Research Article
Print
Research Article
Species delimitation, molecular phylogeny and historical biogeography of the sweetlips fish (Perciformes, Haemulidae)
expand article infoEhsan Damadi, Faezeh Yazdani Moghaddam, Mehdi Ghanbarifardi§
‡ Ferdowsi University of Mashhad, Mashhad, Iran
§ University of Sistan and Baluchestan, Zahedan, Iran
Open Access

Abstract

The subfamily Plectorhinchinae (sweetlips) is composed of poorly-known species with high commercially and ecologically values that exhibit phenotypic plasticity and various morphologies. Few studies have assessed the validity of sweetlips, intergeneric relationships and evolutionary survey in this subfamily, which have not yet been resolved. This study investigated the DNA sequences of (1) the mitochondrial COI gene to delimit species, and (2) two mitochondrial (COI and Cyt b), and one nuclear (RAG1) markers to infer phylogenetic relationships and evolutionary and biogeographic history. The molecular results could differentiate Diagramma punctatum from the other species, but failed to distinguish D. labiosum as a distinct species with considerably lower genetic distances for the COI (0.53%) and Cyt b (0.51%) markers. However, additional taxonomic investigations are required to shed light on this issue. All previously described nominal species of sweetlips in the northwest Indian Ocean were found to be well supported. The monophyly of Plectorhinchus is not supported and Diagramma pictum and D. punctatum should be assigned to the genus Plectorhinchus. The biogeographic history of Plectorhinchinae likely originated in the Indo-Pacific ca. 34 Ma (30–39 Ma; late Eocene/ middle Oligocene) and subsequently colonised the Western Indian Ocean and the Central Indo-Pacific. Maximum diversification within the subfamily occurred from the middle Miocene to Pliocene, coinciding with dispersal and vicariance events. Diversification was probably driven by both biological and geographical factors.

Key Words

biogeography, mito-nuclear, molecular systematics, Plectorhinchus

Introduction

Accurate species delimitation and phylogenetic reconstruction are vital to understand biodiversity assessments, conservation management, evolutionary patterns, and processes (Traldi et al. 2020; McCord et al. 2021). Members of the family Haemulidae Gill, 1885 are dominant in shallow coral or rocky reefs of tropical and subtropical waters worldwide (McKay 2001; Lindeman and Toxey 2003; Nelson et al. 2016). They also have commercial and ecological values for most nations (Rocha et al. 2008). The family comprises two subfamilies Haemulinae Gill, 1885 and Plectorhinchinae Jordan & Thompson, 1912, 21 genera and 136 species (Fricke et al. 2022). The Plectorhinchinae commonly known as sweetlips, inhabit the shallow coastal waters of the Indo-West Pacific and Eastern Atlantic, which currently comprises 42 recognised species (Fricke et al. 2022). The species are classified into four genera (Diagramma Oken, 1817, Genyatremus Gill, 1862, Parapristipoma Bleeker, 1873 and Plectorhinchus Lacépède, 1801), of which Plectorhinchus is the most speciose (Fricke et al. 2022); however, previous molecular studies have shown that Genyatremus does not belong to the Plectorhinchinae (Bernardi et al. 2008; Sanciangco et al. 2011; Tavera et al. 2018). Ever since the validation of Diagramma and Plectorhinchus, the identification of haemulid species, including several cases of re-descriptions and re-allocations has been intensively debated (Johnson et al. 2001; Johnson and Wilmer 2015; Tavera et al. 2018; Damadi et al. 2020). In reality, further research on phylogenetic relationships and taxonomic treatment which, in the past, has been less identified in these genera of Haemulidae is required to figure out the actual number of species because of nominal and complex species as well as the intergeneric relationships between these groups of fishes (Johnson and Wilmer 2015; Tavera et al. 2018). Moreover, many haemulid species manifested morphological diversity in colouration and markings with developmental stages, for example, Plectorhinchus species, or population variation associated with changes in environmental conditions, such as Diagramma pictum (McKay 1983; Liang et al. 2016). Thus, adopting a multiple analytical approach to DNA sequence data can be promising (Padial et al. 2010; Carstens et al. 2013) because it evaluates the morpho-molecular diversity to understand their phylogenetic relationships and lowers the possibility of homoplasy (Nei and Kumar 2000; Tavera et al. 2011; Varón-González et al. 2020). The genera Diagramma and Plectorhinchus comprise 35 valid species worldwide (Fricke et al. 2022). Recent molecular studies confirmed that the subfamily Plectorhinchinae is monophyletic, but the phylogenetic relationships amongst some genera are challenged by those retrieved by morphological studies (Sanciangco et al. 2011; Tavera et al. 2018). Johnson et al. (2001) examined the genus Diagramma and five geographically constrained subspecies recognised within D. pictum (Thunberg, 1792), based on a morphological approach: D. p. pictum from Indo-Malay Archipelago to Central Indo-Pacific excluding Australia (type locality: Japan); D. p. labiosum from northern Australia to South New Guinea (type locality: Australia, Queensland); D. p. cineraseens from the Persian Gulf to the Bay of Bengal (type locality: Sri Lanka, Trincomalee); D. p. punctatum from the Red Sea (type locality: northern Red Sea) and D. p. centurio occurring along the African coast (type locality: Seychelles). The taxonomic position of Diagramma members has been challenged (Allen 1997; Johnson et al. 2001; Heemstra and Heemstra 2004; Bogorodsky et al. 2014; Johnson and Wilmer 2015; Bray 2017; Golani and Fricke 2018). Recently, many subspecies have been promoted to the status of species rank (Fricke et al. 2022). After examining several species, Tavera et al. (2018) proposed to classify them under Plectorhinchus, but it is still necessary to investigate the remaining nominal species (D. labiosum Macleay, 1883, D. punctatum Cuvier, 1830 and D. melanacrum Johnson & Randall, 2001). Of 42 species, 11 species are morphologically known from the northwest Indian Ocean, including Diagramma pictum, D. punctatum, Plectorhinchus albovittatus (Rüppell 1838), P. flavomaculatus (Cuvier, 1830), P. playfairi (Pellegrin, 1914), P. gaterinus (Forsskål, 1775), P. gibbosus (Lacepède, 1802), P. pictus (Tortonese, 1936), P. sordidus (Klunzinger, 1870), P. schotaf (Forsskål, 1775) and P. makranensis Damadi, Yazdani Moghaddam, Ghassemzadeh & Ghanbarifardi, 2020. A comprehensive analysis of all species in this region has not been done. Some of these species have limited distribution, such as D. punctatum (Red Sea) and Plectorhinchus makranensis (the Persian Gulf to the Gulf of Oman) (Zajonz et al. 2019; Damadi et al. 2020). In addition, few genetic studies have explored haemulid species, including DNA barcoding analysis in a few taxa from the northwest Indian Ocean (Asgharian et al. 2011; Rabaoui et al. 2019). The main goals of this study are as follows: (1) to test the validity of the northwest Indian Ocean taxa proposed by morphological classification, (2) to utilise the wide geographic coverage of sweetlips species using species-delimitation algorithms to shed light on their molecular operational taxonomic units (MOTUs), (3) to reconstruct robust phylogenetic trees using two mtDNA (COI and Cyt b and one ncDNA (RAG1) markers to infer the phylogenetic relationships and the systematic position of Diagramma pictum and D. punctatum and (4) to estimate divergence times between genera and species and reconstruct ancestral ranges to test the relative importance of historical geology, biology and ecological factors for the diversification within major clades of the subfamily Plectorhinchinae.

Materials and methods

Sample collection and specimen examination

A total of 254 individuals belonging to 29 species, including both our new samples and the sequences taken from GenBank and the Barcode of Life Data systems (BOLD systems) were analysed (Suppl. material 1). A total of 44 new sequences of three molecular markers were used. The specimens were photographed and then a small piece of right pectoral fin of each individual was excised and preserved in absolute alcohol. The whole samples were stored in formaldehyde for long-term storage. All collected fish specimens were identified as described by Randall (1995). Voucher specimens are deposited in the Zoological Museum, Ferdowsi University of Mashhad (ZMFUM), Iran content-type="institution" (Suppl. material 1).

DNA extraction, amplification and sequencing

Genomic DNA was extracted from the muscle tissue following the BioGene kit protocol. The reactions volume of polymerase chain reaction (PCR) was 25 μl, including 2 μl template DNA, 9.5 μl ddH2O, 12.5 μl Master Mix and 0.5 μl of each forward and reversed primer with 10 μM concentration. Two mitochondrial and one nuclear gene were amplified as follows: 617 bp of the CO1 was amplified using universal primers CO1LBC_F and CO1HBC_R (Ward et al. 2005), 1100 bp of the Cyt b was amplified using UniF and UniR primers (Orrell et al. 2002) and 1440 bp of the nuclear gene RAG1 was amplified using nested PCRs product (1st PCR: 2510F and RAG1R1; 2nd PCR: RAG1F1 and RAG1R2 primers) (López et al. 2004). PCR conditions for the COI, Cyt b, and Rag1 genes were as reported by Sanciangco et al. (2011). The PCR products were evaluated with 1.5% agarose gel for band characterisation, purified and then sent for sequencing at Microsynth Company, Switzerland. Ocyurus chrysurus (Bloch, 1791) and Rhomboplites aurorubens (Cuvier, 1829) from the family Lutjanidae Gill, 1861 were used as outgroups, since Lutjanid species has a close phylogenetic relationship with Haemulidae (Betancur-Rghton et al. 2013; Near et al. 2013).

Phylogenetic analysis and population structure

The quality of new sequences was probed using the CLC Genomics Workbench v.3.6.5 software (QIAGEN Bioinformatics). The sequences were aligned in MAFFT v.7.463 and edited with BIOEDIT v.7.2 software (Hall 1999). Genetic interspecific distances of the mitochondrial COI gene were calculated by MEGA X (Kumar et al. 2018) using the Kimura two-parameter (K2P) model. The best-fit nucleotide substitution model was identified by the Akaike Information Criterion (AIC) in the jModelTest v.2 (Posada 2008). AIC was TIM2+I+G for COI, GTR+I+G for Cyt b, GTR+I+G for the two mtDNA regions and TIM2+I+G for the mitochondrial and nuclear markers. Maximum Likelihood (ML) analysis was performed using RAxMLv.7.2.6 (Stamatakis 2014) with 1000 replicate bootstrap. Maximum Likelihood (ML) analysis was performed using RAxMLv.7.2.6 (Stamatakis 2014) with 1000 replicate bootstrap. Bayesian Inference (BI) was run in MrBayes v.3.2.7 (Ronquist et al. 2012) with two independent Markov Chain Monte Carlo for 50,000,000 generations. The first 25% of trees were discarded as part of the burn-in phase. The remaining trees were combined to generate a 50% majority rule consensus tree. The posterior probability and bootstrap values were calculated from the BI and ML for each clade and values above 0.95 and 75 were deemed as significant supports, respectively. A haplotype network was constructed using the median-joining method in the PopART v.1.7 package (Leigh and Bryant 2015). The parsimony informative and variable sites, as well as haplotype diversity were calculated using DnaSP v.5.10 to test patterns of population structure in D. pictum and D. labiosum (Librado and Rozas 2009). Initially, genetic divergence between the putative species of Diagramma was estimated and the analysis of molecular variance (AMOVA) as ΦST values was performed in GenAlEx v.6.5 (Peakall and Smouse 2006) based on pairwise comparisons. Demographic changes in the effective population size were investigated for both species using two neutrality tests (Fu’s Fs and Tajima’s D) and the Ramos-Onsins and Rozas test R2 in DnaSP v.5.10 (Librado and Rozas 2009). The time elapsed since most recent population expansion was calculated herein for Diagramma pictum and D. labiosum (Rogers and Harpending 1992) using the τ = 2μt value, where τ is the time elapsed since the start of expansion in years, μ is the mutation rate per generation for the sequence (the number of base pairs per divergence rate (1% or 2% Mya). generation time (in years) and t is the time elapsed since the beginning of expansion in years. As reported by Grandcourt et al. (2011), the generation time is 2.9 years for Diagramma pictum. The mutation rate reported for the mtDNA (COI) in previous reef fish studies ranged from 1% to 2% per million years (Bowen et al. 2001).

Divergence time estimation

A calibrated time-tree of Plectorhinchinae was estimated from the combined data set (COI, Cyt b and RAG1: 3157 total bp) in BEAST v.2.6.2 (Bouckaert et al. 2019). Three independent MCMC chains were run for 20 million generations each with sampling every 1,000 generations. Runs were implemented using an uncorrelated lognormal relaxed clock model to calculate rate heterogeneity and the Yule prior method for the tree prior. Convergence of runs and the effective sample size (> 200) were checked in Tracer v.1.6 (Rambaut 2016). The three runs were combined by LogCombiner v.2.6.2 (Bouckaert et al. 2019). The first 25% of trees were discarded in the TreeAnnotator v.2.6.2 (Bouckaert et al. 2019). The consensus tree with divergence time estimates was visualised using FigTree v.1.4.4 (Rambaut, 2016). Due to a lack of fossil record for Haemulidae, two calibration points were used following Near et al. (2013) and Tavera et al. (2018): (1) the Haemulidae-Lutjanidae split in early Eocene (mean 52.5, 95% HPD interval 57.2–46.8) and (2) the Parapristipoma-Plectorhinchus split in the late Eocene (mean 34.5 Ma, 95% HPD interval 42.6–27 Ma).

Ancestral area reconstruction

The ancestral area in Plectorhinchinae was inferred using the BioGeoBEARS analysis in RASP v.4 (Yu et al. 2020). Geographic distribution of each species was derived from literature (Fricke et al. 2022) and sampling of our ranges (Suppl. material 1). Four biogeographic regions were defined, based on present-day distribution and genetic diversification reported in previous studies (Ma et al. 2016; Stiller et al. 2022): Western Atlantic (WA); Eastern Atlantic (EA); Western Indian Ocean (WIO); Central Indo-Pacific (CIP). The input trees of BEAST analysis were used to infer the ancestral area probability. Six different biogeographic models were evaluated to select the best-fit using the AIC cumulative weight (AICc_wt).

Molecular species delimitation

Molecular species delimitation was used from four different methods, including; (A): the Automatic Barcode Gap Discovery (ABGD) (Puillandre et al. 2012) uses the pairwise genetic distance to detect a barcode gap while assuring that intra and interspecific genetic distances do not overlap and are executed in the website (http://wwwabi.snv.jussieu.fr/public/abgd) with prior intraspecific divergence (0.001 to 0.1), steps = 10 and a relative gap width of 1 using the Kimura 2-parameter model (Kimura (K80)); (B): Assemble Species by Automatic Partitioning (ASAP) (Puillandre et al. 2021), based on the distance threshold and a hierarchical clustering algorithm to create a list of best partitions. The partition score combines two metrics: the barcode gap width and the probability of panmixia (p-value), in which a lower score indicates a better partition. The method that was done using the website https://bioinfo.mnhn.fr/abi/public/asap with the Kimura (K80) model; (C): Bayesian Poisson Tree Process (bPTP) (Zhang et al. 2013), a method that estimates the number of mutations between two branching events using the branch length information of the input phylogeny. The bPTP approach was adopted on its website (http://species.hits.org/ptp/) using a non-ultrametric tree of ML inference as the input, 500,000 MCMC generations and 10% burn-in; (D): Generalised Mixed Yule Coalescent (GMYC) (Pons et al. 2006) method was run on its website (http://species.h-its.org/gmyc), using the ultrametric tree produced by BEAST v.2.6.2 (Bouckaert et al. 2019).

Results

MOTU delimitation and genetic diversity

We assembled a total of 217 COI sequences (34 newly developed plus 183 archived; 617 bp) from 29 nominal Plectorhinchinae species. D. punctatum and Plectorhinchus pictus were sequenced for the first time. Novel COI barcodes were provided for Plectorhinchus flavomaculatus, P. gaterinus and P. gibbosus for the first time from the Northern Indian Ocean (NIO). Of 29 nominal species studied, the coalescent-based delimitations (bPTP, GMYC) yielded 27 MOTUs, whereas the distance-based methods (ABGD and ASAP) yielded 26 MOTUs (Fig. 1). ABGD and ASAP identified Plectorhinchus lessonii, and P. diagrammus and P. albovittatus and P. caeruleonothus as a MOTU. bPTP and GMYC methods considered Plectorhinchus lessonii and P. diagrammus as one MOTU. Additionally, haplotype networks, based on the COI gene, identified two main distinct haplogroups (Fig. 2A). The haplogroup I was composed of Diagramma pictum from the NIO to WP, which was separated by 15 mutation steps from haplogroups II, represented by D. punctatum from the Red Sea. In the first haplogroups, samples of D. pictum were mixed with haplotypes of D. labiosum from Australia. As suggested by the genetic distance method that uses the K2P model, the highest interspecific genetic distance was found between Plectorhinchus macrolepis and P. schotaf (21.3%). In comparison, the lowest distance was observed between D. pictum and D. labiosum (0.53%) and P. lessonii and P. diagrammus (1.2%) (Table 1). The highest intraspecific genetic variation was observed in P. gibbosus (0.69%), while values close to 0 to 0.09% were detected in most of the Diagramma and Plectorhinchus species (Table 1). Based on all delimitation analyses, D. labiosum specimens from Queensland, Australia (type locality) are genetically analogous to D. pictum and there is no evidence of their genetic differentiation. Cyt b sequence of a single D. labiosum specimen (DQ784750), deposited at the GenBank of Australia with D. pictum specimens in the phylogenetic tree, was identified as a monophyletic lineage (genetic divergence: 0.51%) (data are not shown). Seventeen haplotypes were detected based on COI sequences between D. labiosum and D. pictum (Table 2). Seventeen haplotypes were detected, based on COI sequences between D. labiosum and D. pictum (Table 2). The highest haplotype and nucleotide diversity indices were found in the samples of D. labiosum (Australia) (Hd = 0.600, π = 0.0021), while specimens of D. pictum (Hd = 0.560: NIO, π = 0.0018: WP) demonstrated the lowest values (Hd = 0.579, π = 0.001) (Table 2), where the diversity indicators were not influenced by sample size reduction. ΦST value revealed a non-significant level between Australia (D. labiosum) and IWP, except Australia (D. pictum) (ΦST = –1.783) (Table 2). Test of neutrality found negative and non-significant Tajima’s D values, while Fu’s Fs (-8.801) and R2 (0.038) results were significant (Table 2). These significant values (Fu’s Fs, R2), in combination with the estimates of τ for both species, indicate a shallow population history with recent expansion (Table 2). The unimodal mismatch distribution for the pooled sample of D. pictum supported a recent historical expansion (Fig. 3).

Table 1.

Genetic K2P distances (%) of the Diagramma and Plectorhinchus species.

Species Maximum intra-group distance Nearest-Neighbour (NN) Distance to NN
Diagramma pictum 0.17 D. labiosum 0.53
D. punctatum 0.00 P. centurio 3.45
Plectorhinchus albovittatus 0.00 P. caeruleonothus 2.31
P. caeruleonothus 0.19 P. albovittatus 2.31
P. centurio 0.35 D. punctatum 3.45
P. chaetodonoides 0.43 P. centurio 8.75
P. chrysotaenia 0.00 P. playfairi 16.30
P. chubbi 0.08 P. sordidus 3.52
P. cinctus 0.64 P. gibbosus 13.44
P. diagrammus 0.00 P. lessonii 1.20
P. flavomaculatus 0.31 P. makranensis 12.62
P. gaterinus 0.28 P. caeruleonothus 12.18
P. gibbosus 0.69 P. cinctus 13.44
P. lessonii 0.08 P. diagrammus 1.20
P. lineatus 0.06 P. polytaenia 7.17
P. macrolepis 0.00 P. plagiodesmus 13.84
P. macrospilus 0.00 P. lessonii 4.10
P. makranensis 0.01 P. schotaf 5.23
P. orientalis P. vittatus 7.97
P. pictus 0.00 P. cinctus 13.55
P. picus 0.08 D. punctatum 8.38
P. plagiodesmus 0.00 P. pictus 13.24
P. playfairi 0.16 P. chubbi 4.61
P. polytaenia 0.00 P. lineatus 7.17
P. schotaf 0.13 P. makranensis 5.23
P. sordidus 0.10 P. chubbi 3.52
P. unicolor 0.23 P. chubbi 5.80
P. vittatus P. orientalis 7.97
Table 2.

Molecular diversity indices for D. pictum and D. labiosum based on mitochondrial DNA (COI) sequence data. Numbers in bold denotes significance (P < 0.05) and for Fu’s Fs (P < 0.02).

Statistics D. pictum D. labiosum
Japan to Indonesia (WP) Iran to Bangladesh (NIO) Australia
(WP)
All samples
Number of individuals (N) N = 44 N = 26 N = 7 N = 76
Expansion time (years) 17,647-35,294
Number of haplotypes (Hn) 8 6 3 17
Haplotype diversity (Hd ± SD) 0.579 ± 0.075 0.560 ± 0.073 0.600 ± 0.215 0.624 ± 0.051
Nucleotide diversity (π ± SD) 0.0018 ± 0.0011 0.0021 ± 0.0016 0.0021 ± 0.0017 0.0024 ± 0.0018
Tajima’s D -1.218 -1.636 -1.233 -1.783
Fu’s FS -3.739 -1.067 -0.189 -8.801
R2 0.058 0.060 0.062 0.038
ΦST Australia 0.084 0.072
Iran to Bangladesh 0.049
Figure 1. 

Bayesian Inference tree, based on the COI showing the results of species delimitation algorithms. The vertical bars indicate species delimitation analyses. The black circles indicate nodes with supports (ML bootstrap BP ≥ 75% and BI probability PP ≥ 0.95), orange circles (BP ≤ 75% and BI PP ≥ 0.95) and blue circles (PP ≤ 0.95 and ML BP ≤ 75%).

Figure 2. 

Median‐joining network obtained of 617 bp of mitochondrial COI for Diagramma species (A). Map depicting the geographical distribution of 17 haplotypes (B). Circle sizes are relative to haplotype frequency. Numbers between haplotypes represent mutational steps between haplotypes. Colour circles indicate species of D. pictum (red: Persian Gulf to Bangladesh, green: China and Indonesia) and D. labiosum (orange: Australia), D. punctatum (black: Red Sea).

Figure 3. 

Mismatch distribution inferred from partial mtDNA COI for the pooled dataset of D. pictum.

Phylogenetic relationships

We analysed 254 COI DNA sequences (617 bp) which show 325 variable sites; the Cyt b gene consisting of 1117 bp showed 295 variable positions; and Rag1 consisting of 1440 bp showed 184 variable sites. Neither gaps nor stop codons were observed in the final alignment. ML and BI analyses with the combined dataset (mtDNA and nuclear) revealed that the phylogenetic relationships of sweetlips species had the highest resolution (Fig. 4). The ML and BI phylogenetic trees, based on two mtDNA (Cyt b + COI) and one nuclear (Rag1) gene, had a congruent topology, with a slight difference in their support values. The phylogenetic relationships observed in the analysis of the concatenated data (Cyt b + COI + Rag1) revealed that Parapristipoma, as a basal position, is the sister clade of the Plectorhinchus species with high bootstrap and posterior probabilities (Fig. 4). Within the Plectorhinchus clade, two well-supported subclades were formed: (1) The first clusters the most WIO species (subclade I): Plectorhinchus schotaf, P. makranensis, P. playfairi, P. sordidus, P. chubbi, P. unicolor, P. plagiodesmus, P. gibbosus, P. macrolepis, P. chrysotaenia, P. cinctus and P. pictus; (2) the second clusters the most CIP species: P. gaterinus, P. chaetodonoides, P. picus, P. albovittatus, P. lessonii, P. lineata, P. macrospilus, P. diagrammus, P. polytaenia, P. vittatus, P. orientalis, P. caeruleonothus along with D. pictum and D. punctatum indicating that the Diagramma species is nested within Plectorhinchus and should be recognised as Plectorhinchus pictum and Plectorhinchus punctatum (subclade II) (PP = 1; BS = 100%).

Figure 4. 

Phylogeny recovered by the Bayesian Inference (BI) and Maximum Likelihood (ML) analyses for the Cyt b, COI and RAG1 dataset. The black circles indicate nodes with supports (ML bootstrap BP ≥ 75% and BI probability PP ≥ 0.95), orange circles (BP ≤ 75% and BI PP ≥ 0.95) and blue circles (PP ≤ 0.95 and ML BP ≤ 75%).

Divergence times and ancestral area reconstructions

The Haemulidae diverged from Lutjanidae ~ 49 Ma (early to middle Eocene, 95% HPD: 43.6–55.2 Ma) and then Plectorhinchinae divided into two main clades (A–B) and two subclades (I–II) (Fig. 5). The first split created Clade A containing Parapristipoma at ~ 35 Ma (late Eocene, 95% HPD: 32.4–37.1 Ma). Clade B diverged at ~ 21 Ma (early Miocene, 95% HPD: 18.4–23.5 Ma) into two subclades. Within each subclade, there were two periods of maximum diversification from the Miocene to Pliocene epochs. The best-fit biogeographic models recovered for the ancestral range reconstruction was the BAYAREALIKE (AICc_wt: 0.71). As suggested by the ancestral area reconstruction using the combined dataset (Cyt b + COI + Rag1), 17 dispersal and eight vicariance events occurred during the evolution of the studied Plectorhinchinae. The deepest node of early Plectorhinchinae constructed probably has its origin in the IWP with the dispersal and vicariance events occurring from the late Eocene to early Oligocene epoch (Fig. 6). However, more comprehensive taxonomic sampling of the Parapristipoma species can help find a definitive point for the origin of the subfamily Plectorhinchinae. The ancestors of the earliest Plectorhinchus lineages were supposed to be of WIO origin with vicariance and dispersal events taking place in the Miocene. Within subclade I, there seems to be a burst of species diversification in the WIO at 20 Mya. Within subclade II, the vicariance event affected speciation in the CIP as the ancestral area at 15 Mya.

Figure 5. 

Time-calibrated phylogeny, based on combined mitochondrial (COI, Cyt b) and nuclear (RAG1) genes, inferred from the BEAST analysis of the Plectorhinchinae. Numbers in the horizontal purple bars represent the 95% Highest Posterior Density intervals for each estimated event. The black circles (pp ≥ 0.95) and orange circles (pp < 0.95). A: Parapristipoma, B: Plectorhinchus.

Figure 6. 

Ancestral area reconstruction obtained using BioGeoBEARS for Plectorhinchinae. Biogeographic areas are shown in the circles and the colours in the squares indicate the current species distribution.

Discussion

This study represents the first attempt to delimit the species boundaries and evolutionary and biogeographic history of marine fish species in Plectorhinchinae. The present study corroborates the findings of the previous molecular studies that show paraphyly of the Plectorhinchus without the inclusion of Diagramma (Sanciangco et al. 2011; Tavera et al. 2018) and contrast with morphological data (Konchina 1976; Carpenter and Niem 2001; Nelson et al. 2016) that supported this genus monophyly.

Molecular phylogeny of Plectorhinchus

According to the molecular data, Plectorhinchus is divided into two subclades (Fig. 4), which display high concordance with the division proposed by Tavera et al. (2018). It can reinforce the phylogenetic value of the geographic distribution and body-colour pattern in distinguishing Plectorhinchus lineages. The first group (subclade I) mostly consists of species with uniformly dark colours (Plectorhinchus schotaf, P. makranensis, P. playfairi, P. sordidus, P. chubbi, P. unicolor, P. plagiodesmus, P. gibbosus, P. macrolepis and P. chrysotaenia) or black pots on the upper sides of the body (P. cinctus and P. pictus), along with more profound body depth (Randall 1995, 1997; McKay 2001), which are those typically distributed in the WIO (Fig. 4) (Fricke et al. 2022). The second group (subclade II) was comprised of species with spots on the whole body (P. gaterinus, P. chaetodonoides and P. picus) or horizontal stripes alongside (P. albovittatus, P. lessonii, P. lineata, P. macrospilus; P. diagrammus, P. polytaenia, P. vittatus and P. orientalis), except for P. caeruleonothus and Diagramma species (Fig. 4) (Randall 1995, 1997; Johnson et al. 2001; McKay 2001). The species includes representatives with shorter bodies from the CIP (Fricke et al. 2022). The colour variation observed in subclade II compared to subclade I can be attributed to the substrate type (Randall et al. 1997; McKay 2001; Johnson et al. 2001). The colour variation of the members of subclade II inhabiting the coral reef may be linked to social communication, mimicry and camouflage defence mechanisms to avoid predation adjusted to their habitat (Randall 1998; McMillan et al. 1999).

Systematics and species boundaries of Plectorhinchus

In our analyses, specimens from 29 nominal species were delimited between 26 and 27 consensus MOTUs. In general, low interspecific genetic divergences were observed in some of the MOTUs within Plectorhinchus, which focuses on the barcoding analysis in closely-related species. This situation was formerly reported in haemulid within Anisotremus (Bernardi et al. 2008) and Haemulon (Rocha et al. 2008). In the meanwhile, the fact that P. lessonii and P. diagrammus, along with P. albovittatus and P. caeruleonothus, grouped together in all the delimitation analyses, exhibits their very recent divergence (< 1 Ma) compared to other Plectorhinchus members. The morphological variability, including minor differences in colouration between D. pictum and D. labiosum determines taxonomic classification into subspecies or species status (Allen 1997; Johnson et al. 2001; Bogorodsky et al. 2014; Bray 2017; Fricke et al. 2022). Our molecular analyses did not yield any genetic evidence to confirm D. labiosum as a distinct species (Figs 1, 2, Table 1). A relatively low level of genetic divergence between D. pictum and D. labiosum found in both markers (COI and Cyt b) may indicate a recent speciation event (Table 1). These distances are considerably shorter than those reported in haemulid species (Johnson and Wilmer 2015; Damadi et al. 2020) and other related species of marine fishes, which is at least 2% between closely-related species (Ward et al. 2005; Hsu et al. 2007). Demographic tests (Fu’s, R2 and mismatch distribution) indicate that D. labiosum and D. pictum share a common ancestor and expansion events over the last few thousand years coincided with the Holocene glaciations that restricted larval dispersal and ocean circulation deflection (Yokoyama et al. 2001; Wilson 2013; Rosser et al. 2020). These features might have reduced coral reef habitats and colonisation from northwest Australia, especially in the reef fish (Ovenden et al. 2002; Salini et al. 2006; Sulaiman and Ovenden 2010). The incongruence between colour variation and the level of genetic divergence is not rare in fish and could be attributed to some other reasons: (1) phenotypic plasticity rather than reproductive isolation, (2) incomplete lineage sorting and (3) introgression through hybridisation (Bowen et al. 2006; Schultz et al. 2007; DiBattista et al. 2012; DiBattista et al. 2015; Bernal et al. 2017). There are several lines of evidence that favour the incomplete lineage sorting over other scenarios; however, there is still a paucity of evidence and further analyses using genomic data are warranted. Firstly, D. labiosum and D. pictum showed extensive mtDNA sharing and coalescence of haplotypes, which could be interpreted as evidence of a very recent divergence (Figs 1, 2). Secondly, their distribution is allopatric (D. labiosum: Australia and D. pictum: IWP, except Australia) (Fig. 2B) (Fricke et al. 2022). Hybridisation events and lineage sharing usually occur in sympatric and parapatric areas rather than in allopatric areas due to the greater opportunities for hybridisation in the secondary contact zones (Hobbs et al. 2009; DiBattista et al. 2016). In contrast, incomplete lineage may take place in both sympatric and allopatric areas (DiBattista et al. 2012; DiBattista et al. 2015). Thirdly, regional colour morphs usually represent minor molecular divergences due to recent genetic divergence, which may not be related to ecological and behavioural adaptations (Taylor and Hellberg 2003; Drew et al. 2008; Drew et al. 2010). Fourthly, the Diagramma species are distinguished only based on colour patterns (Johnson et al. 2001; Heemstra and Heemstra 2004; Bogorodsky et al. 2014) which evolve more quickly than the other morphological traits and genetic characters (Bowen et al. 2006; Schultz et al. 2007). Colouration could be an important criterion to distinguish closely-related species in haemulid fishes (Johnson and Wilmer 2015; Damadi et al. 2021). Fifthly, the most recent divergence occurred between D. pictum and D. labiosum around 28–56 million years ago (Table 1), which is more recent than the complete lineage sorting time. A divergence time of 2.9 Mya seems sufficient for these lineages for complete sorting (1× generation time) (Grandcourt et al. 2011).

Biogeography and temporal diversification

Our biogeographic findings exhibit a complex history highlighting the role of both biological and geographical factors in the diversification and evolution within the Haemulids. The first diversification, i.e. the split of Haemulids from Lutjanids as a sister lineage in the Eocene, could be related to the ice sheet formed in the Antarctic and reduced atmospheric CO2 level (Zachos et al. 2001; Coxall et al. 2005; Lear et al. 2008). These climatic and oceanographic changes resulted in a sea-level drop, upwelling and extinction of warmer climate fauna (Zachos et al. 2001; Livermore et al. 2007; Ma et al. 2016). This extinction formed vacant niches and the emerging Indo-Australian Archipelago probably gave rise to new reef habitats and high species-richness in the cooler oceans (Cowman and Bellwood 2011; Ma et al. 2016). This divergence time corresponds to the one shown by Near et al. (2012) for the same historical event in Haemulids. The historical biogeography suggests that the common ancestor of Plectorhinchinae might have its origin in the IWP and subsequently spread and diversified in the western Indian Ocean and the Central Indo-Pacific. This scenario could be firmly demonstrated by complete biogeographic sampling of some taxa and extra genetic/genomic analyses. For Plectorhinchinae, divergence was first observed in East Atlantic (EA) species (Parapristipoma) with the fewest species from the Plectorhinchus lineage, approximately 33.8 Mya, which coincided with the onset of the terminal Tethyan event (TTE) that disconnected West Tethyan and Arabian marine biodiversity hotspots (Adams et al.1983; Harzhauser et al. 2007). The diminishing connectivity between two regions might have set the scene for vicariant splits, which appear to be linked to the early diversification of this clade. This divergence time is also congruent with the EA and IWP divergence in other species of reef fish, particularly the Groupers and Boxfishes (Santini et al. 2013; Ma et al. 2016). Within the Plectorhinchus IWP Clade, two factors, the Indo-Australian Archipelago (IAA) uplift and major global cooling, were associated with the diversification of sweetlips subclades. These factors were inferred from other reef fish in the Miocene period (Zachos et al. 2001; Bellwood and Wainwright 2002; Shevenell et al. 2004; Cowman and Bellwood 2011; Ma et al. 2016). An estimate of divergence times of each Plectorhinchus subclade suggests further diversification occurred during the Miocene that slow indicated by the long branches, followed by slower diversification in Pliocene to Pleistocene epochs (Fig. 5). The IAA uplift during the early to mid-Miocene in subclade II (CIP Plectorhinchus) created new islands, coinciding with the lowering of sea level, leading to novel shallow reef habitat discontinuity. During Plio-Pleistocene, Diagramma species diverged from each other around 4.1 to 2.3 Mya (Fig. 5) and D. punctatum, an endemic species, formed in the Red Sea, coinciding probably with high temperature and salinity, sea-level changes during glacial cycles and the narrow Strait of Bab al Mandab. These factors led to a decrease in unsuitable habitat and the water flow through this strait (Rihm and Henke 1998; Riegl and Purkis 2012; Stevens et al. 2014; Torquato et al. 2019) that has been documented in marine fish (Froukh and Kochzius 2007; DiBattista et al. 2013; Bogorodsky et al. 2017). The major global cooling could lay the ground for emergence of new environmental niches likely driven by adaptive radiations in subclade I (WIO) during the Miocene (Zachos et al. 2001; Shevenell et al. 2004).

Conclusions

The findings of the present study on sweetlips (Plectorhynchinae) not only shed light on the interrelationships and evolutionary history, but also opened new avenues for research. The results of this study provide a framework to shed light on the probable origin and factors involved in the diversification of the IWP Plectorhynchinae. The contemporary species have a common ancestor that comes from the present day IWP and the current geographical distribution appears to be related to vicariance and dispersal events. It indicates that Diagramma species were divided 4.1 to 2.3 Mya ago into ancestral Diagramma punctatum in the Indian Ocean, directly followed by the WP and recently isolated species. These findings raise the possibility that recent genetic divergence have given rise to colour morphotypes as emerging species on coral reefs.

Acknowledgements

We greatly thank Dr. Joseph D. DiBattista and Dr. Brian W. Bowen for their valuable comments on population genetics. The authors are especially grateful to Mohamed Awad for providing tissue specimens from Red Sea (Egypt). We would like to thank local fishermen for their help in collecting specimens. Funding support for this research was provided by Ferdowsi University of Mashhad and University of Sistan and Baluchestan in Iran (grant number 46162).

References

  • Adams CG, Gentry AW, Whybrow PJ (1983) Dating the terminal Tethyan event. Utrecht Micropaleontol. Bulletins 30: 273–298.
  • Allen R (1997) Marine fishes of tropical Australia and southast Asia. 3rd edn. , Western Australian Museum, Perth, 1292 pp.
  • Asgharian H, Sahafi HH, Ardalan AA, Shekarriz S, Elahi E (2011) Cytochrome c oxidase subunit 1 barcode data of fish of the Nayband National Park in the Persian Gulf and analysis using meta data flag several cryptic species. Molecular Ecology Resources 11(3): 461–472. https://doi.org/10.1111/j.1755-0998.2011.02989.x
  • Bellwood DR, Wainwright PW (2002) The History and Biogeography of Fishes on Coral Reefs. Coral Reef Fishes: Dynamics and Diversity in A Complex Ecosystem (Ed. by P. F. Sale). Academic Press, London, 5–32. https://doi.org/10.1016/B978-012615185-5/50003-7
  • Bernal MA, Gaither MR, Simison WB, Rocha LA (2017) Introgression and selection shaped the evolutionary history of sympatric sister-species of coral reef fishes (genus: Haemulon). Molecular Ecology 26(2): 639–652. https://doi.org/10.1111/mec.13937
  • Bernardi G, Alva-Campbell YR, Gasparini JL, Floeter SR (2008) Molecular ecology, speciation, and evolution of the reef fish genus Anisotremus. Molecular Phylogenetics and Evolution 48(3): 929–935. https://doi.org/10.1016/j.ympev.2008.05.011
  • Bogorodsky SV, Alpermann TJ, Mal AO, Gabr MH (2014) Survey of demersal fishes from southern Saudi Arabia, with five new records for the Red Sea. Zootaxa 3852(4): 401–437. https://doi.org/10.11646/zootaxa.3852.4.1
  • Bogorodsky SV, Iwatsuki Y, Amir SA, Mal AO, Alpermann TJ (2017) Morphological and molecular divergence between Crenidens crenidens (Forsskål, 1775) and C. indicus Day, 1873 (Perciformes: Sparidae) and notes on a Red Sea endemic lineage of C. crenidens. Marine Biodiversity 47(4): 1273–1285. https://doi.org/10.1007/s12526-017-0764-6
  • Bouckaert R, Vaughan TG, Barido-Sottani J, Duchêne S, Fourment M, Gavryushkina A, Heled J, Jones G, Kühnert D, De Maio N, Matschiner M, Mendes FK, Müller NF, Ogilvie HA, du Plessis L, Popinga A, Rambaut A, Rasmussen D, Siveroni I, Suchard MA, Wu C-H, Xie D, Zhang C, Stadler T, Drummond AJ (2019) BEAST 2.5: An advanced software platform for Bayesian evolutionary analysis. PLoS Computational Biology 15(4): e1006650. https://doi.org/10.1371/journal.pcbi.1006650
  • Bowen BW, Muss A, Rocha LA, Grant WS (2006) Shallow mtDNA coalescence in Atlantic pygmy angelfishes (genus Centropyge) indicates a recent invasion from the Indian Ocean. The Journal of Heredity 97(1): 1–12. https://doi.org/10.1093/jhered/esj006
  • Bray DJ (2017) “Diagramma pictum labiosum”. Fishes of Australia. Museums Victoria.
  • Carpenter KE, Niem VH (2001) FAO species identification guide for fishery purposes. The living marine resources of the Western Central Pacific. Vol. 6. Bony fishes part 4 (Labridae to Latimeriidae), estuarine crocodiles, sea turtles, sea snakes and marine mammals. FAO Library, 1–4.
  • Cowman PF, Bellwood DR (2011) Coral reefs as drivers of cladogenesis: expanding coral reefs, cryptic extinction events, and the development of biodiversity hotspots. Journal of Evolutionary Biology 24: 2543–2562. https://doi/full/10.1111/j.1420-9101.2011.02391.x
  • Coxall HK, Wilson PA, Pälike H, Lear CH, Backman J (2005) Rapid stepwise onset of Antarctic glaciation and deeper calcite compensation in the Pacific Ocean. Nature 433(7021): 53–57. https://doi.org/10.1038/nature03135
  • Damadi E, Yazdani Moghaddam FY, Ghassemzadeh F, Ghanbarifardi M (2020) Plectorhinchus makranensis (Teleostei, Haemulidae), a new species of sweetlips from the Persian Gulf and the Gulf of Oman. ZooKeys 980: 141–154. https://doi.org/10.3897/zookeys.980.50934
  • Damadi E, Yazdani Moghaddam FY, Ghassemzadeh F, Ghanbarifardi M (2021) Aspects of morphometry, distribution, and key identification of the genus Pomadasys (Perciformes: Haemulidae) from the Persian Gulf and Gulf of Oman with descriptions of new records. Thalassas. International Journal of Material Science 37: 671–682. https://doi.org/10.1007/s41208-021-00308-1
  • DiBattista JD, Waldrop E, Bowen BW, Schultz JK, Gaither MR, Pyle RL, Rocha LA (2012) Twisted sister species of pygmy angelfishes: Discordance between taxonomy, coloration, and phylogenetics. Coral Reefs 31(3): 839–851. https://doi.org/10.1007/s00338-012-0907-y
  • DiBattista JD, Berumen ML, Gaither MR, Rocha LA, Eble JA, Choat JH, Craig MT, Skillings DJ, Bowen BW (2013) After continents divide: Comparative phylogeography of reef fishes from the Red Sea and Indian Ocean. Journal of Biogeography 40(6): 1170–1181. https://doi.org/10.1111/jbi.12068
  • DiBattista JD, Waldrop E, Rocha LA, Craig MT, Berumen ML, Bowen BW (2015) Blinded by the bright: A lack of congruence between color morphs, phylogeography and taxonomy for a cosmopolitan Indo Pacific butterflyfish, Chaetodon auriga. Journal of Biogeography 42(10): 1919–1929. https://doi.org/10.1111/jbi.12572
  • DiBattista JD, Whitney J, Craig MT, Hobbs JPA, Rocha LA, Feldheim KA, Berumen ML, Bowen BW (2016) Surgeons and suture zones: Hybridization among four surgeonfish species in the Indo-Pacific with variable evolutionary outcomes. Molecular Phylogenetics and Evolution 101: 203–215. https://doi.org/10.1016/j.ympev.2016.04.036
  • Drew J, Allen G, Erdmann M (2010) Congruence between mitochondrial genes and color morphs in a coral reef fish: Population variability in the Indo-Pacific damselfish Chrysiptera rex (Snyder, 1909). Coral Reefs 29(2): 439–444. https://doi.org/10.1007/s00338-010-0586-5
  • Froukh T, Kochzius M (2007) Genetic population structure of the endemic fourline wrasse (Larabicus quadrilineatus) suggests limited larval dispersal distances in the Red Sea. Molecular Ecology 16: 1359–1367. https://10.1111/j.1365-294X.2007.03236.x
  • Golani D, Fricke R (2018) Checklist of the Red Sea Fishes with delineation of the Gulf of Suez, Gulf of Aqaba, endemism and Lessepsian migrants. Zootaxa 4509(1): 1–215. https://doi.org/10.11646/zootaxa.4509.1.1
  • Grandcourt EM, Al Abdessalaam TZ, Francis F, Al Shamsi AT (2011) Reproductive biology and implications for management of the painted sweetlips Diagramma pictum in the southern Arabian Gulf. Journal of Fish Biology 79(3): 615–632. https://doi.org/10.1111/j.1095-8649.2011.03044.x
  • Hall TA (1999) BioEdit: A user-friendly biological sequences alignment editor and analysis program for Windows 95/98/NT. Nucleic Acids Symposium Series 41: 95–98.
  • Harzhauser M, Kroh A, Mandic O, Piller WE, Gohlich U, Reuter M, Berning B (2007) Biogeographic responses to geodynamics: A key study all around the Oligo-Miocene Tethyan Seaway. Zoologischer Anzeiger 246(4): 241–256. https://doi.org/10.1016/j.jcz.2007.05.001
  • Heemstra PC, Heemstra E (2004) Coastal fishes of southern Africa. NISC and SAIAB. [I–xxiv +] 488 pp.
  • Hobbs JPA, Frisch AJ, Allen GR, van Herwerden L (2009) Marine hybrid hotspot at Indo-Pacific biogeographical border. Biology Letters 5(2): 258–261. https://doi.org/10.1098/rsbl.2008.0561
  • Hsu KC, Chen JP, Shao KT (2007) Molecular phylogeny of Chaetodon (Teleostei: Chaetodontidae) in the Indo-West Pacific: evolution in geminate species pairs and species groups. The Raffles Bulletin of Zoology 14: 77–86.
  • Johnson JW, Wilmer JW (2015) Plectorhinchus caeruleonothus a new species of sweetlips (Perciformes: Haemulidae) from northern Australia and the resurrection of P. unicolor (Macleay, 1883) species previously confused with P. schotaf (Forsskål, 1775). Zootaxa 3985: 491–522. https://doi.org/10.11646/zootaxa.3985.4.2
  • Johnson JW, Randall JE, Chenoweth SF (2001) Diagramma melanacrum new species of haemulid fish from Indonesia Borneo and the Philippines with a generic review. Memoirs of the Queensland Museum 46: 657–676.
  • Konchina Y (1976) The systematics and distribution of the grunts family (Pomadasyidae). Journal of Ichthyology 16: 18–35.
  • Kumar S, Stecher G, Li M, Knyaz C, Tamura K (2018) MEGA X: Molecular evolutionary genetics analysis across computing platforms. Molecular Biology and Evolution 35(6): 1547–1549. https://doi.org/10.1093/molbev/msy096
  • Lear CH, Bailey TR, Pearson PN, Coxall HK, Rosenthal Y (2008) Cooling and ice growth across the Eocene-Oligocene transition. Geology 36(3): 251–254. https://doi.org/10.1130/G24584A.1
  • Liang R, Wang C, Zou Q, Zhou A, Zhou M (2016) Molecular phylogenetic relationships of some common sweetlips (Haemulidae: Plectorhynchinae) and the synonyms controversy of two Plectorhinchus species. Mitochondrial DNA Part A 27(3): 2209–2214.
  • Lindeman KC, Toxey C (2003) Haemulidae – Grunts. In: Carpenter KE (Ed.) The living marine resources of the Western Central Atlantic. FAO species identification guide for fishery purposes and American Society of Ichthyologists and Herpetologists Special Publication No. 5, FAO, Rome, Italy, 1375–2127.
  • Livermore R, Hillenbrand CD, Meredith M, Eagles G (2007) Drake Passage and Cenozoic climate: An open and shut case? Geochemistry, Geophysics, Geosystems 8(1): 1–11. https://doi.org/10.1029/2005GC001224
  • Ma KY, Craig MT, Choat JH, van Herwerden L (2016) The historical biogeography of groupers: Clade diversification patterns and processes. Molecular Phylogenetics and Evolution 100: 21–30. https://doi.org/10.1016/j.ympev.2016.02.012
  • McCord CL, Nash CM, Cooper WJ, Westneat MW (2021) Phylogeny of the damselfishes (Pomacentridae) and patterns of asymmetrical diversification in body size and feeding ecology. PLoS ONE 16(10): e0258889. https://doi.org/10.1371/journal.pone.0258889
  • McKay RJ (1983) Haemulidae. FAO species identification sheets for fisheries purposes. Western Indian Ocean (Fishing Area 51), 2–10.
  • McKay RJ (2001) Haemulidae (= Pomadasyidae). Grunts (also sweetlips, rubberlips, hotlips, and velvetchins). In: Carpenter KE, Niem VH (Eds) FAO Species Identification Guide for Fishery Purposes. The Living Marine Resources of the Western Central Pacific. Vol. 5. Bony Fishes Part 3 (Menidae to Pomacentridae). FAO, Rome, 2961–2989.
  • McMillan WO, Weigt LA, Palumbi SR (1999) Color pattern evolution, assortative mating, and genetic differentiation in brightly colored butterflyfishes (Chaetodontidae). Evolution; International Journal of Organic Evolution 53(1): 247–260. https://doi.org/10.2307/2640937.
  • Near TJ, Eytan RI, Dornburg A, Kuhn KL, Moore JA, Davis MP, Wainwright PC, Friedman M, Smith WL (2012) Resolution of ray-finned fish phylogeny and timing of diversification. Proceedings of the National Academy of Sciences. 109(34): 13698–13703. https://doi.org/10.1073/pnas.1206625109
  • Near TJ, Dornburg A, Eytan RI, Keck BP, Smith WL, Kuhn KL, Wainwright PC (2013) Phylogeny and tempo of diversification in the superradiation of spiny-rayed fishes. Proceedings of the National Academy of Sciences of the United States of America 110(31): 12738–12743. https://doi.org/10.1073/pnas.1304661110
  • Nei M, Kumar S (2000) Molecular evolution and phylogenetics. Oxford University Press, USA, 1–350.
  • Orrell TM, Carpenter KE, Musick JA, Graves JE (2002) Phylogenetic and biogeographic analysis of the Sparidae (Perciformes: Percoidei) from cytochrome b sequences. Copeia 3: 618–631. https://doi.org/10.1643/0045-8511(2002)002[0618: PABAOT]2.0.CO;2
  • Ovenden JR, Lloyd J, Newman SJ, Keenan CP, Slater LS (2002) Spatial genetic subdivision between northern Australian and southeast Asian populations of Pristipomoides multidens: A tropical marine reef fish species. Fisheries Research 59(1–2): 57–69. https://doi.org/10.1016/S0165-7836(01)00415-5
  • Pons J, Barraclough TG, Gomez-Zurita J, Cardoso A, Duran DP, Hazell S, Kamoun S, Sumlin WD, Vogler AP (2006) Sequence-based species delimitation for the DNA taxonomy of undescribed insects. Systematic Biology 55(4): 595–60. https://doi.org/10.1080/10635150600852011
  • Rabaoui L, Yacoubi L, Sanna D, Casu M, Scarpa F, Lin YJ, Shen K-N, Clardy TR, Arculeo M, Qurban MA (2019) DNA barcoding of marine fishes from Saudi Arabian waters of the Gulf. Journal of Fish Biology 95(5): 1286–1297. https://doi.org/10.1111/jfb.14130
  • Rambaut A (2016) FigTree v1. 4.4. Institute of Evolutionary Biology, University of Edinburgh. [cited 2019 Oct 18]
  • Randall JE (1995) Coastal fishes of Oman. Crawford House Publishing Pty Ltd Bathurst, 439 pp.
  • Randall JE, Allen GR, Steene RC (1997) Fishes of the Great Barrier Reef and Coral Sea. University of Hawaii Press, 594 pp.
  • Randall JE (1998) Zoogeography of shore fishes of the Indo-Pacific region. Zoological Studies (Taipei, Taiwan) 37: 227–268.
  • Riegl BM, Purkis SJ (2012) Coral reefs of the Gulf: adaptation to climatic extremes in the world’s hottest sea. In: Coral reefs of the Gulf. Springer, Dordrecht, 1–4. https://doi.org/10.1007/978-94-007-3008-3_1
  • Rihm R, Henke CH (1998) Geophysical studies on early tectonic controls on Red Sea rifting, opening and segmentation. In Sedimentation and Tectonics in Rift Basins Red Sea: Gulf of Aden. Springer, Dordrecht, 29–49. https://doi.org/10.1007/978-94-011-4930-3_3
  • Rocha LA, Lindeman KC, Rocha CR, Lessios HA (2008) Historical biogeography and speciation in the reef fish genus Haemulon (Teleostei: Haemulidae). Molecular Biology and Evolution 48: 918–928. https://doi.org/10.1016/j.ympev.2008.05.024
  • Ronquist F, Teslenko M, Van Der Mark P, Ayres DL, Darling A, Hohna S, Larget BL, Liu L, Suchard MA, Huelsenbeck JP (2012) MRBAYES 3.2: Efficient Bayesian phylogenetic inference and model selection across a large model space. Systematic Biology 61(3): 539–542. https://doi.org/10.1093/sysbio/sys029
  • Rosser NL, Edyvane K, Malina AC, Underwood JN, Johnson MS (2020) Geography and spawning season drive genetic divergence among populations of the hard coral Acropora tenuis from Indonesia and Western Australia. Coral Reefs 39(4): 989–999. https://doi.org/10.1007/s00338-020-01923-9
  • Salini JP, Ovenden JR, Street R, Pendre R (2006) Genetic population structure of red snappers (Lutjanus malabaricus Bloch & Schneider, 1801 and Lutjanus erythropterus Bloch, 1790) in central and eastern Indonesia and northern Australia. Journal of Fish Biology 68(sb): 217–234. https://doi.org/10.1111/j.0022-1112.2006.001060.x
  • Sanciangco MD, Rocha LA, Carpenter KE (2011) A molecular phylogeny of the Grunts (Perciformes: Haemulidae) inferred using mitochondrial and nuclear genes. Zootaxa 2966(1): 37–50. https://doi.org/10.11646/zootaxa.2966.1.4
  • Santini F, Sorenson L, Marcroft T, Dornburg A, Alfaro ME (2013) A multilocus molecular phylogeny of boxfishes (Aracanidae, Ostraciidae; Tetraodontiformes). Molecular Phylogenetics and Evolution 66(1): 153–160. https://doi.org/10.1016/j.ympev.2012.09.022
  • Schultz JK, Pyle RL, DeMartini E, Bowen BW (2007) Genetic connectivity among color morphs and Pacific archipelagos for the flame angelfish, Centropyge loriculus. Marine Biology 151: 167–175. https://doi.org/10.1007/s00227-006-0471-5
  • Stiller J, Short G, Hamilton H, Saarman N, Longo S, Wainwright P, Rouse GW, Simison WB (2022) Phylogenomic analysis of Syngnathidae reveals novel relationships, origins of endemic diversity and variable diversification rates. BMC Biology 20(1): 1–21. https://doi.org/10.1186/s12915-022-01271-w
  • Sulaiman ZH, Ovenden JR (2010) Population genetic evidence for the east-west division of the narrow-barred Spanish mackerel (Scomberomorus commerson, Perciformes: Teleostei) along Wallace’s Line. Biodiversity and Conservation 19(2): 563–574. https://doi.org/10.1007/s10531-009-9699-y
  • Tavera JJ, Pizarro AA, De la Cruz Agüero J, Balart EF (2011) Phylogeny and reclassification of the species of two neotropical grunt genera, Anisotremus and Genyatremus (Perciformes: Haemulidae), based on morphological evidence. Journal of Zoological Systematics and Evolutionary Research 49(4): 315–323. https://doi.org/10.1111/j.1439-0469.2011.00622.x
  • Tavera J, Acero A, Wainwright PC (2018) Multilocus phylogeny, divergence times, and a major role for the benthic-to-pelagic axis in the diversification of grunts (Haemulidae). Molecular Phylogenetics and Evolution 121: 212–223. https://doi.org/10.1016/j.ympev.2017.12.032
  • Torquato F, Range P, Ben-Hamadou R, Sigsgaard EE, Thomsen PF, Riera R, Berumen ML, Burt JA, Feary DA, Marshell A, D’Agostino D, DiBattista JD, Møller PR (2019) Consequences of marine barriers for genetic diversity of the coral specialist yellowbar angelfish from the Northwestern Indian Ocean. Ecology and Evolution 9(19): 11215–11226. https://doi.org/10.1002/ece3.5622
  • Traldi JB, Vicari MR, de Fátima Martinez J, Blanco DR, Lui RL, Azambuja M, de Almeida RB, de Cássia Malimpensa G, da Costa Silva GJ, Oliveira C, Pavanelli CS, Filho OM (2020) Recent Apareiodon species evolutionary divergence (Characiformes: Parodontidae) evidenced by chromosomal and molecular inference. Zoologischer Anzeiger 289: 166–176. https://doi.org/10.1016/j.jcz.2020.10.010
  • Varón-González C, Whelan S, Klingenberg CP (2020) Estimating phylogenies from shape and similar multidimensional data: Why it is not reliable. Systematic Biology 69(5): 863–883. https://doi.org/10.1093/sysbio/syaa003
  • Ward RD, Zemlak TS, Innes HB, Last RP, Hebert PDN (2005) DNA barcoding Australia’s Fish species. Philosophical Transactions of the Royal Society of London. Proceedings. Biological Sciences 360(1462): 1847–1857. https://doi.org/10.1098/rstb.2005.1716
  • Wilson B (2013) The biogeography of the Australian North West Shelf: Environmental change and life’s response. Newnes, 415 pp.
  • Yokoyama Y, Purcell A, Lambeck K, Johnston P (2001) Shore-line reconstruction around Australia during the last glacial maximum and late glacial stage. Quaternary International 83: 9–18. https://doi.org/10.1016/S1040-6182(01)00028-3
  • Yu Y, Blair C, He X (2020) RASP 4: ancestral state reconstruction tool for multiple genes and characters. Molecular Biology and Evolution 37(2): 604–606. https://doi.org/10.1093/molbev/msz257
  • Zachos JC, Shackleton NJ, Revenaugh JS, Palike H, Flower BP (2001) Climate response to orbital forcing across the Oligocene-Miocene boundary. Science 292(5515): 274–278. https://doi.org/10.1126/science.1058288

Supplementary material

Supplementary material 1 

Sampling information and GenBank accession numbers for the specimens included in the phylogenetic analyses

Ehsan Damadi, Faezeh Yazdani Moghaddam, Mehdi Ghanbarifardi

Data type: table (excel file)

This dataset is made available under the Open Database License (http://opendatacommons.org/licenses/odbl/1.0/). The Open Database License (ODbL) is a license agreement intended to allow users to freely share, modify, and use this Dataset while maintaining this same freedom for others, provided that the original source and author(s) are credited.
Download file (28.93 kb)
login to comment