Você está na página 1de 6

Fisheries Research 105 (2010) 172–177

Contents lists available at ScienceDirect

Fisheries Research
journal homepage: www.elsevier.com/locate/fishres

Population genetic structure of dolphinfish (Coryphaena hippurus) in the Gulf of


California, using microsatellite loci
Miguel A. Tripp-Valdez a , Francisco J. García de León a , Sofía Ortega-García b , Daniel Lluch-Cota a ,
Juana López-Martínez a , Pedro Cruz a,∗
a
Centro de Investigaciones Biológicas de Noroeste (CIBNOR), Mar Bermejo 195, Col. Playa Palo de Santa Rita, La Paz, B.C.S. 23096, Mexico
b
Centro Interdisciplinario de Ciencias Marinas, Instituto Politécnico Nacional (CICIMAR), N Av. Instituto Politécnico Nacional s/n, Col. Playa Palo de Santa Rita,
La Paz, B.C.S. 23096, Mexico

a r t i c l e i n f o a b s t r a c t

Article history: We assayed genetic variations at five microsatellite loci of dolphinfish captured at five sites in 2005
Received 15 January 2010 and eight sites in 2006 to detect genetic population structure of dolphinfish Coryphaena hippurus in the
Received in revised form 29 March 2010 Gulf of California and surrounding waters. Results show high genetic variation, similar to other pelagic
Accepted 30 March 2010
fishes with large populations. Pairwise FST values and hierarchical AMOVA detected subtle but significant
spatial and temporal heterogeneity, mainly in samples of 2005 and some of 2006. However, Bayesian
Keywords:
assignment analysis failed to detect genetic differentiation, which was also supported by the Mantel test
Dolphinfish
and gene flow estimates among the sampled sites. This suggests that, despite the slight heterogeneity
Microsatellite
Population structure
detected in this region, the dolphinfish forms a single panmictic population with high genetic variation
Gulf of California and gene flow.
© 2010 Elsevier B.V. All rights reserved.

1. Introduction differentiation between populations (Graves, 1998). However,


genetic heterogeneity has been found in some species after ana-
The dolphinfish Coryphaena hippurus, also known as dorado lyzing polymorphic microsatellite markers over an extensive area,
or mahi–mahi, is a migratory pelagic fish inhabiting tropical and as in the case of striped marlin (McDowell and Graves, 2008) and
subtropical regions of all oceans (Palko et al., 1982). It is found yellowfin tuna in the Pacific Ocean (Appleyard et al., 2001) and in
from the northern Pacific coast of the Baja California Peninsula a more restricted area, as the case of bluefin tuna in the Mediter-
to the Gulf of Tehuantepec and inside the Gulf of California. This ranean (Carlsson et al., 2004) and Atlantic herring in Norway (Shaw
species grows quickly, reach sexual maturity at 50 cm fork length et al., 1999). This suggests that some biological or oceanographic
and has the ability to spawn several times a year (Beardsley, factors limit genetic flow.
1967). In the Pacific area of Mexico, molecular markers were used for
By Mexican law, this species is reserved exclusively for the recre- observing population structure of dolphinfish. Rocha-Olivares et
ational fishing industry within 50 nautical miles of the coast. Given al. (2006) found genetic heterogeneity in samples from Hawaii,
high market demand, this species is also captured by artisanal fish- Sinaloa, and Baja California Sur using RFLPs of the mitochondrial
ermen in many parts of Mexico, especially in the northwest. This gene NADH1. Diaz-Jaimes et al. (2006) failed to detect any seasonal
has led authorities to consider modifying fisheries laws so dolphin- and spatial genetic differences in samples of coastal Baja Califor-
fish can be exploited by commercial fishing cooperatives. Whether nia Sur, Sinaloa, Sonora and Chiapas using a 750 bp fragment of
dolphinfish is a single population or segregates into local popula- this gene. Consequently, it is not certain whether exploitation of
tions on the west coast of Mexico will be useful for management dolphinfish in one area affects the whole population. If discrete
plans of this resource. Studies of genetic diversity of marine popula- populations are present, this knowledge is useful to establish man-
tions are important tools for determining whether local stocks are agement plans for each population.
distinct or determining the degree of flow of individuals between We analysed five microsatellite loci of dolphinfish captured in
stocks (Ward, 2000). 2005 and 2006 from the Gulf of California and the Pacific coast
Dolphinfish, like other pelagic fishes, have high mobility, large of Mexico. Using microsatellite loci provided better resolution for
population size and high reproductive capacity that limit genetic detecting population structure in moderately large geographical
areas, even in species with capacity for high dispersion (Waples,
1998). The main objective was to determine if dolphinfish in the
∗ Corresponding author. Tel.: +52 612 123 8484x3345; fax: +52 612 125 3625. Pacific coastal waters of Mexico form relatively discrete popula-
E-mail address: pcruz@cibnor.mx (P. Cruz). tions or a single panmictic population.

0165-7836/$ – see front matter © 2010 Elsevier B.V. All rights reserved.
doi:10.1016/j.fishres.2010.03.023
M.A. Tripp-Valdez et al. / Fisheries Research 105 (2010) 172–177 173

2.2. DNA extraction and amplification

Genomic DNA was extracted using DNAeasy Tissue Kit (Qiagen).


Five microsatellite loci registered in the GenBank (Robert Chap-
man; South Carolina Department of Natural Resources, Charleston
SC., pers. comm.) were amplified (Chi002, Chi008, Chi008a,
Chi023 and Chi032; Table 1). Amplification of the loci DNA took
place in 12.5 ␮l reactions with the following concentrations: 1×
buffer, 3 mM MgCl2 , 0.2 mM dNTP, 0.003 mM of each primer and
0.02 U ␮l−1 Taq polymerase. We used the same PCR conditions as
in Yoshida et al. (2005) to reduce stutter bands. This consisted of
an initial denaturing period of 2 min at 94 ◦ C, five cycles at 94 ◦ C for
5 s, annealing and extension for 60 s at 59 ◦ C, 20 cycles at 94 ◦ C for
0 s, annealing and extension for 60 s at 59 ◦ C, and a final extension
for 60 min at 59 ◦ C. Amplified fragments were run in denaturizing
6% polyacrylamide electrophoresis at 2500 V, 100 mA and 90 W for
∼2.5 h. Alleles were observed using SYBR Gold nucleic acid gel stain
(Rodzen et al., 1998) and a scanner (FMBIO II Multi-viewer, Hitachi
Software Engineering). Alleles were sized using a sequencing lad-
der (Sequitherm Cycle Sequencing kit–Epicentre Tech) using the
PUC 19 plasmid and M13 primer.

2.3. Statistical analyses

Levels of genetic variability, expressed in number of alleles (NA ),


observed heterozygosity (HO ) and expected heterozygosity (HE )
Fig. 1. Sampling sites of the dolphinfish Coryphaena hippurus. Sites with a triangle,
were obtained for each locus from every sample using the soft-
collected only in 2005; sites with a circle collected only in 2006; sites with a square, ware GenAlEx 6.1 (Peakall and Smouse, 2006). Deviations from the
collected in both years, n is sample size: a samples collected in 2005 and b samples Hardy–Weinberg equilibrium were examined for each population
collected in 2006. at each locus by calculating the fixation index FIS with software
Genepop 4.0 (Raymond and Rousset, 1995) with 10,000 dememo-
2. Materials and methods rizations and 1000 iterations. The null hypothesis of independence
between loci (no linkage disequilibrium) was tested using the
2.1. Tissue collection same software. Levels of genetic variation were compared among
sampled sites using the Kruskal–Wallis ANOVA, which is a nonpara-
Muscle tissue (∼2 g) was collected from 731 dolphinfish over metric alternative to one-way ANOVA, using software STATISTICA
2 years (2005 and 2006) in the Gulf of California and three out- v. 7 (StatSoft, 2004).
side sites (see Fig. 1). Three sites were sampled in 2005 and 2006 Homogeneity of microsatellite allele frequencies between all
(Cabo San Lucas, Loreto and Guaymas), two sites were sampled pairs of samples in both years was assessed using Fisher’s exact
only in 2005 (Puerto Libertad, Topolobampo), and five sites were test with software Genepop 4.0 (Raymond and Rousset, 1995)
sampled only in 2006 (Mazatlan, Nayarit, La Paz, Punta Lobos and with 10,000 dememorizations, 500 batches and 5000 iterations
an open ocean sample incidentally collected from a tuna ves- per batch. Sample pairs that did not differ in allele frequency were
sel (10◦ 44 N, 131◦ 12 W). The samples were collected from June pooled. Levels of genetic differentiation between all sampling sites
through October (summer), with the exception of Nayarit, which were analysed by calculating Wright’s pairwise FST statistics (Weir
was collected in January. Individuals were captured by local fish- and Cockerham, 1984). A global AMOVA test including all sam-
ermen at all locations except the recreational fishing fleet at Cabo ples was conducted to obtain the global FST value. Hierarchical
San Lucas and La Paz and the incidental capture of the open ocean genetic structuring of the samples was undertaken by assessing
sample. Tissue samples were preserved in 80% ethanol until DNA the relative contribution among groups, within groups, and within
extraction. populations (AMOVA). For this, the sample sites that did not show

Table 1
Characteristics of the five microsatellite loci for the dolphinfish Coryphaena hippurus.

Locus name Size (pb) Primers Sequence Accession key

Chi002 104 F:GAAAAACTCACACGG TCACTTG (CT)8 (CA)2 CG(CA)3 AY135025


R:GGCTTGCCAACCTGA GATTA

Chi008 134 F:ATTGATGAGGGTTCAGACGG (TC)6 (TGTC)3 TGTA(TGTC)3 (TC)3 AY189832


R:GGCAGCAGTTCAGGAGGTTA

Chi008A 106 F:GGGCTCATGACACAAATT CC (TG)12 AY135026


R:CCAAACATGTGAGTGCTGCT

Chi023 139 F:GATGGGAGACTCCAA CCTGA (TG)2 AG(TG)3 TCAG(TG)3 CAGTAG(TG)3 TCAG(TG)3 AY135027
R:CCCATCTTGTGGAGGTTGAT

Chi037 82 F:GATATCAGGCCTCCT GCT TG (TG)8 AY135028


R:GGGATTGGT TCCCTCACTCT

Source: Robert Chapman (South Carolina Department of Natural Resources, Charleston SC, pers. comm.).
174 M.A. Tripp-Valdez et al. / Fisheries Research 105 (2010) 172–177

genetic differentiation with pairwise FST analysis were pooled and The global AMOVA test that included all samples resulted in a
compared with the rest of the samples. All the analyses were per- small but significant value for FST (FST = 0.01; P = 0.0). For the hierar-
formed using software Arlequin 3.0 (Excoffier et al., 2005) using chical AMOVA test, we pooled the samples that showed no genetic
10,000 permutation in each case. In all cases with multiple tests, differentiation from values for FST (Loreto, Cabo San Lucas 2006,
significance levels were adjusted using the sequential Bonferroni Guaymas 2006, La Paz 2006 and Mazatlan 2006) into a single group
technique (Rice, 1989). and compared this with the other samples. Results indicated sig-
We also used a Bayesian approach for assignment of individuals nificant differences among tested groups (FCT = 0.01; P = 0.00) and
to samples for each year with software Structure 2.3.1 (Pritchard within populations (FST = 0.01; P = 0.00), but no significant differ-
et al., 2000) to complement the results obtained with F statistics. ences among populations within groups (FSC = −0.004; P = 0.99).
For this, we selected the admixture model and the correlated allele These results denote a subtle heterogeneity over time within the
frequencies between populations. The Markov Chain Monte Carlo Gulf and subtle spatial heterogeneity with samples in surrounding
consisted of 105 steps with a burn-in period of 25,000 steps. We waters.
explored a range of K from 1 to 8 with 5 runs of each K value. The Bayesian assignment test made with the Structure software
The Mantel test was used to test correlation between genetic showed that, after five iterations, the higher value of K were for K = 1
distances, expressed as FST /(1 − FST ), and geographic distance, (log likelihood ln P(K) for K = 1 was −16,022, from which the ln P(K)
expressed as logKm with the program ISOLDE (part of Genepop; values began to fall and variance increased). These results suggest
Raymond and Rousset, 1995). that dolphinfish has no genetically differentiated groups within the
We used the software Migrate 2.1.3 (Beerli, 2008) to infer the area that we sampled. According to the Mantel test, there was no
population size parameter  (4NE ), where NE is the effective pop- significant correlation between genetic and geographic distance in
ulation size and  is the mutation rate per site) and migration the sampled locations (R2 = 0.005; P = 0.32).
rate (M, m/), where m is the immigration rate per generation.
For this analysis, we used the maximum likelihood approach with
3.3. Effective population size and migrate rates
the infinite-alleles model. Values of FST were used as the starting
parameters for the estimation of  and M. The ML was run for
Maximun likelihood estimates of the values of  made with
ten short and three long chains with 10,000 and 100,000 recorded
the Migrate software ranged from 1.05 (Ocean site) to 3.97
genealogies, respectively, with a burn-in of 10,000 genealogies for
(Loreto); these values were translated to effective population sizes
each chain. We ran the software two times for verifying consistency
(NE = /4) from 2625 to 9925, assuming a microsatellite mutation
of results.
rate of 10−4 per locus per generation occurs (Table 4).
In general, migration rates estimated from Migrate software
3. Results
were small, with values from 0.06 to 4.04. Gene flow in the sam-
3.1. Genetic diversity ples is asymmetrical, where the higher values of M are found
among samples inside the Gulf of California. Samples from out-
The mean number of alleles (NA ) was between 11.4 and 17.6 side the Gulf (Punta Lobos and Ocean) appear to migrant into the
at the five loci. The observed heterozygosity (HO ) ranged from Gulf.
0.80 to 0.93 and expected heterozygosity (HE ) ranged from 0.8
to 0.86 (Table 2). After the sequential Bonferroni correction, no 4. Discussion
loci were in linkage disequilibrium in either year (P > 0.05), which
indicates that the assortment of alleles is independent at the five 4.1. Genetic variability
loci. Of 65 Hardy–Weinberg equilibrium tests, using FIS (data not
shown) for every sampled site from both years with five loci, only This is the first study of genetic variation using microsatel-
10 tests showed disequilibrium, seven of them with heterozy- lite markers of the dolphinfish. Analyses of five microsatellite loci
gote excess and three with heterozygote deficits (Table 2). The revealed high levels of genetic variation (expressed in number of
Kruskal–Wallis test showed that there was no statistical difference alleles and heterozygosity) during the 2-year study. These results
in values of NA (P = 0.48), HO (P = 0.10) and HE (P = 0.91) between are comparable in variability to the report by DeWoody and Avise
locations. (2000) for other marine fishes (NA : 19.9, HE : 0.77) and pelagic fishes
of commercial importance, such as the swordfish Xiphias gladus
3.2. Population structure (NA : 14.3–17.5, HE : 0.771–0.784 in Muths et al., 2009), bigeye tuna
Thunnus obesus (NA : 8.7–20.4, HE : 0.897–0.909 in Gonzalez et al.,
After sequential Bonferroni correction, the exact test for genetic 2008), and sardine Sardina pilcharus (NA : 27.4–31, HE : 0.944–0.953
differentiation in allele frequencies revealed only two exact tests in Gonzalez and Zardoya, 2007), which have large populations and
with insignificant values (P > 0.025), These were Loreto 2005 ver- high gene flow. The Hardy–Weinberg disequilibrium is common
sus Loreto 2006, which were pooled for subsequent analyses in a in many marine fishes, but deviations to equilibrium generally
single “Loreto” sample, and Cabo San Lucas 2006 versus Guaymas prevail over heterozygotic deficits (Appleyard et al., 2001; Diaz-
2006, which was treated separately because of the different place of Jaimes and Uribe-Alcocer, 2006; Ruzzante et al., 1996), which
origin. Pairwise multi-locus FST analysis showed small but highly are a consequence of factors involving the fish’s reproductive
significant values in 41 out of 66 comparisons (FST from −0.017 systems, presence of null alleles, or a Wahlund effect (reduc-
to 0.031; P < 0.002; Table 3). From table, significant values for FST tion of heterozygosity in a population caused by subpopulation
occurred mainly between samples collected in 2005, between them structure). In the dolphinfish, we found more deviations toward
and samples the collected in 2006. The samples from 2006 collected heterozygous excess, possibly as a consequence of the mixture and
at the more geographically distant locations (Nayarit 2006, Punta reproduction of individuals of different cohorts that could have
Lobos 2006 and Ocean 2006) had significant differences in values different allelic composition. However, we could not find a clear
for FST and with some other 2006 localities inside the Gulf. The other tendency of Hardy–Weinberg disequilibrium in a specific loci or
samples collected in 2006 (Cabo San Lucas 2006, Guaymas 2006, La location, so there is a chance that those deviations of the equilib-
Paz 2006 and Mazatlan 2006) had smaller or insignificant values rium are a consequence of stochastic events in allelic frequency
for FST (P > 0.002). distributions.
M.A. Tripp-Valdez et al. / Fisheries Research 105 (2010) 172–177 175

Table 2
Summary statistics for five microsatellite loci in 13 samples of the dolphinfish Coryphaena hippurus.

Sample Locus

chi002 chi008 chi008a chi023 chi037 Mean population

Cabo San Lucas 2005 N 57 58 59 58 58 58


NA 13 26 15 25 9 17.6
HO 0.82 0.90 0.93 0.93 0.86 0.89
HE 0.78 0.89 0.82 0.93 0.82 0.85

Loreto 2005 N 55 54 55 54 55 55
NA 13 19 12 31 16 18
HO 0.98 0.94 0.89 0.91 0.93 0.93
HE 0.80 0.85 0.83 0.92 0.83 0.85

Guaymas 2005 N 62 63 61 61 63 62
NA 11 19 19 25 12 17.2
HO 0.92 0.92 0.70 0.98 0.92 0.89
HE 0.85 0.88 0.86 0.93 0.82 0.86

Puerto Libertad 2005 N 32 31 31 31 32 31.4


NA 14 13 16 12 11 13.2
HO 0.78 0.65 0.87 0.87 0.91 0.81
HE 0.86 0.82 0.91 0.89 0.80 0.86

Topolobampo 2005 N 73 70 72 69 72 71.2


NA 13 21 14 20 12 16
HO 0.77 0.90 0.86 0.96 0.83 0.86
HE 0.76 0.85 0.86 0.89 0.77 0.82

Cabo San Lucas 2006 N 56 68 63 47 58 58.4


NA 6 14 12 15 11 11.6
HO 0.68 0.85 0.90 0.81 0.78 0.80
HE 0.74 0.88 0.83 0.90 0.75 0.82

Loreto 2006 N 33 32 33 34 32 32.80


NA 7 18 10 16 7 11.60
HO 0.82 0.88 0.91 0.91 0.75 0.85
HE 0.74 0.88 0.85 0.89 0.72 0.82

Guaymas 2006 N 56 52 48 51 48 51
NA 7 15 13 16 10 12.2
HO 0.73 0.87 0.98 0.86 0.85 0.86
HE 0.75 0.86 0.85 0.87 0.78 0.82

Mazatlan 2006 N 71 71 68 64 74 69.6


NA 11 22 13 14 14 14.8
HO 1.00 0.94 0.87 0.97 0.84 0.92
HE 0.79 0.88 0.86 0.89 0.79 0.84

Nayarit 2006 N 78 77 70 73 77 75
NA 9 16 12 16 16 13.8
HO 0.95 0.79 0.76 0.78 0.83 0.82
HE 0.79 0.85 0.79 0.87 0.76 0.81

La Paz 2006 N 41 34 37 41 40 38.6


NA 8 16 10 14 9 11.4
HO 0.85 0.74 0.84 0.93 0.80 0.83
HE 0.71 0.85 0.78 0.90 0.74 0.80

Punta Lobos 2006 N 38 38 35 34 37 36.4


NA 8 17 12 16 8 12.2
HO 0.82 0.87 0.83 0.94 1.00 0.89
HE 0.73 0.84 0.84 0.89 0.72 0.80

Ocean 2006 N 45 48 50 48 48 47.8


NA 9 17 10 15 7 11.6
HO 0.89 0.90 0.92 1.00 0.92 0.92
HE 0.80 0.88 0.82 0.90 0.73 0.82

Mean of loci N 58.08 58.00 56.83 55.42 57.83 57.23


NA 10.17 18.33 13.25 18.25 11.33 14.27
HO 0.84 0.85 0.86 0.91 0.87 0.87
HE 0.78 0.86 0.84 0.90 0.77 0.83

Sample size (N); number of alleles (NA ); observed heterozygosity (HO ); expected heterozygosity (HE ). Data in bold are Hardy–Weinberg disequilibrium for FIS (P < 0.01).

4.2. Population structure showed a slight genetic heterogeneity both in temporal and spatial
scales. However, Bayesian assignment analysis made with Struc-
Small but highly significant values of pairwise FST revealed ture software and an isolation-by-distance test, which correlates
that dolphinfish have subtle genetic differentiation between sam- geographic distance with genetic distance, failed to detect any dif-
pled sites, mainly in the samples collected in 2005 but in some ferent genetic group in the entire data set. This means that even
of the samples of 2006. These results and those from AMOVA when there was some genetic heterogeneity among the samples,
176 M.A. Tripp-Valdez et al. / Fisheries Research 105 (2010) 172–177

Table 3
Pairwise values of FST (below the diagonal) and their respective P values (above the diagonal).

CSL05 LOT GUA05 PLI05 TOP05 CSL06 GUA06 MAZ06 NAY06 LAP06 PLO06 OCE06

CSL05 – 0.000 0.000 0.000 0.000 0.038 0.004 0.000 0.000 0.000 0.000 0.000
LOT 0.019 – 0.000 0.000 0.000 1.000 0.270 0.725 0.011 0.105 0.000 0.001
GUA05 0.027 0.023 – 0.001 0.000 0.000 0.000 0.000 0.000 0.000 0.000 0.000
PLI05 0.020 0.020 0.012 – 0.000 0.373 0.005 0.000 0.000 0.000 0.000 0.000
TOP05 0.024 0.008 0.027 0.015 – 0.974 0.002 0.000 0.000 0.002 0.000 0.000
CSL06 0.004 −0.015 0.010 0.000 −0.005 – 0.997 0.988 0.012 0.999 0.735 0.761
GUA06 0.007 0.000 0.021 0.010 0.007 −0.010 – 0.119 0.002 0.090 0.003 0.006
MAZ06 0.017 −0.002 0.016 0.014 0.009 −0.007 0.001 – 0.000 0.164 0.000 0.004
NAY06 0.021 0.004 0.025 0.019 0.017 0.005 0.007 0.009 – 0.019 0.000 0.000
LAP06 0.016 0.002 0.022 0.018 0.009 −0.017 0.003 0.001 0.006 – 0.061 0.001
PLO06 0.018 0.013 0.031 0.021 0.015 −0.004 0.009 0.011 0.020 0.004 – 0.000
OCE06 0.022 0.006 0.021 0.018 0.010 −0.003 0.006 0.005 0.015 0.009 0.010 –

Bold number of values of FST are significant (P < 0.002; after sequential Bonferroni test); CSL (Cabo San Lucas); LOT (Loreto); GUA (Guaymas); PLI (Puerto Libertad); TOP
(Topolobampo); MAZ (Mazatlan); NAY (Nayarit); LAP (La Paz); PLO (Punta Lobos); OCE (Ocean) sites.

the variability is too high to find any prevalent group of dolpinfish high reproductive capacity and capacity to spawn many times
in this region. This supports the results obtained by Diaz-Jaimes et within the year (Beardsley, 1967) and (4) along the Pacific coast
al. (2006) with mtDNA. They did not detect any population struc- of Mexico, the highest abundance of dolphinfish larvae is present
ture of samples from the Pacific and the Gulf of California from 2002 within the 28 ◦ C isotherm in summer and the 24 ◦ C isotherm in
to 2004. Both results suggest a single population in the study area. spring (Sanchéz-Reyes, 2008). That means that reproduction of dol-
In studies of microsatellites in cichlid fish (Duftner et al., 2006), phinfish is associated with sea surface temperature rather than a
birds (Friesen et al., 2006) and mammals (Cegelski et al., 2003), specific geographic site. This view of the distribution of dolphinfish
data were analysed simultaneously with the Structureı̌s Bayesian is important, given the interannual variability of temperature in the
approach and pairwise FST calculations; both methods revealed a Gulf of California caused by ENSO events (Lluch-Cota, 2004).
well-defined genetic population structure, contrary to what we
found in dolphinfish. The values of FST in our study (<0.027) are 4.3. NE estimate and migration rates
low, compared to cichlid fish (0.04–0.2), birds (0.07–0.2) and mam-
mals (0.07–0.09). This implies that the genetic structure detected Effective population size of dolphinfish, estimated by Migrate
by FST for dolphinfish is too weak for a multi-locus analysis (Struc- software based in  values ranged from 2625 to 9925. Even if there
ture software) to detect, providing it is present. Latch et al. (2006) is no census information on the real population size of dolphinfish
observed that Structure software infers the number of clusters even in this region to compare with estimates of NE , the values derived
when these were not well-differentiated (FST = 0.02–0.03), but with from Migrate software are very similar to population size of big-
results suggesting that FST must be at least 0.05 to reach an accuracy eye tuna (4175–15,175) obtained by Gonzalez et al. (2008) using
>97%. Similar results were obtained for bigeye tuna (Gonzalez et al., Migrate software with the same mutation rate. These authors men-
2008), in which certain values of pairwise FST were very low, but tion that for population size derived from annual catch data NE is
highly significant (FST between 0.001 and 0.027), but the Structure several orders of magnitude lower than the real population size,
software failed to detect any genetically differentiated group, sup- which appears to be common in marine fish (Frankham et al., 2005;
porting the conclusion that there is no genetic structure differences Hauser et al., 2002). This means that the real population size of
in tuna in this area. dolphinfish could be several orders of magnitude larger than the
In this context, it is important to distinguish between low but estimated NE .
biological meaningful genetic differences and a statistical effect Migration rates (M) estimated with Migrate software were small
(Waples, 1998). In our study, significant values of FST found in comparing with other large pelagic fishes, such as striped marlin
both years are too low to suggest genetically different popula- Kajikia audax in the Pacific (immigration rates between 25.1 and
tions. Besides, dolphinfish have biological features that support 71.2; McDowell and Graves, 2008) and bigeye tuna T. obesus (immi-
the hypothesis of a single population in the northern part of Mex- gration rates between 10 and 60; Gonzalez et al., 2008). However,
ico, including: (1) high migratory activity (Kingsford and Defries, these rates demonstrate that there is gene flow between samples
1999); (2) feed on many prey and change their diet according to that prevents isolation of groups. Estimates of migration rates also
the epipelagic environment (Aguilar-Palomino et al., 1998); (3) demonstrate that Cabo San Lucas in 2005 and 2006 seem to be an

Table 4
Maximun likehood estimates of migration rates (M) and effective population size (NE ) of the dolphinfish Coryphaena hippurus.

Location CSL05a LOTa GUA05a PLI05a TOP05a CSL06a GUA06a MAZ06a NAY06a LAP06a PLO06a OCE06a NE

CSL05 1.76 1.91 1.24 2.60 1.85 0.82 1.07 0.18 0.70 0.67 0.40 0.70 4400
LOT 3.27 3.97 1.67 1.25 2.00 2.45 0.89 1.58 1.20 1.20 1.30 2.05 9925
GUA05 2.01 2.69 1.76 1.86 1.42 0.75 0.35 0.91 0.53 0.13 0.89 1.43 4400
PLI05 0.66 0.49 1.80 1.57 2.38 1.63 0.32 1.42 1.50 0.11 0.26 0.26 3925
TOP05 3.39 1.68 1.24 1.86 1.16 0.60 0.39 1.90 0.95 0.50 0.39 1.62 2900
CSL06 0.31 0.21 0.38 0.18 0.33 1.26 1.67 1.67 2.42 1.01 0.83 1.21 3150
GUA06 0.14 1.33 0.89 0.56 0.47 1.36 1.21 1.43 2.18 3.35 1.94 0.80 3025
MAZ06 0.92 2.30 0.68 0.56 0.43 3.56 1.70 1.79 1.80 1.56 0.66 1.05 4475
NAY06 0.15 0.84 0.66 0.51 0.69 1.02 1.31 2.89 1.30 1.46 1.38 0.71 3250
LAP06 0.36 0.69 1.86 0.94 0.76 1.47 4.04 2.86 2.70 1.18 0.83 1.16 2950
PLO06 0.13 0.63 1.09 0.36 0.83 2.92 1.72 0.63 1.39 1.17 1.26 2.20 3150
OCE06 0.06 1.30 0.78 0.71 0.41 2.35 1.27 2.24 1.40 2.18 2.53 1.05 2625
a
Recipient population; values in bold correspond to  estimates of each sample; NE estimated with the formula /4.
M.A. Tripp-Valdez et al. / Fisheries Research 105 (2010) 172–177 177

important center of gene flow. Zúñiga-Flores et al. (2008) observed Cegelski, C.C., Waits, L.P., Anderson, N.J., 2003. Assessing population structure
that, although dolphinfish are found throughout the year near Cabo and gene flow in Montana wolverines (Gulo gulo) using assignment-based
approaches. Mol. Ecol. 12, 2907–2918.
San Lucas, it is more abundant in summer when sea surface tem- DeWoody, J.A., Avise, J.C., 2000. Microsatellite variation in marine, freshwater and
perature inside the Gulf of California increase and forms favorable anadromous fishes compared with other animals. J. Fish. Biol. 56, 461–473.
conditions for aggregation. Since the collection of tissue samples Diaz-Jaimes, P., Uribe-Alcocer, M., 2006. Spatial differentiation in the eastern Pacific
yellowfin tuna revealed by microsatellite variation. Fish. Sci. 72, 590–596.
in our study occurred between June and October, migration rates Diaz-Jaimes, P., Uribe-Alcocer, M., Ortega-Garcia, S., Durand, J.D., 2006. Spatial and
may be a consequence of the movement of dolphinfish to warmer temporal mitochondrial DNA genetic homogeneity of dolphinfish populations
waters. (Coryphaena hippurus) in the eastern central Pacific. Fish. Res. 80, 333–338.
Duftner, N., Sefc, K.M., Koblmuller, S., Nevado, B., Verheyen, E., Phiri, H., Sturmbauer,
Although the results for NE and migration rates for dolphinfish C., 2006. Distinct population structure in a phenotypically homogeneous rock-
are comparable with those reported for other species of pelagic dwelling cichlid fish from lake Tanganyika. Mol. Ecol. 15, 2381–2395.
fishes, such as tuna, sardines and striped marlin (Carlsson et al., Excoffier, L., Laval, G., Schneider, S., 2005. ARLEQUIN. Ver 3.0. An integrated software
package for population genetic data analysis. Evol. Bioinform. Online 1, 47–50.
2004; Gonzalez and Zardoya, 2007; Gonzalez et al., 2008; McDowell
Frankham, R., Ballou, J.D., Briscoe, D., 2005. Introduction to Conservation Genetics.
and Graves, 2008), these data need to be used with caution, espe- Cambridge University Press, Cambridge.
cially if the objective is for managing the resource. Although the Friesen, V.L., Gonzalez, J.A., Cruz-Delgado, F., 2006. Population genetic structure and
coalescent method that uses Migrate software provides more accu- conservation of the Galapagos petrel (Pterodroma phaeopygia). Conserv. Genet.
7, 105–115.
rate and reliable estimates of genetic differentiation than methods Gonzalez, E., Zardoya, R., 2007. Relative role of life-history traits and historical factors
based on Wright’s FST (Beerli and Felsenstein, 2001; Gonzalez et al., in shaping genetic population structure of sardines (Sardina pilchardus). BMC
2008), the estimates obtained with this software could have large Evol. Biol. 7, 197.
Gonzalez, E.G., Beerli, P., Zardoya, R., 2008. Genetic structuring and migration pat-
variations by a few orders of magnitude, depending on the muta- terns of Atlantic bigeye tuna, Thunnus obesus (Lowe, 1839). BMC Evol. Biol. 8,
tion rate that is used (Carlsson et al., 2004). Given these concerns, 252.
the results for NE and gene flow are very important for management Graves, J.E., 1998. Molecular insights into the population structures of cosmopolitan
marine fishes. J. Hered. 89, 427–437.
of dolphinfish, there is still no reliable information on how individ- Hauser, L., Adcock, G.J., Smith, P.J., Bernal Ramãrez, J.H., Carvalho, G.R., 2002. Loss
uals are moving inside the Gulf of California (Zuñiga-Flores, Centro of microsatellite diversity and low effective population size in an overexploited
Interdisciplinario de Ciencias Marinas-IPN, La Paz, B.C.S., Mexico, population of New Zealand snapper (Pagrus auratus). Proc. Natl. Acad. Sci. U.S.A.
99, 11742–11747.
pers. comm.). Kingsford, M.J., Defries, A., 1999. The ecology and fishery for Coryphaena spp. in the
Even when we found temporal and spatial genetic heterogeneity waters around Australia and New Zealand. Sci. Mar. 63, 267–275.
among samples, no obvious grouping of dolphinfish was identified. Latch, E.K., Dharmarajan, G., Glaubitz, J.C., Rhodes, J.R.O.E., 2006. Relative perfor-
mance of Bayesian clustering software for inferring population substructre and
This means that under the genetic concept of stock, which estab-
individual assignment at low levels of population differentiation. Conserv. Gen.
lishes that a stock is a reproductively isolated unit and is genetically 7, 295–302.
different from other units (Begg and Waldman, 1999; Ward, 2000), Lluch-Cota, S.E., 2004. Marine Ecosystems of the North Pacific. Pices Spec. Pub., Gulf
dolphinfish in the Gulf of California and surrounding waters form a of California, pp. 1–7, #1.
McDowell, J.R., Graves, J.E., 2008. Population structure of striped marlin (Kajikia
single panmictic stock with high genetic variability, comparable to audax) in the Pacific Ocean based on analysis of microsatellite and mitochondrial
other commercially exploited pelagic fish. This information should DNA. Can. J. Aquat. Fish. Sci. 65, 1307–1320.
be considered before a management plan is established because any Muths, D., Grewe, P., Jean, C., Bourjea, J., 2009. Genetic population structure of the
swordfish (Xiphias gladius) in the southwest Indian Ocean: sex-biased differ-
local fishery could have potential effects on the entire population entiation, congruency between markers and its incidence in a way of stock
and all areas where dolphinfish are captured. assessment. Fish. Res. 97, 263–269.
Palko, J.B., Beardsley, G.L., Richards, W.J., 1982. Synopsis of the biological data on
dolphin-fishes, Coryphaena hippurus and Coryphaena equiselis L. NOAA Technical
Acknowledgments Report NMFS Circular 443, FAO Fish Synopsis No. 130, 28.
Peakall, R., Smouse, P.E., 2006. Genalex 6: genetic analysis in excel, population
We thank Robert Chapman of the South Carolina Department genetic software for teaching and research. Mol. Ecol. Notes 6, 288–295.
Pritchard, J.K., Stephens, M., Donnelly, P., 2000. Inference of population structure
of Natural Resources for data on microsatellite sequences; Eloisa using multilocus genotype data. Genetics 155, 945–959.
Herrera–Valdivia, Rufino Morales and Marcela Zuñiga for assistance Raymond, M., Rousset, F., 1995. GENEPOP 4.0 population genetic software for exact
with tissue sampling. Jair Rangel for technical support in laboratory. test and ecumenism. J. Hered. 86, 248–249.
Rice, W., 1989. Analyzing tables of statistical tests. Evolution 43, 223–225.
Funding was provided by the Comisión Nacional de Acuacul-
Rocha-Olivares, A., Bobadilla-Jimenez, M., Ortega-Garcia, S., Saavedra-Sotelo, N.,
tura y Pesca (CONAPESCA grant 956-1), Centro de Investigaciones Sandoval-Castillo, J.R., 2006. Mitochondrial variability of dolphinfish Coryphaena
Biológicas del Noroeste and the State Governments of Sonora and hippurus populations in the Pacific Ocean. Cien. Mar. 32, 569–578.
Rodzen, J.A., Agresti, J., Tranah, G., May, B., 1998. Agarose overlay allow simplified
Sinaloa.
staining of polyacrilamide gels. Biotechnology 24, 584.
Ruzzante, D.E., Taggart, C.T., Cook, D., 1996. Spatial and temporal variation in the
References genetic composition of a larval cod (Gadus morhua) aggregation: cohort contri-
bution and genetic stability. Can. J. Fish. Aquat. Sci. 53, 2695–2705.
Sanchéz-Reyes, N.A., 2008. Distribución de larvas de dorado Coryphaena hippurus
Aguilar-Palomino, B., Galvan-Magaña, F., Abitia-Cárdenas, L., Muhlia-Melo, A.,
(Linnaeus, 1758) y Coryphaena equiselis (Linnaeus, 1758) en el Pacífico oriental
Rodriguez-Romero, J., 1998. Feeding aspects of the dolphin Coryphaena hippu-
mexicano. Master’s Thesis. Centro Interdisciplinario de Ciencias Marinas-IPN, La
rus Linnaeus, 1758 in Cabo San Lucas, Baja California Sur, Mexico. Cien. Mar. 24,
Paz, B.C.S., Mexico.
253–265.
Shaw, P.W., Turan, C., Wright, J.M., O’Connell, M., Carvalho, G.R., 1999. Microsatellite
Appleyard, S.A., Grewe, P.M., Innes, B.H., Ward, R.D., 2001. Population structure of
DNA analysis of population structure in Atlantic herring (Cuplea harengus), with
yellowfin tuna (Thunnus albacares) in the western Pacific Ocean, inferred from
direct comparison to allozyme and mtDNA RFLP analyses. Heredity 83, 490–499.
microsatellite loci. Mar. Biol. 139, 383–393.
StatSoft, 2004. STATISTICA, Version 7. StatSoft, www.statsoft.com.
Beardsley Jr., G.L., 1967. Age, growth, and reproduction of the dolphin, Coryphaena
Waples, R.S., 1998. Separating the wheat from the chaff: Patterns of genetic differ-
hippurus, in the Straits of Florida. Copeia 2, 441–451.
entiation in high gene flow species. J. Hered. 89, 438–450.
Beerli, P., 2008. Migrate-N Version 2.4., http://popgen.scs.fsu.edu/Migrate-n.html.
Ward, R.D., 2000. Genetics in fisheries management. Hydrobiologia 420, 191–201.
Beerli, P., Felsenstein, J., 2001. Maximum likelihood estimation of a migration
Weir, B.S., Cockerham, C.C., 1984. Estimating F-statistics for the analysis of popula-
matrix and effective population sizes in n subpopulations by using a coalescent
tion structure. Evolution 38, 1358–1370.
approach. Proc. Natl. Acad. Sci. U.S.A. 98, 4563–4568.
Yoshida, K., Nakagawa, N., Wada, S., 2005. Multiplex PCR system applied for
Begg, G.A., Waldman, J.R., 1999. An holistic approach to fish stock identification. Fish.
analysing microsatellite loci of Schlegel’s black rockfish, Sebastes schlegeli. Mol.
Res. 43, 35–44.
Ecol. Notes 5, 416–418.
Carlsson, J., McDowell, J.R., Diaz-Jaimes, P., Carlsson, J.E.L., Boles, S.B., Gold, J.R.,
Zúñiga-Flores, M.S., Ortega-García, S., Klett-Traulsen, A., 2008. Interannual and sea-
Graves, J.E., 2004. Microsatellite and mitochondrial DNA analyses of Atlantic
sonal variation of dolphinfish (Coryphaena hippurus) catch rates in the southern
bluefin tuna (Thunnus thynnus thynnus) population structure in the Mediter-
Gulf of California, Mexico. Fish. Res. 94, 13–17.
ranean Sea. Mol. Ecol. 13, 3345–3356.

Você também pode gostar