Next Article in Journal
Analysis of Chemical Composition, Amino Acid Content, and Rumen Degradation Characteristics of Six Organic Feeds
Next Article in Special Issue
What Is in Your Shark Fin Soup? Probably an Endangered Shark Species and a Bit of Mercury
Previous Article in Journal
Assessment of the Vaccination Program against Cystic Echinococcosis in Sheep in the Pehuenche Community of Central Chile
Previous Article in Special Issue
Length at Maturity, Sex Ratio, and Proportions of Maturity of the Giant Electric Ray, Narcine entemedor, in Its Septentrional Distribution
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Integrated Taxonomy Revealed Genetic Differences in Morphologically Similar and Non-Sympatric Scoliodon macrorhynchos and S. laticaudus

1
Institute of Ocean and Earth Sciences, Universiti Malaya, Kuala Lumpur 50603, Malaysia
2
Institute of Advanced Studies, Universiti Malaya, Kuala Lumpur 50603, Malaysia
3
CSIRO National Research Collections Australia, Australia National Fish Collection, Hobart, TAS 7001, Australia
4
Institute of Biological Sciences, Universiti Malaya, Kuala Lumpur 50603, Malaysia
5
Florida Museum of Natural History, Dickinson Hall, Gainesville, FL 32601, USA
6
Aow Thai Marine Ecology Centre, Bangkok 10100, Thailand
*
Authors to whom correspondence should be addressed.
Animals 2022, 12(6), 681; https://doi.org/10.3390/ani12060681
Submission received: 27 January 2022 / Revised: 3 March 2022 / Accepted: 5 March 2022 / Published: 8 March 2022
(This article belongs to the Special Issue Sharks and Skates: Ecology, Distribution and Conservation)

Abstract

:

Simple Summary

In this study, the species identities of similar-looking coastal spadenose sharks from different areas were clarified by adding new molecular markers and more individual body measurements, including animals from the Malaysian Peninsula that had not been examined previously. Collective evidence showed that there are two genetically distinct species that do not overlap in their spatial occurrence. The Malacca Strait acts as a boundary delineating the distribution range of the Pacific spadenose shark Scoliodon macrorhynchos to the east and, of the Northern Indian Ocean, S. laticaudus to the west. In addition, the need to determine the species status of Scoliodon animals from Indonesian waters was identified. The present study reinforced the need to rely on comprehensive genetic information in addition to external characteristics to assess the species identities and distribution range for small sharks and rays that have apparent contiguous coastal distribution with limited dispersal abilities.

Abstract

Previous examination of the mitochondrial NADH2 gene and morphological characteristics led to the resurrection of Scoliodon macrorhynchos as a second valid species in the genus, in addition to S. laticaudus. This study applied an integrated taxonomic approach to revisit the classification of the genus Scoliodon based on new materials from the Malaysian Peninsula, Malaysian Borneo and Eastern Bay of Bengal. Mitochondrial DNA data suggested the possibility of three species of Scoliodon in the Indo-West Pacific, while the nuclear DNA data showed partially concordant results with a monophyletic clade of S. macrorhynchos and paraphyletic clades of S. laticaudus and S. cf. laticaudus from the Malacca Strait. Morphological, meristic and dental characteristics overlapped between the three putative species. Collective molecular and morphological evidence suggested that the differences that exist among the non-sympatric species of Scoliodon are consistent with isolation by distance, and Scoliodon macrorhynchos remains as a valid species, while S. cf. laticaudus is assigned as S. laticaudus. The Malacca Strait acts as a spatial delineator in separating the Pacific S. macrorhynchos (including South China Sea) from the Northern Indian Ocean S. laticaudus. Future taxonomic work should focus on clarifying the taxonomic status of Scoliodon from the Indonesian waters.

1. Introduction

The genus Scoliodon was proposed by Müller and Henle [1] for S. laticaudus Müller and Henle [2]. Within the family Carcharhinidae, this genus is distinguished from other genera by its clasper and cranial morphology and very shallowly concave post-ventral caudal margin [3]. The genus Scoliodon is morphologically similar to hammerhead sharks (family Sphyrnidae) in a number of proportional body measurements but is placed in Carcharhinidae, as it does not have the laterally expanded head that is characteristic of hammerheads [4]. The genus sits within the subfamily Scoliodontinae and differs from the other genera within the subfamily, i.e., Rhizoprionodon and Loxodon, by having a greatly depressed, trowel-shaped head, broader and more triangular pectoral fins and a more posteriorly located first dorsal fin (free rear tip about over mid-bases of the caudal fin) [5].
Scoliodon has long been considered to be a monotypic genus until White et al. [5] resurrected S. macrorhynchos [6] as a second species within the genus. Scoliodon laticaudus is common along the insular shelf extending from the Northern Indian Ocean to Northeastern Africa [7]. Scoliodon macrorhynchos is known from Southeast Asia from Taiwan and China to Indonesia and Sarawak, Malaysia [5]. A possible third species was also reported from the Bay of Bengal by White et al. [5] and Naylor et al. [8] based on NADH2 sequence data. These authors suggested that Carcharhias (Physodon) muelleri Müller and Henle [2], described from Bengal may be an available name for this species, but in the absence of specimens, this species was not formally resurrected.
The spadenose shark is one of the smallest carcharhinid species, attaining a maximum total length of 74 cm [9], occurring in shallow muddy and sandy bottom habitats [10]. Nearshore elasmobranchs generally have limited dispersal capabilities [4]. For instance, the bambooshark Chiloscyllium punctatum [11] and the stingray Neotrygon species [12], both of which are small, show regional population subdivisions with limited genetic mixing throughout the Indo-West Pacific. When geographic barriers and the lack of suitable contiguous habitats are combined with a proclivity not to disperse, allopatric speciation becomes more likely. These factors influenced the redescription of S. macrorhynchos from the Eastern South China Sea and the suggestion that S. muelleri from the Bay of Bengal might also be a distinct species [5].
White et al. [5] found that S. macrorhynchos and S. laticaudus showed high intraspecific variations from morphometric data (as high as 5.2% in some head and snout measurements) but low interspecific variations; only a limited number of morphometric measurements differed between the two species, with partly overlapping ranges. For the molecular analysis, the interspecific genetic distance of the NADH dehydrogenase subunit 2 (NADH2) gene between S. macrorhynchos and S. laticaudus was about 3%. This degree of divergence falls at the borderline of “intra” versus “inter”-specific genetic variations in sharks and rays. Mobula kuhlii and M. eregoodoo were viewed as one species based on their close genetic distance (interspecific distance < 1.5%) but viewed as distinct species based on morphological data [13]. Hypanus berthalutzae was viewed as a distinct species from other closely related Hypanus species based on genetics (interspecific distance 0.82–3.11%), morphology, and ecological niche modeling data [14]. These examples highlight the challenge of distinguishing similar-looking but potentially distinct species, such as those in the genus Scoliodon.
Reliance on mitochondrial DNA (mtDNA) alone in elucidating phylogenetic relationships among closely related species has been called into question. Reviews by Galtier et al. [15] and Balloux [16] presented some of the limitations associated with reliance on mitochondrial data. The concerns raised arose from limited cases of non-maternally transmitted mtDNA that may call into question the assumption of reduced within-individual diversity [17,18,19], non-neutral evolution through selection [20,21,22], and the nonconstant mutation rate in mtDNA [23,24,25]. While these concerns may not necessarily be applicable in the representation of within-species history for Scoliodon, the genetic basis for delineating S. macrorhynchos as a separate species from S. laticaudus [5] merits a critical review using more representative sampling with additional markers.
In this study, both nuclear and mtDNA markers were used in addition to morphological data sample specimens across known geographical range of Scoliodon to clarify the phylogenetic relationships for the group. We included specimens of Scoliodon from the Malacca Strait, the west coast of Peninsula Malaysia, that had not been previously examined. The fine-scale contemporary distribution range of the Scoliodon genus, especially in the Indo-Malaya region, and knowledge gaps were discussed.

2. Materials and Methods

2.1. Sample Collection

Specimens of Scoliodon were acquired at fish landing sites located in the Malacca Strait on the west coast of Peninsular Malaysia, i.e., Hutan Melintang (3°52′13.6″ N 100°55′39.3″ E), Sungai Besar (3°40′15.2″ N 100°58′52.3″ E), and Pasir Penambang (3°21′03.9″ N 101°15′07.0″ E), henceforth labeled as S. cf. laticaudus and S. macrorhynchos from two landing sites in Sarawak in Malaysian Borneo, i.e., Kuching (1°34′04.7″ N, 110°22′45.8″ E) and Mukah (2°53′50.6″ N, 112°05′45.6″ E). Tissue samples were taken from a random subset of specimens (10 each from Malacca Strait and from Sarawak) and stored in 95% alcohol prior to molecular analyses, while the whole specimens were fixed using 10% formalin and store in 70% alcohol. A subset of specimens, 21 from Malacca Strait and 13 from Sarawak, was preserved whole and retained for subsequent morphological analysis by one of us (KCL). Eleven whole specimens of S. cf. laticaudus were also collected from the Ranong fishing port in Thailand, Eastern Bay of Bengal, during recent surveys of that landing site [26].

2.2. Molecular Analyses

Two mitochondrial DNA (cytochrome oxidase subunit 1 (COI) and NADH dehydrogenase subunit 2 (NADH2) regions) were used in molecular species identification and seven nuclear genes following Aschliman et al. [27] DNA (actin-like protein (ACT), kelch repeat and BTB domain-containing protein 2 (KBTBD2), prospero homeobox protein 1 (PROX1), recombination activating gene 1 (RAG1), recombination activating gene 2 (RAG2), sec1 family domain-containing protein 2 (SCFD2), and transducer of ERBB2.1 (TOB1) region) were used to verify the taxonomic assignment using mitochondrial DNA. DNA was extracted using 10% Chelex resin incubated for two minutes at 60 °C, followed by 25 min at 103 °C (modified from Hyde et al. [28]). Extracted DNA was subjected to Polymerase Chain Reaction (PCR) to amplify all targeted DNA markers. PCR were carried out either using iTaqTM Plus DNA Polymerase (iNtRON Biotechnology, INC., Seongnam-si, Korea) or MyTaqTM Red Mix (Bioline, London, United Kingdom) in 20 µL of reaction mix containing 2 µL of 10x PCR buffer; 0.5 µL of dNTP mixture (2.5 mM each); 1 µL of 10-pmol primer (both primers); 1.25 unit of Taq DNA polymerase; 1 µL of 50-pg–1.0-µg DNA templates; and top up with molecular-grade water or 10 µL of MyTaqTM Red Mix premix (mixture of 10x PCR buffer, dNTPs, and Taq polymerase); 1 µL of 10-pmol primer (both primers); 1 µL of 50-pg–1.0-µg DNA templates; and top up with molecular-grade water, respectively. The PCR cycles for mitochondrial DNA comprised of 2-min initial denaturation at 94 °C, followed by 30 cycles of 20 s at 94 °C, 20 s at 44 °C (COI) or 52 °C (NADH2), and 1 min at 72 °C and, subsequently, a final extension of 5 min at 72 °C. The PCR cycles for nuclear DNA comprised 3-min initial denaturation at 95 °C, followed by 35 cycles of 15 s at 95 °C, 15 s at 52–60 °C, and 1 min at 72 °C and, subsequently, a final extension of 5 min at 72 °C. Touchdown PCR with annealing temperature that decreased 0.3 °C/cycle from 68 °C to 58 °C was performed on PROX1 due to the amplification of nonspecific DNA at all tested temperatures between 45 and 60 °C. The primer sets used for all the targeted regions are listed in Table 1. All PCR products were examined using 1% agarose in TAE buffer prior to the Sanger sequencing service at Apical Scientific Sdn Bhd (Selangor, Malaysia).

2.3. Phylogenetic Analysis

Sequences were reviewed manually using BioEdit [31], aligned using ClustalX [32], and finally, trimmed using BioEdit [31]. They were all submitted to the NCBI GenBank database, with the accession numbers provided in Supplementary Table S1. The following analyses applied to individual marker, as well as grouped markers by mitochondrial DNA and nuclear DNA. The aligned sequences were subjected to the best model search based on Akaike Information Criterion (AIC) and Bayesian Information Criterion (BIC) for Maximum Likelihood (ML) and Bayesian Inference (BI) analyses, respectively, using Kakusan v 3 [33], as shown in Supplementary Table S2. The generated files were subsequently used for phylogenetic tree construction using Treefinder for ML [34] and MrBayes for BI [35]. The ML analyses were performed with 1000 bootstrap replicates. The Bayesian analyses were initiated with a random starting tree and two parallel runs, each of which consisted of running four chains of Markov chain Monte Carlo (MCMC) iterations for 2,000,000 generations (sampled every 100th generation for each chain). The convergence and burn-in from “sump” commands in MrBayes were used to evaluate likelihood values for post-analysis trees and parameters. Five thousand trees generated were discarded as burn-in (where the likelihood values were stabilized prior before the burn-in), and the remaining trees after burn-in were used to calculate the posterior probabilities using the “sumt” command.
The finalized ML and BI phylogenetic trees were processed via Figtree v 1.3.1 (http://tree.bio.ed.ac.uk/software/figtree/, accessed on 1 October 2014). For mitochondrial DNA, sequences of closely related species Loxodon macrorhinus and Lamiopsis tephrodes were used as outgroups. Sequences of Sphyrna lewini and Rhizoprionodon acutus, on the other hand, were used as the outgroup for nuclear DNA, as the sequences for Loxodon and Lamiopsis were not available. As such, the sequence of S. lewini was added to mitochondrial DNA analyses to facilitate the comparison between mitochondrial and nuclear DNA. Some other sequences available in the National Center for Biotechnology Information (NCBI) GenBank and Barcode of Life Data (BOLD) systems were also used in the tree construction for comparison (Supplementary Table S3). Uncorrected p-distance was calculated using PAUP* 40b10 software [36] to evaluate the genetic divergence among the sampled Scoliodon species by sampling areas.
We tested species delimitation using a multispecies coalescent analysis implemented in ASTRAL 5.7.7 [37,38] and BPP 4.3 [39,40,41]. In the ASTRAL analysis, two hundred gene trees were searched under ML + rapid bootstrap for each of the genes using raxmlGUI 1.5 beta [42]. All generated gene trees were combined manually as input into ASTRAL to generate a ASTRAL species tree and normalized quartet score. The normalized quartet score refers to the proportion of gene trees that matched with the species tree; a higher score indicates greater agreement between gene trees and species tree. In the BPP analysis, we performed an unguided species delimitation analysis (A11) to test if the Scoliodon species can be assigned as a single species. We set multiple theta (population size) and tau (divergence time) combinations using the inverse gamma prior to IG (2, X), with X being 0.1, 0.01, and 0.001. Each analysis was performed twice to confirm the stability of the results.

2.4. Morphological and Meristic Data

Measurement terminology followed Compagno [3,4,43], who assigned names and abbreviations to measurements often indicated by descriptive phrases (example: snout to upper caudal origin = precaudal length = PRC). Dentitional terms generally followed Compagno [3,43,44]. Vertebral terminology, method of counting, and vertebral ratios followed Springer and Garrick [45] and Compagno [3,43,44].
A total of 83 morphometric measurements were obtained from 74 Scoliodon specimens from a range of locations encompassing a large proportion of the geographic range of the three ‘species’ types: S. laticaudus, S. cf. laticaudus, and S. macrorhynchos (Figure 1). A total of 8 specimens of S. laticaudus (India); 32 specimens of S. cf. laticaudus (including the S. muelleri holotype from ‘Bengal’, Malacca Strait, and the Ranong fishing port in the Andaman Sea); and 34 specimens of S. macrorhynchos (Hong Kong, Indonesia, Borneo, and Taiwan) were measured in full (Table 2). Vertebral counts were taken from radiographs of 13 specimens of S. cf. laticaudus and 13 specimens of S. macrorhynchos. Counts were obtained separately for the trunk (monospondylous), precaudal (monospondylous and diplospondylous to the origin of upper lobe of caudal fin), and caudal (centra of the caudal fin) vertebrae. Tooth row counts were taken in situ or from excised jaws of 7 specimens of S. laticaudus, 5 specimens of S. cf. laticaudus, and 8 specimens of S. macrorhynchos.

2.5. Multivariate Analyses

Morphometric measurements, as % total length (TL), were subjected to nonmetric multidimensional scaling (MDS) ordination (Primer v 7.0 package, Quest Research Limited, Auckland, New Zealand) to determine whether significant differences between putative species exist or whether intraspecific variations of a single species is a factor. One-way Analyses of Similarity (ANOSIM) were employed to test whether morphometric measurements differed significantly among size classes. Similarity Percentages (SIMPER) were employed when a pairwise ANOSIM result was significant at p < 0.05 to determine what characters contributed most to the observed differences. To determine if significant differences between size classes exist, samples were allocated to one of four arbitrary size classes: (1) <249 mm TL, (2) 250–299 mm TL, (3) 300–399 mm TL, and (4) >400 mm TL. Morphometric measurements were analyzed without transformation since the preliminary analyses revealed that the stress levels were acceptable (i.e., <0.3) for MDS analyses (see Clarke and Gorley [46]). Several measurements, associated with the clasper and trunk and abdomen heights and widths, were not available for measurement for all individuals, so these characters were excluded from the MDS analysis.

2.6. Museum Holdings

Collection details for the 74 Scoliodon specimens examined are provided in Supplementary Data S1. Specimens are referred to by the following prefixes for their registration numbers: BMNH, British Museum of Natural History, London; IPPS, Sarawak Fisheries Research Institute, Bintawa, Malaysia; CSIRO, Australian National Fish Collection, Hobart; RMNH, Rikjsmuseum van Natuurlkjke Histoire, Leiden; and MNHN, Museum National d’Histoire Naturelle, Paris, France.

3. Results

3.1. Molecular Analysis

Using the NADH2 and COI mitochondrial DNA sequences (Figure 2 and Supplementary Figure S1a,b), three monophyletic groups with moderate-to-full support bootstrap values (ML 58.3—100/BI 68—100) were identified based on sampling locations, i.e., Scoliodon laticaudus from the Indian Ocean (based on samples from the west coast of India), Scoliodon macrorhynchos from South China Sea (Kuching and Mukah, both localities in Sarawak, which were grouped with samples from China and Taiwan), and a possible third species from the Malacca Strait, tentatively labeled as S. cf. laticaudus, were grouped with samples from Bangladesh, Myanmar, and Thailand. The uncorrected p-distances among these three monophyletic groups ranged from 0.61 to 3.06% for COI, 2.98 to 4.23% for NADH2, and 2.12 to 3.19% for the combined mitochondrial DNA (Table 3).
The estimated trees for Scoliodon species using nuclear DNA (Figure 2 and Figure 3 and Supplementary Figure S1) showed partial agreement with those using mitochondrial DNA. Three out of five individual nuclear DNA gene trees indicated monophyly of the Scoliodon genus (Prox1, RAG1, and TOB1) (Supplementary Figure S1c–i). Topologies of concatenated nuclear DNA estimated tree showed two monophyletic groups, S. macrorhynchos and S. laticaudusS. cf. laticaudus groups (Figure 3). The uncorrected p-distance for nuclear DNA among the three monophyletic groups identified from mitochondrial DNA ranged from 0 to 0.91% (mean 0.2%) (Table 3).
The species tree estimated in ASTRAL for both mitochondrial and nuclear DNA were topologically congruent with their respective gene trees and had a normalized quartet score of 0.81 and 0.61, respectively (Figure 4). The BPP run supported both estimations from traditional phylogenetic analyses (Bayesian and ML) and ASTRAL. Specifically, the BPP run on mtDNA supported the separation of Scoliodon into three separate species with a probability of 1 under any combination of the theta and tau priors. The BPP run on nuclear DNA, on the other hand, varied depending on the theta and tau prior settings; settings of theta at 0.1 regardless of tau prior supported the monospecificity of Scoliodon (probability > 0.99), theta at 0.001 in combinations with tau at 0.01 and at 0.001 supported separation into three species (probability 0.65–0.88), and other settings in between supported the combination of S. laticaudus and S. cf. laticaudus as a separate group from S. macrorhynchos (probability 0.51–0.61).

3.2. Morphology and Meristics

No nonoverlapping morphometric ranges were found between the three putative ‘species’ of Scoliodon. Likewise, vertebral counts strongly overlapped between the three ‘species’. No dental morphological differences were detected between the three Scoliodon ‘species’.
The MDS analysis of the measured Scoliodon specimens showed considerable overlaps among the three ‘species’ (Figure 5a). Measurements of the limited S. laticaudus samples were highly variable but generally fell within the two overlapping clusters of S. macrorhynchos and S. cf. laticaudus animals. ANOSIM showed that the ‘species’ were significantly different overall (p < 0.01) although the global R2 value was very low (0.24). Similarly, pairwise comparisons between the three ‘species’ were also significantly different (p < 0.01) but with low R2 values (0.18–0.42).
When the same ordination plot was coded by size class (1 ≤ 250 mm TL, 2 = 250–299 mm TL, 3 = 300–399 mm TL, and 4 ≥ 400 mm TL), the samples for each size class formed only partially overlapping groups, with the smallest specimens to the left of the plot and the largest to the right (Figure 5b). ANOSIM showed that the size classes significantly different overall (p < 0.01), and with a higher global R2 value (0.54). All pairwise comparisons of size classes were also significantly different (p < 0.01), with generally higher R2 values (0.3–0.96). The measurements shown by SIMPER to be the most responsible for the differences between the size classes were pre-anal length, pectoral–pelvic space, pre-pectoral length, pre-pelvic length, head length, and pre-first dorsal length.

4. Discussion

Based on a combination of nuclear and mitochondrial markers, the evidence supports the split proposed by White et al. [5]. Evidence from mtDNA suggests genetic isolation among the three ‘species’ types; S. laticaudus from India is a separate species from S. macrorhynchos from Sarawak, Malaysian Borneo that appears to cluster with samples from China and Taiwan. Evidence from the pooled nuclear markers group S. cf. laticaudus (from Malacca Strait) with S. laticaudus. Both molecular and morphological data presented suggest that any differences that exist among the species of Scoliodon are consistent with isolation by distance. We found no evidence of sympatry among any of the three ‘species’. Presently, we cautiously recommend that S. cf. laticaudus of the Malacca Strait be assigned as S. laticaudus. These results and the updated distributional range of the Scoliodon species are discussed below.

4.1. Taxonomic Conclusions and Recommendations

The decision to resurrect S. macrorhynchos as distinct from S. laticaudus was primarily based on the NADH2 sequence data obtained in White et al. [5]. Recent studies have highlighted that the use of single mitochondrial markers alone to distinguish between species can be questionable, especially in light of discordant species trees using mitochondrial and nuclear DNA (for example, Chimaera ogilbyi in Finucci et al. [47], freshwater snail genus Cipangopaludina in Hirano et al. [48], and terrapins (family Emydidae) in Wiens et al. [49]). In the case of Scoliodon, there is considerable concordance between mitochondrial and nuclear signals to support the conclusion of White et al. [5], i.e., the resurrection of S. macrorhynchos as a valid species and separate from S. laticaudus from India.
Phylogenetic and species trees using combined mitochondrial markers group S. macrorhynchos from Sarawak Borneo and from China together, but the same cannot be said for nuclear markers due to the nonavailability of China sequences. Both mitochondrial and nuclear phylogenetic trees mostly support S. macrorhynchos from Sarawak Borneo as separate from S. cf. laticaudus from the Malacca Strait. The discordance between mitochondrial and nuclear signals arises regarding the relationship of S. cf. laticaudus and S. laticaudus. Ambiguity in individual nuclear signals underscores the need to use multiple genes to infer species relationship, and concatenated nuclear signals provisionally group Scoliodon individuals from the Malacca Strait as S. laticaudus. In addition to congruence between mitochondrial and nuclear data, congruence between molecular and morphological characteristics has also been employed to delimit species (e.g., Finucci et al. [47] and Petean et al. [14]). For Scoliodon, White et al. [5] documented only mean differences in several morphometric characteristics but with ranges partially overlapping, i.e., head length, pre-pectoral length, lower labial furrow length, and second dorsal fin origin to anal fin origin. The more comprehensive morphological data presented in this study did not find any nonoverlapping morphological differences in the Scoliodon specimens examined. However, given the high intraspecific variability in measurements from S. laticaudus (Figure 5a), measurements from additional individuals across a broad distribution range are important to clarify the morphological distinctions between S. laticaudus and S. macrorhynchos.
The available molecular evidence delimits the Malacca Strait as the easternmost boundary for the range of S. laticaudus, thereby extending the distribution of the species based on the most recent International Union for Conservation of Nature (IUCN) assessment [10]. The Malay Peninsula appears to serve as a contemporary physical barrier between the two species. This pattern has been seen for a number of coastal-associated species with limited dispersal abilities, such as bamboosharks [11], guitarfishes [50], groupers [51], sea snails [52], and a number of mangrove species [53]. The molecular differences between morphologically similar but non-sympatric S. macrorhynchos and S. laticaudus suggest a relatively recent divergence due to geographical isolation with limited mixing that drove allopatric speciation, which is feasible given the complexity of the past geological history of the Sundaland region [54]. Further population genetic studies to corroborate this will help shed light on the evolutionary history and biogeography of the species.
Another important aspect to investigate for Scoliodon is the population genetic structure. Scoliodon is one of the top landed sharks in terms of both abundance and biomass in surveyed areas within Malaysia [55,56]. A strong coastal affiliation [7] and limited dispersal due to small size are traits that likely promote genetic differentiation and, thus, increase their vulnerability to localized fishing impacts. A similar pattern of a fine-scale population structure has been revealed for a similar small-sized benthic coastal shark, Chiloscyllium punctatum, that is subject to high fishing pressure in the Southeast Asian region [11,57]. Further investigation into the genetic structure of Scoliodon in Southeast Asia and Indian waters is warranted given the high fishing pressure exerted [58].

4.2. Geographic Range

Distributional ranges for species are often based on a combination of literature sources and expert opinions; therefore, validating some occurrences can be difficult. Since Scoliodon is herein confirmed with two valid species, notwithstanding the possibility of another in the Bay of Bengal, it is important to critically investigate the full distributional range for S. laticaudus and S. macrorhynchos. The identity of Scoliodon at locations without genetic sequences is putatively assigned as either S. cf. laticaudus or S. cf. macrorhynchos using the Malay Peninsula as the genus distribution break. The resulting distributional range is displayed in Figure 6, with questionable occurrences noted. Investigation of the range is discussed below in an east to west direction.
Off Japan, S. cf. macrorhynchos has been recorded as a rare occurrence from Kochi Prefecture [59] (as S. sorrakowah). Although listed as occurring off the Pacific coast of Southern Japan by Nakaya [60] and Nakabo [61], it is noticeably absent from checklists of coastal fishes in prefectures on the Pacific coast of Southern Japan, e.g., Mie [62], Kagoshima [63], and Nagasaki [64]. Furthermore, nine specimens of Scoliodon deposited in Japanese collections with geographic data were caught in either China, Taiwan, or Vietnam (via http://science-net.kahaku.go.jp/, accessed on 28 February 2022). The distribution off Southern Japan appears to be erroneous and should not be included in the range of this species. It has not been previously recorded from South Korea, but Cho et al. [65] reported on a single specimen collected from a Yeosu fish market, Busan in 1995 identified as S. laticaudus and supposedly caught from the South Sea of Korea. Off China, Wang [66] noted that S. macrorhynchos was abundant off Wenzhou in Southern Zhejiang Province in late spring and early summer but rarely caught in the northern part of the province. Zhu et al. [67] also recorded S. macrorhynchos (identified as S. laticaudus) from Zhejiang Province but throughout much of the year. Lam and Sadovy de Micheson [68] found that Scoliodon, identified as S. laticaudus, was the most abundant shark species present during comprehensive market surveys off the Fujian, Hainan, and Guangdong Provinces of China, as well as off Hong Kong. Likewise, Ebert et al. [69] noted that this species was very abundant in fisheries catches around Taiwan.
Naylor et al. [8] provided numerous NADH2 sequences from specimens caught off Vietnam recorded during local ichthyofaunal surveys. Orlov [70] listed Scoliodon spp. as one of the pelagic predators found in marine waters off Cambodia, which likely refers wholly or in part to S. cf. macrorhynchos. Deechum [71] and Springer [72] included records of Scoliodon (identified as S. laticaudus) from the Gulf of Thailand. No Scoliodon individuals were recorded during comprehensive ichthyofaunal surveys along the east coast of Peninsular Malaysia ([56] Lim et al., unpublished data) but are caught in high abundance in the waters of the west coast of Peninsular Malaysia. As verified by Compagno et al. [73], Scoliodon was largely absent in the Philippines. A recent listing of this species in the Philippines elasmobranch identification guide by Alava et al. [74] was likely based on an old record of misidentified Loxodon or Rhizoprionodon. In Malaysian Borneo, none were recorded from off Sabah from multiple fish surveys, but S. macrorhynchos is caught in high abundance off Sarawak ([75] Lim et al., unpublished data, and Manjaji-Matsumoto pers. comm.). Scoliodon was not recorded in shark catches off Bintan Island in the Riau Archipelago of Indonesia just to the southeast of Singapore [76].
In Indonesia, S. cf. macrorhynchos appears to be restricted to Kalimantan [75] and around the river outflows of Eastern Sumatra that flow into the Malacca Strait [77]. It has not been recorded in the literature from West Sumatra or from recent landing site surveys (Fahmi, pers. comm.). Although Bleeker [6] described S. macrorhynchos from a juvenile specimen from off Batavia (= Jakarta), which would have likely been caught locally, it has not been recorded off Java in surveys over the last half a century (e.g., Widodo et al. [78] and Widodo and Mahiswara [79]). Springer [72] also listed a specimen deposited at the Smithsonian Institute (USNM 72479) from Batavia (= Jakarta, West Java). This specimen was collected by Owen Bryant and William Palmer in 1909 during a natural history specimen collection trip [80]. Despite being the most abundant species found in recent surveys of the Muara Baru fishing port in Jakarta [81], these were caught in South Kalimantan and only landed in Jakarta. Due to the lack of accurate baseline information, it is not possible to determine whether Scoliodon has been extirpated from Javan waters due to overexploitation.
Arunrugstichai et al. [26] recorded S. laticaudus as one of the most abundant shark species landed off the Andaman Coast of Thailand. Psomadakis et al. [82] stated that this species is found in coastal waters and lower reaches of the rivers in Myanmar. Jit et al. [83] recorded it as the most abundant shark species based on surveys of two landing centers in Bangladesh, i.e., Chittagong and Cox’s Bazar. Scoliodon laticaudus is abundant off the Indian coastline, with verified records from all coastal states (from east to west): Andaman and Nicobar archipelago [84,85], West Bengal [86], Orissa [87,88], Andhra Pradesh [89], Tamil Nadu [90], Kerala [91], Karnataka [92,93], Goa [94], Maharashtra [95,96], and Gujarat [97]. Scoliodon laticaudus has not been recorded from the Indian union territory of Lakshadweep (formerly Laccadive Archipelago) nor further south in the Maldives or Chagos Archipelago. The presence of S. laticaudus off Sri Lanka is less clear. Some checklists have included this species from Sri Lankan waters, e.g., Misra [98] (as Scoliodon sorrakowah), Mendis [99] (as Carcharias laticaudus), and De Silva [100]. However, recent surveys of 15 fish markets around Sri Lanka recorded no Scoliodon [101]. Likewise, Moron et al. [102] did not include this species as present off the west coast of Sri Lanka. Given that Scoliodon is usually found in abundance where it occurs, its absence is notable in these studies. Thus, it may be absent from Sri Lankan waters or restricted to only the northern part of Sri Lanka around Palk Bay and the Gulf of Mannar, where it is known to be abundant on the respective Indian coastlines. Off Pakistan, S. cf. laticaudus was recorded from the coasts of the Sindh Province (Misra [103] as S. sorrakowah) and a single specimen recorded during port surveys at Jiwani in Westernmost Balochistan Province, close to the Iranian border [104].
The range of Scoliodon has recently included the Persian Gulf and parts of East Africa [7,9]. Bishop [105] and Sivasubramanian and Ibrahim [106] recorded it from off Kuwait and Qatar, respectively, but more recent comprehensive surveys of these locations, as well as of Bahrain and the United Arab Emirates, did not record any S. laticaudus in fisheries landings [107,108,109]. Amojil et al. [110] included this species as only possibly occurring in the Persian Gulf due to the lack of verifiable records. Scoliodon cf. laticaudus was not recorded during comprehensive surveys of fish landing sites in Oman [111,112]. It was also not recorded from catches of Russian trawlers operating off the entire Yemen coast (including Socotra Island) between 1985 and 1990 [113] or in a recent comprehensive survey of the fish fauna of Socotra Islands [114].
Scoliodon cf. laticaudus was included as part of the marine fauna of Somalia [115] and reported as rare in the Somali shark fishery [116]. Although included in a species catalog of Kenya [117], surveys of catches in small-scale fisheries off Kenya over the last decade have not recorded any individuals of this species ([118] B. Kiilu, pers. comm.). Compagno [4] included Tanzania in the range for S. cf. laticaudus and also included it as present in Mozambique [119]. However, this species has not been recorded from fishery bycatches in recent years in either Mozambique or Tanzania (A. Marshall, S. Pierce, C. Rohner, and D. Ebert, pers. comm.). The presence of Scoliodon in the fauna of East Africa from Somalia to Mozambique is dubious. Where S. laticaudus is found, they are typically caught in high numbers and common in coastal waters. It is more likely that they are misidentifications of similar species, e.g., Rhizoprionodon acutus, which was previously referred to as Scoliodon walbeehmi throughout the Indo-West Pacific before being synonymized. Thus, the East Africa distribution of S. laticaudus is treated as dubious.
The present distribution delineation is mostly consistent with the recently published IUCN assessment for S. laticaudus [10] and S. macrorhynchos [120]. In a largely contiguous coastline distribution of Scoliodon (Figure 6), we noted two contemporary spatial ‘breaks’, i.e., along the east coast of the Malaysian Peninsula and off the Sabah coastline of Northeastern Borneo. These breaks could be due to the presence of unsuitable bottom habitats for the species (Manjaji-Matsumoto, pers. comm.) and also reflect the complex evolutionary history of the Sundaland region. Notably, the presence and taxonomic status of Scoliodon in the Indonesian region, especially along Eastern Sumatra and along the Kalimantan coastline (Figure 6), needs to be investigated using an integrative approach, i.e., molecular and morphological analyses. It was hypothesized that animals along Eastern Sumatra are S. laticaudus, while those in Kalimantan waters are S. macrorhynchos, with the Karimata Strait acting as a physical and/or genetic barrier—this is consistent with evidence presented for the genetic structure seen for C. punctatum [11].

5. Conclusions

Collective evidence from mitochondrial DNA, nuclear DNA, and morphological analyses clearly supports the previous resurrection of S. macrorhynchos as distinct species from S. laticaudus. Genetic distinctiveness between the two species is likely a product of isolation by distance with the Malaysian Peninsula acting as a physical barrier. The identity of Scoliodon from Indonesian waters remained unverified and should be the focus for future taxonomic studies. Both Scoliodon species are currently classified as “near threatened” in the IUCN Red List. With the new evidence from this study, we recommend updating the distribution information of these species and investigating the taxonomic status of Scoliodon animals from Indonesian coastal waters.

Supplementary Materials

The following supporting information can be downloaded at www.mdpi.com/article/10.3390/ani12060681/s1: Table S1: Genetic samples used in this study with locality data and GenBank accession numbers for each of the mitochondrial and nuclear markers. Table S2: Best model selected for maximum likelihood and Bayesian inference analysis according to each marker and the combined markers. Table S3: NCBI GenBank and Barcode of Life Data (BOLD) Systems accession number of the reference sequences used in the analyses. Data S1: Collection data for all specimens of Scoliodon examined in this study. Figure S1a: COI gene phylogenetic relationships of Scoliodon species (phylogram). The bootstrap values (ML/BI) are shown at branches. Sequence names in bold were from the present study. Figure S1b: NADH2 gene phylogenetic relationships of Scoliodon species (phylogram). The bootstrap values (ML/BI) are shown at branches. Sequence names in bold were from the present study. Figure S1c: ACT phylogenetic relationships of Scoliodon species (phylogram). The bootstrap values (ML/BI) are shown at branches. Figure S1d: KBTBD2 phylogenetic relationships of Scoliodon species (phylogram). The bootstrap values (ML/BI) are shown at branches. Figure S1e: PROX1 phylogenetic relationships of Scoliodon species (phylogram). The bootstrap values (ML/BI) are shown at branches. Figure S1f: RAG1 phylogenetic relationships of Scoliodon species (phylogram). The bootstrap values (ML/BI) are shown at branches. Figure S1g: RAG2 phylogenetic relationships of Scoliodon species (phylogram). The bootstrap values (ML/BI) are shown at branches. Figure S1h: SCFD1 phylogenetic relationships of Scoliodon species (phylogram). The bootstrap values (ML/BI) are shown at branches. Figure S1i: TOB1 phylogenetic relationships of Scoliodon species (phylogram). The bootstrap values (ML/BI) are shown at branches.

Author Contributions

Designed the study: K.C.L., W.T.W., A.Y.H.T. and K.-H.L. Performed the field work: K.C.L., K.-H.L. and S.A. Conducted the statistical analysis of the data: K.C.L. and W.T.W. Conducted the molecular analyses: K.C.L. and G.J.P.N. Captured morphological and meristic data: W.T.W. and S.A. Wrote the manuscript: K.C.L., W.T.W. and A.Y.H.T. All authors have read and agreed to the published version of the manuscript.

Funding

Financial support for the lead author K.C.L. was provided by the University Malaya Research Grant RP018C-16SUS and the University Malaya Research Fund Assistance BK018-2015; for K.-H.L., funding for rental the facility instruments of molecular analyses in the IOES was provided by the World Wide Fund-Malaysia PV049-2019, the APC was provided by Third Institute of Oceanography, State Oceanic Administration, China IF004-2022 and the UM Research Grant RU009H-2020; for W.T.W., by the CSIRO National Research Collections Australia; and for G.J.P.N., funding for some of the molecular analyses was provided through a US National Science Foundation Division of Environmental Biology Award (#1132229). Financial support to collect specimens was provided by the Australian Centre for Agricultural Research (Indonesian projects FIS/2000/006 and FIS/2003/037), National Science Foundation (Borneo projects DEB 0103640 and DEB 0542846), Murdoch University (internal funding for Taiwan collection trip), CSIRO, and the University of Hong Kong (discretionary funding for guest lecture and collection trip in Hong Kong), Centre for Biodiversity in Peninsular Thailand, and Trocadero group funding (to S.A. for Thailand collection).

Institutional Review Board Statement

The specimens of the spadenose sharks used in this study were all taken from fish market surveys of Southeast Asia between 2001 and 2016. All specimens were dead prior to the surveys being undertaken. The fishing port surveys in Thailand were conducted by SA as part of his Master’s Thesis program through the Prince of Songkhla University, and permission to collect data was granted in accordance with the Thailand Department of Fisheries. The fishing port surveys in Malaysia were conducted by KCL as part of his PhD Thesis program through the University of Malaya. No permission to collect data from Peninsular Malaysia was needed at the time of the study, and permission to collect data from Sarawak waters was granted by the Fisheries Research Institute Sarawak, Department of Fisheries Malaysia. All comparative materials used in this study were deposited in museum collections around the world and were borrowed with official loan documentation from these collections. No live animals were collected or killed during this study. All applicable international, national, and/or institutional guidelines for the care and use of animals were followed by the authors.

Data Availability Statement

Ranges of the morphological data obtained in this study are provided in Table 2. The raw morphological data generated during this current study are available from the corresponding author on reasonable request. All sequences used in this study have been deposited in GenBank, and the related accession numbers are provided in the related figure and text sections.

Acknowledgments

Thanks to J. Caira (University of Connecticut); K. Jensen (University of Kansas); P. Last, G. Yearsley, and J. Stevens (CSIRO); Mabel Manjaji-Matsumoto (Universiti Malaysia Sabah); Annie Lim (Fisheries Biosecurity Centre Sarawak); Fahmi (Indonesian Institute of Sciences); V. Lam and S. Lea (University of Hong Kong); and Dharmadi (Research Centre for Capture Fisheries, Jakarta) for their valuable work in the field. We would also like to acknowledge J. Pogonoski (CSIRO) for capturing the meristic data, H. O’Neill (CSIRO) for providing editorial comments, A. Graham (CSIRO) for providing collection information and registering specimens, and L. Conboy (CSIRO) for image preparation. We would also like to thank the following museum staff for allowing accessing and assisting with specimen examination: M. van Oijen and R. de Ruiter at the Rijksmuseum van Natuurlijke Histoire (RMNH) in Leiden; R. Causse, B. Séret, G. Duhamel, and P. Pruvost at the Muséum national d’Histoire naturelle (MNHN) in Paris; and P. Campbell and J. Maclaine at the British Museum of Natural History (BMNH) in London.

Conflicts of Interest

The authors declare that they have no conflicts of interest.

References

  1. Müller, J.; Henle, F.G.J. Gattungen der Haifische und Rochen nach einer von ihm mit Hrn. Henle unternommenen gemeinschaftlichen Arbeit über die Naturgeschichte der Knorpelfische. Ber. Königlichen Preuss. Akad. Wiss. Berl. 1837, 1837, 111–118. [Google Scholar]
  2. Müller, J.; Henle, F.G.J. Systematische Beschreibung der Plagiostomen; Veit und Comp: Berlin, Germany, 1938–1941; (pp. 1–28 published in 1838, 29–102 in 1839, 103–200 in 1841). [Google Scholar]
  3. Compagno, L.J.V. Sharks of the Order Carcharhiniformes; The Blackburn Press: Caldwell, NJ, USA, 1988. [Google Scholar]
  4. Compagno, L.J.V. FAO Species Catalogue. Volume 4, Sharks of the World. An Annotated and Illustrated Catalogue of Shark Species Known to Date; FAO Fisheries Synopsis No 125; FAO: Rome, Italy, 1984. [Google Scholar]
  5. White, W.T.; Last, P.R.; Naylor, G.J.P. Scoliodon macrorhynchos (Bleeker, 1852), a second species of spadenose shark from the Western Pacific (Carcharhiniformes: Carcharhinidae). In Descriptions of New Australian Chondrichthyans, CSIRO Marine and Atmospheric Research Paper 32; Last, P.R., White, W.T., Pogonoski, J.J., Eds.; CSIRO: Hobart, Australia, 2010; pp. 61–76. [Google Scholar]
  6. Bleeker, P. Bijdrage tot de kennis der Plagiostomen van den Indischen Archipel. Verh. Batav. Genoots. Kuns. 1852, 24, 1–92. [Google Scholar]
  7. Simpfendorfer, C. Scoliodon laticaudus. The IUCN Red List of Threatened Species. 2009. Available online: https://doi.org/10.2305/IUCN.UK.2009-2.RLTS.T39383A10188364.en (accessed on 14 February 2020).
  8. Naylor, G.J.P.; Caira, J.N.; Jensen, K.; Rosana, K.A.M.; White, W.T.; Last, P.R. A DNA sequence-based approach to the identification of shark and ray species and its implications for global elasmobranch diversity and parasitology. Bull. Am. Nat. Hist. Mus. 2012, 367, 1–263. [Google Scholar] [CrossRef]
  9. Ebert, D.A.; Fowler, S.; Compagno, L. Sharks of the World: A Fully Illustrated Guide; Wild Nature Press: Plymouth, UK, 2013. [Google Scholar]
  10. Dulvy, N.K.; Simpfendorfer, C.; Akhilesh, K.V.; Derrick, D.; Elhassan, I.; Fernando, D.; Haque, A.B.; Jabado, R.W.; Maung, A.; Valinassab, T.; et al. Scoliodon laticaudus. The IUCN Red List of Threatened Species. 2021. Available online: https://doi.org/10.2305/IUCN.UK.2021-2.RLTS.T169234201A173436322.en (accessed on 14 December 2021).
  11. Lim, K.C.; Then, A.Y.; Wee, A.K.S.; Sade, A.; Rumpet, R.; Loh, K.-H. Brown banded bamboo shark (Chiloscyllium punctatum) shows high genetic diversity and differentiation in Malaysian waters. Sci. Rep. 2021, 11, 14874. [Google Scholar] [CrossRef] [PubMed]
  12. Puckridge, M.; Last, P.R.; White, W.T.; Andreakis, N. Phylogeography of the Indo-West Pacific maskrays (Dasyatidae, Neotrygon): A complex example of chondrichthyan radiation in the Cenozoic. Ecol. Evol. 2013, 3, 217–232. [Google Scholar] [CrossRef]
  13. White, W.T.; Corrigan, S.; Yang, L.; Henderson, A.C.; Bazinet, A.L.; Swofford, D.L.; Naylor, G.J. Phylogeny of the manta and devilrays (Chondrichthyes: Mobulidae), with an updated taxonomic arrangement for the family. Zool. J. Linnean. Soc. 2018, 182, 50–75. [Google Scholar] [CrossRef] [Green Version]
  14. Petean, F.F.; Naylor, G.J.; Lima, S.M. Integrative taxonomy identifies a new stingray species of the genus Hypanus Rafinesque, 1818 (Dasyatidae, Myliobatiformes), from the Tropical Southwestern Atlantic. J. Fish. Biol. 2020, 97, 1120–1142. [Google Scholar] [CrossRef]
  15. Galtier, N.; Nabholz, B.; Glémin, S.; Hurst, G.D.D. Mitochondrial DNA as a marker of molecular diversity: A reappraisal. Mol. Ecol. 2009, 18, 4541–4550. [Google Scholar] [CrossRef]
  16. Balloux, F. The worm in the fruit of the mitochondrial DNA tree. Heredity 2010, 104, 419–420. [Google Scholar] [CrossRef]
  17. Awadalla, P.; Eyre-Walker, A.; Maynard-Smith, J. Linkage disequilibrium and recombination in hominid mitochondrial DNA. Science 1999, 286, 2524–2525. [Google Scholar] [CrossRef] [Green Version]
  18. Eyre-Walker, A.; Smith, N.H.; Maynard-Smith, J. How clonal are human mitochondria? Proc. R. Soc. B Biol. Sci. 1999, 266, 477–483. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  19. Hagelberg, E.; Goldman, N.; Lió, P.; Whelan, S.; Schiefenhöel, W.; Clegg, J.B.; Bowden, D.K. Evidence for mitochondrial DNA recombination in a human population of island Melanesia. Proc. R. Soc. B Biol. Sci. 1999, 266, 485–492. [Google Scholar] [CrossRef] [Green Version]
  20. Rand, D.M. The units of selection on mitochondrial DNA. Annu. Rev. Ecol. Syst. 2001, 32, 415–448. [Google Scholar] [CrossRef]
  21. Bazin, E.; Glémin, S.; Galtier, N. Population size does not influence mitochondrial genetic diversity in animals. Science 2006, 312, 570–572. [Google Scholar] [CrossRef] [PubMed]
  22. Castoe, T.A.; de Koning, A.J.; Kim, H.M.; Gu, W.; Noonan, B.P.; Naylor, G.; Jiang, Z.J.; Parkinson, C.L.; Pollock, D.D. Evidence for an ancient adaptive episode of convergent molecular evolution. Proc. Natl. Acad. Sci. USA 2009, 106, 8986–8991. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  23. Xu, W.; Jameson, D.; Tang, B.; Higgs, P.G. The relationship between the rate of molecular evolution and the rate of genome rearrangement in animal mitochondrial genomes. J. Mol. Evol. 2006, 63, 375–392. [Google Scholar] [CrossRef] [PubMed]
  24. Nabholz, B.; Glémin, S.; Galtier, N. Strong variations of mitochondrial mutation rate across mammals—The longevity hypothesis. Mol. Biol. Evol. 2008, 25, 120–130. [Google Scholar] [CrossRef] [Green Version]
  25. Nabholz, B.; Glémin, S.; Galtier, N. The erratic mitochondrial clock: Variations of mutation rate, not population size, affect mtDNA diversity across mammals and birds. BMC Evol. Biol. 2009, 9, 54. [Google Scholar] [CrossRef] [Green Version]
  26. Arunrugstichai, S.; True, J.D.; White, W.T. Catch composition and aspects of the biology of sharks caught by Thai commercial fisheries in the Andaman Sea. J. Fish. Biol. 2018, 92, 1487–1504. [Google Scholar] [CrossRef]
  27. Aschliman, N.C.; Cleason, K.M.; McEachran, J.D. Phylogeny of batoidea. In Biology of Sharks and Their Relatives, 2nd ed.; Carrier, J.C., Musick, J.A., Heithaus, M.R., Eds.; CRC Press: Boca Raton, FL, USA, 2012; pp. 57–95. [Google Scholar]
  28. Hyde, J.R.; Lynn, E.; Humphreys, R., Jr.; Musyl, M.; West, A.P.; Vetter, R. Shipboard identification of fish eggs and larvae by multiplex PCR, and a description of fertilized eggs of blue marlin, short bill spearfish, and wahoo. Mar. Ecol. Prog. Ser. 2005, 286, 269–277. [Google Scholar] [CrossRef]
  29. Ward, R.D.; Holmes, B.H.; White, W.T.; Last, P.R. DNA barcoding Australasian chondrichthyans: Results and potential uses in conservation. Mar. Freshwat. Res. 2008, 59, 57–71. [Google Scholar] [CrossRef]
  30. Naylor, G.J.P.; Ryburn, J.A.; Fedrigo, O.; Lopez, J.A. Phylogenetic relationships among the major lineages of modern elasmobranchs. In Reproductive Biology and Phylogeny: Sharks, Skates, Stingrays, and Chimaeras; Hamlett, W.C., Jamieson, B.G.M., Eds.; Science Publishers Inc.: Plymouth, UK, 2005; pp. 1–25. [Google Scholar]
  31. Hall, T.A. BioEdit: A user-friendly biological sequence alignment editor and analysis program for Windows 95/98/NT. Nucleic Acids Symp. Ser. 1999, 41, 95–98. [Google Scholar]
  32. Thompson, J.D.; Gibson, T.J.; Plewniak, F.; Jeanmougin, F.; Higgins, D.G. The CLUSTAL_X windows interface: Flexible strategies for multiple sequence alignment aided by quality analysis tools. Nucleic Acids Res. 1997, 25, 4876–4882. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  33. Tanabe, A.S. Kakusan: A computer program to automate the selection of a nucleotide substitution model and the configuration of a mixed model on multilocus data. Mol. Ecol. Notes 2007, 7, 962–964. [Google Scholar] [CrossRef]
  34. Jobb, G.; von Haeseler, A.; Strimmer, K. Treefinder: A powerful graphical analysis environment for molecular phylogenetics. BMC Evol. Biol. 2004, 4, 18. [Google Scholar] [CrossRef] [Green Version]
  35. Huelsenbeck, J.P.; Ronquist, F. MrBayes: Bayesian Inference of Phylogenetic Trees. Bioinformatics 2001, 17, 754–755. [Google Scholar] [CrossRef] [Green Version]
  36. Swofford, D.L. PAUP*: Phylogenetic Analysis Using Parsimony (* and Other Methods), Version 4; Sinauer Associates: Sunderland, MA, USA, 2002. [Google Scholar]
  37. Zhang, C.; Rabiee, M.; Sayyari, E.; Mirarab, S. ASTRAL-III: Polynomial time species tree reconstruction from partially resolved gene trees. BMC Bioinform. 2018, 19, 153. [Google Scholar] [CrossRef] [Green Version]
  38. Rabiee, M.; Sayyari, E.; Mirarab, S. Multi-Allele Species Reconstruction Using ASTRAL. Mol. Phylogenet. Evol. 2019, 130, 286–296. [Google Scholar] [CrossRef]
  39. Yang, Z.; Rannala, B. Unguided species delimitation using DNA sequence data from multiple loci. Mol. Biol. Evol. 2014, 31, 3125–3135. [Google Scholar] [CrossRef] [Green Version]
  40. Rannala, B.; Yang, Z. Efficient Bayesian species tree inference under the multispecies coalescent. Syst. Biol. 2017, 66, 823–842. [Google Scholar] [CrossRef] [Green Version]
  41. Flouri, T.; Jiao, X.; Rannala, B.; Yang, Z. Species tree inference with BPP using genomic sequences and the multispecies coalescent. Mol. Biol. Evol. 2018, 35, 2585–2593. [Google Scholar] [CrossRef] [PubMed]
  42. Silvestro, D.; Michalak, I. raxmlGUI: A graphical front-end for RAxML. Org. Divers. Evol. 2012, 12, 335–337. [Google Scholar] [CrossRef]
  43. Compagno, L.J.V. Sharks of the World: An Annotated and Illustrated Catalogue of Shark Species Known to Date, Volume 2. Bullhead, Mackerel and Carpet Sharks (Heterodontiformes, Lamniformes and Orectolobiformes); FAO: Rome, Italy, 2001. [Google Scholar]
  44. Compagno, L.J.V. Carcharhinoid Sharks: Morphology, Systematics and Phylogeny. Ph.D. Thesis, Stanford University, Stanford, CA, USA, 1979. [Google Scholar]
  45. Springer, V.G.; Garrick, J.A.F. A survey of vertebral numbers in sharks. Proc. U. S. Natl. Mus. 1964, 116, 73–96. [Google Scholar] [CrossRef]
  46. Clarke, K.R.; Gorley, R.N. PRIMER v6 User Manual/Tutorial; Primer-E Ltd.: Plymouth, UK, 2006. [Google Scholar]
  47. Finucci, B.; White, W.T.; Kemper, J.M.; Naylor, G.J. Redescription of Chimaera ogilbyi (Chimaeriformes; Chimaeridae) from the Indo-Australian region. Zootaxa 2018, 4375, 191–210. [Google Scholar] [CrossRef] [PubMed]
  48. Hirano, T.; Saito, T.; Tsunamoto, Y.; Koseki, J.; Ye, B.; Miura, O.; Suyama, Y.; Chiba, S. Enigmatic incongruence between mtDNA and nDNA revealed by multi-locus phylogenomic analyses in freshwater snails. Sci. Rep. 2019, 9, 6223. [Google Scholar] [CrossRef] [PubMed]
  49. Wiens, J.J.; Kuczynski, C.A.; Stephens, P.R. Discordant mitochondrial and nuclear gene phylogenies in emydid turtles: Implications for speciation and conservation. Biol. J. Linn. Soc. 2010, 99, 445–461. [Google Scholar] [CrossRef] [Green Version]
  50. Giles, J.L.; Riginos, C.; Naylor, G.J.; Dharmadi; Ovenden, J.R. Genetic and phenotypic diversity in the wedgefish Rhynchobatus australiae, a threatened ray of high value in the shark fin trade. Mar. Ecol. Prog. Ser. 2016, 548, 165–180. [Google Scholar] [CrossRef] [Green Version]
  51. Ma, K.Y.; van Herverden, L.; Newman, S.J.; Brumen, M.L.; Choat, J.H.; Chu, K.H.; de Mitcheson, Y.S. Contrasting population genetic structure in three aggregating groupers (Percoidei: Epinephelidae) in the Indo-West Pacific: The importance of reproductive mode. BMC Evol. Biol. 2018, 18, 180. [Google Scholar] [CrossRef]
  52. Reid, D.G.; Lal, K.; Mackenzie-Dodds, J.; Kaligis, F.; Littlewood, D.T.J.; Williams, S.T. Comparative phylogeography and species boundaries in Echinolittorina snails in the central Indo-West Pacific. J. Biogeogr. 2006, 33, 990–1006. [Google Scholar] [CrossRef]
  53. Mantiquilla, J.A.; Shiao, M.S.; Shih, H.C.; Chen, W.H.; Chiang, Y.C. A review on the genetic structure of ecologically and economically important mangrove species in the Indo-West Pacific. Ecol. Genet. Genom. 2021, 18, 100078. [Google Scholar] [CrossRef]
  54. Hall, R. Cenozoic geological and plate tectonic evolution of SE Asia and the SW Pacific: Computer-based reconstructions, model and animations. J. Asian Earth Sci. 2002, 20, 353–431. [Google Scholar] [CrossRef]
  55. Ahmad, A.; Abdul Haris Hilmi, A.A.; Ismail, I. Implementation of the National Plan of Action for Conservation and Management of Shark Resources in Malaysia (Malaysia NPOA-Shark); Terminal Report; SEAFDEC/MFRDMD: Kuala Terengganu, Malaysia, 2015. [Google Scholar]
  56. Arai, T.; Azri, A. Diversity, occurrence and conservation of sharks in the southern South China Sea. PLoS ONE 2019, 14, e0213864. [Google Scholar] [CrossRef] [PubMed]
  57. Fahmi; Tibbetts, I.R.; Bennett, M.B.; Ali, A.; Krajangdara, T.; Dudgeon, C.L. Population structure of the brown-banded bamboo shark, Chiloscyllium punctatum and its relation to fisheries management in the Indo-Malay region. Fish. Res. 2021, 240, 105972. [Google Scholar] [CrossRef]
  58. FAO. FAO Yearbook. Fishery and Aquaculture Statistics 2016; FAO: Rome, Italy, 2018; 104p. [Google Scholar]
  59. Kamohara, T. Revised catalogue of fishes of Kochi Prefecture, Japan. Rep. USA Mar. Biol. Stat. 1964, 11, 1–99. [Google Scholar]
  60. Nakaya, K. Carcharhinidae. In The Fishes of the Japanese Archipelago; Masuda, H., Amaoka, K., Araga, C., Uyeno, T., Yoshino, T., Eds.; Tokai University Press: Tokyo, Japan, 1984; pp. 5–6. [Google Scholar]
  61. Nakabo, T. Fishes of Japan with Pictorial Keys to the Species, 3rd ed.; Tokai University Press: Hadano, Japan, 2013. [Google Scholar]
  62. Okada, Y.; Mori, K. Descriptions and figures of marine fishes obtained at Mie Prefecture, the middle of Honshu, Japan. J. Fac. Fish. Pref. Univ. Mie 1958, 3, 1–39. [Google Scholar]
  63. Kamohara, T. List of fishes from Amami-Oshima and adjacent regions, Kagoshima Prefecture, Japan. Rep. USA Mar. Biol. Stat. 1957, 4, 1–65. [Google Scholar]
  64. Shinohara, G.; Matsuura, K.; Shirai, S. Fishes of Tachibana Bay, Nagasaki, Japan. Mem. Nat. Sci. Mus. Tokyo 1998, 30, 105–138. [Google Scholar]
  65. Cho, H.G.; Kweon, S.M.; Kim, B.J. New record of the spadenose shark, Scoliodon laticaudus (Carcharhiniformes: Carcharhinidae) from South Sea, Korea. Korean J. Ichthyol. 2014, 26, 336–339. [Google Scholar]
  66. Wang, K.F. Preliminary notes on the fishes of Chekiang (Elasmobranches). Contr. Biol. Lab. Sci. Soc. China Zool. Ser. 1933, 9, 87–117. [Google Scholar]
  67. Zhu, J.F.; Dai, X.J.; Li, Y. Preliminary study on biological characteristics of spadenose shark, Scoliodon laticaudus, caught from coastal waters of Zhejiang Province. J. Shanghai Fish. Inst. 2008, 17, 635–639. [Google Scholar]
  68. Lam, V.Y.Y.; Sadovy de Micheson, Y. The sharks of South East Asia—Unknown, unmonitored and unmanaged. Fish Fish. 2010, 12, 51–74. [Google Scholar] [CrossRef]
  69. Ebert, D.A.; White, W.T.; Ho, H.C.; Last, P.R.; Nakaya, K.; Séret, B.; Straube, N.; Naylor, G.J.P.; de Carvalho, M.R. An annotated checklist of the chondrichthyans of Taiwan. Zootaxa 2013, 3752, 279–386. [Google Scholar] [CrossRef] [PubMed]
  70. Orlov, A.M. Brief review of the marine ichthyofauna of Cambodia. J. Ichthyol. 1995, 35, 81–87. [Google Scholar]
  71. Deechum, W. Species Compositions and Some Biological Aspects of Sharks and Rays from the Gulf of Thailand and Andaman Landing Sites. Ph.D. Thesis, Prince of Songkla University, Songkla, Thailand, 2009. [Google Scholar]
  72. Springer, V.G. A revision of the carcharhinid shark genera Scoliodon, Loxodon, and Rhizoprionodon. Proc. U. S. Nat. Mus. 1964, 115, 559–632. [Google Scholar] [CrossRef] [Green Version]
  73. Compagno, L.J.V.; Last, P.R.; Stevens, J.D.; Alava, M.N.R. Checklist of Philippine chondrichthyes. CSIRO Mar. Lab. Rep. 2005, 243, 103. [Google Scholar]
  74. Alava, M.N.R.; Gaudiano, J.P.A.; Utzurrum, J.T.; Capuli, E.E.; Aquino, M.T.R.; Luchvez-Maypa, M.M.A.; Santos, M.D. Pating Ka Ba? In An Identification Guide to Sharks, Batoids and Chimaeras of the Philippines; Department of Agriculture Bureau of Fisheries and Aquatic Resources—National Fisheries Research and Development Institute and the Marine Wildlife Watch of the Philippines: Metro Manila, Philippines, 2014; 200p. [Google Scholar]
  75. Last, P.R.; White, W.T.; Caira, J.N.; Jensen, K.; Lim, A.P.K.; Manjaji-Matsumoto, B.M.; Naylor, G.J.P.; Pogonoski, J.J.; Stevens, J.D.; Yearsley, G.K. Sharks and Rays of Borneo; CSIRO Publishing: Melbourne, Australia, 2010. [Google Scholar]
  76. Emiliya, P.A.; Putra, R.D. Identifikasi jenis hiu Hasil Tangkapan Nelayan di Pulau Bintan Provinsi Kepulauan Riau [Identification of the Type Shark Fishermen Catch in Bintan Island Riau Islands Province]; Project Report; Universitas Maritim Raja Ali Haji: Kota Tanjung Pinang, Indonesia, 2017. [Google Scholar]
  77. Teshima, K.; Ahmad, M.; Mizue, K. Studies on sharks—XIV. Reproduction in the Telok Anson shark collected from Perak River, Malaysia. Jpn. J. Ichthyol. 1978, 25, 181–189. [Google Scholar]
  78. Widodo, J.; Pralampita, W.A.; Chodriyah, U. Length-weight relationships and condition factors of sharks landed from the Indian Ocean south of Java, Bali, and Lombok, Indonesia. In Proceedings of the First Annual Meeting on Artisanal Shark and Rays Fisheries in East Indonesia: Their Socio-Economic and Fishery Characteristics and relationship to Australian Resources, Perth, Australia, 4–5 April 2002. [Google Scholar]
  79. Widodo, A.A.; Mahiswara, M. Sumberdaya ikan cucut (hiu) yang tertangkap nelayan di perairan Laut Jawa [The shark resource caught by fishermen in Java Sea]. J. Iktiol. Indones. 2007, 7, 15–21. [Google Scholar]
  80. Bean, B.A.; Weed, A.C. Notes on a collection of fishes from Java, made by Owen Bryant & William Palmer in 1909. Proc. U. S. Nat. Mus. 1912, 42, 587–611. [Google Scholar]
  81. White, W.T. Catch composition and reproductive biology of whaler sharks (Carcharhiniformes: Carcharhinidae) caught by fisheries in Indonesia. J. Fish Biol. 2007, 71, 1512–1540. [Google Scholar] [CrossRef]
  82. Psomadakis, P.N.; Thein, H.; Russell, B.C.; Tun, M.T. Field Identification Guide to the Living Marine Resources of Myanmar, FAO Species Identification Guide for Fishery Purposes; FAO: Rome, Italy, 2019. [Google Scholar]
  83. Jit, R.B.; Alam, M.F.; Rhaman, M.G.; Singha, N.K.; Akhtar, A. Landing trends, species composition and percentage composition of sharks and rays in Chittagong and Cox’s Bazar, Bangladesh. Int. J. Adv. Res. Biol. Sci. 2014, 1, 81–93. [Google Scholar]
  84. Kumar, R.R.; Venu, S.; Akhilesh, K.V.; Bineesh, K.K.; Rajan, P.T. First report of four deep-sea chondrichthyans (Elasmobranchii and Holocephali) from Andaman waters, India with an updated checklist from the region. Acta Ichthyol. Piscat. 2018, 48, 289–301. [Google Scholar] [CrossRef] [Green Version]
  85. Tyabi, Z.; Jabado, R.W.; Sutaria, D. New records of sharks (Elasmobranchii) from the Andaman and Nicobar Archipelago in India with notes on current checklists. Biodivers. Data J. 2018, 6, e28593. [Google Scholar] [CrossRef] [PubMed]
  86. Sen, S.; Dash, G.; Mukherjee, I. An overview of elasmobranch fisheries of West Bengal in 2018. Mar. Fish Infor. Serv. Tech. Ext. Ser. 2018, 238, 18–22. [Google Scholar]
  87. Barman, R.P.; Mishra, S.S.; Kar, S.; Mukherjee, P.; Saren, S.C. Marine and estuarine fish fauna of Orissa. Rec. Zool. Surv. India Occas. Pap. 2007, 260, 1–186. [Google Scholar]
  88. Talwar, P.K. A contribution to the taxonomy of Rhizoprionodon oligolinx Springer 1964: An important component of the shark fishery of Orissa, India. Indian J. Fish. 1974, 21, 604–607. [Google Scholar]
  89. Rao, S.C.V.S. Scientific, common and local names of commercially important edible marine fin and shell fishes of Andhra Pradesh. Mar. Fish. Infor. Serv. Tech. Ext. Ser. 1991, 108, 1–10. [Google Scholar]
  90. Joshi, K.K.; Sreeram, M.P.; Zacharia, P.U.; Abdussamad, E.M.; Varghese, M.; Habeeb Mohammed, O.M.M.J.; Jayabalan, K.; Kanthan, K.P.; Kannan, K.; Sreekumar, K.M.; et al. Check list of fishes of the Gulf of Mannar ecosystem, Tamil Nadu, India. J. Mar. Biol. Assoc. India 2016, 58, 34–54. [Google Scholar] [CrossRef] [Green Version]
  91. Bineesh, K.K.; Gopalakrishnan, A.; Akhilesh, K.V.; Sajeela, K.A.; Abdussamad, E.M.; Pillai, N.G.K.; Basheer, V.S.; Jena, J.K.; Ward, R.D. DNA barcoding reveals species composition of sharks and rays in the Indian commercial fishery. Mitochondrial DNA A 2016, 28, 458–472. [Google Scholar] [CrossRef]
  92. Kulkarni, G.N.; Shanbhogue, S.L.; Udupa, K.S. Length-weight relationship of Scoliodon laticaudus Muller and Henle and Carcharhinus limbatus (Muller and Henle), from Dakshina Kannada coast. Indian J. Fish. 1988, 35, 300–301. [Google Scholar]
  93. Veena, S.; Thomas, S.; Raje, S.G.; Durgekar, R. Case of leucism in the spadenose shark, Scoliodon laticaudus (Muller & Henle, 1838) from Mangalore, Karnataka. Indian J. Fish. 2011, 58, 109–112. [Google Scholar]
  94. Pillai, P.P.; Parakkal, B. Pelagic sharks in the Indian seas their exploitation, trade, management and conservation. CMFRI Spec. Publ. 2000, 70, 1–95. [Google Scholar]
  95. Nair, R.V.; Appukuttan, K.K.; Rajapandian, M.E. On the systematics and identity of four pelagic sharks of the family Carcharhinidae from Indian region. Indian J. Fish. 1974, 21, 220–232. [Google Scholar]
  96. Mathew, C.J.; Devaraj, M. The biology and population dynamics of the spadenose shark Scoliodon laticaudus in the coastal waters of Maharastra State, India. Indian J. Fish. 1997, 44, 11–27. [Google Scholar]
  97. Fofandi, M.D.; Zala, M.; Koya, M. Observations on selected biological aspects of the spadenose shark (Scoliodon laticaudus Müller & Henle, 1838), landed along Saurashtra coast. Indian J. Fish. 2013, 60, 51–54. [Google Scholar]
  98. Misra, K.S. A check list of the fishes of India, Burma & Ceylon. Part I. Elasmobranchii and Holocephali. Rec. Indian Mus. 1947, 45, 1–46. [Google Scholar]
  99. Mendis, A.S. Fishes of Ceylon: A catalogue, key & bibliography. Fish. Res. Stat. Bull. 1954, 2, 1–222. [Google Scholar]
  100. De Silva, R.I. Taxonomy and status of the sharks and rays of Sri Lanka. Fauna Sri Lanka 2006, 2006, 294–301. [Google Scholar]
  101. Fernando, D.; Bown, R.M.K.; Tanna, A.; Gobiraj, R.; Ralicki, H.; Jockusch, E.L.; Ebert, D.A.; Jensen, K.; Caira, J.N. New insights into the identities of the elasmobranch fauna of Sri Lanka. Zootaxa 2019, 4585, 201–238. [Google Scholar] [CrossRef] [Green Version]
  102. Moron, J.; Bertrand, B.; Last, P. A check-list of sharks and rays of western Sri Lanka. J. Mar. Biol. Assoc. India 1998, 40, 142–157. [Google Scholar]
  103. Misra, K.S. An aid to the identification of the common commercial fishes of India and Pakistan. Rec. Indian Mus. 1962, 57, 1–320. [Google Scholar]
  104. Gore, M.; Waqas, U.; Khan, M.M.; Ahmad, E.; Baloch, A.S.; Baloch, A.R. A first account of the elasmobranch fishery of Balochistan, south-west Pakistan. West. Indian Ocean. J. Mar. Sci. 2019, 18, 95–105. [Google Scholar] [CrossRef] [Green Version]
  105. Bishop, J.M. History and current checklist of Kuwait’s ichthyofauna. J. Arid. Environ. 2003, 54, 237–256. [Google Scholar] [CrossRef]
  106. Sivasubramanian, K.; Ibrahim, M.A. Common Fishes of Qatar. Scientific Atlas of Qatar 1; Doha Modern Printing Press: Doha, Qatar, 1982. [Google Scholar]
  107. Moore, A.B.M.; McCarthy, I.D.; Carvalho, G.R.; Peirce, R. Species, sex, size and male maturity composition of previously unreported elasmobranch landings in Kuwait, Qatar and Abu Dhabi Emirate. J. Fish Biol. 2012, 80, 1619–1642. [Google Scholar] [CrossRef] [PubMed]
  108. Moore, A.B.M.; Peirce, R. Composition of elasmobranch landings in Bahrain. Afr. J. Mar. Sci. 2013, 35, 593–596. [Google Scholar] [CrossRef]
  109. Jabado, R.W.; Al Ghais, S.M.; Hamza, W.; Shivji, M.S.; Henderson, A.C. Shark diversity in the Arabian/Persian Gulf higher than previously thought: Insights based on species composition of shark landings in the United Arab Emirates. Mar. Biodiv. 2015, 45, 719–731. [Google Scholar] [CrossRef]
  110. Almojil, D.K.; Moore, A.B.M.; White, W.T. Sharks and Rays of the Arabian/Persian Gulf; MBG (INT) Ltd.: London, UK, 2015. [Google Scholar]
  111. Henderson, A.C.; McIlwain, J.L.; Al-Oufi, H.S.; Al-Sheili, S. The Sultanate of Oman shark fishery: Species composition, seasonality and diversity. Fish. Res. 2007, 86, 159–168. [Google Scholar] [CrossRef]
  112. Henderson, A.C.; McIlwain, J.L.; Al-Oufi, H.S.; Al-Sheile, S.; Al-Abri, N. Size distributions and sex ratios of sharks caught by Oman’s artisanal fishery. Afr. J. Mar. Sci. 2009, 31, 233–239. [Google Scholar] [CrossRef]
  113. Al Sakaff, H.; Esseen, M. Occurrence and distribution of fish species off Yemen (Gulf of Aden and Arabian Sea). Naga ICLARM Q. 1999, 22, 43–47. [Google Scholar]
  114. Zajonz, U.; Lavergne, E.; Bogorodsky, S.V.; Saeed, F.N.; Aideed, M.S.; Krupp, F. Coastal fish diversity of the Socotra Archipelago, Yemen. Zootaxa 2019, 4636, 1–108. [Google Scholar] [CrossRef] [Green Version]
  115. Sommer, C.; Schneider, W.; Poutiers, J.M. Living Marine Resources of Somalia. FAO Species Identification Field Guide for Fishery Purposes; FAO: Rome, Italy, 1996. [Google Scholar]
  116. Marshall, N.T. The Somali shark fishery in the Gulf of Aden and the Western Indian Ocean. In Trade in Sharks and Shark Products in the Western Indian and Southeast Atlantic Oceans; Marshall, N.T., Barnes, R., Eds.; TRAFFIC East/Southern Africa: Nairobi, Kenya, 1997; pp. 24–30. [Google Scholar]
  117. Anam, R.; Mostarda, E. Field Identification Guide to the Living Marine Resources of Kenya. FAO Species Identification Field Guide for Fishery Purposes; FAO: Rome, Italy, 2012. [Google Scholar]
  118. Kiilu, B.K.; Ndegwa, S. Shark Bycatch—Small Scale Tuna Fishery Interactions along the Kenyan Coast; IOTC-2013-WPEB09-13; Indian Ocean Tuna Commission (IOTC): Victoria, Seychelles, 2013. [Google Scholar]
  119. Fischer, W.; Sousa, I.; Silva, C.; de Freitas, A.; Poutiers, J.M.; Schneider, W.; Borges, T.C.; Feral, J.P.; Massinga, A. Fichas FAO de Identificaçao de Espécies para Actividades de Pesca. Guia de Campo Das Espécies Comerciais Marinhas e de águas Salobras de Moçambique; FAO: Rome, Italy, 1990. [Google Scholar]
  120. Rigby, C.L.; Bin Ali, A.; Bineesh, K.K.; Chen, X.; Derrick, D.; Dharmadi Ebert, D.A.; Fahmi Fernando, D.; Gautama, D.A.; Haque, A.B.; Herman, K.; et al. Scoliodon macrorhynchos. The IUCN Red List of Threatened Species. 2020, p. e.T169233669A169233911. Available online: https://doi.org/10.2305/IUCN.UK.2020-3.RLTS.T169233669A169233911.en (accessed on 13 December 2021).
Figure 1. Lateral view of Scoliodon ‘species’. (a) S. macrorhynchos IPPS WWPLAL#1 (adult male 426 mm TL, fresh), (b) S. cf. laticaudus CSIRO H 8401-09 (adult male 394 mm TL), and (c) S. laticaudus MNHN 1123 (female 524 mm TL, preserved).
Figure 1. Lateral view of Scoliodon ‘species’. (a) S. macrorhynchos IPPS WWPLAL#1 (adult male 426 mm TL, fresh), (b) S. cf. laticaudus CSIRO H 8401-09 (adult male 394 mm TL), and (c) S. laticaudus MNHN 1123 (female 524 mm TL, preserved).
Animals 12 00681 g001
Figure 2. NADH2COI gene mid-point rooting phylogenetic relationships of Scoliodon ‘species’ (phylogram). The bootstrap values (ML/BI) are shown at branches. Sequence names in bold are from the present study.
Figure 2. NADH2COI gene mid-point rooting phylogenetic relationships of Scoliodon ‘species’ (phylogram). The bootstrap values (ML/BI) are shown at branches. Sequence names in bold are from the present study.
Animals 12 00681 g002
Figure 3. Nuclear gene mid-point rooting phylogenetic relationships of Scoliodon ‘species’ (phylogram). The bootstrap values (ML/BI) are shown at branches.
Figure 3. Nuclear gene mid-point rooting phylogenetic relationships of Scoliodon ‘species’ (phylogram). The bootstrap values (ML/BI) are shown at branches.
Animals 12 00681 g003
Figure 4. ASTRAL species tree of Scoliodon species for (a) mtDNA and (b) nuclear DNA.
Figure 4. ASTRAL species tree of Scoliodon species for (a) mtDNA and (b) nuclear DNA.
Animals 12 00681 g004
Figure 5. Nonmetric multidimensional scaling (MDS) ordination of Scoliodon ‘species’ morphometric percentages (% TL): (a) coded by species and (b) coded by size class.
Figure 5. Nonmetric multidimensional scaling (MDS) ordination of Scoliodon ‘species’ morphometric percentages (% TL): (a) coded by species and (b) coded by size class.
Animals 12 00681 g005
Figure 6. Map of the Indo-West Pacific region showing the refined range of Scoliodon species based on the materials examined and a critical examination of the literature. Dubious range locations are highlighted with a question mark. Red = S. laticaudus, blue = S. macrorhynchos, and green = Scoliodon sp. (verification needed).
Figure 6. Map of the Indo-West Pacific region showing the refined range of Scoliodon species based on the materials examined and a critical examination of the literature. Dubious range locations are highlighted with a question mark. Red = S. laticaudus, blue = S. macrorhynchos, and green = Scoliodon sp. (verification needed).
Animals 12 00681 g006
Table 1. Primers used in this study and their references.
Table 1. Primers used in this study and their references.
MarkerForward Primer (5′–3′)Reverse Primer (5′–3′)References
COIFishF2—TCG ACT AAT CAT AAA GAT ATC GGC ACFishR2—ACT TCA GGG TGA CCG AAG AAT CAG AAWard et al. [29]
NADH2ILEM—AAG GAG CAG TTT GAT AGA GTASNM—AAC GCT TAG CTG TTA ATT AANaylor et al. [30]
ACTACT-F—GCT TTC ATC TCC TTC GGC AGT TTGACT-R—CCA CTG GTA ATT GGG ATA CTT GGCDesign based on GN’s sequence of sample GN1680
KBTBD2KBT-F—CTC AGT ATC TAT CTT CAG TCC TTG GCKBT-R—GCT CTT ACA CAG GGA TCA GAG TAG CDesign based on GN’s sequence of sample GN1680
PROX1PRO1-F—AAT TCT TCA AGG GAA AGT GCC CAA GPRO1-R—CAG ACT GCT CCG ACG AGT TTT TGDesign based on GN’s sequence of sample GN1680
RAG1RAG1-F—CTT ATT CAA ACC ATC AAC AAC ACA ACARAG1-R—CTG CAT GAC TGC TTC CAA CTC ATCDesign based on GN’s sequence of sample GN1680
RAG2RAG2-F—TCA GAA TCA AAC AGC CTC ATT TAC CRAG2-R—TTA ATT TCA TTG GAC CAT TCT GGG GDesign based on GN’s sequence of sample GN1680
SCFD2SCFD-F—AGG TGA AAG CGG TAT TTG TGG TGSCFD-R—TGA GCT GCA GAA CTT CAA ACA TAGDesign based on GN’s sequence of sample GN1680
TOB1TOB1-F—ATA TGA AGG TCA CTG GTA TCC AGA CTOB1-R—GAA AAC AAA CTC CTT GGC ATT GGG ADesign based on GN’s sequence of sample GN1680
Table 2. Ranges of proportional dimensions as percentages of the total length for the three ‘species’ of Scoliodon.
Table 2. Ranges of proportional dimensions as percentages of the total length for the three ‘species’ of Scoliodon.
S. laticaudusS. cf. laticaudusS. macrorhynchos
n = 8n = 32n = 34
Min.Max.Min.Max.Min.Max.
Total length (mm)169524239490227562
Precaudal length75.678.075.378.073.678.0
Pre-second dorsal length62.665.462.966.761.566.5
Pre-first dorsal length35.138.833.037.733.038.1
Head length21.529.121.526.321.325.6
Head length (horiz)21.028.621.025.120.925.0
Pre-branchial length17.123.517.120.416.520.7
Pre-branchial length (horiz)16.622.616.519.816.019.5
Preorbital length8.912.68.911.78.511.6
Preorbital length (horiz)8.111.37.910.87.010.7
Preoral length7.111.17.110.47.210.4
Pre-narial length6.69.16.68.76.28.4
Pre-narial length (horiz)5.98.25.68.14.87.8
Pre-pectoral length22.126.421.526.620.126.2
Pre-pelvic length43.948.443.950.243.849.2
Snout–vent length45.949.245.951.445.450.6
Preanal length56.759.956.762.054.860.4
Interdorsal space16.121.717.921.717.922.2
Dorsal-caudal space7.29.37.29.97.29.4
Pectoral–pelvic space16.719.716.920.716.821.6
Pelvic–anal space5.29.05.611.14.88.7
Anal–caudal space6.49.16.48.86.49.1
Eye length1.52.21.62.51.32.4
Eye height1.32.51.32.51.52.2
Interorbital space7.411.27.49.87.510.3
Nostril width1.42.01.52.31.42.3
Internarial space4.96.94.96.74.96.5
Anterior nasal flap length0.20.60.20.50.20.6
Mouth length4.55.64.14.93.55.2
Mouth width6.07.65.37.65.77.6
Upper labial furrow length0.20.60.10.50.10.5
Lower labial furrow length0.81.20.21.50.31.4
First gill slit height2.33.12.34.12.24.0
Second gill slit height2.33.62.12.62.23.2
Third gill slit height2.43.82.24.72.34.4
Fourth gill slit height2.43.72.02.82.43.3
Fifth gill slit height2.23.22.13.32.33.1
Intergill length4.65.94.65.44.56.4
Head height6.110.27.79.97.010.6
Trunk height7.910.88.310.87.813.1
Abdomen height7.511.210.011.49.413.9
Tail height6.310.27.09.47.511.3
Caudal peduncle height3.94.53.84.54.05.0
Head width7.39.46.99.97.910.8
Trunk width6.48.56.58.86.211.8
Abdomen width5.27.14.96.95.28.9
Tail width4.15.64.25.64.66.5
Caudal peduncle width1.92.72.33.52.23.7
Pectoral length10.212.19.811.69.911.7
Pectoral anterior margin9.512.19.411.59.211.9
Pectoral base4.56.65.26.44.86.6
Pectoral height7.810.37.410.37.510.1
Pectoral inner margin5.26.24.66.44.36.2
Pectoral posterior margin6.310.66.812.56.89.8
Pelvic length7.38.77.18.96.98.3
Pelvic anterior margin4.75.44.36.04.35.6
Pelvic base4.75.64.36.34.36.1
Pelvic height3.24.32.34.42.74.2
Pelvic inner margin length2.23.72.13.92.23.5
Pelvic posterior margin length3.45.33.45.33.85.1
Clasper outer length6.09.04.510.24.010.0
Clasper inner length8.411.86.412.46.512.1
Clasper base width1.01.40.61.70.61.4
First dorsal length13.315.613.315.712.915.5
First dorsal anterior margin11.113.511.814.311.214.6
First dorsal base8.910.98.911.48.811.0
First dorsal height6.68.65.88.86.59.0
First dorsal inner margin3.85.13.95.33.54.9
First dorsal posterior margin6.79.25.79.06.28.9
Second dorsal length7.59.37.39.16.98.6
Second dorsal Anterior margin4.15.53.45.53.45.0
Second dorsal base4.04.83.24.93.24.8
Second dorsal height1.72.21.22.41.32.0
Second dorsal inner margin3.24.73.85.13.34.8
Second dorsal posterior margin3.85.33.94.93.64.7
Anal length11.413.59.613.710.814.1
Anal anterior margin5.16.74.17.04.97.8
Anal base8.010.36.110.37.211.2
Anal height2.83.72.23.62.63.8
Anal Inner margin3.03.93.03.92.84.1
Anal posterior margin6.68.45.88.86.58.9
Dorsal caudal margin22.024.921.624.621.925.6
Pre-ventral caudal margin8.510.27.810.78.010.5
Lower post-ventral caudal margin3.44.72.95.02.94.8
Upper post-ventral caudal margin9.511.58.911.09.112.3
Caudal fork width5.47.55.46.85.97.1
Caudal fork length7.89.78.09.87.89.8
Subterminal caudal margin3.95.63.14.73.15.3
Subterminal caudal width2.63.42.63.52.73.4
Terminal caudal margin4.57.44.86.84.97.3
Terminal caudal lobe7.68.96.88.67.29.3
Second dorsal origin4.66.93.06.95.29.1
Second dorsal insertion0.52.20.62.00.62.7
Mid-base first dorsal fin to pectoral insertion10.912.710.513.411.014.6
Mid-base first dorsal fin to pelvic origin4.46.24.67.94.47.6
First dorsal insertion to pelvic mid-base2.83.92.85.41.95.1
Pelvic mid-base to second dorsal origin12.918.113.618.113.519.0
Table 3. Genetic distance range (mean, in percent) among monophyletic groups in mitochondrial DNA and nuclear DNA phylogenetic trees. Slat—Scoliodon laticaudus, Scflat—S. cf. laticaudus, and Smac—S. macrorhynchos.
Table 3. Genetic distance range (mean, in percent) among monophyletic groups in mitochondrial DNA and nuclear DNA phylogenetic trees. Slat—Scoliodon laticaudus, Scflat—S. cf. laticaudus, and Smac—S. macrorhynchos.
Slat-ScflatSlat-SmacScflat-Smac
COI0.82 (0.61–1.53)2.35 (1.99–2.75)2.29 (2.14–3.06)
NADH23.05 (2.98–3.27)3.06 (2.98–3.26)3.64 (3.46–4.23)
Mitochondrial2.16 (2.12–2.18)2.82 (2.71–2.89)3.05 (2.95–3.18)
ACT0.10 (0.00–0.25)0.50 (0.50–0.50)0.50 (0.25–0.74)
KBTBD20.00 (0.00–0.00)0.22 (0.22–0.22)0.22 (0.22–0.22)
PROX10.00 (0.00–0.00)0.02 (0.00–0.11)0.02 (0.00–0.11)
RAG10.12 (0.12–0.12)0.12 (0.12–0.12)0.02 (0.00–0.12)
RAG20.54 (0.45–0.61)0.91 (0.91–0.91)0.58 (0.45–0.61)
SCFD20.13 (0.00–0.21)0.21 (0.21–0.21)0.17 (0.00–0.42)
TOB10.00 (0.00–0.00)0.00 (0.00–0.00)0.00 (0.00–0.00)
Nuclear0.12 (0.10–0.14)0.25 (0.25–0.25)0.19 (0.16–0.21)
Publisher’s Note: MDPI stays neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Share and Cite

MDPI and ACS Style

Lim, K.C.; White, W.T.; Then, A.Y.H.; Naylor, G.J.P.; Arunrugstichai, S.; Loh, K.-H. Integrated Taxonomy Revealed Genetic Differences in Morphologically Similar and Non-Sympatric Scoliodon macrorhynchos and S. laticaudus. Animals 2022, 12, 681. https://doi.org/10.3390/ani12060681

AMA Style

Lim KC, White WT, Then AYH, Naylor GJP, Arunrugstichai S, Loh K-H. Integrated Taxonomy Revealed Genetic Differences in Morphologically Similar and Non-Sympatric Scoliodon macrorhynchos and S. laticaudus. Animals. 2022; 12(6):681. https://doi.org/10.3390/ani12060681

Chicago/Turabian Style

Lim, Kean Chong, William T. White, Amy Y. H. Then, Gavin J. P. Naylor, Sirachai Arunrugstichai, and Kar-Hoe Loh. 2022. "Integrated Taxonomy Revealed Genetic Differences in Morphologically Similar and Non-Sympatric Scoliodon macrorhynchos and S. laticaudus" Animals 12, no. 6: 681. https://doi.org/10.3390/ani12060681

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