Next Article in Journal
Species Composition and Structure of Beetle Associations in Caves of the Częstochowa Upland, Poland
Previous Article in Journal
The Reptile Relocation Industry in Australia: Perspectives from Operators
Previous Article in Special Issue
Native Population Structure beyond Hatchery Introgression in the Endemic Sicilian Trout
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Investigating an Unknown Biodiversity: Evidence of Distinct Lineages of the Endemic Chola Guitarfish Pseudobatos percellens Walbaum, 1792 in the Western Atlantic Ocean

1
Laboratório de Biologia e Genética de Peixes, Instituto de Biociências de Botucatu, UNESP, Botucatu 18618-689, Brazil
2
Acervo Zoológico da Universidade Santa Cecília (AZUSC), Universidade Santa Cecília, UNISANTA, Santos 11045-907, Brazil
3
Marine Vertebrate Evolution and Conservation Lab—EvolVe, Departamento de Biologia, Universidade Federal do Ceará, Fortaleza 60020-181, Brazil
4
Laboratório de Dinâmica de Populações Marinhas, Universidade Federal Rural de Pernambuco, Recife 52171-900, Brazil
*
Author to whom correspondence should be addressed.
Diversity 2023, 15(3), 344; https://doi.org/10.3390/d15030344
Submission received: 17 January 2023 / Revised: 12 February 2023 / Accepted: 22 February 2023 / Published: 28 February 2023

Abstract

:
Anthropogenic actions have affected marine species for a long time, through overexploitation of natural stocks and habitat degradation, influencing the life strategies of several taxa, especially rays and sharks, which have suffered significant population declines in recent years. Therefore, conservation actions and stock management have become paramount. In this regard, chola guitarfish, Pseudobatos percellens, distributed throughout the Brazilian coast, is often commercially fished by local artisanal fleets or as by-catch in shrimp trawl fisheries. Therefore, this study aimed to understand the genetic diversity of P. percellens throughout the Brazilian coast, using single nucleotide polymorphisms (SNPs). Genetic analyses employing 3329 SNPs revealed a hidden biodiversity within P. percellens, with at least one lineage occurring in the Northern and Northeastern regions and another distributed in the Southeastern/Southern Brazilian coast, with high genetic differentiation between them. However, the Discriminant Analysis of Principal Components (DAPC) indicated the presence of in fact three lineages distributed in these regions that must still be better investigated. Therefore, to ensure adequate conservation of chola guitarfish biodiversity, populations must be managed separately along the Brazilian coast. Furthermore, the need for a taxonomic review for this group is noted.

1. Introduction

Dramatic and persistent population declines have been documented for many fish species worldwide in recent decades [1] due to rapid environmental changes caused by anthropogenic impacts, such as habitat loss and pollution, in addition to overfishing [1,2,3,4,5,6], which has resulted in an unprecedented crisis. According to Díaz et al. [7], about one million species are estimated to be threatened with extinction due to the acceleration of this global biodiversity crisis [7,8,9,10,11], with an increasing number of threatened species requiring management. Because of this, methods to assess these species and prioritize conservation actions are now paramount [6,12,13].
Management objectives often focus on distinct populations or independent lineages, aspects that are difficult to determine in some species. In this regard, population genetic methods provide a means of delineating geographic distribution patterns and molecular variations, inferring the processes that generate and maintain these standards and allowing for hypothesis postulations [14,15,16,17,18].
Population genetics plays an important role in the management and conservation of endangered species [19,20], revealing population peculiarities, such as inbreeding and loss of genetic diversity, both of which compromise the viability of wild populations [21]. Maintaining genetic diversity is one of the foundations of conservation genetics, mainly because genetic diversity maintenance provides species plasticity in the face of a series of environmental variables and supports their evolutionary potential [22,23,24].
Sharks and rays (Elasmobranchii) are one of the most vulnerable groups concerning anthropogenic effects, mainly due to their K-strategist life history characteristics, such as low fecundity rates, slow growth, longevity, large size at hatching, and late age at maturity [25,26,27]. These animals play an important role in controlling and regulating the populations of other species, in addition to influencing the structure and function of coastal and oceanic ecosystems and comprising good marine ecosystem health indicators [28,29,30,31].
Currently, about 1485 elasmobranchs species have been identified worldwide (665 sharks and 820 rays or stingrays) [32]. Despite the greater number of rays species, most studies focus on sharks rather than on rays [33,34,35], especially abundant species and/or those affected by industrial fisheries, such as the Tiger Shark [36,37,38,39,40] and Blue Shark [41,42,43,44].
Guitarfish belonging to the Rhinobatidae family that is comprised of three genera, Acroteriobatus, Pseudobatos, and Rhinobatos [45,46]. These genera can be identified based on their external nasal flap morphology and molecular data [45], with Pseudobatos presenting a more restricted distribution compared to the others, occurring only in amphi-American regions.
The chola guitarfish, Pseudobatos percellens Walbaum, 1792, is distributed along the continental shelf of the western Atlantic Ocean, from Panama to Brazil [46,47,48,49]. Marine environments presenting different biological characteristics due to oceanographic, glacial-eustatic, biogeographic, and ecological factors are noted along its almost 9000 km distribution stretch [50,51,52]. Among the different types of barriers, the Amazon River discharge plume [53,54,55] and water temperature [56,57,58,59] are the most noteworthy, as they both alter the physical and chemical components of the North Brazil Shelf Region and can act as barriers for sister species, leading to speciation in different marine species [60,61,62,63].
Among the few available studies on P. percellens, Yokota and Lessa’s study [64] on shark and ray nursery regions in northeastern Brazil is highlighted, indicating separate ecological niches for sharks and spatial overlap with respect to the assessed species. In another study, Caltabellotta et al. [65] conducted an age and growth study, evidencing that parameters were markedly different between P. percellens males and females. Recently, Cruz et al. [66] identified that the P. percellens population in southern Brazil (São Paulo, Paraná, and Santa Catarina) displays low genetic diversity among specimens by employing a mitochondrial marker (Dloop).
Long-term population viability is associated with genetic variability levels [67], that is, if a certain population cannot evolve in response to environmental variable changes, it may face a greater risk of extinction. In addition, anthropic actions and coastal pollution are responsible for degrading habitats in certain distribution areas [68]. This has contributed to the current “Endangered” (EN) assessment of P. percellens as established by the International Union for Conservation of Nature (IUCN) Red List of Threatened Species [69] and “Vulnerable” (VU) according to the Brazilian National List of Threatened Species [70].
Despite its threatened status and presenting metal concentrations in their meat many times above safe limits for human consumption [71], P. percellens is fished commercially in Brazil by artisanal fleets or as non-targeted catch in the shrimp-trawling fishery [72,73,74], and its meat is marketed in several areas of the Brazilian coast [75,76,77,78,79,80].
In this context, this study aimed to understand the genetic diversity and population structure of P. percellens along the Brazilian coast using single nucleotide polymorphisms (SNPs) by double-digest restriction site associated DNA (ddRAD) sequencing as genetic markers. We first questioned what are the genetic variation levels among specimens from the northern, northeastern, and southeastern/southern coast of Brazil? Second, how is genetic variation structured across populations on a local and regional scale? Third, do northern, northeastern and southeastern/southern populations still maintain gene flow? Finally, can biogeographic Brazilian coast barriers influence the distribution of this species? Genetic diversity and biological factors which may influence P. percellens distribution are discussed, testing the hypothesis of the existence of a single lineage throughout the species distribution range along the Brazilian coast.

2. Material and Methods

2.1. Sample Collection and DNA Barcoding

Total DNA from 52 Pseudobatos percellens samples were obtained from five locations in northern and northeastern Brazil, namely in Amapa (AP, N = 03), Pará (PA, N = 05), and Pernambuco (PE, N = 14), and from the southeastern/southern regions, in São Paulo (SP, N = 15) and Paraná (PR, N = 15) (Supplementary Table S1). Sampling was permitted by capture permit no. 13843-1 issued by SISBIO/ICMBio, Brazil. Tissues of all individuals were deposited at the Laboratory of Fish Biology and Genetics fish collection, UNESP in Botucatu, São Paulo, Brazil. The DNA samples were extracted from muscle tissue fragments preserved in 95% ethanol using the Wizard ® Genomic DNA Purification kit (Promega, Madison, WI, USA), following the manufacturer’s instructions.
The COI barcode region was amplified according to Hebert et al. [81] for DNA barcode analyses of each sample. PCR amplicons were visualized on a 1% agarose gel E-Gels (Invitrogen) and bi-directionally sequenced using the BigDye Terminator v.3.1 Cycle Sequencing Kit (Applied Biosystems, Inc.) on an ABI 3130 capillary sequencer, following the manufacturer’s instructions. Sequences were aligned in Geneious 4.8.5 [82]. Aligned consensus sequences were compared with those deposited in the National Center for Biotechnology Information (NCBI) database (http://www.ncbi.nlm.nih.gov/, accessed on 27 August 2022) using the Basic Local Alignment Search Tool—Nucleotide (BLASTn).
Genetic distances were obtained by the Kimura-2-Parameter Model (K2P) method [83]. The final matrix contained 126 sequences, comprising 40 obtained in the present study and 86 extracted from GenBank (ncbi.nlm.nih.gov/genbank, accessed on 3 January 2021). A maximum likelihood phylogenetic reconstruction was applied to construct a tree from the pairwise distances, estimated using the Tamura and Nei [84] substitution model, as implemented in MEGA X [85]. The phylogenetic tree was tested by the bootstrap method with 1000 pseudoreplicates [86] and all analyzed sequences were submitted to GenBank (Accession Nos. OQ255948–OQ255987).

2.2. ddRAD Library Preparation

The SNP dataset was generated using a double digest restriction site associated DNA sequencing (ddRADseq) method following Peterson et al. [87], with modifications as reported by Campos et al. [88]. Briefly, we used 34 μL of a 200 ng/μL DNA solution and 1 μL of each of the restriction enzymes EcoRI (20 U/μL) and MspI (10 U/μL) (New England Biolabs, NEB) and 4 μL of the TANGO buffer, in a total volume of 40 μL at this stage, for sample genomic DNA digestion. A pair of customized adapters for each restriction enzyme was then applied, and the adapter P1 (3 nM—EcoRI) and adapter P2 (6 nM—MSPI) were bound to 31.5 μL of the digestion product. The ligation reaction of the adapters was developed by using 2 μL of the adapter for each of the enzymes, 4 μL of T4 Ligase Buffer 1X (Promega), and 0.5 μL of the enzyme T4 Ligase (Promega), with a final volume of 40 μL. The samples were finally incubated at 23 °C for 30 min, 65 °C for 10 min, and 63°C for 90 s, and then the temperature was reduced by 2 °C every 90 s until 23 °C. The ligation reaction samples were then purified.
The indexing reaction was performed with the insertion of the complement sequence of Nextera® Index Primers (Illumina, San Diego, CA, USA) S500 and N700 (Nextera DNA CD Indexes—96 indices, 96 samples) in each sample, to perform indexing with the Nextera® DNA Sample Preparation Kit (Illumina). The indexing reaction contained 15 μL of the ligation product and 5 μL of the S500 and N700 indices, and each sample contained a unique index combination with 25 μL of Phusion High-Fidelity PCR Master Mix (Thermo Scientific) for a final volume of 50 μL. The indexing reaction was followed by consecutive thermocycler steps, with an initial step at 72 °C for 3 min followed by two denaturation steps at 95 °C for 30 s, followed by 16 cycles at 95 °C for 30 s, an annealing step at 55 °C for 30 s, an extension step at 72 °C for 30 s, and a final extension step at 72 °C for 5 min, remaining at 4 °C to infinity.
Final sample concentration was standardized to a final volume of 10 ng/μL per sample and were then pooled and purified. After pool quantification, size selection was performed to select the size of the fragments of interest using the Wizard® SV Gel and PCR Clean-Up System kit (Promega, USA) in 1% agarose gels. The pool was applied to the gel, with the selection of the fragment of interest between 300 and 500 base pairs (bp), and the fragment was then purified. The library was quantified by real-time PCR (qPCR) to determine concentrations and sequenced with single-end reads of 150 bp in a NGS Illumina Nextseq500 platform.
After sequencing, the programs FastQC [89] and MultiQC [90] were used to assess the quality of the readings and adapter detection of the raw reads. Low quality base adapters and reads (quality score < 20) were removed using the TRIMMOMATIC v.o.32 Program [91]. In addition, in silico digestion was performed according to Driller et al. [92] and all reads were trimmed to 140 bp again using the TRIMMOMATIC v.o.32. The samples were then processed using the de novo pipeline from STACKS v2.55 [93] employing parameter values m = 3, M = 2, and n = 1. At the end, the STACKS population pipeline was used with the application of three SNP filters, the first to select SNPs that occurred in at least 80% of analyzed individuals (r = 0.80) per population, the second to exclude SNPs that presented a minor allele frequency (MAF) value <0.05, and the third, to exclude very high levels of heterozygosity, where SNPs with maximum observed heterozygosity greater than 0.65 were filtered. The plink v1.9 software [94] was used to filter loci that significantly (p < 0.05) deviated from the Hardy–Weinberg Equilibrium and were in linkage disequilibrium (R2 > 0.4).

2.3. Population Genomics Analysis

Private alleles were calculated using the Stacks population program. The ARLEQUIN v3.5.2.2 software [95] was used to estimate the observed heterozygosity (Ho), expected heterozygosity (He), inbreeding coefficient (FIS), and probability of deviation from the Hardy–Weinberg Equilibrium (HWE p-value), calculated by using an exact test. Deviations from the Hardy–Weinberg Equilibrium (HWE) in the datasets were assessed with the Global Bartlett tests, performed between observed and expected heterozygosity available in the ADEGENET v package. 2.1.1 of the R Program [96].
Genetic differences were analyzed using pairwise FST values calculated in the Arlequin v.3.5.2.2 software [95]. The number of existing clusters among P. percellens samples was investigated applying a Bayesian Analysis using the STRUCTURE Software [97], and the k-value was estimated considering k = 1 to k = 7, with 500,000 MCMC, burn-in of 10%, 10 independent runs per K and was performed using an admixture model, with the assumption of correlated frequencies. Isolation by distance (IBD) by a Mantel test, was tested by computing the regression of FST/1-FST on geographic distances and the level of significance determined by performing a test with ISOLDE in GENEPOP 4.2 [98] based on 1000 randomization.
Identification of the number of genetic clusters (k) was performed by the Puechmaille Method [99], using the Structure Selector [100]. Based on the number of detected structure clusters, the Fst pairwise was analyzed and a Discriminant Analysis of Principal Components (DAPC) was conducted using the Adegenet v.2.2.1 R package [96]. Clusters were generated using the find.clusters function, which uses the k-means algorithm to generate clusters for various k values, and the Bayesian Information Criterion (BIC) was then considered to determine the most likely k value.
For a better understanding of the P. percellens group results, dendrograms were generated applying two different approaches, the Maximum Likelihood (ML) and the Bayesian Inference (BI). The ML approach was performed using the RAxML 8.2.8 software [101] applying the rapid bootstrapping option with 1000 replicates. Since the general time reversible model and its variants are the only available options in RAxML, we replaced the best-fit models observed in the jModelTest by the GTR+Γ. In the same way, the BI analyses were performed in MrBayes 3.2 [102] using four simultaneous runs, each with four Markov chains (T = 0.2). The analyses began with randomly chosen trees, being extended through 50 million generations, with sampling every 5000 generations and using a burn-in fraction of 25 % of the trees. An explorative network analysis using the NeighborNet Algorithm implemented in Splitstree v.4.14 [103] was also performed.

3. Results

The forty specimens used in this study were molecularly identified by DNA barcoding to confirm their previous morphological identification. The obtained barcode sequences ranged in length from 510 to 644 bases pairs, revealing high similarity (>99%) with the P. percellens sequences deposited in Genbank. The interspecific genetic distances values ranged from 1.3% between P. percellens and Pseudobatos lentiginosus, to 21.4% in Pseudobatos horkelii and Rhinobatus rhinobatus (Table S2).
The Maximum Likelihood (ML) tree based on the DNA barcoding identified two main groups, one for Pseudobatos and the other for Rhinobatos. A first clade with P. percellens from the Brazilian coast was observed in the Pseudobatos group for the specimens used in this study, followed by a second clade containing P. lentiginosus individuals from Mexico, and finally, a third clade consisting of P. horkelii from Brazil (Figure S1).
After confirming specimen identification, a ddRAD library was developed for the 52 P. percellens samples, resulting in 86,319,471 raw data reads and ranging from 207,416 to 3,044,397 reads per sample (Table S1). After quality filtering, 70,757,837 reads were maintained, ranging from 175,933 to 2,578,145. All reads were standardized at 140 bp, and a total of 3329 high-quality SNPs were used for subsequent analyses after quality filtering, with only 36% of data missing. These SNPs were selected by loci genotyped in >75 % of the individuals and showing an MAF > 0.05, among which 1779 transitions and 1550 transversions were detected (Figure S2).
Observed heterozygosity (Ho) ranged from 0.281 (SP) to 0.577 (AP) among the five analyzed populations (Table 1), while expected heterozygosity (He) ranged from 0.239 (SP) to 0.5487 (AP). The differences between observed and expected heterozygosity variations in all sampling locations were statistically significant (Bartlett’s K-square = 573.47, df = 1, p-value < 2.2 × 10−16), where Ho (General Ho = 0.260) was statistically different from He (general He = 0.348), suggesting that the global population deviated from the HWE.
The average FIS values were negative in all the populations, ranging from −0.542 (AP) to −0.088 (PE) (Table 1). The average nucleotide diversity (π) value across all populations ranged from 0.031 (SP) to 0.115 (AP) (Table 1).
High levels of population genetic differentiation among geographical P. percellens sample regions were also detected. The highest mean FST values were found between Amapá and São Paulo (FST = 0.764), and the lowest between São Paulo and Paraná (FST = 0.043), indicating that P. percellens populations in the Brazilian coast are genetically different, with two extant lineages (Figure S3a, Table S3). The Mantel test demonstrated a correlation between geographic distance and fixation index values (R2 = 0.5904) with P = 0.001.
The STRUCTURE software analyses including all 52 samples best supported a value of k = 2 genetic clusters, based on the Evanno Method ((ln (P) D) 3886 ± 43,141) (Figure 1), which correspond to Amapá, Pará, and Pernambuco for Cluster 1 and São Paulo and Paraná for Cluster 2. The performed DAPC involving the analyzed geographical regions indicated a clear separation between Amapá, Pará, and Pernambuco, whereas a clear overlap between São Paulo and Paraná was observed (Figure 2).
The analyzed DAPC identified an optimum of k = 3 genetically distinct clusters (Figure 3a). All individuals displayed a 100 % membership probability to the group to which they were assigned. All clusters contained between 7 and 30 individuals, and some clusters spanned two geographical populations, with Cluster 1 including samples from Amapá and Pará, Cluster 2 with all samples from Pernambuco and one sample of Amapá, and Cluster 3 exclusively composed of all samples from São Paulo and Paraná, thus revealing a high genetic relatedness among those accessions (Figure 3b). The clustering was supported by results of the pairwise FST Analysis between pairs of clusters (Figure S3b). The highest FST values were detected between Cluster 1 and Cluster 3 (FST = 0.694) and between Cluster 2 and Cluster 3 (FST = 0.601), indicating a high genetic differentiation among clusters of accessions. The lowest FST was obtained between Clusters 1 and 2 (FST = 0.106).
The divMigrate Analysis unveiled a high gene flow in the southeastern region, between São Paulo and Paraná, and a weak gene flow in the northern and northeastern regions, between Amapá, Pará, and Pernambuco (Figure S4). In addition, a very limited and unexpected gene flow (<0.08) was found between the southeast and northern and northeastern regions.
For a more accurate investigation of the existence of two possible P. percellens lineages, the selection of SNPs that genotyped 100% of all individuals of the two lineages was performed, thus maintaining 1.448 SNPs. Both Bayesian Inference (BI) and Maximum Likelihood (ML) analyses supported the division of P. percellens into two groups representing the northern and northeastern (Amapá, Pará, and Pernambuco) and southeastern/southern regions (São Paulo and Paraná) (Figure 4A,B). The group I lineage included populations from Amapá, Pará, and Pernambuco, totaling 22 samples, while the group II lineage, with 30 samples, included populations from São Paulo and Paraná.

4. Discussion

This study was developed with SNP markers through ddRAD sequencing in individuals of five local chola guitarfish P. percellens populations, providing a new perspective into genetic differentiation patterns along the Brazilian coast for the first time. The analyses revealed at least one lineage occurring in the northern and northeastern region and another distributed along the southeastern/southern Brazilian coast, with high genetic differentiation between them. The DAPC analysis, however, pointed out the presence of three lineages distributed in these regions.
Pseudobatos percellens is an endemic species in the western Atlantic Ocean, occurring from Panamá to Brazil, and is generally found in sympatry with P. horkelli on the southeastern continental shelf from Rio de Janeiro to Mar del Plata (Argentina) [58,66,104,105]. Another species belonging to the same genus, P. lentiginosus Garman 1880, occurs in the western north Atlantic region, although reports of occurrence in the northernmost region of Brazil are noted [106], although they still requiring confirmation [58,107].
Due to the occurrence of sympatric species belonging to the same genus in the region, the specimens used in this study were previously analyzed by DNA barcoding and compared to P. lentiginosus and P. horkelli sequences and to species belonging to the Rhinobatos genus deposited in Genbank. The results confirmed that all the analyzed specimens corresponded to P. percellens, without any taxonomic conflict with other species described for the Pseudobatos genus. It is also important to note that the DNA barcode in fishes does not often reveal genetic peculiarities or lineages within groups, more specifically in elasmobranchs. For example, shark species belonging to the Squalus genus [108] or freshwater Potamotrygon stingrays [109] present a high taxonomic complexity that is not well elucidated by solely employing the COI gene, and requires the use of other complementary genetic markers.
The genome-wide SNPs data obtained herein when analyzing P. percellens indicated a high degree of genetic differentiation between specimens from the northern and northeastern regions in comparison to specimens from southeastern/southern Brazil. These results may partially reflect the biological behavior of P. percellens, as this species is known to perform seasonal migrations or short-distance movements [50], supported by the Mantel test and isolation by distance model of population structure. These movements over time could be reflected in the complex speciation nature of this group, which ranges from the presence of continuous intragroup variations without reproductive isolation, to complete and irreversible reproductive isolation between groups [110]. Nevertheless, the extent of the effects of these seasonal movements on a long-term evolutionary time scale remain uncertain [110].

Factors Influencing Dispersal: Biogeography Barriers along the Brazilian Coast

The pairwise FST estimations and DAPC plots implied that the São Paulo and Paraná groups exhibit higher genetic similarity than evidenced between Amapá, Pará, and Pernambuco samples, revealing genome differences in these regions. These results were somehow expected, as the sampled regions in São Paulo (Santos/SP) and Paraná (Pontal do Paraná/PR) are geographically close (<270 km). In addition, these regions share similar environmental conditions, such as salinity and sediments and are located in or close to the Cananéia–Iguape–Paranaguá Estuarine Lagoon Complex [111].
Fish are well known for their phenotypic plasticity related to specific environmental changes, which results in thermal performance curve shifts with chronic changes in temperature, potentially leading to a compensation for negative temperature effects [112]. Temperature differences have been identified as an important biogeographic barrier (thermal insulation) for several marine fish species, even more so as the north coast region of Brazil is home to more temperate waters when compared to the south region [56,57,113,114,115,116,117,118].
Temperature is directly associated with Elasmobranchii biodiversity in the American continent, with the greatest richness observed in the Atlantic Tropical Zone, between the tropics of Cancer and Capricorn [119]. Considering latitudinal and vertical temperature water column variations, species with wide spatial distributions present greater physiological tolerance and, thus, manage to overcome thermal barriers, while restricted species generally display low adaptability to variations [52,120]. This particularity has been observed for different shark and ray species in the southwestern Atlantic [56].
Environmental temperature variation effects may be considered with regards to the distribution patterns of P. horkelii, an endemic ray species that occurs in the Argentine Zoogeographic Province, although it has also been recorded in northeastern Brazil [56,57,58,121]. Therefore, temperature differences may play an important role in separating the two genetic lineages identified herein in P. percellens. The results indicate that the lineages belonging to group I may be more adapted to the warmer tropical waters found in the north and northeast coastal Brazilian regions, while individuals from group II seem to be found in cooler water temperatures, in south and southeast Brazil.
The sampled Pará and Amapá regions are relatively close and located in neighboring states, while Pernambuco is located at a greater geographical distance (over 1200 km). Regardless of the geographic distance between the sampled locations, the northern and northeastern Brazil regions are considered more heterogeneous environments, due to unique characteristics as a result of the Plume in the Amazon and Orinoco River (PAOR) [53,54]. This specific environment is considered an important haline barrier for many local coastal fishes [55,60,61,63,122,123].
The results evidencing the structuring of P. percellens populations on both sides of the Amazon Estuary, that is, between Amapá/Pará and Pernambuco, may suggest that this haline barrier is unlikely to be responsible for the differentiation observed herein. In addition, the simple presence of the population located in northern Brazil under the direct influence of the Amazon River discharge and partially within the PAOR demonstrated that this species can withstand salinity changes, which is also confirmed by P. percellens records in estuarine areas throughout its distribution range [68,124,125,126,127].
Ocean currents may also influence the distribution of these specimens throughout the Brazilian coast and contribute to the separation of the two lineages. In this regard, the southern Equatorial Current (SEC) can be strongly considered among the acting factors in the region, contributing to species dispersion. When the SEC reaches the South American continent, it splits into the North Brazilian Current (NBC—northwards) and the Brazilian Current (BC—southwards) [128]. The NBC is an intense western current that curves back on itself after passing the mouth of the Amazon River [129]. The BC is a weak western boundary current carrying warm subtropical water, flowing south along the Brazilian coast, and plays an important role in determining water circulation in the southern Bay of Brazil (SBB). This biogeographical region comprises the area between Cabo Frio (Rio de Janeiro/RJ) and Cabo de Santa Marta (Santa Catarina/SC), where the BC meets the Falklands/Malvinas Cold Current (FMCC), a branch of the Antarctic Circumpolar Current at the Atlantic Subtropical Convergence (ASC), which limits the Argentine Zoogeographic Province (AZP) [130,131,132].
At the Vitória-Trindade Seamount Chain region, the BC is also influenced by the Central Water of the South Atlantic (CWSA), penetrating over the shelf, reaching coastal areas with a thermocline between 10 and 15 m, retracting during the winter, and consequently being replaced by typically tropical waters. This region thus represents an important faunal transition area, comprising the southern limit of occurrence of several tropical fish species and the northern limit of species from temperate regions, thus leading to population or lineage differentiation, as reported herein and also detected for Sardinella brasiliensis [54,113,114,130,133].

5. Conclusions

Pseudobatos is a Rhinobatidae family genus of rays comprising only nine species [134], together presenting an amphi-American distribution. Pseudobatos percellens presents the widest distribution area in the Atlantic Ocean, including individuals inhabiting the maximum recorded depth of 162 m, almost twice the area of the congeneric species P. horkelii and P. lentiginosus in this ocean [58]. This wide P. percellens distribution differs from the patterns observed in other species belonging to this family, which generally present a more restricted distribution pattern. This may be considered an indication of the existence of cryptic undescribed species in this group [107], corroborating the results reported in the present study.
Our findings emphasize the need for studies on the regional biodiversity of this group of fishes, in order to understand different distribution patterns and their limiting factors, especially in species presenting active dispersion processes, as noted for Chondrichthyes, carried out mainly during adult life stages [46,52,116,135,136,137]. In this context, the knowledge of biogeographic species limits and the distribution of their respective populations, as well as their genetic diversity patterns and connectivity, provide data to define management and conservation priority areas. This is especially important when species are distributed in areas comprising various management authorities and threat levels of connectivity, as in the case of P. percellens, P. horkelii, and P. lentiginosus [52,56,136].
Moreover, the low gene flow between components belonging to group I found in north and northeast Brazil, and to group II, which inhabit the southeastern/southern regions, is noteworthy, due to several aforementioned factors, and may prevent the occurrence of shared haplotypes. In this case, the genetic drift could lead to genetic differences between individuals from these regions, leading to reciprocal lineages and ultimately, speciation, as pointed by Avise [138].
Finally, the findings reported herein revealed a hidden biodiversity within P. percellens and support the need for a taxonomic revision of this group, with the recommendation of including areas not sampled in the present study. In addition, the identification of a hidden diversity within P. percellens should be used to improve the management of the Neotropical ichthyofauna that inhabits modified and degraded regions impacted by pollution and habitat fragmentation and are subject to fisheries pressure. Therefore, considering our results employing genome-wide SNP data, two distinct lineages were identified for P. percellens on the Brazilian coast that should, therefore, be managed separately to ensure adequate biodiversity conservation.

Supplementary Materials

The following supporting information can be downloaded at: https://www.mdpi.com/article/10.3390/d15030344/s1, Figure S1: The Maximum Likelihood tree of the Pseudobatos and Rhinobatos specimens, based on mitochondrial cytochrome c oxidase subunit I gene sequences under the K2P Model. The asterisks* represent the 40 samples sequenced in this study; Figure S2: Identification of 3.329 SNPS (in blue) in transitions (in grey) and transversions (in red) characterized in Pseudobatos percellens; Figure S3: Heatmap depicting the pairwise Fst Distance matrices using the 52 Pseudobatos percellens sample dataset with 3.329 SNPs, between geographical populations (a) and based on clustering (b). Non-significant p-value >0.05 represented by x. AP, Amapá; PA, Pará; PE, Pernambuco; SP, São Paulo; PR, Paraná; Figure S4: Contemporary gene flow based on number of migrants (Nm), illustrating relationships based on geographic distribution for the P. percellens samples. Arrows indicate the migration direction, while stronger gene flows are represented by larger numbers, as well as thicker and darker colored lines. Fine and clear lines represent limited gene flow with a coefficient < 0.3. AP, Amapá; PA, Pará; PE, Pernambuco; SP, São Paulo; PR, Paraná. Table S1: Summary of the sequencing and processing of 52 Pseudobatos percellens samples. Asterisks* represent samples not sequenced in the DNA barcoding analysis; Table S2: Genetic distances (Kimura-2-Parameter—K2P) based on COI sequences among Pseudobatos and Rhinobatos species (below the diagonal) and standard errors (above the diagonal). Numbers in bold represent the intraspecific K2P genetic distances; Table S3: Matrix of Pairwise FST (below the diagonal) and p-values (above the diagonal divide) among geographical state of Pseudobatos percellens populations based on data from 3.329 SNPs.

Author Contributions

Conceptualization, V.P.C., M.M.R., B.C.S., C.O., R.L. and F.F.; methodology, V.P.C., B.R.B. and B.C.S.; validation, V.P.C. and B.R.B.; formal analysis, V.P.C., B.R.B., B.C.S. and N.N.C.D.C.; investigation, V.P.C., M.M.R., P.C. and B.C.S.; resources, V.P.C., C.O. and F.F.; writing—original draft preparation, V.P.C., M.M.R., C.O., R.L. and F.F.; writing—review and editing, P.C., C.O., R.L. and F.F.; supervision, F.F; project administration C.O., R.L. and F.F.; funding acquisition, F.F. All authors have read and agreed to the published version of the manuscript.

Funding

This research was funded by the Conselho Nacional de Desenvolvimento Científico e Tecnológico (CNPq, grants to V.P.C., 107761/2019-0 to A.A.A.). C.O. received financial support from the Fundação de Amparo à Pesquisa do Estado de São Paulo (FAPESP grants 2018/20610-1, 2016/09204-6, 2014/26508-3), and Conselho Nacional de Desenvolvimento Científico e Tecnológico (CNPq proc. 306054/2006-0 (C.O.)).

Institutional Review Board Statement

All samples were collected in strict accordance with the Brazilian Federal Animal Ethics Committee (SISBIO 13843–1) regulations, and all analyses followed the International Guidelines for Animal Experiments, as authorized by CEEAA IBB/UNESP, protocol number 556.

Informed Consent Statement

Not applicable.

Data Availability Statement

The data presented in the present study was deposited and made openly available through the GenBank genetic sequence database, accession Nos. OQ255948–OQ255987).

Acknowledgments

We would like to thank the UNESP Biosciences Institute for infrastructure use. This study was supported by grants from the Conselho Nacional de Desenvolvimento Científico e Tecnológico (CNPq, grants to V.P.C., 107761/2019-0). C.O. received financial support from the Fundação de Amparo à Pesquisa do Estado de São Paulo (FAPESP grants 2018/20610-1, 2016/09204-6), and Conselho Nacional de Desenvolvimento Científico e Tecnológico (CNPq proc. 306054/2006-0 (C.O.)).

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Yan, H.F.; Kyne, P.M.; Jabado, R.W.; Leeney, R.H.; Davidson, L.N.; Derrick, D.H.; Finucci, B.; Freckleton, R.P.; Fordham, S.V.; Dulvy, N.K. Overfishing and habitat loss drive range contraction of iconic marine fishes to near extinction. Sci. Adv. 2021, 7, eabb6026. [Google Scholar] [CrossRef] [PubMed]
  2. Kappel, C.V. Losing pieces of the puzzle: Threats to marine, estuarine, and diadromous species. Front. Ecol. Environ. 2005, 3, 275–282. [Google Scholar] [CrossRef]
  3. Rosser, A.M.; Mainka, S.A. Overexploitation and species extinctions. Conserv. Biol. 2002, 16, 584–586. [Google Scholar] [CrossRef]
  4. Sala, O.E.; Stuart Chapin, F.I.; Armesto, J.J.; Berlow, E.; Bloomfield, J.; Dirzo, R.; Huber-Sanwald, E.; Huenneke, L.F.; Jackson, R.B.; Kinzig, A.; et al. Global biodiversity scenarios for the year 2100. Science 2000, 287, 1770–1774. [Google Scholar] [CrossRef]
  5. Tilman, D.; Clark, M.; Williams, D.R.; Kimmel, K.; Polasky, S.; Packer, C. Future threats to biodiversity and pathways to their prevention. Nature 2017, 546, 73–81. [Google Scholar] [CrossRef]
  6. Dulvy, N.K.; Pacoureau, N.; Rigby, C.L.; Pollom, R.A.; Jabado, R.W.; Ebert, D.A.; Simpfendorfer, C.A. Overfishing drives over one-third of all sharks and rays toward a global extinction crisis. Curr. Biol. 2021, 31, 4773–4787. [Google Scholar] [CrossRef]
  7. Díaz, S.; Zafra-Calvo, N.; Purvis, A.; Verburg, P.H.; Obura, D.; Leadley, P.; Zanne, A.E. Set ambitious goals for biodiversity and sustainability. Science 2020, 370, 411–413. [Google Scholar] [CrossRef]
  8. Hooper, D.U.; Adair, E.C.; Cardinale, B.J.; Byrnes, J.E.; Hungate, B.A.; Matulich, K.L.; O’Connor, M.I. A global synthesis reveals biodiversity loss as a major driver of ecosystem change. Nature 2012, 486, 105–108. [Google Scholar] [CrossRef]
  9. Pasquaud, S.; Vasconcelos, R.P.; França, S.; Henriques, S.; Costa, M.J.; Cabral, H. Worldwide patterns of fish biodiversity in estuaries: Effect of global vs. local factors. Estuar. Coast. Shelf Sci. 2015, 154, 122–128. [Google Scholar] [CrossRef]
  10. Habibullah, M.S.; Din, B.H.; Tan, S.H.; Zahid, H. Impact of climate change on biodiversity loss: Global evidence. Environ. Sci. Pollut. Res. 2022, 29, 1073–1086. [Google Scholar] [CrossRef]
  11. Reader, M.O.; Eppinga, M.B.; de Boer, H.J.; Damm, A.; Petchey, O.L.; Santos, M.J. Biodiversity mediates relationships between anthropogenic drivers and ecosystem services across global mountain, island and delta systems. Glob. Environ. Change 2023, 78, 102612. [Google Scholar] [CrossRef]
  12. Liu, J.; Slik, F.; Zheng, S.; Lindenmayer, D.B. Undescribed species have higher extinction risk than known species. Conserv. Lett. 2022, 15, e12876. [Google Scholar] [CrossRef]
  13. Chowdhury, S. Threatened species could be more vulnerable to climate change in tropical countries. Sci. Total Environ. 2023, 858, 159989. [Google Scholar] [CrossRef] [PubMed]
  14. Bertorelle, G.; Bruford, M.W.; Hauffe, H.C.; Rizzoli, A.; Vernesi, C. Population Genetics for Animal Conservation; United States of America by Cambridge University Press: New York, NY, USA, 2009. [Google Scholar]
  15. Allendorf, F.W. Genetics and the conservation of natural populations: Allozymes to genomes. Mol. Ecol. 2017, 26, 420–430. [Google Scholar] [CrossRef] [Green Version]
  16. Hanson, J.O.; Rhodes, J.R.; Butchart, S.H.; Buchanan, G.M.; Rondinini, C.; Ficetola, G.F.; Fuller, R.A. Global conservation of species’ niches. Nature 2020, 580, 232–234. [Google Scholar] [CrossRef]
  17. De Kort, H.; Prunier, J.G.; Ducatez, S.; Honnay, O.; Baguette, M.; Stevens, V.M.; Blanchet, S. Life history, climate and biogeography interactively affect worldwide genetic diversity of plant and animal populations. Nat. Commun. 2021, 12, 516. [Google Scholar] [CrossRef] [PubMed]
  18. Stronen, A.V.; Norman, A.J.; Vander Wal, E.; Paquet, P.C. The relevance of genetic structure in ecotype designation and conservation management. Evol. Appl. 2022, 15, 185–202. [Google Scholar] [CrossRef] [PubMed]
  19. DeSalle, R.; Amato, G. The Expansion of Conservation Genetics. In Conservation Genetics in the Age of Genomics; Amato, G., DeSalle, R., Ryder, O., Rosenbaum, H., Eds.; Publisher Columbia University Press: New York, NY, USA, 2009; Volume 1, pp. 1–24. [Google Scholar] [CrossRef]
  20. Frankham, R. Challenges and opportunities of genetic approaches to biological conservation. Biol. Conserv. 2010, 143, 1919–1927. [Google Scholar] [CrossRef]
  21. Frankham, R.; Ballou, J.D.; Dudash, M.R.; Eldridge, M.D.; Fenster, C.B.; Lacy, R.C.; Ryder, O.A. Implications of different species concepts for conserving biodiversity. Biol. Conserv. 2012, 153, 25–31. [Google Scholar] [CrossRef]
  22. Bernatchez, L. On the maintenance of genetic variation and adaptation to environmental change: Considerations from population genomics in fishes. J. Fish Biol. 2016, 89, 2519–2556. [Google Scholar] [CrossRef]
  23. Coates, D.J.; Byrne, M.; Moritz, C. Genetic diversity and conservation units: Dealing with the species-population continuum in the age of genomics. Front. Ecol. Evol. 2018, 6, 165. [Google Scholar] [CrossRef] [Green Version]
  24. Willi, Y.; Kristensen, T.N.; Sgrò, C.M.; Weeks, A.R.; Ørsted, M.; Hoffmann, A.A. Conservation genetics as a management tool: The five best-supported paradigms to assist the management of threatened species. Proc. Natl. Acad. Sci. USA 2022, 119, e2105076119. [Google Scholar] [CrossRef]
  25. Cortés, E. Life history patterns and correlations in sharks. Rev. Fish. Sci. 2000, 8, 299–344. [Google Scholar] [CrossRef]
  26. Stevens, J.D.; Bonfil, R.; Dulvy, N.K.; Walker, P.A. The effects of fishing on sharks, rays, and chimaeras (chondrichthyans), and the implications for marine ecosystems. ICES J. Mar. Sci. 2000, 57, 476–494. [Google Scholar] [CrossRef]
  27. Whitenack, L.B.; Kim, S.L.; Sibert, E.C. Bridging the gap between chondrichthyan paleobiology and biology. In Biology of Sharks and Their Relatives, 3rd ed.; Carrier, J.C., Simpfendorfer, C.A., Heithaus, M.R., Yopak, K.E., Eds.; CRC Press: Boca Raton, FL, USA, 2022; pp. 1–29. [Google Scholar]
  28. Ferretti, F.; Worm, B.; Britten, G.L.; Heithaus, M.R.; Lotze, H.K. Patterns and ecosystem consequences of shark declines in the ocean. Ecol. Lett. 2010, 13, 1055–1071. [Google Scholar] [CrossRef] [PubMed]
  29. Navia, A.F.; Cortés, E.; Mejía-Falla, P.A. Topological analysis of the ecological importance of elasmobranch fishes: A food web study on the Gulf of Tortugas, Colombia. Ecol. Modell. 2010, 221, 2918–2926. [Google Scholar] [CrossRef]
  30. Bornatowski, H.; Navia, A.F.; Braga, R.R.; Abilhoa, V.; Corrêa, M.F.M. Ecological importance of sharks and rays in a structural foodweb analysis in southern Brazil. ICES J. Mar. Sci. 2014, 71, 1586–1592. [Google Scholar] [CrossRef] [Green Version]
  31. Rupp, A.; Bornatowski, H. Food web model to assess the fishing impacts and ecological role of elasmobranchs in a coastal ecosystem of Southern Brazil. Environ. Biol. Fishes 2021, 104, 905–921. [Google Scholar] [CrossRef]
  32. Froese, R.; Pauly, D. Editors. FishBase. World Wide Web Electronic Publication. 2022. Available online: www.fishbase.org,version (accessed on 1 August 2022).
  33. Castillo-Páez, A.; Sosa-Nishizaki, O.; Sandoval-Castillo, J.; Galván-Magaña, F.; Blanco-Parra, M.D.P.; Rocha-Olivares, A. Strong population structure and shallow mitochondrial phylogeny in the banded guitarfish, Zapteryx exasperata (Jordan y Gilbert, 1880), from the northern Mexican Pacific. J. Hered. 2014, 105, 91–100. [Google Scholar] [CrossRef] [Green Version]
  34. Dulvy, N.K.; Fowler, S.L.; Musick, J.A.; Cavanagh, R.D.; Kyne, P.M.; Harrison, L.R.; White, W.T. Extinction risk and conservation of the world’s sharks and rays. Elife 2014, 3, e00590. [Google Scholar] [CrossRef] [Green Version]
  35. Shiffman, D.S.; Bittick, S.J.; Cashion, M.S.; Colla, S.R.; Coristine, L.E.; Derrick, D.H.; Dulvy, N.K. Inaccurate and biased global media coverage underlies public misunderstanding of shark conservation threats and solutions. Iscience 2020, 23, 101205. [Google Scholar] [CrossRef]
  36. Cardeñosa, D.; Hyde, J.; Caballero, S. Genetic diversity and population structure of the pelagic thresher shark (Alopias pelagicus) in the Pacific Ocean: Evidence for two evolutionarily significant units. PLoS ONE 2014, 9, e110193. [Google Scholar] [CrossRef] [Green Version]
  37. Carmo, C.B.; Ferrette, B.L.; Camargo, S.M.; Roxo, F.F.; Coelho, R.; Garla, R.C.; Mendonça, F.F. A new map of the tiger shark (Galeocerdo cuvier) genetic population structure in the western Atlantic Ocean: Hypothesis of an equatorial convergence centre. Aquat. Conserv. 2019, 29, 760–772. [Google Scholar] [CrossRef]
  38. Bargnesi, F.; Lucrezi, S.; Ferretti, F. Opportunities from citizen science for shark conservation, with a focus on the Mediterranean Sea. Eur. Zool. J. 2020, 87, 20–34. [Google Scholar] [CrossRef] [Green Version]
  39. Bernard, A.M.; Finnegan, K.A.; Pavinski Bitar, P.; Stanhope, M.J.; Shivji, M.S. Genomic assessment of global population structure in a highly migratory and habitat versatile apex predator, the tiger shark (Galeocerdo cuvier). J. Hered. 2021, 112, 497–507. [Google Scholar] [CrossRef]
  40. Hammerschlag, N.; McDonnell, L.H.; Rider, M.J.; Street, G.M.; Hazen, E.L.; Natanson, L.J.; Kirtman, B. Ocean warming alters the distributional range, migratory timing, and spatial protections of an apex predator, the tiger shark (Galeocerdo cuvier). Glob. Change Biol. 2022, 28, 1990–2005. [Google Scholar] [CrossRef] [PubMed]
  41. Veríssimo, A.; Sampaio, Í.; McDowell, J.R.; Alexandrino, P.; Mucientes, G.; Queiroz, N.; Noble, L.R. World without borders—Genetic population structure of a highly migratory marine predator, the blue shark (Prionace glauca). Ecol. Evol. 2017, 7, 4768–4781. [Google Scholar] [CrossRef] [Green Version]
  42. Bailleul, D.; Mackenzie, A.; Sacchi, O.; Poisson, F.; Bierne, N.; Arnaud-Haond, S. Large-scale genetic panmixia in the blue shark (Prionace glauca): A single worldwide population, or a genetic lag-time effect of the “grey zone” of differentiation? Evol. Appl. 2018, 11, 614–630. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  43. Boldrocchi, G.; Storai, T. Data-mining social media platforms highlights conservation action for the Mediterranean Critically Endangered blue shark Prionace glauca. Aquat. Conserv. 2021, 31, 3087–3099. [Google Scholar] [CrossRef]
  44. Godínez-Padilla, C.J.; Castillo-Géniz, J.L.; Hernandez de la Torre, B.; González-Ania, L.V.; Román-Verdesoto, M.H. Marine-climate interactions with the blue shark (Prionace glauca) catches in the western coast of Baja California Peninsula, Mexico. Fish. Oceanogr. 2022, 31, 291–318. [Google Scholar] [CrossRef]
  45. Last, P.R.; Henderson, A.C.; Naylor, G.J. Acroteriobatus omanensis (Batoidea: Rhinobatidae), a new guitarfish from the Gulf of Oman. Zootaxa 2016, 4144, 276–286. [Google Scholar] [CrossRef] [PubMed]
  46. Last, P.R.; Seret, B.; Naylor, G.J. A new species of guitarfish, Rhinobatos borneensis sp. nov. with a redefinition of the family-level classification in the order Rhinopristiformes (Chondrichthyes: Batoidea). Zootaxa 2016, 4117, 451–475. [Google Scholar] [CrossRef] [PubMed]
  47. Figueiredo, J.L. Manual dos Peixes Marinhos do Sudeste do Brasil. I. Introdução. Cações, Raias e Quimeras. Museu de Zoologia; Universidade de São Paulo: São Paulo, Brasil, 1977; p. 104. [Google Scholar]
  48. Cervigón, F.; Alcalá, A. Los peces marinos de Venezuela, Vol. I. Segunda edición; Fundación Museo del Mar: Caracas, Venezuela, 1999; p. 425. [Google Scholar]
  49. McEachran, J.D.; de Carvalho, M.R. Batoid fishes. In The Living Marine Resources of the Western Central Atlantic; Kent E. Carpenter: Rome, Italy, 2002. [Google Scholar]
  50. Rocha, F.; Gadig, O.B.F. Reproductive biology of the guitarfish Rhinobatos percellens (Chondrichthyes, Rhinobatidae) from the São Paulo Coast, Brazil, western South Atlantic Ocean. J. Fish Biol. 2013, 82, 306–317. [Google Scholar] [CrossRef]
  51. Pinheiro, H.T.; Bernardi, G.; Simon, T.; Joyeux, J.C.; Macieira, R.M.; Gasparini, J.L.; Rocha, L.A. Island biogeography of marine organisms. Nature 2017, 549, 82–85. [Google Scholar] [CrossRef]
  52. Hirschfeld, M.; Dudgeon, C.; Sheaves, M.; Barnett, A. Barriers in a sea of elasmobranchs: From fishing for populations to testing hypotheses in population genetics. Glob. Ecol. Biogeogr. 2021, 30, 2147–2163. [Google Scholar] [CrossRef]
  53. Sullivan-Sealy, K.; Bustamante, G. Setting Geographic Priorities for Marine Conservation in Latin America and the Caribbean; The Nature Conservancy: Arlington, VA, USA, 1999; p. 125. [Google Scholar]
  54. Miloslavich, P.; Klein, E.; Díaz, J.M.; Hernandez, C.E.; Bigatti, G.; Campos, L.; Martín, A. Marine biodiversity in the Atlantic and Pacific coasts of South America: Knowledge and gaps. PLoS ONE 2011, 6, e14631. [Google Scholar] [CrossRef] [Green Version]
  55. Giachini-Tosetto, E.; Bertrand, A.; Neumann-Leitão, S.; Nogueira Júnior, M. The Amazon River plume, a barrier to animal dispersal in the Western Tropical Atlantic. Sci. Rep. 2022, 12, 537. [Google Scholar] [CrossRef]
  56. Menni, R.C.; Jaureguizar, A.J.; Stehmann, M.F.; Lucifora, L.O. Marine biodiversity at the community level: Zoogeography of sharks, skates, rays and chimaeras in the southwestern Atlantic. Biodivers. Conserv. 2010, 19, 775–796. [Google Scholar] [CrossRef]
  57. Caires, S.A.; Wuddivira, M.N.; Bekele, I. Assessing the Temporal Stability of Spatial Patterns of Soil Apparent Electrical Conductivity using Geophysical Methods; International Agrophysics: Lublin, Poland, 2014; Volume 28, pp. 423–433. [Google Scholar] [CrossRef]
  58. Séret, B.; Last, P.R.; Naylor, G.J.P. Guitarfishes. Family Rhinobatidae. Rays of the World; Cornell University Press: Ithaca, NY, USA, 2016; pp. 77–109. [Google Scholar]
  59. Morley, S.A.; Abele, D.; Barnes, D.K.; Cárdenas, C.A.; Cotté, C.; Gutt, J.; Henley, S.F.; Höfer, J.; Hughes, K.A.; Martin, S.M.; et al. Global drivers on Southern Ocean ecosystems: Changing physical environments and anthropogenic pressures in an earth system. Front. Mar. Sci. 2020, 7, 547188. [Google Scholar] [CrossRef]
  60. Rocha, L.A. Patterns of distribution and processes of speciation in Brazilian reef fishes. J. Biogeogr. 2003, 30, 1161–1171. [Google Scholar] [CrossRef] [Green Version]
  61. Floeter, S.R.; Rocha, L.A.; Robertson, D.R.; Joyeux, J.C.; Smith-Vaniz, W.F.; Wirtz, P.; Bernardi, G. Atlantic reef fish biogeography and evolution. J. Biogeogr. 2008, 35, 22–47. [Google Scholar] [CrossRef] [Green Version]
  62. Rodríguez-Rey, G.T.; Carvalho Filho, A.; De Araújo, M.E.; Solé-Cava, A.M. Evolutionary history of Bathygobius (Perciformes: Gobiidae) in the Atlantic biogeographic provinces: A new endemic species and old mitochondrial lineages. Zool. J. Linn. Soc. 2018, 182, 360–384. [Google Scholar] [CrossRef]
  63. Araujo, G.S.; Rocha, L.A.; Lastrucci, N.S.; Luiz, O.J.; Di Dario, F.; Floeter, S.R. The Amazon-Orinoco Barrier as a driver of reef-fish speciation in the Western Atlantic through time. J. Biogeogr. 2022, 49, 1407–1419. [Google Scholar] [CrossRef]
  64. Yokota, L.; Lessa, R.P. A nursery area for sharks and rays in Northeastern Brazil. Environ. Biol. Fishes 1998, 75, 349–360. [Google Scholar] [CrossRef]
  65. Caltabellotta, F.P.; Siders, Z.A.; Murie, D.J.; Motta, F.S.; Cailliet, G.M.; Gadig, O.B. Age and growth of thre endemic threatened guitarfishes Pseudobatos horkelii, P. percellens and Zapteryx brevirostris in the western South Atlantic Ocean. J. Fish Biol. 2019, 95, 1236–1248. [Google Scholar] [CrossRef] [PubMed]
  66. Cruz, V.P.; Adachi, A.M.C.L.; Oliveira, P.H.; Ribeiro, G.S.; Paim, F.G.; Souza, B.C.; Foresti, F. Genetic diversity in two threatened species of guitarfish (Elasmobranchii: Rhinobatidae) from the Brazilian and Argentinian coasts: An alert for conservation. Neotrop. Ichthyol. 2021, 19, 1–15. [Google Scholar] [CrossRef]
  67. Barrett, S.C.; Kohn, J.R. Genetic and evolutionary consequences of small population size in plants: Implications for Conservation. In Genetics and Conservation of Rare Plants; Oxford University Press: New York, NY, USA, 1991; pp. 3–30. [Google Scholar]
  68. Gonçalves-Silva, F.; Vianna, M. Use of a species-rich and degraded tropical estuary by Elasmobranchs. Braz. J. Oceanogr. 2018, 66, 339–346. [Google Scholar] [CrossRef]
  69. Pollom, R.; Charvet, P.; Avalos, C.; Blanco-Parra, M.P.; Derrick, D.; Espinoza, E.; Faria, V.; Herman, K.; Mejía-Falla, P.A.; Motta, F.; et al. Pseudobatos percellens. The IUCN Red List of Threatened Species 2020-1.2020. E.T161373A887217. Available online: https://www.iucnredlist.org (accessed on 21 March 2021).
  70. Instituto Chico Mendes de Conservação da Biodiversidade, ICMBio. 2022. Available online: https://icmbio.gov.br (accessed on 18 October 2022).
  71. Martins, M.F.; Costa, P.G.; Gadig, O.B.; Bianchini, A. Metal contamination in threatened elasmobranchs from an impacted urban coast. Sci. Total Environ. 2021, 757, 143803. [Google Scholar] [CrossRef]
  72. Costa, L.; Chaves, P.D.T. D C. Elasmobrânquios capturados pela pesca artesanal na costa sul do Paraná e norte de Santa Catarina, Brasil. Biota Neotrop. 2006, 6, 1–10. [Google Scholar] [CrossRef] [Green Version]
  73. Rodrigues, A.F.S.; de Sousa Rangel, B.; Wosnick, N.; Bornatowski, H.; Santos, J.L.; Moreira, R.G.; de Amorim, A.F. Report of injuries in batoids caught in small-scale fisheries: Implications for management plans. Oecol. Aust. 2018, 23, 78–89. [Google Scholar] [CrossRef] [Green Version]
  74. Guimarães, N.; da Costa, F.M.; de Oliveira, F.P.; de Alcantara, A.A.; de Matos, J.G.M.V. Ocorrência de Hypanus americanus (Hildebrand and Schroeder, 1928) e Pseudobatos percellens (Walbaum, 1792) como descarte das pescas artesanais da Praia Grande da Ilha de Itacuruçá, Rio de Janeiro, Brasil. Acta Sci. Tech. 2021, 9, 81–85. [Google Scholar] [CrossRef]
  75. Mariguela, T.C.; De-Franco, B.; Almeida, T.V.V.; Mendonça, F.F.; Gadig, O.B.F.; Foresti, F.; Oliveira, C. Identification of guitarfish species Rhinobatos percellens, R. horkelii, and Zapteryx brevirostris (Chondrichthyes) using mitochondrial genes and RFLP technique. Conserv. Genet. Resour. 2009, 1, 393–396. [Google Scholar] [CrossRef]
  76. De-Franco, B.; Mendonça, F.F.; Hashimoto, D.T.; Porto-Foresti, F.; Oliveira, C.; Foresti, F. Forensic identification of the guitarfish species Rhinobatos horkelli, R. percellens and Zapteryx brevirostris using multiplex-PCR. Mol. Ecol. Resour. 2010, 10, 197–199. [Google Scholar] [CrossRef] [PubMed]
  77. Alexandre de-Franco, B.; Fernandes Mendonça, F.; Oliveira, C.; Foresti, F. Illegal trade of the guitarfish Rhinobatos horkelii on the coasts of central and southern Brazil: Genetic identification to aid conservation. Aquat. Conserv. 2012, 22, 272–276. [Google Scholar] [CrossRef]
  78. Ferrette, B.L.; Domingues, R.R.; Rotundo, M.M.; Miranda, M.P.; Bunholi, I.V.; De Biasi, J.B.; Oliveira, C.; Foresti, F.; Mendonça, F.F. DNA Barcode reveals the bycatch of endangered batoids species in the Southwest Atlantic: Implications for sustainable fisheries management and conservation efforts. Genes 2019, 10, 304. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  79. Bernardo, C.; de Lima Adachi, A.M.C.; da Cruz, V.P.; Foresti, F.; Loose, R.H.; Bornatowski, H. The label “Cação” is a shark or a ray and can be a threatened species! Elasmobranch trade in Southern Brazil unveiled by DNA barcoding. Mar. Pol. 2020, 116, 103920. [Google Scholar] [CrossRef]
  80. Alvarenga, M.; Solé-Cava, A.M.; Henning, F. What’s in a name? Phylogenetic species identification reveals extensive trade of endangered guitarfishes and sharks. Biol. Conserv. 2021, 257, 109119. [Google Scholar] [CrossRef]
  81. Hebert, P.D.; Ratnasingham, S.; De Waard, J.R. Barcoding animal life: Cytochrome c oxidase subunit 1 divergences among closely related species. Proc. R. Soc. B Biol. Sci. 2003, 270, S96–S99. [Google Scholar] [CrossRef] [Green Version]
  82. Kearse, M.; Moir, R.; Wilson, A.; Stones-Havas, S.; Cheung, M.; Sturrock, S.; Drummond, A. Geneious Basic: An integrated and extendable desktop software platform for the organization and analysis of sequence data. Bioinformatics 2012, 28, 1647–1649. [Google Scholar] [CrossRef] [Green Version]
  83. Kimura, M. A simple method for estimating evolutionary rates of base substitutions through comparative studies of nucleotide sequences. J. Mol. Evol. 1980, 16, 111–120. [Google Scholar] [CrossRef]
  84. Tamura, K.; Nei, M. Estimation of the number of nucleotide substitutions in the control region of mitochondrial DNA in humans and chimpanzees. Mol. Biol. Evol. 1993, 10, 512–526. [Google Scholar] [PubMed] [Green Version]
  85. Kumar, S.; Stecher, G.; Li, M.; Knyaz, C.; Tamura, K. MEGA X: Molecular evolutionary genetics analysis across computing platforms. Mol. Biol. Evol. 2018, 35, 1547. [Google Scholar] [CrossRef]
  86. Felsenstein, J. Confidence limits on phylogenies: An approach using the bootstrap. Evolution 1985, 39, 783–791. [Google Scholar] [CrossRef]
  87. Peterson, B.K.; Weber, J.N.; Kay, E.H.; Fisher, H.S.; Hoekstra, H.E. Double digest RADseq: An inexpensive method for de novo SNP discovery and genotyping in model and non-model species. PLoS ONE 2012, 7, e37135. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  88. Campos, M.; Conn, J.E.; Alonso, D.P.; Vinetz, J.M.; Emerson, K.J.; Ribolla, P.E.M. Microgeographical structure in the major Neotropical malaria vector Anopheles darlingi using microsatellites and SNP markers. Parasit. Vectors 2017, 10, 1–8. [Google Scholar] [CrossRef] [Green Version]
  89. Andrews, S. FastQC: A quality Control Tool for High Throughput Sequence Data. 2010. Available online: https://www.bioinformatics.babraham.ac.uk/projects/fastqc/ (accessed on 1 January 2021).
  90. Ewels, P.; Magnusson, M.; Lundin, S.; Käller, M. MultiQC: Summarize analysis results for multiple tools and samples in a single report. Bioinformatics 2016, 32, 3047–3048. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  91. Bolger, A.M.; Lohse, M.; Usadel, B. Trimmomatic: A flexible trimmer for Illumina sequence data. Bioinformatics 2014, 30, 2114–2120. [Google Scholar] [CrossRef] [Green Version]
  92. Driller, M.; Vilaca, S.T.; Arantes, L.S.; Carrasco-Valenzuela, T.; Heeger, F.; Chevallier, D.; Mazzoni, C.J. Optimization of ddRAD-like data leads to high quality sets of reduced representation single copy orthologs (R2SCOs) in a sea turtle multi-species analysis. BioRxiv 2020, 1–22. [Google Scholar] [CrossRef] [Green Version]
  93. Rochette, N.C.; Rivera-Colón, A.G.; Catchen, J.M. Stacks 2: Analytical methods for paired-end sequencing improve RADseq-based population genomics. Mol. Ecol. 2019, 28, 4737–4754. [Google Scholar] [CrossRef]
  94. Chang, C.C.; Chow, C.C.; Tellier, L.C.; Vattikuti, S.; Purcell, S.M.; Lee, J.J. Second-generation PLINK: Rising to the challenge of larger and richer datasets. Gigascience 2015, 4, s13742-015. [Google Scholar] [CrossRef]
  95. Excoffier, L.; Lischer, H.E. Arlequin suite ver 3.5: A new series of programs to perform population genetics analyses under Linux and Windows. Mol. Ecol. Resour. 2010, 10, 564–567. [Google Scholar] [CrossRef]
  96. Jombart, T.; Ahmed, I. adegenet 1.3-1: New tools for the analysis of genome-wide SNP data. Bioinformatics 2011, 27, 3070–3071. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  97. Pritchard, J.K.; Stephens, M.; Donnelly, P. Inference of population structure using multilocus genotype data. Genetics 2000, 155, 945–959. [Google Scholar] [CrossRef] [PubMed]
  98. Rousset, F. GENEPOP’007: A complete re-implementation of the GENEPOP software for Windows and Linux. Mol. Ecol. Resour. 2008, 8, 103–106. [Google Scholar] [CrossRef] [PubMed]
  99. Puechmaille, S.J. The program structure does not reliably recover the correct population structure when sampling is uneven: Subsampling and new estimators alleviate the problem. Mol. Ecol. Resour. 2016, 16, 608–627. [Google Scholar] [CrossRef]
  100. Li, Y.L.; Liu, J.X. StructureSelector: A web-based software to select and visualize the optimal number of clusters using multiple methods. Mol. Ecol. Resour. 2018, 18, 176–177. [Google Scholar] [CrossRef]
  101. Stamatakis, A. RAxML version 8: A tool for phylogenetic analysis and post-analysis of large phylogenies. Bioinformatics 2014, 30, 1312–1313. [Google Scholar] [CrossRef] [Green Version]
  102. Ronquist, F.; Teslenko, M.; Van Der Mark, P.; Ayres, D.L.; Darling, A.; Höhna, S.; Huelsenbeck, J.P. MrBayes 3.2: Efficient Bayesian phylogenetic inference and model choice across a large model space. Syst. Biol. 2012, 61, 539–542. [Google Scholar] [CrossRef] [Green Version]
  103. Menni, R.; Stehmann, M. Distribution, environment and biology of batoid fishes off Argentina, Uruguay and Brazil. A review. Rev. Mus. Argent. Cienc. Nat. Nueva Ser. 2000, 2, 69–109. [Google Scholar] [CrossRef]
  104. Miranda, L.D.; Vooren, C.M. Captura e esforço da pesca de elasmobrânquios demersais no sul do Brasil nos anos de 1975 a 1997. Frente Marít. 2003, 19, 217–231. [Google Scholar]
  105. Ehemann, N.R.; del V González-González, L. First record of a single-clasper specimen of Pseudobatos percellens (Elasmobranchii: Rhinopristiformes: Rhinobatidae) from the Caribbean Sea, Venezuela. Acta Ichthyol. Piscat. 2018, 48, 235–240. [Google Scholar] [CrossRef] [Green Version]
  106. Lessa, R.P.; Barreto, R.R.; Quaggio, A.L.C.; Valença, L.R.; Santana, F.; Yokota, L.; Gianetti, M.D. Levantamento das espécies de elasmobrânquios capturados por aparelhos-de-pesca que atuam no berçário de Caiçara do Norte (RN). Arquiv. Cienc. Mar. 2008, 41, 58–64. [Google Scholar]
  107. Rosa, R.S.; Gadig, O.B.F. Conhecimento da diversidade dos Chondrichthyes marinhos no Brasil: A contribuição de José Lima de Figueiredo. Arq. Zool. 2014, 45, 89–104. [Google Scholar] [CrossRef] [Green Version]
  108. Ferrari, A.; Di Crescenzo, S.; Cariani, A.; Crobe, V.; Benvenuto, A.; Piattoni, F.; Tinti, F. Puzzling over spurdogs: Molecular taxonomy assessment of the Squalus species in the Strait of Sicily. Eur. Zool. J. 2021, 88, 181–190. [Google Scholar] [CrossRef]
  109. Toffoli, D.; Hrbek, T.; Araújo, M.L.G.D.; Almeida, M.P.D.; Charvet-Almeida, P.; Farias, I.P. A test of the utility of DNA barcoding in the radiation of the freshwater stingray genus Potamotrygon (Potamotrygonidae, Myliobatiformes). Genet. Mol. Biol. 2008, 31, 324–336. [Google Scholar] [CrossRef] [Green Version]
  110. Hendry, A.P.; Bolnick, D.I.; Berner, D.; Peichel, C.L. Along the speciation continuum in sticklebacks. J. Fish Biol. 2009, 75, 2000–2036. [Google Scholar] [CrossRef]
  111. Prado, H.M.; Schlindwein, M.N.; Murrieta, R.S.S.; Nascimento, D.R.D.; Souza, E.P.D.; Cunha-Lignon, M.; Contente, R.F. The Valo Grande channel in the Cananéia-Iguape estuary-lagoon complex (SP, Brazil): Environmental history, ecology, and future perspectives. Ambient. Soc. 2019, 22, 1–22. [Google Scholar] [CrossRef]
  112. Ferrol-Schulte, D.; Wolff, M.; Ferse, S.; Glaser, M. Sustainable Livelihoods Approach in tropical coastal and marine social–ecological systems: A review. Mar. Pol. 2013, 42, 253–258. [Google Scholar] [CrossRef]
  113. Castro, R.M.C.; Menezes, N.A. Estudo Diagnóstico de Diversidade de Peixes do Estado de São Paulo. In Biodiversidade do Estado de São Paulo, Brasil: Síntese do Conhecimento ao Final do Século XX; Joly, C.A., Bicudo, C.E.M., Eds.; Vertebrados-6; FAPESP: São Paulo, Brazil, 1998; pp. 1–13. [Google Scholar]
  114. Menezes, N.A.; Buckup, P.A.; de Figueiredo, J.L.; de Moura, R.L. (Eds.) Catálogo das Espécies de Peixes Marinhos do Brasil; Museu de Zoologia da Universidade de São Paulo: São Paulo, Brazil, 2003; Volume 1. [Google Scholar]
  115. Schlaff, A.M.; Heupel, M.R.; Simpfendorfer, C.A. Influence of environmental factors on shark and ray movement, behaviour and habitat use: A review. Rev. Fish Biol. Fish. 2014, 24, 1089–1103. [Google Scholar] [CrossRef]
  116. Machado, L.F.; de Souza Damasceno, J.; Bertoncini, Á.A.; Farro, A.P.C.; Hostim-Silva, M.; Oliveira, C. Population genetic structure and demographic history of the spadefish, Chaetodipterus faber (Ephippidae) from Southwestern Atlantic. J. Exp. Mar. Biol. Ecol. 2017, 487, 45–52. [Google Scholar] [CrossRef] [Green Version]
  117. Little, A.G.; Loughland, I.; Seebacher, F. What do warming waters mean for fish physiology and fisheries? J. Fish Biol. 2020, 97, 328–340. [Google Scholar] [CrossRef]
  118. Lindmark, M.; Audzijonyte, A.; Blanchard, J.L.; Gårdmark, A. Temperature impacts on fish physiology and resource abundance lead to faster growth but smaller fish sizes and yields under warming. Glob. Change Biol. 2022, 28, 6239–6253. [Google Scholar] [CrossRef] [PubMed]
  119. Carrillo-Briceño, J.D.; Villafaña, J.A.; De Gracia, C.; Flores-Alcívar, F.F.; Kindlimann, R.; Abella, J. Diversity and paleoenvironmental implications of an elasmobranch assemblage from the Oligocene–Miocene boundary of Ecuador. PeerJ. 2020, 8, e9051. [Google Scholar] [CrossRef] [PubMed]
  120. Dennis, A.B.; Hellberg, M.E. Ecological partitioning among parapatric cryptic species. Mol. Ecol. 2010, 19, 3206–3225. [Google Scholar] [CrossRef] [PubMed]
  121. Barreto, G.C.; Di Domenico, M.; Medeiros, R.P. Human dimensions of marine protected areas and small-scale fisheries management: A review of the interpretations. Mar. Pol. 2020, 119, 104040. [Google Scholar] [CrossRef]
  122. Collette, B.B.; Rützler, K. Reef fishes over sponge bottoms off the mouth of the Amazon river. In Proceedings of the Third International Coral Reef Symposium, Miami, FL, USA, 1 May 1977; Volume 1, pp. 305–310. [Google Scholar]
  123. Luiz, O.J.; Madin, J.S.; Robertson, D.R.; Rocha, L.A.; Wirtz, P.; Floeter, S.R. Ecological traits influencing range expansion across large oceanic dispersal barriers: Insights from tropical Atlantic reef fishes. Proc. R. Soc. B Biol. Sci. 2012, 279, 1033–1040. [Google Scholar] [CrossRef]
  124. Bornatowski, H.; Abilhoa, V.; Charvet-Almeida, P. Elasmobranchs of the Paraná Coast, southern Brazil, south-western Atlantic. Mar. Biodiv. Rec. 2009, 158, 1–6. [Google Scholar] [CrossRef]
  125. Hackradt, C.W.; Félix-Hackradt, F.C.; Pichler, H.A.; Spach, H.L.; Santos, L.D.O. Factors influencing spatial patterns of the ichthyofauna of low energy estuarine beaches in southern Brazil. J. Mar. Biol. Assoc. UK 2011, 91, 1345–1357. [Google Scholar] [CrossRef]
  126. Carmo, W.P.; Bornatowski, H.; Oliveira, E.C.; Fávaro, L.L. Diet of the chola guitarfish, Rhinobatos percellens (Rhinobatidae), in the Paranaguá Estuarine complex. An. Acad. Bras. Ciên. 2015, 87, 721–731. [Google Scholar] [CrossRef] [Green Version]
  127. Possatto, F.E.; Broadhurst, M.K.; Spach, H.L.; Winemiller, K.O.; Millar, R.B.; Santos, K.M.; Lamour, M.R. Mapping the spatio-temporal distribution of threatened batoids to improve conservation in a subtropical estuary. J. Fish Biol. 2016, 89, 1098–1104. [Google Scholar] [CrossRef]
  128. Fratantoni, D.M.; Glickson, D.A. North Brazil Current ring generation and evolution observed with SeaWiFS. J. Phys. Oceanogr. 2002, 32, 1058–1074. [Google Scholar] [CrossRef]
  129. Roberts, A.M. Effect of Heinrich Events on Ocean Circulation and Sediment Input on the Nordeste Brazilian Continental Margin; University of Southampton: Southampton, UK, 2017. [Google Scholar]
  130. Pires-Vanin, A.M.S.; Rossi-Wongtschowski, C.L.D.B.; Aidar, E.; Mesquita, H.D.S.; Soares, L.S.H.; Katsuragawa, M.; Matsuura, Y. Estrutura e função do ecossistema de plataforma continental do Atlântico Sul brasileiro: Síntese dos resultados. Publicação Espec. Inst. Oceanográfico 1993, 10, 217–231. [Google Scholar]
  131. Braga, E.D.S.; Niencheski, L.F.H. Composição das Massas de Água e Seus Potenciais Produtivos na Área Entre o Cabo de São Tomé (RJ) e o Chuí (RS); O ambiente Oceanografia da plataforma continental e do talude na região Sudeste-Sul do Brasil: São Paulo, Brasil, 2006. [Google Scholar]
  132. Palma, E.D.; Matano, R.P.; Piola, A.R. A numerical study of the Southwestern Atlantic Shelf circulation: Stratified ocean response to local and offshore forcing. J. Geophys. Res. Ocean. 2008, 113, C11. [Google Scholar] [CrossRef]
  133. Rotundo, M.M.; Severino-Rodrigues, E.; Barrella, W.; Petrere-Junior, M.; Ramires, M. Checklist of marine demersal fishes captured by the pair trawl fisheries in Southern (RJ-SC) Brazil. Biota Neotrop. 2019, 19, e20170432. [Google Scholar] [CrossRef] [Green Version]
  134. Rutledge, K.M. A new guitarfish of the genus Pseudobatos (Batoidea: Rhinobatidae) with key to the guitarfishes of the Gulf of California. Copeia 2019, 107, 451–463. [Google Scholar] [CrossRef]
  135. Weigmann, S. Reply to Borsa 2017: Comment on ‘Annotated checklist of the living sharks, batoids and chimaeras (Chondrichthyes) of the world, with a focus on biogeographical diversity by Weigmann (2016)’. J. Fish Biol. 2017, 90, 1176–1181. [Google Scholar] [CrossRef] [PubMed]
  136. Boussarie, G.; Momigliano, P.; Robbins, W.D.; Bonnin, L.; Cornu, J.F.; Fauvelot, C.; Vigliola, L. Identifying barriers to gene flow and hierarchical conservation units from seascape genomics: A modelling framework applied to a marine predator. Ecography 2022, 2022, 06158. [Google Scholar] [CrossRef]
  137. Bunholi, I.V.; da Silva Ferrette, B.L.; Domingues, R.R.; Rotundo, M.M.; Cuevas, J.M.; García, M.; Mendonça, F.F. Multilocus phylogeography of the endemic and endangered angular angelshark (Squatina guggenheim) in the Southwest Atlantic Ocean. Hydrobiologia 2022, 849, 2177–2192. [Google Scholar] [CrossRef]
  138. Avise, J.C. Phylogeography: The History and Formation of Species; Harvard University Press: Cambridge, MA, USA, 2000. [Google Scholar]
Figure 1. Bayesian Analysis of Genetic Similarity among geographical Pseudobatos percellens populations performed on STRUCTURE results (k = 2) for 52 individuals from 3329 SNPs.
Figure 1. Bayesian Analysis of Genetic Similarity among geographical Pseudobatos percellens populations performed on STRUCTURE results (k = 2) for 52 individuals from 3329 SNPs.
Diversity 15 00344 g001
Figure 2. Scatterplots of the Discriminant Analysis of Principal Components (DAPC) om 3.329 SNPs from Pseudobatos percellens. Dots represent individuals and colored ellipses correspond to geographical populations. The Principal Component Analysis (PCA) and DA (Significance tests) eigenvalue representatives are highlighted in the inset.
Figure 2. Scatterplots of the Discriminant Analysis of Principal Components (DAPC) om 3.329 SNPs from Pseudobatos percellens. Dots represent individuals and colored ellipses correspond to geographical populations. The Principal Component Analysis (PCA) and DA (Significance tests) eigenvalue representatives are highlighted in the inset.
Diversity 15 00344 g002
Figure 3. Inference of the number of clusters obtained in the Discriminant Analysis of Principal Components (DAPC) performed on the 3.329 SNPs dataset from Pseudobatos percellens. (A) DAPC scatterplot distinguishing between the three genetic clusters and FST pairwise values detected between them. The Principal Component Analysis (PCA) and DA (Significance tests) eigenvalue representatives are highlighted in the inset. (B) Subdivision of the individuals based on the DAPC membership probabilities into three genetic clusters (k = 3). The same color for individuals indicates that they belong to the same cluster.
Figure 3. Inference of the number of clusters obtained in the Discriminant Analysis of Principal Components (DAPC) performed on the 3.329 SNPs dataset from Pseudobatos percellens. (A) DAPC scatterplot distinguishing between the three genetic clusters and FST pairwise values detected between them. The Principal Component Analysis (PCA) and DA (Significance tests) eigenvalue representatives are highlighted in the inset. (B) Subdivision of the individuals based on the DAPC membership probabilities into three genetic clusters (k = 3). The same color for individuals indicates that they belong to the same cluster.
Diversity 15 00344 g003
Figure 4. (A) Map of the analyzed Pseudobatos percellens Brazilian coast sampling sites, highlighted along the Southern Equatorial Current (SEC), region of the mouth of the Amazon River, and Victoria-Trindade Seamount Chain. (B) Phylogenetic tree based on Bayesian Inference with 1.448 SNPS identified in P. percellens, and posterior probabilities (≥0.98) and Bootstrap values (≥57) obtained from the Maximum Likelihood Analysis employing the same matrix. (C) NeighborNet splits based on the uncorrected p-distance with the groups identified in the samples AP, Amapá; PA, Pará; PE, Pernambuco; SP, São Paulo; PR, Paraná.
Figure 4. (A) Map of the analyzed Pseudobatos percellens Brazilian coast sampling sites, highlighted along the Southern Equatorial Current (SEC), region of the mouth of the Amazon River, and Victoria-Trindade Seamount Chain. (B) Phylogenetic tree based on Bayesian Inference with 1.448 SNPS identified in P. percellens, and posterior probabilities (≥0.98) and Bootstrap values (≥57) obtained from the Maximum Likelihood Analysis employing the same matrix. (C) NeighborNet splits based on the uncorrected p-distance with the groups identified in the samples AP, Amapá; PA, Pará; PE, Pernambuco; SP, São Paulo; PR, Paraná.
Diversity 15 00344 g004
Table 1. Genetic diversity statistics for Pseudobatos percellens using 3.329 SNPs. ID: sampling localities, N: number of individuals per location, Ap: number of private alleles, Np: number of polymorphic loci, Ho: observed heterozygosity, He: expected heterozygosity, FIS: inbreeding coefficient, (π): nucleotide diversity.
Table 1. Genetic diversity statistics for Pseudobatos percellens using 3.329 SNPs. ID: sampling localities, N: number of individuals per location, Ap: number of private alleles, Np: number of polymorphic loci, Ho: observed heterozygosity, He: expected heterozygosity, FIS: inbreeding coefficient, (π): nucleotide diversity.
LocalitiesIDNApNpHoHeFIS(π)
AmapáAP03832160.5770.487−0.5420.115 ± 0.067
ParáPA051914480.4710.415−0.2160.108 ± 0.057
PernambucoPE143893310.3150.298−0.0880.102 ± 0.050
São PauloSP151401150.2810.239−0.2130.031 ± 0.015
ParanáPR151921580.3330.273−0.2330.038 ± 0.019
Disclaimer/Publisher’s Note: The statements, opinions and data contained in all publications are solely those of the individual author(s) and contributor(s) and not of MDPI and/or the editor(s). MDPI and/or the editor(s) disclaim responsibility for any injury to people or property resulting from any ideas, methods, instructions or products referred to in the content.

Share and Cite

MDPI and ACS Style

Cruz, V.P.; Rotundo, M.M.; Charvet, P.; Boza, B.R.; Souza, B.C.; Cerqueira, N.N.C.D.; Oliveira, C.; Lessa, R.; Foresti, F. Investigating an Unknown Biodiversity: Evidence of Distinct Lineages of the Endemic Chola Guitarfish Pseudobatos percellens Walbaum, 1792 in the Western Atlantic Ocean. Diversity 2023, 15, 344. https://doi.org/10.3390/d15030344

AMA Style

Cruz VP, Rotundo MM, Charvet P, Boza BR, Souza BC, Cerqueira NNCD, Oliveira C, Lessa R, Foresti F. Investigating an Unknown Biodiversity: Evidence of Distinct Lineages of the Endemic Chola Guitarfish Pseudobatos percellens Walbaum, 1792 in the Western Atlantic Ocean. Diversity. 2023; 15(3):344. https://doi.org/10.3390/d15030344

Chicago/Turabian Style

Cruz, Vanessa P., Matheus M. Rotundo, Patrícia Charvet, Beatriz R. Boza, Bruno C. Souza, Najila N. C. D. Cerqueira, Claudio Oliveira, Rosângela Lessa, and Fausto Foresti. 2023. "Investigating an Unknown Biodiversity: Evidence of Distinct Lineages of the Endemic Chola Guitarfish Pseudobatos percellens Walbaum, 1792 in the Western Atlantic Ocean" Diversity 15, no. 3: 344. https://doi.org/10.3390/d15030344

Note that from the first issue of 2016, this journal uses article numbers instead of page numbers. See further details here.

Article Metrics

Back to TopTop