Plan
Comptes Rendus

Molecular biology and genetics/Biologie et génétique moléculaires
Phylogenetic inference in Odontesthes and Atherina (Teleostei: Atheriniformes) with insights into ecological adaptation
[Inférence phylogénétique chez Odontesthes et Atherina (Teleostei : Atheriniformes) avec un aperçu de l’adaptation écologique]
Comptes Rendus. Biologies, Volume 334 (2011) no. 4, pp. 273-281.

Résumé

This contribution provides an insight into Atheriniformes systematics based on four mitochondrial regions: 12S rRNA, cytb, COI and control region (2794 bp in total). In the Atherinopsoidei (New World silversides), comparisons among five species of Odontesthes, O. argentinensis, O. bonariensis, O. smitti, O. hatcheri and O. incisa revealed a putative marine-freshwater pairing pattern of Odontesthes species, possibly driven by sea level fluctuations of South American waters. This study represents the first data on molecular phylogeny of Odontesthes species that can be of usefulness to biodiversity conservation policies. In the Atherinoidei (Old World silversides), Atherina boyeri was corroborated as a species complex constituted by a marine form, a marine with dark spots form and a brackish form. Concretely, Odontesthes and Atherina may represent geographically replicated models to study genetic adaptation and speciation of marine species to brackish and freshwater habitats. In addition, phylogenetic analyses supported Odontesthes and Atherina as monophyletic taxa and their separation into two differentiated suborders Atherinopsoidei and Atherinoidei, respectively.

Métadonnées
Reçu le :
Accepté le :
Publié le :
DOI : 10.1016/j.crvi.2011.03.004
Mots clés : Odontesthes, Atherina, Molecular phylogeny, 12S rRNA, Cyt b, COI, Control region
Sandra Heras 1 ; María Inés Roldán 1

1 Laboratori d’Ictiologia Genètica, Departament de Biologia, Universitat de Girona. Campus Montilivi, 17071 Girona, Spain
@article{CRBIOL_2011__334_4_273_0,
     author = {Sandra Heras and Mar{\'\i}a In\'es Rold\'an},
     title = {Phylogenetic inference in {\protect\emph{Odontesthes}} and {\protect\emph{Atherina}} {(Teleostei:} {Atheriniformes)} with insights into ecological adaptation},
     journal = {Comptes Rendus. Biologies},
     pages = {273--281},
     publisher = {Elsevier},
     volume = {334},
     number = {4},
     year = {2011},
     doi = {10.1016/j.crvi.2011.03.004},
     language = {en},
}
TY  - JOUR
AU  - Sandra Heras
AU  - María Inés Roldán
TI  - Phylogenetic inference in Odontesthes and Atherina (Teleostei: Atheriniformes) with insights into ecological adaptation
JO  - Comptes Rendus. Biologies
PY  - 2011
SP  - 273
EP  - 281
VL  - 334
IS  - 4
PB  - Elsevier
DO  - 10.1016/j.crvi.2011.03.004
LA  - en
ID  - CRBIOL_2011__334_4_273_0
ER  - 
%0 Journal Article
%A Sandra Heras
%A María Inés Roldán
%T Phylogenetic inference in Odontesthes and Atherina (Teleostei: Atheriniformes) with insights into ecological adaptation
%J Comptes Rendus. Biologies
%D 2011
%P 273-281
%V 334
%N 4
%I Elsevier
%R 10.1016/j.crvi.2011.03.004
%G en
%F CRBIOL_2011__334_4_273_0
Sandra Heras; María Inés Roldán. Phylogenetic inference in Odontesthes and Atherina (Teleostei: Atheriniformes) with insights into ecological adaptation. Comptes Rendus. Biologies, Volume 334 (2011) no. 4, pp. 273-281. doi : 10.1016/j.crvi.2011.03.004. https://comptes-rendus.academie-sciences.fr/biologies/articles/10.1016/j.crvi.2011.03.004/

Version originale du texte intégral

1 Introduction

Although relationships within the large and complex series of Atherinomorpha fishes (Teleostei) have been widely examined [1], there are still many uncertainties. Within this series, the order Atheriniformes (Table 1) comprises the respective suborders Atherinoidei (including the Atherinidae – Old World silversides) and Atherinopsoidei (including the Atherinopsidae – New World silversides). However, their separation into two differentiated suborders has not been still genetically verified. In addition, despite their significant biological and commercial value, the genetic relationships among these fishes remain largely unknown.

Table 1

Specimens and abbreviation codes, location of collection sites, sample size, haplotypes and frequencies (in parenthesis) for each mitochondrial region.

Haplotypes
Species Code Locality n Phe + 12S rRNA Cytb COI Thr + Pro + CR Combined markers
Atherinopsoidei
 Atherinopsidae
  O. argentinensis Oarg Mar Chiquita lagoon, Argentina 1 1(1) 1(1) 1(1) 1(1) 1(1)
Mar del Plata, Argentina 2 1(2) 1(2) 2(1), 3(1) 2(1), 3(1) 2(1), 3(1)
  O. smitti Osmi Mar del Plata, Argentina 3 2(2), 3(1) 1(3) 4(2), 5(1) 4(1), 5(1), 6(1) 4(1), 5(1), 6(1)
  O. incisa Oin Mar del Plata, Argentina 3 4(1), 5(1), 6(1) 1(3) 6(1), 7(1), 8(1) 7(1), 8(1), 9(1) 7(1), 8(1), 9(1)
  O. bonariensis Obon Gómez lagoon, Argentina 3 7(2), 8(1) 2(3) 9(2), 10(1) 10(1), 11(1), 12(1) 10(1), 11(1), 12(1)
  O. hatcheri Ohat Nahuel Huapi lake, Argentina 1 9(1) 3(1) 11(1) 13(1) 13(1)
Atherinoidei
 Atherinidae
  A. hepsetus Ath Calella de Palafrugell, Spain 4 10(4) 4(4) 12(1), 13(2), 14(1) 14(1), 15(1), 16(1), 17(1) 14(1), 15(1), 16(1), 17(1)
  A. boyeri Atb Mar Menor coastal lagoon, Spain 4 11(3), 12(1) 5(3), 6(1) 15(3), 16(1) 18(1), 19(1), 20(2) 18(1), 19(1), 20(2)
 Melanoteniidae
  Melanotaenia sp Mel Unknown 4 13(4) 7(4) 17(4) 21(4) 21(4)

The genus Odontesthes, commonly known as “pejerreyes”, represents most of the New World silversides, being exclusive to South American waters (Fig. 1) [2]. Valued in fisheries and aquaculture, they have attracted sport fisheries as well [3]. Species of primary interest in the southwestern Atlantic include O. argentinensis Valenciennes, 1835, with commercial and sport fisheries in estuaries and coasts from Brazil, Uruguay and Argentina; O. bonariensis Valenciennes, 1835, a freshwater fish distributed in lakes and lagoons from Province of Buenos Aires (Argentina) and Rio Grande do Sul (Brazil); O. hatcheri Eigenmann, 1909, inhabiting Patagonian lakes and rivers of Argentina and Chile [2]; O. smitti Lahille, 1929, distributed along the Atlantic coast of Uruguay and Argentina and passing through Strait of Magellan to the Pacific coast of Chile [4,5]; and O. incisa Jenyns, 1841, distributed in the Atlantic coast from Dos Patos lagoon (Brazil) to Golfo Nuevo (Argentina) [5]. Molecular phylogenetic studies have been restricted to O. argentinensis (allozymes [6]; microsatellites and mtDNA control region [7]), where a species-wide morphological uniformity [2] contrasts with the incipient estuarine and coastal species suggested by these genetic studies. According to Froese [8], there is the need of deep revisions and identification catalogues of South American fishes, especially those from freshwater. In spite of this, few molecular studies of Odontesthes species exist and there are no studies on their molecular systematics.

Fig. 1

Distribution and sampling locations of Odontesthes species analyzed. 1: Mar Chiquita lagoon; 2: Mar del Plata; 3: Gómez lagoon; 4: Nahuel Huapi lake.

Within Atherinidae, genus Atherina is represented in Mediterranean waters by two species with commercial interest [9,10]: A. hepsetus Linnaeus, 1758, a typical marine silverside and Atherina boyeri Risso, 1810, which is distributed in lagoons, estuaries and marine environments. Systematic questions concern the relationships between these species, and the existence of a species complex within A. boyeri constituted by a marine non-punctuated form, a marine form with dark spots along the lateral line and a lagoon form [11–19].

As molecular markers, we used four regions of the mitochondrial genome to examine intra- and interspecific relationships within Odontesthes, providing a first molecular contribution to the understanding of their and Atherina phylogenetic relationships. We also clarify intergeneric relationships within Atheriniformes and report inter- and intrafamilial genetic distances for future taxonomical comparisons.

2 Material and methods

2.1 Sampling

Details of the 21 specimens collected for this study include individuals of two genera and seven species; Melanotaenia sp., belonging to the same suborder as Atherina genus (Atherinoidei), was also analyzed (Table 1). Morphological identification was based on Bauchot [9], Cousseau and Perrotta [5] and Dyer [2]. Muscle tissue was excised and conserved in 95% ethanol at the Laboratori d’Ictiologia Genètica collection.

2.2 PCR and sequencing

DNA isolation and polymerase chain reaction of phenylalanine transfer RNA (Phe), 12S rRNA, cytochrome b (cytb), cytochrome c oxidase subunit I (COI), threonine transfer RNA (Thr), proline transfer RNA (Pro) and control region (CR) were carried out in 50 μl reaction volume following Heras et al. [20]. The Phe and 12S rRNA region were amplified using the following pair sets of primers: L15927-Thr: 5′-AGA GCG TCG GTC TTG TAA TCC G-3′ [21] + 12ASR-H: 5′-ATA GTG GGG TAT CTA ATC CCA GTT-3′ [22] and L1091: 5′-CAA ACT GGG ATT AGA TAC CCC ACT AT-3′ [23] + H1478: 5′-TGA CTG CAG AGG GTG ACG GGC GGT GTG T-3′ [23] for Odontesthes spp., L15927-Thr + H1358-12S: 5′-CGA CGG CGG TAT ATA GGC-3′ [21] and L1091 + H1478 for Atherina spp. and DLHR: 5′-CAT CTG GTT CTT ACT TCA GG-3′, reverse of DLH, [24] + H1358-12S and L1091 + H1478 for Melanotaenia sp. For cytb amplifications, we used L14850-CYB: 5′-GCC TGA TGA AAC TTT GGC TC-3′ and H15560-CYB: 5′-TAG GCA AAT AGG AAG TAT CA-3′ [25] for all species. FishF2 5′-TCG ACT AAT CAT AAA GAT ATC GGC AC-3 and FishR1 5′-TAG ACT TCT GGG TGG CCA AAG AAT CA-3′ [26] were applied for COI PCRs for all species and finally, L15927-Thr and CSBDH: 5′-TGA ATT AGG AAC CAG ATG CCA G-3′ [27] for that of Thr, Pro and CR. Amplifications were verified on 1% agarose gel with ethidium bromide (0.5 mg ml−1) and cleaned in a GFX PCR DNA and Gel Band Purification Kit (Amersham Biosciences, Little Chalfont, UK). Sequencing was performed by BigDye terminator v1.1 Cycle Sequencing Kit (Applied Biosystems) according to manufacturer's instructions and fragments were read with an ABI PRISM 3130 Genetic Analyzer (Applied Biosystems) at our laboratory. The nucleotide sequence of each gene in all species was verified from both strands. The primers (3.2 μM) used for forward and reverse strands sequencing were the same as for PCR.

2.3 Genetic analyses

Sequences were processed using SeqScape v2.5 (Applied Biosystems) and final alignments and sequence editing were carried out with BioEdit v7.0.4.1 [28]. DAMBE v4.5.61 [29] plots of the accumulated transitions and transversions vs Tamura and Nei [30] genetic distance were employed to test for nucleotide saturation. Tamura and Nei [30] mean distance values between species groups were obtained by MEGA v4 [31] for all mitochondrial regions separately. The partition homogeneity test (α = 0.05) was conducted with PAUP* v4.0b10 [32] to examine the homogeneity or incongruence between mitochondrial regions to be combined. Phylogenetic relationships were inferred by: neighbor-joining (NJ) analysis based on Tamura and Nei model employing MEGA and based on a maximum likelihood (ML) distance matrix employing PAUP*; maximum parsimony (MP) analysis and ML analysis both performed using PAUP* with an heuristic search and TBR branch-swapping algorithm with 10 random sequence addition; and lastly, Bayesian inference by MrBayes v3.1.2 [33]. Modeltest v3.7 [34] selected GTR + I + G as the best-fit model under Akaike information criterion for ML analyses and MrModeltest v2.2 [35] estimated GTR + I + G as the evolutionary model to run MrBayes program. Metropolis-coupled Markov chain Montecarlo (MCMC) analysis in MrBayes was achieved with four chains of 1 × 106 generations sampled every 100th and discarded 25% of samples as burn-in. A consensus tree with branch length and clade credibility (posterior probability) was generated with the 75% remaining samples. Robustness of trees was tested using bootstrap analysis [36] with 1000 replicates except for Bayesian analysis, which employed posterior probability for clade credibility. Unclear relationships inside Acanthopterygii regarding Atheriniformes were reported before [37] but available information indicate that Paracanthopterygii are their common ancestor, consequently, Gadus morhua (Paracanthopterygii - GenBank accession no. X99772) was used as outgroup species for all analyses.

3 Results

3.1 Haplotype diversity

Sequence alignment of Phe + 12S rRNA (908 bp), cytb (702 bp), COI (651 bp) and Thr + Pro + CR (533 bp) produced 13 (GenBank accession no. GQ352651–GQ352663), 7 (GenBank accession no. GQ352666–GQ352672), 17 (GenBank accession no. GQ352673–GQ352689) and 21 (GenBank accession nos. GQ352690–GQ352710) haplotypes, respectively (Table 1). Note that O. argentinensis (GQ352664), O. smitti (GQ352665) and O. incisa (GQ352666) shared haplotype 1 for cytb indicating the low evolutionary signal of this molecular marker for discerning among these taxa and suggesting its elimination for latter analyses. In contrast, no shared haplotypes between species were detected for the rest of mitochondrial genes. Furthermore, no clear association between haplotypes and environmental distribution for O. argentinensis was observed because of shared haplotypes detected between individuals from Mar Chiquita lagoon (estuarine) and Mar del Plata (open sea) in Phe + 12S rRNA and cytb (Table 1).

3.2 Genetic divergence

No saturation was detected for used markers (data not shown). As expected, the pairwise Tamura and Nei distance matrix generated from each dataset (Table 2) identified much higher distances for intergeneric comparisons. All genera appeared almost equally separated among them for Phe + 12SrRNA and COI, but for Thr + Pro + CR, the proximity of Atherina and Melanotaenia versus Odontesthes was apparent. The lowest divergence values were those between O. argentinensisO. bonariensis and O. smittiO. hatcheri for all molecular markers, indicating close genetic relationship between these paired taxa (Table 2).

Table 2

Tamura and Nei [30] mean genetic distances and standard error for each sample and three mitochondrial markers. Species codes are as in Table 1.

Phe + 12S rRNA COI Thr + Pro + CR
Within Odontesthes
 Oarg 0 0.0031 ± 0.0017 0.0116 ± 0.0042
 Osmi 0.0011 ± 0.0011 0.0015 ± 0.0014 0.0131 ± 0.0042
 Oin 0.0022 ± 0.0013 0.0083 ± 0.0029 0.0087 ± 0.0034
 Obon 0.0011 ± 0.0011 0.0031 ± 0.0020 0.0237 ± 0.0061
Between Odontesthes
 Oarg-Osmi 0.0119 ± 0.0037 0.0302 ± 0.0067 0.0693 ± 0.0119
 Oarg-Oin 0.0105 ± 0.0033 0.0516 ± 0.0083 0.0654 ± 0.0112
 Oarg-Obon 0.0017 ± 0.0012 0.0046 ± 0.0020 0.0253 ± 0.0054
 Oarg-Ohat 0.0113 ± 0.0037 0.0294 ± 0.0065 0.0680 ± 0.0119
 Osmi-Oin 0.0058 ± 0.0023 0.0405 ± 0.0078 0.0540 ± 0.0101
 Osmi-Obon 0.0136 ± 0.0039 0.0291 ± 0.0066 0.0703 ± 0.0111
 Osmi-Ohat 0.0039 ± 0.0020 0.0101 ± 0.0039 0.0204 ± 0.0054
 Oin-Obon 0.0100 ± 0.0032 0.0516 ± 0.0083 0.0629 ± 0.0102
 Oin-Ohat 0.0075 ± 0.0027 0.0439 ± 0.0081 0.0546 ± 0.0106
 Obon-Ohat 0.0130 ± 0.0040 0.0283 ± 0.0066 0.0704 ± 0.0115
Within Atherina
 Ath 0 0.0062 ± 0.0024 0.0084 ± 0.0028
 Atb 0.0022 ± 0.0016 0.0015 ± 0.0015 0.0054 ± 0.0026
Between Atherina
 Atb-Ath 0.0479 ± 0.0076 0.1297 ± 0.0151 0.2162 ± 0.0226
Odontesthes vs Atherina
 Oarg-Atb 0.1643 ± 0.0150 0.2336 ± 0.0224 0.5740 ± 0.0522
 Oarg-Ath 0.1685 ± 0.0150 0.2253 ± 0.0217 0.5156 ± 0.0455
 Osmi-Atb 0.1602 ± 0.0150 0.2173 ± 0.0212 0.5937 ± 0.0542
 Osmi-Ath 0.1643 ± 0.0151 0.2153 ± 0.0205 0.5633 ± 0.0488
 Oin-Atb 0.1636 ± 0.0149 0.2185 ± 0.0211 0.6114 ± 0.0561
 Oin-Ath 0.1677 ± 0.0150 0.2233 ± 0.0213 0.5664 ± 0.0492
 Obon-Atb 0.1652 ± 0.0152 0.2302 ± 0.0222 0.5775 ± 0.0521
 Obon-Ath 0.1694 ± 0.0152 0.2221 ± 0.0213 0.5248 ± 0.0463
 Ohat-Atb 0.1637 ± 0.0150 0.2253 ± 0.0221 0.5977 ± 0.0543
 Ohat-Ath 0.1678 ± 0.0151 0.2174 ± 0.0206 0.5731 ± 0.0503
Odontesthes vs Melanotaenia
 Oarg-Mel 0.1714 ± 0.0163 0.2219 ± 0.0215 0.5595 ± 0.0496
 Osmi-Mel 0.1706 ± 0.0161 0.2234 ± 0.0210 0.5972 ± 0.0538
 Oin-Mel 0.1708 ± 0.0160 0.2207 ± 0.0201 0.6093 ± 0.0548
 Obon-Mel 0.1737 ± 0.0162 0.2206 ± 0.0211 0.5576 ± 0.0486
 Ohat-Mel 0.1726 ± 0.0162 0.2199 ± 0.0206 0.5904 ± 0.0525
Atherina vs Melanotaenia
 Atb-Mel 0.1586 ± 0.0149 0.2096 ± 0.0206 0.4240 ± 0.0369
 Ath-Mel 0.1772 ± 0.0160 0.2068 ± 0.0207 0.4040 ± 0.0345
Odontesthes vs Atherina vs Melanotaenia
Odontesthes spp-Atherina spp 0.1660 ± 0.0147 0.2237 ± 0.0193 0.5706 ± 0.0463
Odontesthes spp-Melanotaenia sp 0.1717 ± 0.0163 0.2214 ± 0.0202 0.5816 ± 0.0506
Atherina spp-Melanotaenia sp 0.1710 ± 0.0150 0.2085 ± 0.0191 0.4155 ± 0.0328

3.3 Phylogenetic analysis

Partition homogeneity test on 2794 bp excluded cytb (P < 0.0001) as noted above (Table 1), establishing the congruence of combining Phe + 12S rRNA, COI and Thr + Pro + CR, and generating a consensus alignment of 2092 bp (P = 0.7900). Phylogenetic analysis generated similar topologies of the trees indicating two major different lineages or phylogroups (Fig. 2). One lineage comprised all Odontesthes spp. haplotypes (100% robustness) clearly supporting a monophyletic group. It is noteworthy that haplotypes of O. argentinensis from Mar del Plata open sea (Oarg2 and Oarg3) clustered together. In all the analyses, O. argentinensis and O. bonariensis had minima distance values and shared a common ancestor (99% of bootstrap support). Similarly, O. smitti and O. hatcheri were closely related (99% of bootstrap support) and grouped with O. incisa haplotypes (81% of bootstrap support). A second lineage was formed by A. boyeri and A. hepsetus (100% robustness) which strongly clustered with Melanotaenia sp (100% of bootstrap support).

Fig. 2

Maximum parsimony tree with combined markers based on 573 parsimony informative sites. The numbers in nodes indicate ≥ 60 bootstrap value. The values below branches (in italics) indicate the number of mutational steps (1641 in total). Haplotype codes are as in Table 1. Gadus morhua was used as outgroup species.

We constructed additional phylogenies on Atherina spp., incorporating comparable data of 12S rRNA (Fig. 3A) (42 haplotypes from 165 sequences) and control region (Fig. 3B) (300 haplotypes from 453 sequences) from GenBank [13,16,17,38] with those of our study. In this analysis, our A. hepsetus haplotypes clustered with those from previous authors and grouped with A. presbyter. In this analysis, our A. boyeri haplotypes matched the lagoon-estuarine type and was clearly differentiated from the marine types punctuated (P) and non-punctuated (NP) that emerged into two diverged lineages.

Fig. 3

Comparative study in Atherina spp. A: Neighbor-joining tree of 12S rRNA (351 bp) based on Tamura and Nei [30] distances using 31 sequences from this study, and 165 sequences from other authors. B: Neighbor-joining tree of control region (381 bp) based on Tamura and Nei [30] distances using 31 sequences from this study and 453 sequences from other authors. The numbers on nodes indicate ≥ 60 bootstrap values. Triangle sizes are proportional to the number of haplotypes present in the cluster. NP = non-punctuated; P = punctuated. Gadus morhua was used as outgroup species.

4 Discussion

4.1 Odontesthes phylogeny

In corroborating the monophyly of the genus Odontesthes analysing five species from marine and freshwater environments, our results support the morphological analyses revised by Dyer [2] based on anatomical characters. For all mitochondrial regions, genetic distances between Odontesthes species were relatively low in comparison with those relative to Atherina species, indicative that genus Odontesthes radiated as recently as cichlid fishes in Lake Victoria [39]. Genetic distances in COI fit the most frequent value between congeneric fishes previously reported by Ward ([40]; D ≈ 0.035). Likewise, the sympatric distribution of the marine species O. argentinensis, O. smitti and O. incisa, could be explained as a secondary contact after allopatric speciation as indicated by the splitting in three different nodes (Fig. 2).

Within the Odontesthes cluster (Fig. 2), O. argentinensis from Mar del Plata open sea (Oarg2 and Oarg3) grouped together and split off from that of Mar Chiquita estuarine lagoon (Oarg1), suggesting the occurrence of a barrier to gene flow [41] between these two environments. These findings are consistent with genetic studies using allozymes [6] and microsatellites and mitochondrial control region [42] and morphological landmarks [43], each indicating differentiated populations or even incipient species of O. argentinensis from marine and estuarine environments along Brazilian and Uruguayan coasts, respectively. Moreover, spawning grounds and timing in the two forms of O. argentinensis are different, as well as their habitat preference. Temporal and spatial separation of their life cycle implying an ecological divergence [6] may contribute to create one or several mechanisms producing complete reproductive isolation between them [44]. Moreover, there is no geographical barrier between Mar Chiquita lagoon and Mar del Plata open sea in Argentina separating the lagoon-estuarine form from the marine one, like in the Brazilian area studied by [6]. However, our reduced number of individuals and haplotypes and the low genetic distances, characteristic of intraspecific levels, do not allow us to establish two valid species. Nonetheless, bottlenecks or founder effects may determine the differentiation in O. argentinensis populations of brackish environments coming from an ancestral marine population [6], provoking a feasible beginning of ecological speciation [45]. We consider O. argentinensis as an interesting model for divergent adaptive selection studies associated to reproductive isolation mechanisms, consequently, the preservation of the marine and estuarine-lagoon populations must be taken into account for biodiversity conservation policies in the Mar Chiquita lagoon, designated a UNESCO World Biosphere reserve in 1996 [46].

Genetic distances detected between O. bonariensis (freshwater fish) and O. argentinensis for all molecular markers were within the range of Odontesthes intraspecific levels (Table 2). Moreover, O. bonariensis and O. argentinensis comprised a common lineage in all phylogenetical analyses (Fig. 2) consistent with their shared morphological characters [47]. In addition, Tombari et al. [48] pointed out that O. bonariensis and O. argentinensis were difficult to distinguish morphologically with no significant morphometric differentiation. Moreover, Tejedor [49] observed natural interbreeding between these two taxa in the La Salada Grande lagoon (General Madariaga, Argentina). O. bonariensis, a freshwater fish would be the product of the speciation of a common marine ancestor shared with O. argentinensis [49] as well as with O. perugiae species complex [7]. Experimental works that could demonstrate speciation at present time in nature had been considered unrealizable due to the assumption of a long time of isolation requirements. However, Hendry et al. [50] proved how the reproductive isolation could evolve in a fast way when new and different habitats were colonized. In that work, reproductive isolation was detected at only after 13 generations, representing a short evolution period of 56 years, between two populations of Oncorhynchus nerka coming from a common freshwater reproductive source that was introduced in different environments (coastal marine and river). Silversides exhibit a great ability for invading and speciating in vacant niches [51] like the Laguna de Gómez, where O. bonariensis lives, that have a connection towards Atlantic Sea through Río Salado, in front of Argentinean coastal waters where O. argentinensis occurs. Therefore, the high genetic variability of the Odontesthes marine-brackish ancestor would have preadapted them to colonize and speciate rapidly into new habitats [45,52]. Colonization of new estuarine habitats, with very unstable physical-chemical conditions and biological factors, would favour a fast adaptation and reproductive isolation by means of natural selection [53].

Similarly, our analyses revealed a close relationship between O. smitti (marine) and O. hatcheri (freshwater) (Table 2; Fig. 2), both joining O. incisa (marine) in a common cluster (Fig. 2), although Dyer [2] argued that O. hatcheri was the sister group for the entire genus Odontesthes. Despite the occurrence of spontaneous hybridization cases in captivity suggesting the proximity of O. bonariensis and O. hatcheri, both freshwater fishes were easily discriminated with all mitochondrial genes used in this study as well as in a previous study with RFLPs [54,55].

According to Bamber and Henderson [51], most atherinid fishes including genus Odontesthes from Patagonia, where O. hatcheri occurs, had mainly originated from marine immigrants. A rapid uplift of the Andes is speculated to have occurred at later Miocene (10–5 mya) [56] and until Pleistocene-Holocene (1.8 mya–10,000 ya) there was a mass of water at the continental platform due to marine transgressions (Paranense Sea, Patagonian Sea) [47,57]. Thus, O. hatcheri may have diverged just before the sea level descended at the one currently known in that area. In addition, phylogenetic analyses suggest a marine–freshwater pairing pattern in Odontesthes and an expected freshwater reciprocal fish for O. incisa, a hypothesis awaiting to be verified when the complete genus will be analyzed. Therefore, the vicariant segregation of a marine and freshwater species could be a result of both tectonic plates and glacial intervals provoking fluctuations in the sea level that depicted the major part of the variability at marine species from the Atlantic Ocean [58]. As was suggested by O. perugiae complex [45], change of the sea level during Pleistocene-Holocene could have originated the lineages derived from the sea in South America matching with the recent radiation of Odontesthes [59]. Given that vicariant patterns of speciation in marine fishes are difficult to detect due to insufficient phylogenetic analyses [60], Odontesthes phylogeny offers a great opportunity to study biogeographic processes in South America [61].

4.2 Atherina phylogeny

Recently A. boyeri as a single taxon has been questioned and a complex of forms has been proposed including: (1) two types, marine and lagoon [11–13,18] and (2a) three species, A. boyeri (marine non-punctuated), A. punctata (marine punctuated) and A. lagunae (lagoon) [14,15,19]; (2b) three species, marine non-punctuated form, marine punctuated form and A. boyeri (lagoon form) [17]. In the recent revision by Kottelat and Freyhof [62], the individuals named as A. lagunae by Trabelsi et al. [14,15] were in fact A. boyeri based on morphological diagnostic characters that were consistent with Francisco et al. [17] conclusions. In addition, A. punctata is an unavailable name due to its homonymy with A. punctata Bennet, 1833 and thus, the 21 available synonyms of A. boyeri should be considered and examined before naming this taxon [62,63]. To verify these proposals, we calculated mean Tamura and Nei [30] genetic distances in 12S rRNA (351 bp, 42 haplotypes) and control region (381 bp, 300 haplotypes) among all Atherina species and types to determine their degree of genetic differentiation. Genetic distances between the different types of A. boyeri (lagoon - marine punctuated were D = 0.0713 ± 0.0143, D = 0.2028 ± 0.0243; lagoon - marine non-punctuated were D = 0.0457 ± 0.0106, D = 0.1996 ± 0.0240; marine punctuated - marine non-punctuated were D = 0.0527 ± 0.0121, D = 0.1904 ± 0.0233), are noticeable higher than those detected between two recognized species of genus Atherina, A. hepsetus and A. presbyter (D = 0.0179 ± 0.0057, D = 0.1133 ± 0.0164). Consequently, given that species status of A. hepsetus and A. presbyter has not been questioned, these results together with the three phylogroups detected (Fig. 3) support the three different species associated with lagoon and marine environments of A. boyeri previously proposed constituted by: 1) the marine form without the presence of spots; 2) the marine form with dark spots along the lateral line, restricted to western Mediterranean Sea [19]; and 3) the brackish form from estuaries and lagoons. In this way, isolation that had operated over populations exploiting different niches [64] or environments with different salinities [52] would support the occurrence of ecological speciation. Thanks to silverside plasticity, that would imply a selection of generalist genotypes to confront a wide range of conditions [42], the ancestor of A. boyeri could have adapted to radiate repeatedly across vacant habitats, thus colonizing, estuaries and lagoons and acquiring also the ability to invade freshwater biotopes [51]. Thus, this specific variation (meristics, size, particular morphological structures, etc.) of some species as a result of a response to the ambience [52] due to both fluctuations of the Mediterranean Sea levels during Pleistocene age and the Messinian crisis [17] probably originated the species complex in A. boyeri.

4.3 Atheriniformes phylogeny

Genetic distances for COI between the three families analyzed, Atherinopsidae (Odontesthes), Atherinidae (Atherina) and Melanotaeniidae (Melanotaenia) (D = 0.2085–0.2237; Table 2) fit into the mean value between families of the same order of fishes (D = 0.2256; [40]). In addition, Atherinoidei was stated as a monophyletic group (Fig. 2). Nelson [65] included Atherinidae in the suborder Atherinoidei with two recognized groups, Old World and New World silversides as subfamilies. More recently, Nelson [1] revised the classification in the light of recent morphoanatomical systematic works [47,66], where: (1) Old World silversides are retained in the Atherinidae family in suborder Atherinoidei (Table 1); and (2) the family Atherinopsidae was created for New World silversides within a new suborder Atherinopsoidei (Table 1). Our results support the new taxonomical category assignments showing the closeness, both in the means of phylogenetic analyses and genetic distances, of Atherina and Melanotaenia (Atherinoidei), compared to Odontesthes (Atherinopsoidei) (Table 2; Fig. 2).

5 Conclusions

  • 1 The close genetic relationships between Odontesthes species were indicative that this genus has radiated in relatively recent time;
  • 2 The status of A. boyeri needs to be revised and three species are corroborated corresponding to a marine non-punctuated form, a marine punctuated form and a brackish form;
  • 3 Odontesthes and Atherina are recognized as monophyletic genera;
  • 4 The monophyly of suborders Atherinopsoidei and Atherinoidei is recognized.

As molecular data accumulate, more refined phylogenetic analyses will became possible. Meanwhile, our genetic data caste light on some taxonomical aspects of Atheriniformes. However, the whole Odontesthes phylogeny based on molecular data still remains unknown and further studies including other species of Odontesthes are required. In addition, we recommend studies on reproductive biology and population genetics within O. argentinensis and between O. argentinenesisO. bonariensis and O. smittiO. hatcheri in order to estimate interbreeding levels, and to corroborate or reject the incipient marine–freshwater dichotomy suggested in this work. Odontesthes and Atherina may represent geographically replicated ideal models to study genetic adaptation and speciation of marine fish to brackish and freshwater habitats.

Disclosure of interest

The authors declare that they have no conflicts of interest concerning this article.

Acknowledgments

The participation of Patricia Noguera, Juan Roldán, Martí Cortey, María Berta Cousseau, Mariano González-Castro, Mercedes González and Asunción Andreu made the collections possible. We would like to thank Fred Utter for helpful comments on the manuscript and corrections of the English style. Funding was provided to SHM by BRAE predoctoral fellowship and to MIR by 910305 grant from Universitat de Girona.


Bibliographie

[1] J.S. Nelson Fishes of the world, John Wiley and Sons, New York, 2006

[2] B.S. Dyer Systematic revision of the South American silversides (Teleostei, Atheriniformes), Biocell, Volume 30 (2006), pp. 69-88

[3] B.S. Dyer Atherinopsidae (Neotropical silversides) (R.E. Reis; S.O. Kullander; C.J. Ferraris, eds.), Checklist of the Freshwater Fishes of South and Central America, Edipucrs, Porto Alegre, 2003, pp. 515-525

[4] B.S. Dyer; A.E. Gosztonyi Phylogenetic revision of the South American subgenus Austromenidia Hubbs, 1918 (Teleostei, Atherinopsidae, Odontesthes) and a study of meristic variation, Rev. Biol. Mar. Oceanogr., Volume 34 (1999), pp. 211-232

[5] M.B. Cousseau; R.G. Perrotta Peces marinos de Argentina: biología, distribución, pesca, INIDEP, Mar del Plata, Argentina, 2003

[6] L.B. Beheregaray; J.A. Levy Population genetics of the silverside Odontesthes argentinensis (Teleostei, Atherinopsidae): evidence for speciation in an estuary of Southern Brazil, Copeia, Volume 2 (2000), pp. 441-447

[7] L.B. Beheregaray; P. Sunnucks Microsatellite loci isolated from Odontesthes argentinensis and the O. perugiae species group and their use in other South American silverside fish, Mol. Ecol, Volume 9 (2000), pp. 629-644

[8] R. Froese The good, the bad, and the ugly: a critical look at species and their institutions from a user's perspective, Rev. Fish Biol. Fisheries, Volume 9 (1999), pp. 375-378

[9] M.L. Bauchot Poissons osseux (W. Fischer; M.L. Bauchot; M. Schneider, eds.), Fiches FAO d’identification pour les besoins de la pêche, (rev. 1), Méditerranée et mer Noire, Zone de pêche 37, 2, Commission des Communautés Européennes and FAO, Rome, 1987, pp. 967-969

[10] A. Pallaoro; J. Dulcic; I. Jardas; M. Kraljevic Some biological parameters of the Mediterranean sand smelt Atherina (Atherina) hepsetus, Linnaeus, 1758 (Pisces: Atherinidae) from the middle eastern Adriatic (Croatian coast), J. Appl. Ichthyol, Volume 23 (2007), pp. 189-192

[11] B. Focant; E. Rosecchi; A.J. Crivelli Attempt at biochemical characterization of sand smelt Atherina boyeri Risso, 1810 (Pisces, Atherinidae) populations from the Camargue (Rhône delta, France), Comp. Biochem. Physiol. B, Volume 122 (1999), pp. 261-267

[12] E. Klossa-Kilia; M. Prassa; V. Papasotiropoulos; S. Alahiotis; G. Kilias Mitochondrial DNA diversity in Atherina boyeri populations as determined by RFLP analysis of three mtDNA segments, Heredity, Volume 89 (2002), pp. 363-370

[13] E. Klossa-Kilia; V. Papasotiropoulos; G. Tryfonopoulos; S. Alahiotis; G. Kilias Phylogenetic relationships of Atherina hepsetus and Atherina boyeri (Pisces: Atherinidae) populations from Greece, based on mtDNA sequences, Biol. J. Linn. Soc., Volume 92 (2007), pp. 151-161

[14] M. Trabelsi; E. Faure; J.P. Quignard; M. Boussaïd; B. Focant; F. Mâamouri Atherina punctata and Atherina lagunae (Pisces, Atherinidae), new species in the Mediterranean Sea, 1. Biometric investigations of three atherinid species, C. R. Biol., Volume 325 (2002), pp. 967-975

[15] M. Trabelsi; A. Gilles; C. Fleury; F. Mâamouri; J.P. Quignard; E. Faure Atherina punctata and Atherina lagunae (Pisces, Atherinidae), new species found in the Mediterranean Sea, 2. Molecular investigations of three atherinid species, C. R. Biol., Volume 325 (2002), pp. 1119-1128

[16] L. Astolfi; I. Dupanloup; R. Rossi; P.M. Bisol; E. Faure; L. Congiu Mitochondrial variability of sand smelt (Atherina boyeri, Risso, 1810) populations from North Mediterranean coastal lagoons, Mar. Ecol. Pro. Ser., Volume 297 (2005), pp. 233-243

[17] S.M. Francisco; L. Congiu; S. Stefanni; R. Castilho; A. Brito; P.P. Ivanova; A. Levy; H. Cabral; G. Kilias; I. Doadrio; V.C. Almada Phylogenetic relationships of the North-eastern Atlantic and Mediterranean forms of Atherina (Pisces, Atherinidae), Mol. Phylogenet. Evol., Volume 48 (2008), pp. 782-788

[18] S. Kraitsek; E. Klossa-Kilia; V. Papasotiropoulos; S.N. Alahiotis; G. Kilias Genetic divergence among marine and lagoon Atherina boyeri populations in Greece using mtDNA analysis, Biochem. Genet., Volume 46 (2008), pp. 781-798

[19] V. Milana; L. Sola; L. Congiu; A.R. Rossi Mitochondrial DNA in Atherina (Teleostei, Atheriniformes): differential distribution of an intergenic spacer in lagoon and marine forms of Atherina boyeri, J. Fish Biol., Volume 73 (2008), pp. 1216-1227

[20] S. Heras; M.I. Roldán; M. González Castro Molecular phylogeny of Mugilidae fishes revised, Rev. Fish Biol. Fisheries, Volume 19 (2009), pp. 217-231

[21] M. Miya; M. Nishida Use of mitogenomic information in Teleostean molecular phylogenetics: a tree-based exploration under the Maximum-Parsimony optimality criterion, Mol. Phylogenet. Evol., Volume 17 (2000), pp. 437-455

[22] S. Palumbi; A. Martin; S. Romano; W.O. McMillan; L. Stice; G. Grabowski The simple fool's guide to PCR, version 2.0. Honolulu: Department of Zoology and Kewalo Marine Laboratory, University of Hawaii, USA, 1991

[23] T.D. Kocher; W.K. Thomas; A. Meyer; S.V. Edwards; S. Paabo; F.X. Villablanca; A.C. Wilson Dynamics of mitochondrial DNA evolution in animals: amplification and sequencing with conserved primers, Proc. Natl. Acad. Sci. U S A, Volume 86 (1989), pp. 6196-6200

[24] R. Tiedemann; J. Harder; C. Gmeiner; E. Haase Mitochondrial DNA sequence patterns of Harbour porpoises (Phocoena phocoena) from the North and the Baltic Sea, Z. Säugetierkunde, Volume 61 (1996), pp. 104-111

[25] M. Miya; M. Nishida Organization of the mitochondrial genome of a deep-sea fish, Gonostoma gracile (Teleostei: Stomiiformes): first example of transfer RNA gene rearrangements in bony fishes, Mar. Biotechnol., Volume 1 (1999), pp. 416-426

[26] R.D. Ward; T.S. Zemlak; B.H. Innes; P.R. Last; P.D.N. Hebert DNA barcoding Australia's fish species, Philos. Trans. R. Soc. B-Biol. Sci., Volume 360 (2005), pp. 1847-1857

[27] J.R. Alvarado Bremer; J. Mejuto; A.J. Baker Mitochondrial DNA control region sequences indicate extensive mixing of swordfish (Xiphias gladius L.) populations in the Atlantic Ocean, Can. J. Fish. Aquat. Sci., Volume 52 (1995), pp. 1720-1732

[28] T.A. Hall BioEdit: a user-friendly biological sequence alignment editor and analysis program for Windows 95/98/NT, Nucl. Acids Sym. Ser., Volume 41 (1999), pp. 95-98

[29] X. Xia; Z. Xie DAMBE: data analysis in molecular biology and evolution, J. Hered., Volume 92 (2001), pp. 371-373

[30] K. Tamura; M. Nei Estimation of the number of nucleotide substitutions in the control region of mitochondrial DNA in humans and chimpanzees, Mol. Biol. Evol., Volume 10 (1993), pp. 512-526

[31] K. Tamura; J. Dudley; M. Nei; S. Kumar MEGA4: Molecular Evolutionary Genetics Analysis (MEGA) software version 4.0, Mol. Biol. Evol., Volume 24 (2007), pp. 1596-1599

[32] D.L. Swofford PAUP* Phylogenetic analysis using parsimony (*and other methods), version 4, Sinauer Associates, Sunderland, USA, 2002

[33] F. Ronquist; J.P. Huelsenbeck MrBayes 3: Bayesian phylogenetic inference under mixed models, Bioinformatics, Volume 19 (2003), pp. 1572-1574

[34] D. Posada; K.A. Crandall Modeltest: testing the model of DNA substitution, Bioinformatics, Volume 14 (1998), pp. 817-818

[35] J.A.A. Nylander MrModeltest v2., Program distributed by the author. Evolutionary Biology Centre, Uppsala University, Sweden, 2004

[36] J. Felsenstein Confidence limits on phylogenies: an approach using the bootstrap, Evolution, Volume 39 (1985), pp. 783-791

[37] D.H.E. Setiamarga; M. Miya; Y. Yamanoue; K. Mabuchi; T.P. Satoh; J.G. Inoue; M. Nishida Interrelationships of Atherinomorpha (medakas, flyingfishes, killifishes, silversides, and their relatives): the first evidence based on whole mitogenomes sequences, Mol. Phylogenet. Evol., Volume 49 (2008), pp. 598-605

[38] S.M. Francisco; H. Cabral; M.N. Vieira; V.C. Almada Contrasts in genetic structure and historical demography of marine and riverine populations of Atherina at similar geographical scales, Estuar. Coast. Shelf Sci., Volume 69 (2006), pp. 655-661

[39] A. Meyer; T.D. Kocher; P. Basasibwaki; A.C. Wilson Monophyletic origin of Lake Victoria cichlid fishes suggested by mitochondrial DNA sequences, Nature, Volume 347 (1990), pp. 550-553

[40] R.D. Ward DNA barcode divergence among species and genera of birds and fishes, Mol. Ecol. Res., Volume 9 (2009), pp. 1077-1085

[41] J.W. Sites; J.C. Marshall Delimiting species: a Renaissance issue in systematic biology, Trends Ecol. Evol., Volume 18 (2003), pp. 462-470

[42] L.B. Beheregaray; P. Sunnucks Fine-scale genetic structure, estuarine colonization and incipient speciation in the marine silverside fish Odontesthes argentinensis, Mol. Ecol., Volume 10 (2001), pp. 2849-2866

[43] M.A. Bemvenuti Silversides in South Brazil: morphological and ecological aspects, Biocell, Volume 30 (2006), pp. 111-118

[44] G.F. Turner What is a fish species?, Rev. Fish Biol. Fisheries, Volume 9 (1999), pp. 281-297

[45] L.B. Beheregaray; P. Sunnucks; D.A. Briscoe A rapid fish radiation associated with the last sea-level changes in southern Brazil: the silverside Odontesthes perugiae complex, Proc. R. Soc. Lond. B Biol. Sci., Volume 269 (2002), pp. 65-73

[46] O. Iribarne Reserva de Biosfera Mar Chiquita: Características físicas, biológicas y ecológicas, UNESCO-Editorial Martín, Mar del Plata, 2001

[47] B.S. Dyer Phylogenetic systematics and historical biogeography of the Neotropical silverside family Atherinopsidae (Teleostei, Atheriniformes) (L.R. Malabarba; R.E. Reis; R.P. Vari; Z.M. Lucena; C.A.S. Lucena, eds.), Phylogeny and Classification of Neotropical Fishes, Edipucrs, Porto Alegre, 1998, pp. 519-536

[48] A.D. Tombari; A.V. Volpedo; D.D. Echeverría Development of the sagitta in young and adults of Odontesthes argentinensis (Valenciennes, 1835) and Odontesthes bonariensis (Valenciennes, 1835) from Buenos Aires province, Argentina (Teleostei: Atheriniformes), Rev. Chil. Hist. Nat., Volume 78 (2005), pp. 632-633

[49] D. Tejedor El pejerrey como recurso genético (F. Grosman, ed.), Fundamentos biológicos económicos y sociales para una correcta gestión del recurso pejerrey, Astyanax, Buenos Aires, 2001, pp. 27-31

[50] A.P. Hendry; J.K. Wenburg; P. Bentzen; E.C. Volk; T.P. Quinn Rapid evolution of reproductive isolation in the wild: evidence from introduced salmon, Science, Volume 290 (2000), pp. 516-518

[51] R.N. Bamber; P.A. Henderson Pre-adaptive plasticity in atherinids and the estuarine seat of teleost evolution, J. Fish Biol., Volume 33 (1988), pp. 17-23

[52] G. Cognetti; F. Maltagliati Biodiversity and adaptive mechanisms in brackish water fauna, Mar. Poll. Bull., Volume 40 (2000), pp. 7-14

[53] D.T. Bilton; J. Paula; J.D.D. Bishop Dispersal, genetic differentiation and speciation in estuarine organisms, Estuar Coast, Shelf Sci., Volume 55 (2002), pp. 937-952

[54] C.A. Strüssmann; T. Akaba; K. Ijima; K. Yamaguchi; G. Yoshizaki; F. Takashima Spontaneous hybridization in the laboratory and genetic markers for the identification of hybrids between two atherinid species, Odontesthes bonariensis (Valenciennes, 1835) and Patagonina hatcheri (Eigenmann, 1909), Aquac. Res, Volume 28 (1997), pp. 291-300

[55] G. Yoshizaki; K. Yamaguchi; T. Oota; C.A. Strüssmann; F. Takashima Cloning and characterization of pejerrey mitochondrial DNA and its application for RFLP analysis, J. Fish Biol., Volume 51 (1997), pp. 193-203

[56] C.M. Garzione; G.D. Hoke; J.C. Libarkin; S. Withers; B. MacFadden; J. Eiler; P. Ghosh; A. Mulch Rise of the Andes, Science, Volume 320 (2008), pp. 1304-1307

[57] R.M. Hernández; T.E. Jordan; A. Dalenz Farjat; L. Echavarría; B.D. Idleman; J.H. Reynolds Age, distribution, tectonics, and eustatic controls of the Paranense and Caribbean marine transgressions in southern Bolivia and Argentina, J. South Am. Earth Sci., Volume 19 (2005), pp. 495-512

[58] J. Colborn; R.E. Crabtree; J.B. Shaklee; E. Pfeiler; B.W. Bowen The evolutionary enigma of bonefishes (Albula spp.): cryptic species and ancient separations in a globally distributed shorefish, Evolution, Volume 55 (2001), pp. 807-820

[59] N.R. Lovejoy; J.S. Albert; W.G.R. Crampton Miocene marine incursions and marine/freshwater transitions: evidence from Neotropical fishes, J. South Am. Earth Sci., Volume 21 (2006), pp. 5-13

[60] E.O. Wiley On species and speciation with reference to the fishes, Fish Fish., Volume 3 (2002), pp. 161-170

[61] P. Sunnucks; A.C.C. Wilson; L.B. Beheregaray; K. Zenger SSCP is not so difficult: the application and utility of single-stranded conformation polymorphism in evolutionary biology and molecular ecology, Mol. Ecol, Volume 9 (2000), pp. 1699-1710

[62] M. Kottelat; J. Freyhof Notes on the taxonomy and nomenclature of some European freshwater fishes, Ichthyol. Explor. Freshw., Volume 20 (2009), pp. 75-90

[63] R. Froese, D. Pauly, FishBase: World Wide Web electronic publication, www.fishbase.org, 2010.

[64] E.B. Taylor Species pairs of north temperate freshwater fishes: evolution, taxonomy, and conservation, Rev. Fish Biol. Fisheries, Volume 9 (1999), pp. 299-324

[65] J.S. Nelson Fishes of the world, John Wiley and Sons, New York, 1994

[66] B.S. Dyer; B. Chernoff Phylogenetic relationships among atheriniform fishes (Teleostei: Atherinomorpha, Zool. J. Linn. Soc., Volume 117 (1996), pp. 1-69


Commentaires - Politique


Ces articles pourraient vous intéresser

Surfing among species, populations and morphotypes: Inferring boundaries between two species of new world silversides (Atherinopsidae)

Mariano González-Castro; Juan José Rosso; Ezequiel Mabragaña; ...

C. R. Biol (2016)


Atherina punctata and Atherina lagunae (Pisces, Atherinidae), new species found in the Mediterranean Sea. 2. Molecular investigations of three Atherinid species

Monia Trabelsi; André Gilles; Caroline Fleury; ...

C. R. Biol (2002)


Atherina punctata and Atherina lagunae (Pisces, Atherinidae), new species in the Mediterranean Sea. 1. Biometric investigations of three Atherinid species

Monia Trabelsi; Éric Faure; Jean-Pierre Quignard; ...

C. R. Biol (2002)