synopsis 125, 1137.
Cowen, R.K., Gawarkiewicz, G., Pineda, J., Thorrold, S.R., Werner, F.E., 2007.
Population connectivity in marine systems: an overview. Oceanography 20 (3),
1421.
Davies, C.A., Gosling, E.M., Was, A., Brophy, D., Tysklind, N., 2011. Microsatellite
analysis of albacore tuna (Thunnus alalunga): population genetic structure in the
North-East Atlantic Ocean and Mediterranean Sea. Mar. Biol. 158 (12), 27272740.
https://doi.org/10.1007/s00227-011-1772-x.
de Jong, M.J., de Jong, J.F., Hoelzel, A.R., Janke, A., 2021. SambaR: an R package for
fast, easy and reproducible population-genetic analyses of biallelic SNP data sets.
Molecular Ecology Resources 21 (4), 13691379. https://doi.org/10.1111/1755-
0998.13339.
Do, C., Waples, R.S., Peel, D., Macbeth, G.M., Tillett, B.J., Ovenden, J.R., 2014.
NeEstimator v2: re-implementation of software for the estimation of contemporary
effective population size (Ne) from genetic data. Molecular Ecology Resources 14
(1), 209214. https://doi.org/10.1111/1755-0998.12157.
Dray, S., Dufour, A.-B., 2007. The ade4 package: implementing the duality diagram for
ecologists. J. Stat. Software 22 (4), 120. https://doi.org/10.18637/jss.v022.i04.
Excofer, L., Smouse, P.E., Quattro, J.M., 1992. Analysis of molecular variance inferred
from metric distances among DNA haplotypes: application to human mitochondrial
DNA restriction data. Genetics 131 (2), 479491.
Forde, S., von der Heyden, S., Le Moan, A., Nielsen, E.S., Durholtz, D., Kainge, P.,
Henriques, R., 2023. Management and conservation implications of cryptic
population substructure for two commercially exploited shes (Merluccius spp.) in
southern Africa. Molecular Ecology Resources. https://doi.org/10.1111/1755-
0998.13820.
Fr´
edou, F.L., Hazin, F., Vi˜
nas, J., Oll´
e, J., Hajjej, G., Macias, D., Massa-Gallucci, A., 2021.
Final report of the short-term contract for iccat smtyp for the biological samples
collection for growth, maturity and genetics studies year #3. ICCAT Collective
Volume of Scientic Papers 78 (8), 103131.
Fuentes-Pardo, A.P., Farrell, E.D., Pettersson, M.E., Sprehn, C.G., Andersson, L., 2023.
The genomic basis and environmental correlates of local adaptation in the Atlantic
horse mackerel (Trachurus trachurus). Evol Appl 16 (6), 12011219. https://doi.
org/10.1111/eva.13559.
Goudet, J., 2005. hierfstat, a package for r to compute and test hierarchical F-statistics.
Mol. Ecol. Notes 5 (1), 184186. https://doi.org/10.1111/j.1471-8286.2004.00828.
x.
Hauser, L., Carvalho, G.R., 2008. Paradigm shifts in marine sheries genetics: ugly
hypotheses slain by beautiful facts. Fish Fish. 9 (4), 333362. https://doi.org/
10.1111/j.1467-2979.2008.00299.x.
Hohenlohe, P.A., Funk, W.C., Rajora, O.P., 2021. Population genomics for wildlife
conservation and management. Mol. Ecol. 30 (1), 6282. https://doi.org/10.1111/
mec.15720.
Jombart, T., 2008. adegenet: a R package for the multivariate analysis of genetic
markers. Bioinformatics 24 (11), 14031405. https://doi.org/10.1093/
bioinformatics/btn129.
Jombart, T., Ahmed, I., 2011. Adegenet 1.3-1: new tools for the analysis of genome-wide
SNP data. Bioinformatics 27 (21), 30703071. https://doi.org/10.1093/
bioinformatics/btr521.
Jombart, T., Devillard, S., Balloux, F., 2010. Discriminant analysis of principal
components: a new method for the analysis of genetically structured populations.
BMC Genet. 11 (1), 94. https://doi.org/10.1186/1471-2156-11-94.
Juan-Jord´
a, M.J., Andonegi, E., Murua, H., Ruiz, J., Lourdes-Ramos, M.L., Sabarros, P.H.
S., Bach, P., 2020. In support of the ICCAT ecosystem report card: three ecosystem
indicators to monitor the ecological impacts of purse seine sheries in the tropical
Atlantic Ecoregion. ICCAT Collective Volumes of Scientic Papers 76 (9), 130143.
Kahraman, A.E., G¨
oktürk, D., Yildiz, T., Uzer, U., 2014. Age, growth, and reproductive
biology of atlantic bonito (Sarda sarda bloch, 1793) from the Turkish coasts of the
black sea and the sea of marmara. Turk. J. Zool. 38, 614621. https://doi.org/
10.3906/zoo-1311-25.
Kamvar, Z.N., Tabima, J.F., Grunwald, N.J., 2014. Poppr: an R package for genetic
analysis of populations with clonal, partially clonal, and/or sexual reproduction.
PeerJ 2, e281. https://doi.org/10.7717/peerj.281.
Keenan, K., McGinnity, P., Cross, T.F., Crozier, W.W., Prod¨
ohl, P.A., 2013. diveRsity: an
R package for the estimation and exploration of population genetics parameters and
their associated errors. Methods Ecol. Evol. 4 (8), 782788. https://doi.org/
10.1111/2041-210X.12067.
King, R.A., Stevens, J.R., 2021. Development of SNP markers derived from RAD
sequencing for Atlantic salmon (Salmo salar L.) inhabiting the rivers of southern
England. Conservation Genetics Resources 13 (4), 369373. https://doi.org/
10.1007/s12686-021-01215-6.
Knaus, B.J., Grünwald, N.J., 2017. vcfr: a package to manipulate and visualize variant
call format data in R. Molecular Ecology Resources 17 (1), 4453. https://doi.org/
10.1111/1755-0998.12549.
Koched, W., Alemany, F., Rimel, B., Hattour, A., 2016. Characterization of the spawning
area of tuna species on the northern Tunisian coasts. Sci. Mar. 80 (2), 187198.
https://doi.org/10.3989/scimar.04332.27A.
Li, H., Durbin, R., 2009. Fast and accurate short read alignment with Burrows-Wheeler
transform. Bioinformatics 25 (14), 17541760. https://doi.org/10.1093/
bioinformatics/btp324.
Lucena-Fr´
edou, F., Mourato, B., Fr´
edou, T., Lino, P.G., Mu˜
noz-Lechuga, R., Palma, C.,
Pons, M., 2021. Review of the life history, sheries, and stock assessment for small
tunas in the Atlantic Ocean. Rev. Fish Biol. Fish. 31 (3), 709736. https://doi.org/
10.1007/s11160-021-09666-8.
Luu, K., Bazin, E., Blum, M.G.B., 2017. pcadapt: an R package to perform genome scans
for selection based on principal component analysis. Molecular Ecology Resources
17 (1), 6777. https://doi.org/10.1111/1755-0998.12592.
Macías, D., G´
omez-Vives, M.J., García, S., Ortiz de Urbina, J., 2005. Reproductive
characteristics of Atlantic bonito (Sarda sarda) from the south western Spanish
mediterranean. ICCAT Collective Volume of Scientic Papers 58 (2), 470483.
Marandel, F., Charrier, G., Lamy, J.B., Le Cam, S., Lorance, P., Trenkel, V.M., 2020.
Estimating effective population size using RADseq: effects of SNP selection and
sample size. Ecol. Evol. 10 (4), 19291937. https://doi.org/10.1002/ece3.6016.
Marandel, F., Lorance, P., Berthel´
e, O., Trenkel, V.M., Waples, R.S., Lamy, J.-B., 2019.
Estimating effective population size of large marine populations, is it feasible? Fish
Fish. 20 (1), 189198. https://doi.org/10.1111/faf.12338.
Montes, I., Zarraonaindia, I., Iriondo, M., Grant, W.S., Manzano, C., Cotano, U.,
Estonba, A., 2016. Transcriptome analysis deciphers evolutionary mechanisms
underlying genetic differentiation between coastal and offshore anchovy populations
in the Bay of Biscay. Mar. Biol. 163 (10) https://doi.org/10.1007/s00227-016-2979-
7.
Narum, S.R., Buerkle, C.A., Davey, J.W., Miller, M.R., Hohenlohe, P.A., 2013.
Genotyping-by-sequencing in ecological and conservation genomics. Mol. Ecol. 22
(11), 28412847. https://doi.org/10.1111/mec.12350.
Nielsen, E.S., Hanson, J.O., Carvalho, S.B., Beger, M., Henriques, R., Kershaw, F., von der
Heyden, S., 2023. Molecular ecology meets systematic conservation planning.
Trends Ecol. Evol. 38 (2), 143155. https://doi.org/10.1016/j.tree.2022.09.006.
Oll´
e, J., Macías, D., Saber, S., Jos´
e G´
omez-Vives, M., P´
erez-Bielsa, N., Vi˜
nas, J., 2019.
Genetic analysis reveals the presence of frigate tuna (Auxis thazard) in the bullet tuna
(Auxis rochei) shery of the Iberian Peninsula. Bull. Mar. Sci. 95 (2), 317325.
https://doi.org/10.5343/bms.2018.0049.
Oll´
e, J., Vil`
a-Valls, L., Alvarado-Bremer, J., Cerdenares, G., Duong, T.Y., Hajjej, G.,
Vi˜
nas, J., 2022. Population genetics meets phylogenetics: new insights into the
relationships among members of the genus Euthynnus (family Scombridae).
Hydrobiologia 849 (1), 4762. https://doi.org/10.1007/s10750-021-04707-6.
Oll´
e-Vilanova, J., P´
erez-Bielsa, N., Araguas, R.M., Sanz, N., Saber, S., Macías, D.,
Vi˜
nas, J., 2022. Larval retention and homing behaviour shape the genetic structure
of the Bullet tuna (Auxis rochei) in the Mediterranean Sea. Fishes 7 (5). https://doi.
org/10.3390/shes7050300.
Pascual, M., Rives, B., Schunter, C., Macpherson, E., 2017. Impact of life history traits on
gene ow: a multispecies systematic review across oceanographic barriers in the
Mediterranean Sea. PLoS One 12 (5), e0176419. https://doi.org/10.1371/journal.
pone.0176419.
Patarnello, T., Volckaert, F.A.M.J., Castilho, R., 2007. Pillars of Hercules: is the Atlantic-
Mediterranean transition a phylogeographical break? Mol. Ecol. 16 (21),
44264444.
Petukhova, N.G., 2020. Preliminary assessment of the stock status of atlantic bonito
(Sarda sarda) in the northeastern part of the Atlantic Ocean. J. Ichthyol. 60 (5),
732741. https://doi.org/10.1134/S0032945220050069.
Priv´
e, F., Luu, K., Vilhj´
almsson, B.J., Blum, M.G.B., 2020. Performing highly efcient
genome scans for local adaptation with R package pcadapt version 4. Mol. Biol. Evol.
37 (7), 21532154. https://doi.org/10.1093/molbev/msaa053.
Reiss, H., Hoarau, G., Dickey-Collas, M., Wolff, W.J., 2009. Genetic population structure
of marine sh: mismatch between biological and sheries management units. Fish
Fish. 10 (4), 361395. https://doi.org/10.1111/j.1467-2979.2008.00324.x.
Rey, J.C., Alot, E., Ramos, A., 1984. Sinopsis biol´
ogica del bonito (Sarda sarda) del
Mediterr´
aneo y Atl´
antico este. ICCAT Collective Volume of Scientic Papers 20,
469502.
Rey, J.C., Cort, J.L., 1981. Migraci´
on de bonitos (Sarda sarda) y bacoreta (Euthynnus
alletteratus) entre el Mediterr´
aneo y el Atl´
antico. ICCAT Collective Volume of
Scientic Papers 15 (2), 346347.
Rochette, N.C., Rivera-Col´
on, A.G., Catchen, J.M., 2019. Stacks 2: analytical methods for
paired-end sequencing improve RADseq-based population genomics. Mol. Ecol. 28
(21), 47374754. https://doi.org/10.1111/mec.15253.
Rodríguez-Ezpeleta, N., Díaz-Arce, N., Walter III, J.F., Richardson, D.E., Rooker, J.R.,
Nøttestad, L., Arrizabalaga, H., 2019. Determining natal origin for improved
management of Atlantic bluen tuna. Front. Ecol. Environ. 17 (8), 439444. https://
doi.org/10.1002/fee.2090.
Rodríguez-Roda, J., Di Centa, A., 1981. ´
Area de puesta del atún, melva y bonito en las
costas de Espa˜
na y Marruecos. ICCAT Collective Volume of Scientic Papers 15 (2),
278283.
Sabat´
es, A., Recasens, L., 2001. Seasonal distribution and spawning of small tunas (Auxis
rochei and Sarda sarda) in the northwestern Mediterranean. Sci. Mar. 65 (2), 95100.
SCRS, 2019. Report of the 2019 ICCAT small tunas species group intersessional meeting.
ICCAT Collective volume of scientic papers 76 (7), 189.
Smith, B.L., Lu, C.-P., Garcia-Cortes, B., Vi˜
nas, J., Yeh, S.-Y., Bremer, J.R.A., 2015.
Multilocus bayesian estimates of intra-oceanic genetic differentiation, connectivity,
J. Oll´
e-Vilanova et al.
Marine Environmental Research xxx (xxxx) xxx
9
and admixture in Atlantic Swordsh (Xiphias gladius L.). PLoS One 10 (6). https://
doi.org/10.1371/journal.pone.0127979.
Todorovi´
c, S., Juan-Jord´
a, M.J., Arrizabalaga, H., Murua, H., 2019. Pelagic ecoregions:
operationalizing an ecosystem approach to sheries management in the Atlantic
Ocean. Mar. Pol. 109, 103700 https://doi.org/10.1016/j.marpol.2019.103700.
Turan, C., 2015. Microsatellite DNA reveals genetically different populations of Atlantic
bonito Sarda sarda in the Mediterranean Basin. Biochem. Systemat. Ecol. 63,
174182. https://doi.org/10.1016/j.bse.2015.10.007.
Van Wyngaarden, M., Snelgrove, P.V., DiBacco, C., Hamilton, L.C., Rodriguez-
Ezpeleta, N., Jeffery, N.W., Bradbury, I.R., 2017. Identifying patterns of dispersal,
connectivity and selection in the sea scallop, Placopecten magellanicus, using RADseq-
derived SNPs. Evolutionary Applications 10 (1), 102117. https://doi.org/10.1111/
eva.12432.
Vi˜
nas, J., Alvarado Bremer, J.R., Pla, C., 2004. Phylogeography of the Atlantic bonito
(Sarda sarda) in the northern Mediterranean: the combined effects of historical
vicariance, population expansion, secondary invasion, and isolation by distance.
Mol. Phylogenet. Evol. 33 (1), 3242.
Vi˜
nas, J., Alvarado Bremer, J.R., Pla, C., 2010. Phylogeography and phylogeny of the
epineritic cosmopolitan bonitos of the genus Sarda (Cuvier): inferred patterns of
intra- and inter-oceanic connectivity derived from nuclear and mitochondrial DNA
data. J. Biogeogr. 37 (3), 557570. https://doi.org/10.1111/j.1365-
2699.2009.02225.x.
Waples, R.K., Larson, W.A., Waples, R.S., 2016. Estimating contemporary effective
population size in non-model species using linkage disequilibrium across thousands
of loci. Heredity 117 (4), 233240. https://doi.org/10.1038/hdy.2016.60.
Vi˜
nas, J., Oll´
e Vilanova, J., Hajjej, G., Macías, D., Saber, S., Lino, P.G., Lucena Fr´
edou, F.,
2020. Population genetic of Atlantic bonito in the North east Atlantic and
Mediterranean Sea. ICCAT Collect. Vol. Sci. Pap. 77 (9), 612.
Waples, R.S., 1998. Separating the wheat from the chaff: patterns of genetic
differentiation in high gene ow species. J. Hered. 89 (5), 438450.
Waples, R.S., 2016. Tiny estimates of the Ne/N ratio in marine shes: are they real?
J. Fish. Biol. 89 (6), 24792504. https://doi.org/10.1111/jfb.13143.
Waples, R.S., Grewe, P.M., Bravington, M.W., Hillary, R., Feutry, P., 2018. Robust
estimates of a high N(e)/N ratio in a top marine predator, Southern bluen tuna. Sci.
Adv. 4 (7), eaar7759 https://doi.org/10.1126/sciadv.aar7759.
Waples, R.S., Punt, A.E., Cope, J.M., 2008. Integrating genetic data into management of
marine resources: how can we do it better? Fish Fish. 9 (4), 423449. https://doi.
org/10.1111/j.1467-2979.2008.00303.x.
Whitlock, M.C., Lotterhos, K.E., 2015. Reliable detection of loci responsible for local
adaptation: inference of a null model through trimming the distribution of FST. Am.
Nat. 186 (S1), S24S36. https://doi.org/10.1086/682949.
J. Oll´
e-Vilanova et al.
Article
Full-text available
Atlantic bonito meat has economic potential as an alternative to mackerel consumption. Thus, considering the presence of myoglobin (Mb) in red fish muscles, we report the characterisation of Atlantic bonito Mb compared to Atlantic and Tinker mackerel Mbs since this haemoprotein is implicated in lipid oxidation and fish meat preservation. A plethora of biochemical approaches were employed to purified Mb from Atlantic bonito and determine the autoxidation rate constant (0.189 ± 0.009 h−1), melting temperature (Tm = 72.84 ± 1.02 °C) and pseudoperoxidase activity in different conditions (pH and several cations). Atlantic and Tinker mackerel Mbs showed a lower Tm (∼66.85 °C), while oxyMb autoxidation rate constant was higher for Atlantic mackerel (∼1.08-fold) and lower for Tinker mackerel (∼1.35-fold) compared to Atlantic bonito. This Mb had a Michaelis-Menten constant (Km) of 38.63 ± 1.89 μM, ∼2.49-fold and 2.27-fold lower than Atlantic and Tinker mackerel Mbs, respectively. Atlantic bonito Mb primary structure has 146 amino acid residues with the N-terminal acetylated and 25 amino acid substitutions with respect to Atlantic and Tinker mackerel Mbs. In silico analysis revealed that 7 out of 25 substitutions are close to the haem-pocket, while 18 out of 25 are far from this region. All substitutions, except H20, L70 and L81 are exposed on the protein globular surface. Overall, the results of this research provide new information for future studies that will be useful to the fish industry for preservation of frozen or canned Atlantic bonito meat considering the presence of Mb as a reactive haemoprotein.
Article
Full-text available
Understanding how populations adapt to their environment is increasingly important to prevent biodiversity loss due to overexploitation and climate change. Here we studied the population structure and genetic basis of local adaptation of Atlantic horse mackerel, a commercially and ecologically important marine fish that has one of the widest distributions in the eastern Atlantic. We analyzed whole‐genome sequencing and environmental data of samples collected from the North Sea to North Africa and the western Mediterranean Sea. Our genomic approach indicated low population structure with a major split between the Mediterranean Sea and the Atlantic Ocean and between locations north and south of mid‐Portugal. Populations from the North Sea are the most genetically distinct in the Atlantic. We discovered that most population structure patterns are driven by a few highly differentiated putatively adaptive loci. Seven loci discriminate the North Sea, two the Mediterranean Sea, and a large putative inversion (9.9 Mb) on chromosome 21 underlines the north–south divide and distinguishes North Africa. A genome–environment association analysis indicates that mean seawater temperature and temperature range, or factors correlated to them, are likely the main environmental drivers of local adaptation. Our genomic data broadly support the current stock divisions, but highlight areas of potential mixing, which require further investigation. Moreover, we demonstrate that as few as 17 highly informative SNPs can genetically discriminate the North Sea and North African samples from neighboring populations. Our study highlights the importance of both, life history and climate‐related selective pressures in shaping population structure patterns in marine fish. It also supports that chromosomal rearrangements play a key role in local adaptation with gene flow. This study provides the basis for more accurate delineation of the horse mackerel stocks and paves the way for improving stock assessments.
Article
Full-text available
Background: The bullet tuna (Auxis rochei) is an epipelagic fish with a worldwide distribution that is highly targeted by fisheries. Genetic diversity and population genetics are good indicators of population structure and thus, essential tools for fisheries management. Knowing which factors (biotic and abiotic) might be shaping such structure is crucial for management plans. In the present study, we assessed the population structure of the bullet tuna in the western and central Mediterranean Sea. Methods: We used two types of genetic data: the mitochondrial DNA control region and seven microsatellite loci. The analysis of 431 sequences with a length of 386 bp from the mtDNA CR and the results from 276 individuals were genotyped by seven microsatellite loci. Results: Both types of markers coincided in showing significant genetic differences between an Iberian Peninsula–Strait of Gibraltar stock in comparison with a North African stock. Conclusions: We argue that this differentiation pattern is likely caused by reproductive strategies such as coastal spawning, larval retention, and natal homing behavior. These results should endorse the implementation of management plans for a resource that currently is not being managed. Thus, to ensure sustainability, these new policies should consider the presence of at least two genetically identified stocks.
Article
Full-text available
Integrative and proactive conservation approaches are critical to the long-term persistence of biodiversity. Molecular data can provide important information on evolutionary processes necessary for conserving multiple levels of biodiversity (genes, populations, species, and ecosystems). However, molecular data are rarely used to guide spatial conservation decision-making. Here, we bridge the fields of molecular ecology (ME) and systematic conservation planning (SCP) (the ‘why’) to build a foundation for the inclusion of molecular data into spatial conservation planning tools (the ‘how’), and provide a practical guide for implementing this integrative approach for both conservation planners and molecular ecologists. The proposed framework enhances interdisciplinary capacity, which is crucial to achieving the ambitious global conservation goals envisioned for the next decade.
Article
Full-text available
Euthynnus (family Scombridae) is a genus of marine pelagic fish species with a worldwide distribution that comprises three allopatric species: E. alletteratus, E. affinis and E. lineatus. All of them targeted by artisanal and commercial fisheries. We analyzed 263 individuals from Atlantic and Pacific Oceans using two genetic markers, the mtDNA Control Region (350 bp) and nuclear calmodulin (341 bp). The results obtained challenge the phylogeny of this group. We found a deep genetic divergence, probably at species level, within E. alletteratus, between the North Atlantic-Mediterranean and the Tropical East Atlantic. This deep genetic divergence was tested with several species delimitation methods. This complete phylogeographic association between the North Atlantic and the Tropical East Atlantic support the hypothesis of two cryptic species. In addition, population genetic heterogeneity was detected between the North East Atlantic–Mediterranean and North West Atlantic regions. Our results indicate two scales of differentiation in what is currently considered a single population. Accordingly, for management purposes, the populations of E. alletteratus, should be divided into a minimum of three management units. On the other hand, the high level of differentiation found in E. alletteratus contrasts with the shallow genetic divergence of E. affinis and E. lineatus.
Article
Full-text available
Despite being an important source of wealth and food security for many countries, most of the small tuna stocks in the Atlantic Ocean and Mediterranean Sea remain unassessed. In this study, we summarized the current state of knowledge of this group of species reviewing the information available on life history parameters, stock structure, historical catches, size frequency distributions and current knowledge of stock status. In relation to the life history parameters, data are overall scarce and mainly missing in the Eastern Atlantic where small tunas are relevant in small-scale fisheries. From the 27 defined stocks, only 11 have been quantitatively assessed. From those, the Northwest wahoo and the Southeast little tunny stocks may be experiencing overfishing, deserving priority management attention. Length-based rather than catch-based methods showed a more promising applicability for small tunas, although representative length distributions from the catch are scarce for some stocks. Historical catch time series for small tuna are still incomplete, however, the last two decades are the most accurate and could be considered in future assessment methods. The gaps of knowledge related mainly to life history parameters and historical catches are the main reason why most of these stocks remain unassessed and unmanaged.
Article
Full-text available
The rivers of the Hampshire Basin, southern England contain a genetically unique group of Atlantic salmon that have suffered dramatic declines in numbers over the last 40 years. Knowledge of levels and patterns of genetic diversity is essential for effective management of these vulnerable populations. Using restriction site-associated DNA sequencing (RADseq) data, we describe the development and characterisation of a panel of 94 single nucleotide polymorphism (SNP) loci for salmon from this region and investigate their applicability and variability in both target (i.e. southern English) and non-target populations. The SNP loci will be useful for population genetic and assignment studies on Atlantic salmon within the UK and beyond.
Article
Full-text available
Biodiversity is under threat worldwide. Over the past decade, the field of population genomics has developed across nonmodel organisms, and the results of this research have begun to be applied in conservation and management of wildlife species. Genomics tools can provide precise estimates of basic features of wildlife populations, such as effective population size, inbreeding, demographic history and population structure, that are critical for conservation efforts. Moreover, population genomics studies can identify particular genetic loci and variants responsible for inbreeding depression or adaptation to changing environments, allowing for conservation efforts to estimate the capacity of populations to evolve and adapt in response to environmental change and to manage for adaptive variation. While connections from basic research to applied wildlife conservation have been slow to develop, these connections are increasingly strengthening. Here we review the primary areas in which population genomics approaches can be applied to wildlife conservation and management, highlight examples of how they have been used, and provide recommendations for building on the progress that has been made in this field.
Article
Full-text available
Based on the Russian fishery-biological data, an attempt was made to assess the stock status of Atlantic bonito Sarda sarda in the northeastern part of the Atlantic Ocean using the length-based spawning potential ratio (LBSPR) method. Values of the parameters of the Bertalanffy equation are calculated for Atlantic bonito: the theoretical maximum length of an individual is 75.6 cm and growth coefficient is 0.41. The values of the length where 50 and 95% of the fish are mature are 44.7 and 57.0 cm, respectively. The resulting estimate of the spawning potential ratio (0.28) is lower than the biological target reference point (0.40) and formally indicates an overfishing status of the stock.
Article
Genomic information can aid in the establishment of sustainable management plans for commercially exploited marine fishes, aiding in the long-term conservation of these resources. The southern African hakes (Merluccius capensis and M. paradoxus) are commercially valuable demersal fishes with similar distribution ranges but exhibiting contrasting life histories. Using a comparative framework based on Pool-Seq genome-wide SNP data, we investigated whether the evolutionary processes that shaped extant patterns of diversity and divergence are shared among these two congeneric fishes, or unique to each one. Our findings revealed that M. capensis and M. paradoxus show similar levels of genome-wide diversity, despite different census sizes and life-history features. In addition, M. capensis shows three highly structured geographic populations across the Benguela Current region (one in the northern Benguela and two in the southern Benguela), with no consistent genome-environment associations detected. In contrast, although population structure and outlier analyses suggested panmixia for M. paradoxus, reconstruction of its demographic history suggested the presence of an Atlantic-Indian Ocean subtle substructuring pattern. Therefore, it appears that M. paradoxus might be composed by two highly connected populations, one in the Atlantic and one in the southwest Indian Ocean. The reported similar low levels of genomic diversity, as well as newly discovered genetically distinct populations in both hake species can thus assist in informing and improving conservation and management plans for the commercially important southern African Merluccius.
Article
SNP data sets can be used to infer a wealth of information about natural populations, including information about their structure, genetic diversity, and the presence of loci under selection. However, SNP data analysis can be a time-consuming and challenging process, not in the least because at present many different software packages are needed to execute and depict the wide variety of mainstream population-genetic analyses. Here, we present SambaR, an integrative and user-friendly R package which automates and simplifies quality control and population-genetic analyses of biallelic SNP data sets. SambaR allows users to perform mainstream population-genetic analyses and to generate a wide variety of ready to publish graphs with a minimum number of commands (less than 10). These wrapper commands call functions of existing packages (including adegenet, ape, LEA, poppr, pcadapt and StAMPP) as well as new tools uniquely implemented in SambaR. We tested SambaR on online available SNP data sets and found that SambaR can process data sets of over 100,000 SNPs and hundreds of individuals within hours, given sufficient computing power. Newly developed tools implemented in SambaR facilitate optimization of filter settings, objective interpretation of ordination analyses, enhance comparability of diversity estimates from reduced representation library SNP data sets, and generate reduced SNP panels and structure-like plots with Bayesian population assignment probabilities. SambaR facilitates rapid population genetic analyses on biallelic SNP data sets by removing three major time sinks: file handling, software learning, and data plotting. In addition, SambaR provides a convenient platform for SNP data storage and management, as well as several new utilities, including guidance in setting appropriate data filters. The SambaR source script, manual and example data set are distributed through GitHub: https://github.com/mennodejong1986/SambaR.