Next Article in Journal
Chronic Treatment with Nigella sativa Oil Exerts Antimanic Properties and Reduces Brain Inflammation in Rats
Next Article in Special Issue
Comparative Mitogenome Analyses of Fifteen Ramshorn Snails and Insights into the Phylogeny of Planorbidae (Gastropoda: Hygrophila)
Previous Article in Journal
Rapid Biodistribution of Fluorescent Outer-Membrane Vesicles from the Intestine to Distant Organs via the Blood in Mice
Previous Article in Special Issue
The Complete Mitochondrial Genome and Phylogenetic Analysis of the Freshwater Shellfish Novaculina chinensis (Bivalvia: Pharidae)
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Mitogenomic Characterization and Phylogenetic Placement of African Hind, Cephalopholis taeniops: Shedding Light on the Evolution of Groupers (Serranidae: Epinephelinae)

1
Institute of Fisheries Science, Pukyong National University, Busan 48513, Republic of Korea
2
Institute of Marine Life Science, Pukyong National University, Busan 48513, Republic of Korea
3
Marine Integrated Biomedical Technology Center, National Key Research Institutes in Universities, Pukyong National University, Busan 48513, Republic of Korea
4
Ocean Georesources Research Department, Korea Institute of Ocean Science and Technology, Busan 49111, Republic of Korea
5
Advance Tropical Biodiversity, Genomics, and Conservation Research Group, Department of Biology, Faculty of Science and Technology, Airlangga University, Surabaya 60115, Indonesia
6
Department of Marine, Faculty of Fisheries and Marine, Airlangga University, Surabaya 60115, Indonesia
7
Department of Marine Biology, Pukyong National University, Busan 48513, Republic of Korea
8
International Graduate Program of Fisheries Science, Pukyong National University, Busan 48513, Republic of Korea
*
Author to whom correspondence should be addressed.
Int. J. Mol. Sci. 2024, 25(3), 1822; https://doi.org/10.3390/ijms25031822
Submission received: 20 December 2023 / Revised: 23 January 2024 / Accepted: 31 January 2024 / Published: 2 February 2024

Abstract

:
The global exploration of evolutionary trends in groupers, based on mitogenomes, is currently underway. This research extensively investigates the structure of and variations in Cephalopholis species mitogenomes, along with their phylogenetic relationships, focusing specifically on Cephalopholis taeniops from the Eastern Atlantic Ocean. The generated mitogenome spans 16,572 base pairs and exhibits a gene order analogous to that of the ancestral teleost’s, featuring 13 protein-coding genes (PCGs), two ribosomal RNA genes (rRNAs), 22 transfer RNA genes (tRNAs), and an AT-rich control region. The mitogenome of C. taeniops displays an AT bias (54.99%), aligning with related species. The majority of PCGs in the mitogenome initiate with the start codon ATG, with the exceptions being COI (GTG) and atp6 (TTG). The relative synonymous codon usage analysis revealed the maximum abundance of leucine, proline, serine, and threonine. The nonsynonymous/synonymous ratios were <1, which indicates a strong negative selection among all PCGs of the Cephalopholis species. In C. taeniops, the prevalent transfer RNAs display conventional cloverleaf secondary structures, except for tRNA-serine (GCT), which lacks a dihydrouracil (DHU) stem. A comparative examination of conserved domains and sequence blocks across various Cephalopholis species indicates noteworthy variations in length and nucleotide diversity. Maximum likelihood, neighbor-joining, and Bayesian phylogenetic analyses, employing the concatenated PCGs and a combination of PCGs + rRNAs, distinctly separate all Cephalopholis species, including C. taeniops. Overall, these findings deepen our understanding of evolutionary relationships among serranid groupers, emphasizing the significance of structural considerations in mitogenomic analyses.

1. Introduction

The mitochondrial genome consists of compact, circular molecules, spanning 16–17 kilobase pairs in size, housing highly conserved encoded genes that play an essential role in the vitality of nearly all eukaryotes [1]. The mitogenome is typically composed of 13 protein-coding genes (PCGs), two ribosomal RNAs (rRNAs), 22 transfer RNAs (tRNAs), and two noncoding regions—the control region (CR) and the origin of L-strand replication (OL) [2]. Notably, the gene order within the mitogenome exhibits a remarkable degree of conservation among vertebrates, including teleosts [3]. Due to several advantageous features, such as maternal inheritance and a higher mutation rate, the mitochondrial genome has found extensive use as a powerful tool for conducting phylogenetic and population genetic analyses in vertebrates, with a particular emphasis on fish [4,5]. Consequently, it is imperative to unravel the structure and variability of the mitogenome in any organism to comprehend its functions [6]. While mitogenomic research on fish has been extensively initiated globally, there exists a substantial gap in knowledge within numerous classified groups.
Groupers within the genus Cephalopholis (Serranidae: Epinephelinae) represent primarily tropical and subtropical marine fish species, with a global presence encompassing 25 extant species [7,8]. The majority of these species inhabit reef ecosystems extending from the Red Sea to the Indo-Pacific region, while a subset confines their distribution to the Eastern Pacific and the Western to Eastern Atlantic. Economically significant, these species are actively pursued by commercial, artisanal, and recreational fishermen [9,10]. The African hind or blue-spotted seabass, Cephalopholis taeniops, inhabits sandy or rocky seafloors at depths ranging from 20 to 200 m in the Eastern Atlantic, spanning from Morocco south to Angola, including the Canary Islands (Spain), Cape Verde Islands, and São Tomé and Principe [7]. Subsequently, this species has been documented in the Canary Islands, the Mediterranean Sea in Libya and Israel, as well as in the Southwestern Atlantic Ocean [10,11,12,13,14,15,16]. Beyond the Eastern Atlantic, populations in the Mediterranean and Southwestern Atlantic are considered introduced populations of C. taeniops [16]. Commercially, C. taeniops is predominantly utilized for local human consumption within its distribution area and is frequently exported to Europe from Senegal [7,11]. Despite the scarcity of available data regarding population biology, reproductive strategies, and the actual impact of fishing, C. taeniops is classified as a ‘Least Concern’ species in the IUCN Red List of Threatened Species due to insufficient evidence. Nevertheless, groupers remain highly vulnerable to fishing pressure, owing to life history traits such as longevity, late sexual maturation, and aggregation spawning [17].
Within the Serranidae family, focused investigations on Cephalopholis species have primarily addressed aspects such as physiology and reproductive life history [18,19], genetic connectivity, phylogeography, evolutionary implications through Pleistocene isolation [20,21], and the assembly of chromosome-level genomes alongside transcriptome comparisons [22]. Further, in conjunction with other serranid species, comprehensive examinations of the phylogenetic relationships of Cephalopholis species have involved the utilization of two mitochondrial (16S rRNA and 12S rRNA) and two nuclear (Tmo-4C4 and histone H3) genes [23,24]. Numerous molecular datasets have been generated for serranid species, contributing to the discovery of new species [25], the delineation of cryptic diversity [26], species discrimination in hybrid zones [27], and the elucidation of phylogenetic relationships [28,29]. Nevertheless, the systematic classification of serranids has experienced frequent changes across various taxonomic levels, leading to inconsistencies, a situation rectified through the application of three mitochondrial genes (COI, 16S rRNA, and 12S rRNA) and one nuclear gene (TMO4C4) [30]. The resulting molecular data have been instrumental in estimating phylogeographic patterns, demographic history [31,32,33,34], gaining insights into population dynamics and adaptive radiation [35], detecting biological invasions [36], and contributing to the improvement of fisheries management and grouper conservation [37]. Subsequent advancements in molecular tools have further facilitated the exploration of phylogenomics and population genomics within this fish group, unveiling intraspecific variation, adaptation differentiation, physiological changes, radiations, and the intricate processes of speciation [38,39,40].
Examining the complete mitochondrial genome information of serranids, a comprehensive set of 94 species sequences is presently available in the global GenBank database (https://www.ncbi.nlm.nih.gov, accessed on 12 December 2023). Notably, the majority of these mitogenomes have been generated within the subfamily Epinephelinae and the tribe Epinephelini. The database is further enriched by the inclusion of mitogenomes from nine Cephalopholis species [41,42,43,44,45,46]. Given the demonstrated efficacy of mitochondrial genomes in elucidating phylogenetic inferences within teleosts [47,48,49], there is an urgent need to enhance taxonomic coverage for a more profound understanding of the evolutionary relationships within the targeted group. Novel mitogenomic data not only provide insights into the structural variations in mitochondrial genes (PCGs, rRNAs, tRNAs, and D-loop), but also shed light on their functions [50,51]. Additionally, mitogenomic data prove valuable in environmental DNA metabarcoding studies and contribute to the conservation genetics of teleosts [52].
Furthermore, the imperative need for the meticulous tracking of tropical marine biodiversity extends to both established hotspots and peripheral ecosystems [53]. The study region, situated within the West African hotspot, recognized as one of the global biodiversity hotspots [54], is characterized by a limited understanding of the diversity, origin, and evolution of reef-associated fish, specifically within Cephalopholis species in the Eastern Atlantic Ocean. To overcome these challenges, leveraging genetic information and employing phylogenetic approaches have proven to be successful strategies in elucidating the underlying causes of richness patterns across global marine regions. This involves a comparative assessment of the relative importance of colonization time, the number of colonization events, and diversification rates [55]. The adoption of such integrated approaches is instrumental not only in enhancing our comprehension of these aspects, but also in formulating effective conservation priorities for global marine biodiversity. These priorities, spanning multiple dimensions, play a pivotal role in the context of marine protected areas (MPAs), supporting our capacity to combat and adapt to climate change [56,57]. Therefore, the present study aims to generate a novel mitogenome of the African Hind, C. taeniops, from the Eastern Atlantic Ocean, characterizing its genomic features in comparison with other congeners. The research also conducts cladistic analyses to elucidate the evolutionary relationships of the targeted species with other major lineages of serranids. The genetic data acquired will play a crucial role in validating the population structure of the African Hind and its applications in the field of conservation genetics.

2. Results and Discussion

2.1. Mitogenome Structure and Organization

In the current investigation, we elucidated the mitogenome of C. taeniops, revealing a length of 16,572 base pairs (bp) with GenBank accession no. OQ420715. Notably, the mitogenome of C. taeniops exhibited the shortest length among the Cephalopholis species, ranging from 16,585 bp (Cephalopholis leopardus and Cephalopholis miniata) to 16,771 bp (Cephalopholis boenak). Comprising 13 PCGs, 22 tRNAs, two rRNAs, and an AT-rich control region, the mitogenome of C. taeniops displayed a unique arrangement. The positive strand accommodated 12 PCGs, two rRNAs, and 14 tRNAs, while the negative strand housed ND6 and 8 tRNAs (Table 1, Figure 1). The gene order in C. taeniops, along with other species (C. boenak, C. leopardus, C. miniata, Cephalopholis sexmaculata, Cephalopholis sonnerati, Cephalopholis urodeta, and Cephalopholis spiloparaea), mirrored that of the ancestral teleosts, with the exception of Cephalopholis argus, which exhibited content duplication of trnD and CR, as reported previously [41] (Figure 1).
The mitogenome of C. taeniops demonstrated an AT bias (54.99%), with nucleotide composition comprising 28.93% A, 26.06% T, 16.27% G, and 28.74% C. Similar AT biasness was observed in other Cephalopholis species, ranging from 54.99% (C. taeniops) to 56.94% (C. boenak). In the C. taeniops mitogenome, the AT skew and GC skew were calculated as 0.052 and −0.277, respectively. Comparative analysis with other Cephalopholis mitogenomes revealed an AT skew ranging from 0.026 (C. argus) to 0.061 (C. sonnerati) and a GC skew ranging from −0.282 (C. leopardus and C. sonnerati) to −0.247 (C. argus) (Table S2).
Our investigation further unveiled nine intergenic spacers totaling 64 bp and five overlapping regions spanning 23 bp in the C. taeniops mitogenome. The longest intergenic spacer (37 bp) was identified between tRNA-Asn (N) and tRNA-Cys (C), while the most extensive overlapping region (10 bp) occurred between ATP synthase 8 (atp8) and ATP synthase 6 (atp6) genes. A comparative analysis with other Cephalopholis mitogenomes disclosed intergenic spacer numbers ranging from 9 to 12, with the most extended spacer (37 bp to 40 bp) observed between trnN and trnC (Table S3). Moreover, six overlapping regions were consistently observed in most Cephalopholis mitogenomes, with a maximum length of 10 bp between atp8 and atp6. Notably, an overlapping region (1 bp) between atp6 and cytochrome c oxidase subunit III (COIII) was common in most Cephalopholis species but absent in C. taeniops. Additionally, an unconventional intergenic spacer (3 bp) was noted between tRNA-Thr (T) and tRNA-Pro (P) in C. argus, contrasting with the single base pair overlap present in the same gene boundary in other species (Table S3). The observed genetic variations within the mitogenomes of Cephalopholis species provide valuable insights into their evolutionary processes and energy metabolism, consistent with comparable findings in other fish species [58]. This investigation adds crucial information about the structural characteristics of Cephalopholis mitogenomes, thereby enhancing our comprehension of the functions encoded by these mitogenomes and their constituent genes.

2.2. Protein-Coding Genes

The mitogenome of C. taeniops comprises 13 PCGs, with the shortest length observed in atp8 and the longest in ND5. The total length of C. taeniops PCGs is 11,301 base pairs (bp), constituting 68.19% of the complete mitogenome. In contrast to other Cephalopholis species, the total PCG length ranges from 11,301 bp (C. taeniops) to 11,430 bp (C. argus). Seven additional species (C. boenak, C. leopardus, C. miniata, C. sexmaculata, C. sonnerati, C. urodeta, and C. spiloparaea) maintain an equal PCG length of 11,429 bp in their mitogenomes (Table S2). The PCGs in Cephalopholis species exhibit an AT bias ranging from 54.57% (C. taeniops) to 56.7% (C. argus). AT skews and GC skews in Cephalopholis PCGs vary from −0.060 (C. argus) to −0.012 (C. miniata and C. spiloparaea) and −0.328 (C. leopardus) to −0.277 (C. argus). In the mitogenome of C. taeniops, most PCGs start with ATG, except for COI (GTG) and atp6 (TTG). The COI gene initiates with GTG in all Cephalopholis species, while atp6 starts with TTG in three species and CTG in six others. Notably, an exception is observed with ND4 in C. argus, which starts with GTG instead of ATG. In C. taeniops, six PCGs terminate with TAA, COIII with GGC, and the remaining PCGs exhibit incomplete stop codons. An exception is ND1 in C. boenak, COI in five species, ND5 in C. argus, and ND6 in two species, where TAG serves as the stop codon (Table S4). Nucleotide diversity analysis, using a sliding window approach on concatenated PCGs, yielded an average nucleotide diversity value (Pi) of 0.13191, with 3582 polymorphic sites across all Cephalopholis species (Figure 2A). A saturation analysis indicated non-saturation for both transitions and transversions with increasing Kimura 2-parameter (TN84) divergence values (Figure 2B).

2.3. Substitutions Pattern and Codon Usage

Darwinian selection stands as a pivotal hypothesis in elucidating the evolutionary dynamics of genes under positive selection, playing a crucial role in the divergence of species [59,60,61,62,63]. The examination of synonymous (Ks) and nonsynonymous (Ka) substitution rates within PCGs provides evidence for Darwinian selection and adaptive molecular evolution in vertebrates [64,65]. The Ka/Ks ratio serves as an established indicator of selective pressure and evolutionary relationships at the molecular level, applicable to both homogenous and heterogeneous species [66,67]. In this study, we investigated the evolutionary rates between homologous gene pairs by calculating Ka/Ks substitutions for C. taeniops and comparing them with Cephalopholis species. The Ka/Ks ratio ranged from 0.0109 ± 0.004 in nad5 to 0.1986 ± 0.030 in cox3 and the resulted following order: nad5 < nad4L < nad3 < nad1 < nad6 < nad4 < cox2 < cox1 < atp8 < nad2 < atp6 < cytb < cox3 (Figure 2C). Most PCGs exhibited Ka/Ks values less than 1, indicating a strong negative selection among the studied Cephalopholis species, suggesting that the mutations were replaced by synonymous substitutions (Table S5). A comparative analysis of the Ka/Ks ratio among 13 PCGs of Cephalopholis species showed that all PCGs are evolving under negative selection (Table S6, Figure S1). This observation reflects the influence of natural selection in mitigating deleterious mutations with negative selective coefficients, aligning with general patterns observed in other vertebrates [65,66]. Thus, the comparative analysis of Ka/Ks within Cephalopholis species’ mitogenomes offers a platform for gaining insights into the nuances of natural selection shaping the evolutionary trajectory of species. This analysis aids in unraveling the intricate interplay between mutations and selective pressures, elucidating their collective role in steering the evolution of proteins. The codons corresponding to each amino acid were found to be conserved across all PCGs in the compared Cephalopholis species. An analysis of RSCU unveiled a maximal abundance of leucine, proline, serine, and threonine in the PCGs of C. taeniops, whereas aspartic acid, cysteine, glutamic acid, and tryptophan were less prevalent (Figure 2D). Similar amino acid abundance patterns were observed in other Cephalopholis species, mirroring those in C. taeniops (Table S7). Notably, the RSCU analysis demonstrated a significant decrease in the frequency of the ACG codon in threonine and the AGT codon in serine in C. taeniops (Figure S2). A comparative RSCU analysis further revealed a substantial reduction in the frequency of the GCG codon in alanine across most species (C. argus, C. leopardus, C. miniata, C. sexmaculata, C. sonnerati, C. urodeta, C. spiloparaea), except in C. boenak where ACT was observed in threonine (Table S7).

2.4. Ribosomal RNA and Transfer RNA Genes

The mitogenome of C. taeniops encompasses two ribosomal RNA molecules, namely, 12S rRNA (953 bp) and 16S rRNA (1712 bp), collectively contributing to 16.08% of the entire mitogenome. A comparative analysis with other Cephalopholis species revealed varying lengths of rRNAs, ranging from 2663 bp (C. miniata and C. spiloparaea) to 2679 bp (C. argus). The rRNA genes exhibit AT bias, ranging from 53.03% (C. taeniops) to 54.91% (C. argus) (Table S2). The AT skews and GC skews range from 0.184 (C. argus) to 0.263 (C. boenak) and −0.136 (C. boenak) to −0.084 (C. argus), respectively. In C. taeniops, the cumulative length of tRNA genes is 1565 bp, contributing 9.44% to the entire mitogenome. A comparative assessment with other Cephalopholis species shows the total tRNA length varying from 1562 bp (C. sonnerati) to 1638 bp (C. argus). The tRNA genes of Cephalopholis species display AT bias, ranging from 55.62% (C. boenak) to 57.63% (C. argus), with the AT skew ranging from 0.008 (C. taeniops) to 0.031 (C. boenak) (Table S2). Ribosomes, universally conserved ribonucleoproteins, facilitate the translation of genetic information from mRNAs to proteins. The structural arrangement of these rRNA genes, especially conserved loops, provides crucial insights into the catalytic processes underlying protein synthesis [68]. In C. taeniops, most tRNAs exhibit classic clover-leaf secondary structures, except for trnS1 (GCT), which lacks the DHU stem (Figure S3). Wobble base pairings are observed in 18 tRNAs, with the highest number in trnA and trnE (Figure S3). These pairings occur in various tRNA regions, including the DHU stem, acceptor stem, TψC stem, and anticodon stem, indicating sites with G × U/T pairs for recognition by proteins and other RNAs. The anticodons of all 21 tRNAs exhibited uniformity across all Cephalopholis species, with the exception of tRNA-Ser (S1) (AGC), which was exclusively identified in three species, C. taeniops, C. argus, and C. boenak. Notably, a duplication event of tRNA-Asp (D) in C. argus resulted in both tRNAs featuring the GAC anticodon (Table S8). This allows wobble pairs to play essential roles in diverse biological processes [69,70]. tRNAs, acting as adaptor molecules, facilitate the translation of genetic information into protein sequences by delivering amino acids during translation. Additionally, gene rearrangements of tRNAs and high levels of length heteroplasmy in the WANCY region contribute to illuminating the evolution of mitochondrial genes [71,72].

2.5. Features of Control Region

The CR of C. taeniops spans 873 base pairs (bp), representing 5.27% of the total mitogenome. A comparative analysis across various Cephalopholis species reveals a diverse range of CR lengths, ranging from 813 bp (C. argus) to 1064 bp (C. boenak). The CRs exhibit an AT bias, varying from 62.24% (C. argus) to 70.11% (C. boenak), with AT skews ranging from −0.035 (C. boenak) to 0.103 (C. argus) (Table S2). In scrutinizing the extant mitogenomes of six Cephalopholis species, a comprehensive examination of diverse domains was conducted, following the methodology outlined for C. sonnerati (KC593378) [41]. A comparative analysis reveals the presence of six central conserved domains (CCDs), CSB-A, CSB-B, CSB-C, CSB-D, CSB-E, and CSB-F, along with three conserved sequence blocks (CSBs), CSB-1, CSB-2, and CSB-3, within the CR of C. taeniops and the other six Cephalopholis species (C. leopardus, C. miniata, C. sexmaculata, C. sonnerati, C. urodeta, and C. spiloparaea). This observation aligns with patterns identified in other teleost mitogenomes [6,41]. Among the CCDs, CSB-D stands out as the longest at 33 base pairs, while CSB-A, CSB-B, CSB-C, CSB-E, and CSB-F exhibit lengths of 18 base pairs, 18 base pairs, 28 base pairs, 19 base pairs, and 20 base pairs, respectively (Figure 3). Comparative analyses unveil significant nucleotide variability within CSB-A and CSB-E (four and six base pairs, respectively), while other CCDs remain largely conserved across all Cephalopholis species. Within the CSBs, CSB-1 is the longest at 23 base pairs, with CSB-2 and CSB-3 exhibiting lengths of 18 base pairs and 20 base pairs, respectively (Figure 3). Comparative analyses of CSBs reveal notable nucleotide variability within CSB-3 (four base pairs), while CSB-1 and CSB-2 remain predominantly conserved. Due to unprecedented length variation and heteroplasmy, the CRs of C. argus (KC593377) and C. boenak (KC537759) were not investigated. The CR of C. argus lacks certain sequence elements found in other grouper species, shares no significant sequence similarity, and features an additional tRNA-Asp (D) insertion in the middle of the region. Except for C. sexmaculata (2.7 copy number of consensus 17 bp: CATATATGTATAGTAAC), no other Cephalopholis species’ CRs contain tandem repeats in the extended termination-associated sequences (ETAS) region. The repeat-rich extended termination-associated sequence (ETAS) region emerges as the most dynamically variable segment within the CR, characterized by specific motifs. This variability prompts the formation of stable hairpin loops, hypothesized to function as sequence-specific signals for the termination of mitochondrial DNA (mtDNA) replication [41]. Beyond the conserved attributes, the CRs of Cephalopholis species harbor highly polymorphic sequences, offering a robust tool for distinguishing between species and elucidating population structures in fish, a phenomenon well documented in other fish species, including groupers [73,74]. Furthermore, the intricate mechanisms governing the CR, such as genomic rearrangement through double replications, random loss, dimer-mitogenomes, and non-random loss, contribute significantly to comprehending the structural diversity of mitogenomes and the complexities inherent in the evolution of mitochondrial genomes.

2.6. Genetic Distances and Phylogenetic Relationship

The serranids exhibit an overall mean genetic distance of 25.4% in the current mitogenomic dataset. Among the Cephalopholis species, inter-species genetic distances range from 0.06% (C. miniata and C. spiloparaea) to 20.8% (C. argus and C. spiloparaea). In the four tribes (Epinephelini, Diploprionini, Grammistini, and Liopropomini) within the subfamily Epinephelinae, genetic distances range from 28% to 30.9%. Furthermore, the subfamily Epinephelinae demonstrates genetic distances of 30.2% to 32.5% with other serranid subfamilies, Serraninae and Anthiinae, respectively. The maximum likelihood (ML) topology, constructed using concatenated protein-coding genes, distinctly separates all serranids, elucidating their phylogenetic relationships (Figure 4). Taxa from different taxonomic lineages, both at the tribe and subfamily levels, exhibit clear clustering patterns. Cephalopholis species display a cohesive and monophyletic clustering pattern in the cladistics analysis, aligning with previous evolutionary hypotheses on groupers [24,30]. Compelling evidence supports the consideration of Cephalopholis as a valid genus, as indicated by both the recent cladistic analysis based on mitogenomes and previous screenings using partial mitochondrial and nuclear genes [24,28,30]. Notably, the mitogenome of Aethaloperca rogaa clusters closely with C. argus and C. boenak in the ML phylogeny. The classification of the genus Aethaloperca is contentious, and A. rogaa, based on morphological features, was previously considered to belong to a monotypic genus [7]. However, recent studies propose synapomorphic characters that unite Aethaloperca with Cephalopholis, supporting its placement within the latter. The present mitogenome-based ML phylogeny corroborates this, designating A. rogaa nested under the Cephalopholis clade. Nevertheless, it is essential to subject this hypothesis to scrutiny through a comprehensive examination of additional morphological characteristics pertaining to these groupers and their related counterparts. Within Cephalopholis species, two subclusters emerge in the ML phylogeny. Three species (C. argus, C. boenak, and C. rogaa) cluster together, while the other seven species form a separate subclade. This subclustering does not seem to reflect any distributional distinction of these groupers in marine environments.
Additional studies on taxonomy and in-depth molecular data may provide further clarity on their evolutionary significance in marine ecosystems. Notably, the targeted species, C. taeniops, closely clusters with six other Cephalopholis species (C. leopardus, C. miniata, C. sexmaculata, C. sonnerati, C. spiloparaea, and C. urodeta) (Figure 4). Nevertheless, based on the cladding pattern, it can be asserted that C. taeniops serves as the ancestral species when compared to other closely related Cephalopholis species. In addition to the PCGs, the phylogenetic assessment of the studied serranids was reevaluated by incorporating both PCGs and rRNAs. rRNAs play a crucial role in protein synthesis, and their structures are highly conserved among different organisms [75]. Consequently, accounting for structural considerations becomes particularly important when aligning rRNA genes, especially for phylogenetic analyses, as such inferences necessitate comparisons of homologous characters across diverse sequences [76,77]. The NJ and BA topologies, based on 13 PCGs and 13 PCGs + two rRNAs, reveal a clustering pattern similar to that depicted by the concatenated PCGs alone (Figures S4–S6). Overall, this study underscores the effectiveness of mitochondrial genes in discriminating and elucidating the evolutionary relationships of serranids, as observed in other fish [78,79]. Large-scale genomic data offer valuable insights into time-calibrated phylogeny, adaptation to euryhalinity, and speciation [80]. Nevertheless, the adverse effects of global warming and extreme temperature events have impacted marine biodiversity, ecosystem functions, and services across all ocean basins over the past two decades, leading to significant losses in fisheries revenues and livelihoods in most maritime countries [81]. Recognizing the crucial role that genomic data play in conservation genetics and fish management [82], this study advocates for the generation of additional large-scale genomic data on groupers. This initiative aims to contribute to a more profound understanding of their evolution, diversification, and adaptation in marine environments.

3. Materials and Methods

3.1. Sampling and Species Identification

The Cephalopholis grouper specimen was obtained from the Atlantic Ocean, situated off the coast of Cameroon, Africa (Figure 1). Species identification was corroborated as C. taeniops, relying on morphological characteristics outlined in prior literature [7,11,12]. The body of the species is robust and slightly compressed, displaying a red-orange hue. It is adorned with numerous light blue spots, each surrounded by black margins, and the posterior part of all fins exhibits a slightly darker shade. The head and mouth are large, with a protruding lower jaw that features two robust canines at the anterior of each jaw. The upper arch boasts four knob-like undeveloped gill rakers and four fully developed ones, while the lower arch comprises three knob-like undeveloped gill rakers and 12 fully developed ones. The dorsal fin is continuous and equipped with nine spines, the anal fin features three spines, and the pelvic fin is supported by one spine. The posterior nostril is positioned close to the eye at the upper third level, while the anterior nostril is very close and slightly lower, furnished with a small flap. Muscle tissue from the ventral thoracic region was precisely excised and placed under sterile conditions within the Department of Marine Biology at Pukyong National University in Busan, Republic of Korea. Voucher specimens were accurately preserved in 10% formaldehyde at the Fisheries and Animal Industries (MINEPIA) facility in Yaoundé, Cameroon. Approval for the research protocol, granted by the Institutional Animal Care and Use Committee (IACUC) under the code PKNUIACUC-2022-72 on 15 December 2022, confirms that the use of biological material in the experiments adhered to ethical standards, ensuring that the targeted fish were not subjected to harm by the researchers. The range distribution of C. taeniops is mapped according to the IUCN data (.shp files) and the additional records published in the previous literature [7,10,11,12,13,14,15,16] (Figure 1).

3.2. DNA Extraction, Sequencing, and Assembly

The extraction of genomic DNA was conducted using the AccuPrep® DNA extraction kit from Bioneer, situated in Daejeon, the Republic of Korea, following established standard protocols. The quality and quantity of the genomic DNA were thoroughly assessed using a NanoDrop spectrophotometer (Thermo Fisher Scientific D1000, Waltham, MA, USA). To obtain the complete mitogenome of C. taeniops, sequencing procedures were carried out on the NovaSeq platform at Macrogen (https://dna.macrogen.com/, accessed on 12 December 2023) in Daejeon, the Republic of Korea, facilitated by the Illumina platform. Sequencing libraries were prepared according to the manufacturer’s specifications for the TruSeq Nano DNA High-Throughput Library Prep Kit (Illumina, Inc., San Diego, CA, USA). In summary, 100 ng of genomic DNA underwent fragmentation utilizing adaptive focused acoustic technology (Covaris, Woburn, MA, USA), resulting in double-stranded DNA molecules with blunt ends and 5′-phosphorylation. Following the end-repair step, DNA fragments were size-selected using a bead-based method, modified with the addition of a single ‘A’ base, and ligated with TruSeq DNA UD Indexing adapters. The products were purified and enriched through PCR to generate the final DNA library. Library quantification was carried out using qPCR, following the qPCR Quantification Protocol Guide (KAPA Library Quantification Kits for Illumina Sequencing Platforms), and a quality assessment was performed using Agilent Technologies 4200 TapeStation D1000 screentape (Agilent Technologies, Santa Clara, CA, USA). Paired-end (2 × 150 bp) sequencing was conducted via Macrogen on the NovaSeq platform (Illumina, Inc., San Diego, CA, USA).

3.3. Mitogenome Assembly and Validation of Control Region

The processing of over 20 million raw reads was undertaken using the Cutadapt tool (http://code.google.com/p/cutadapt/, accessed on 12 December 2023) to trim adapters and remove low-quality bases with a Phred quality score (Q score) cutoff of 20. An assembly of the targeted genome from high-quality paired-end next-generation sequencing (NGS) reads was performed using Geneious Prime version 2023.0.1, employing reference mapping with the mitogenome of a closely related species as a reference and utilizing default mapping algorithms. To validate the mitogenome assembly, the alignment of overlapping regions was scrutinized using MEGA X [83]. The boundaries and orientations of individual genes were confirmed through the MITOS v806 (http://mitos.bioinf.uni-leipzig.de, accessed on 12 December 2023) and MitoAnnotator (http://mitofish.aori.u-tokyo.ac.jp/annotation/input/, accessed on 12 December 2023) web servers [84,85]. For the validation of protein-coding genes (PCGs), the translated putative amino acid sequences underwent analysis using the Open Reading Frame Finder web tool (https://www.ncbi.nlm.nih.gov/orffinder/, accessed on 12 December 2023), based on the vertebrate mitochondrial genetic code. Additionally, to confirm the full-length control region, a target-specific primer pair (5′-CGAGCACTAACCTTCCGACC-3′ and 5′-GGCTAAGCAAGGTGTCGTG-3′) was designed for further amplification. The PCR was conducted using the TaKaRa Verity Thermal Cycler with a 1X PCR buffer, 1 U Taq polymerase, 10 pmol primers, 2.5 mM dNTPs, and 1 µL template DNA. Purification of the PCR products was carried out using the AccuPrep® PCR/Gel Purification Kit (Bioneer, Daejeon, Republic of Korea). Subsequently, the amplicons were subjected to amplification with the BigDye® Terminator v3.1 Cycle Sequencing Kit (Applied Biosystems, Foster City, CA, USA) and sequenced in both directions utilizing the ABI PRISM 3730XL DNA analyzer available at Macrogen (https://dna.macrogen.com/, accessed on 12 December 2023), Daejeon, the Republic of Korea. The assembly of the control region with the complete mitogenome involved ensuring the alignment of overlapping regions through MEGA X, after eliminating any noisy segments via the SeqScanner version 1.0 (Applied Biosystems Inc., Foster City, CA, USA). The resulting C. taeniops mitogenome was appropriately submitted to the global GenBank database to acquire a unique accession number.

3.4. Characterization and Comparative Analyses

We utilized the MitoAnnotator (http://mitofish.aori.u-tokyo.ac.jp/annotation/input/, accessed on 12 December 2023) to construct a three-dimensional representation of the generated mitogenome. Our comprehensive comparative analysis aimed to evaluate the mitogenomic architecture and variations in our sequenced data in comparison to eight existing mitogenomes of Cephalopholis species [41,42,43,44,45,46] (Table S1). Due to taxonomic uncertainties found in Eschmeyer’s Catalog of Fishes, GenBank, and previous investigations [7,30], the mitochondrial genome of Aethaloperca rogaa (currently Cephalopholis rogaa) (KC593376) was excluded from the comparative analysis of structure and variation. Manually calculated values included the intergenic spacers between adjacent genes and overlapping regions. Nucleotide compositions within protein-coding genes (PCGs), rRNAs, tRNAs, and the control region (CR) were determined using MEGA X. A sliding window analysis of nucleotide diversity, with a window size of 200 bp and a step size of 25 bp, was conducted using DnaSP6.0 [86]. Base composition skews were computed using established formulas: AT-skew = [A − T]/[A + T] and GC-skew = [G − C]/[G + C] [87]. The saturation of the transition codon of the mitochondrial PCGs based on transition (s) and transversion (v), as well as AT and GC skews, was depicted using DAMBE6 [88]. The validation of the initiation and termination codons for each PCG, along with compliance with the vertebrate mitochondrial genetic code, was performed using MEGA X. The analysis involved the computation of relative synonymous codon usage (RSCU), the relative abundance of amino acids, and the distribution of codons using DnaSP6.0. Subsequently, pairwise tests for synonymous (Ks) and nonsynonymous (Ka) substitutions were conducted between Cephalopholis taeniops and other Cephalopholis species, employing DnaSP6.0. Additionally, the boundaries of rRNA and tRNA genes were confirmed through the utilization of the tRNAscan-SE Search Server 2.0 in conjunction with ARWEN 1.2 [89,90]. The identification of structural domains within the control region was achieved by conducting CLUSTAL X alignments, following previous research [41,91].

3.5. Genetic Distance and Phylogenetic Analyses

Pairwise genetic distances among various taxonomic levels within the Serranidae family were calculated using the Kimura 2-parameter (K2P) method within MEGA X. For the phylogenetic analysis, excluding Cephalopholis species, 16 mitogenomes from representative species across four tribes of the subfamily Epinephelinae (Epinephelini, Diploprionini, Grammistini, and Liopropomini) were retrieved from GenBank [6,41,46,92,93,94] (Table S1). Anthias nicholsi (OP056908) from the subfamily Anthiinae and Serranus papilionaceus (OK054500) from the subfamily Serraninae were designated as outgroup taxa [94,95]. Two concatenated datasets (13 PCGs and 13 PCGs + two rRNAs) were assembled using the iTaxoTools 0.1 tool to elucidate the primary evolutionary relationships among serranids, with a particular focus on Cephalopholis species within the subfamily Epinephelinae and tribe Epinephelini [96]. The neighbor-joining (NJ) phylogeny was constructed using the PCGs dataset and the K2P model through MEGA X. Model selection analysis identified the ‘GTR + G + I’ model as the most suitable, serving as the optimal model for all PCGs and yielding the lowest Bayesian information criterion (BIC) scores. This model selection process was conducted through PartitionFinder 2 on the CIPRES Science Gateway v3.3 and JModelTest v2 [97,98,99]. The maximum likelihood (ML) phylogeny was constructed using the IQ-Tree web server with 1000 bootstrap samples and PhyML 3.0, following the standard protocol [100,101]. A Bayesian (BA) tree was constructed using Mr. Bayes 3.1.2, with nst = 6, involving one cold and three hot metropolis-coupled Markov chain Monte Carlo (MCMC) chains. The analysis ran for 10,000,000 generations, with tree sampling at every 100th generation, and 25% of the samples were discarded as burn-in [102]. The resulting BA tree was visualized using the iTOL v4 web server (https://itol.embl.de/login.cgi, accessed on 12 December 2023) [103].

4. Conclusions

The escalating impacts of global warming and extreme temperature events pose a threat to marine biodiversity, resulting in considerable losses in fisheries worldwide. Groupers (Perciformes: Serranidae) are economically and recreationally valuable reef-associated fish. Beside species invention, the understanding of evolutionary patterns in groupers based on mitogenomes is currently limited on a global scale. This research extensively delves into the structure of and variations in Epinephelinae mitogenomes, emphasizing the complete mitogenome of Cephalopholis taeniops from the Eastern Atlantic Ocean. The study reveals substantial genetic divergence between C. taeniops and its congeners, providing crucial insights into the evolutionary dynamics of groupers. These findings offer valuable resources for further investigations into grouper species identification, conservation genetics, speciation, and other evolutionary biology studies. The acquired genetic knowledge is pivotal for formulating effective conservation strategies in MPAs, safeguarding species diversity, and ensuring the sustainability of marine life, especially within the Serranidae family.

Supplementary Materials

The following supporting information can be downloaded at: https://www.mdpi.com/article/10.3390/ijms25031822/s1.

Author Contributions

Conceptualization: S.K. and H.-W.K.; methodology: S.R.L. and E.-B.K.; software: S.K., A.R.K. and S.R.L.; validation: S.K. and H.-E.K.; formal analysis: S.K., A.R.K. and S.R.L.; investigation: E.-B.K., M.H.F.A. and S.A.; resources: H.-W.K.; data curation: E.-B.K., M.H.F.A. and S.A.; writing—original draft: S.K., H.-E.K. and A.R.K.; writing—review and editing: S.K., H.-E.K. and H.-W.K.; visualization: S.K., M.H.F.A. and S.A.; supervision: S.K. and H.-W.K.; project administration: H.-W.K. and K.K.; funding acquisition: H.-W.K. and K.K. All authors have read and agreed to the published version of the manuscript.

Funding

This research was supported by the Basic Science Research Program through the National Research Foundation of Korea (NRF) funded by the Ministry of Education (2021R1A6A1A03039211).

Institutional Review Board Statement

Not applicable.

Informed Consent Statement

Not applicable.

Data Availability Statement

The genome sequence data that support the findings of this study are openly available in GenBank of NCBI at https://www.ncbi.nlm.nih.gov, under the accession no. OQ420715.

Acknowledgments

The authors would like to extend their appreciation to Fantong Zealous Gietbong from the Ministry of Livestock, Fisheries and Animal Industries (MINEPIA), Yaounde, Cameroon, and Piyumi S. De Alwis and Flandrianto Sih Palimirmo from the Department of Marine Biology at Pukyong National University, Busan, for their invaluable assistance in conducting sampling, laboratory experiments, and data analyses.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Anderson, S.; Bankier, A.T.; Barrell, B.G.; de Bruijn, M.H.; Coulson, A.R.; Drouin, J.; Eperon, I.C.; Nierlich, D.P.; Roe, B.A.; Sanger, F.; et al. Sequence and organization of the human mitochondrial genome. Nature 1981, 290, 457–465. [Google Scholar] [CrossRef]
  2. Boore, J.L. Animal mitochondrial genomes. Nucleic Acids Res. 1999, 27, 1767–1780. [Google Scholar] [CrossRef]
  3. Tzeng, C.S.; Hui, C.F.; Shen, S.C.; Huang, P.C. The complete nucleotide sequence of the Crossostoma lacustre mitochondrial genome: Conservation and variations among vertebrates. Nucleic Acids Res. 1992, 20, 4853–4858. [Google Scholar] [CrossRef]
  4. Harrison, R.G. Animal mitochondrial DNA as a genetic marker in population and evolutionary biology. Trends Ecol. Evol. 1989, 4, 6–11. [Google Scholar] [CrossRef]
  5. Miya, M.; Satoh, T.P.; Nishida, M. The phylogenetic position of toadfishes (order Batrachoidiformes) in the higher ray-finned fish as inferred from partitioned Bayesian analysis of 102 whole mitochondrial genome sequences. Biol. J. Linn. Soc. 2005, 85, 289–306. [Google Scholar] [CrossRef]
  6. Satoh, T.P.; Miya, M.; Mabuchi, K.; Nishida, M. Structure and variation of the mitochondrial genome of fishes. BMC Genom. 2016, 17, 719. [Google Scholar] [CrossRef] [PubMed]
  7. Heemstra, P.C.; Randall, J.E. FAO Species Catalogue. Groupers of the World (Family Serranidae, Subfamily Epinephelinae). An Annotated and Illustrated Catalogue of the Grouper, Rockcod, Hind, Coral Grouper and Lyretail Species Known to Date. FAO Fish. Synop. 1993, 125, 1–382. [Google Scholar]
  8. Fricke, R.; Eschmeyer, W.N.; Van der Laan, R. (Eds.) Eschmeyer’s Catalog of Fishes: Genera, Species. Available online: http://researcharchive.calacademy.org/research/ichthyology/catalog/fishcatmain.asp (accessed on 12 December 2023).
  9. Craig, M.T.; de Mitcheson, Y.S.; Heemstra, P.C. Groupers of the World: A Field and Market Guide; National Inquiry Services Centre: Grahamstown, South Africa, 2011; pp. 1–47. [Google Scholar]
  10. Vella, N.; Vella, A.; Darmanin, S.A. Morphological and genetic analyses of the first record of the Niger Hind, Cephalopholis nigri (Perciformes: Serranidae), in the Mediterranean Sea and of the African Hind, Cephalopholis taeniops, in Malta. Mar. Biodivers. Rec. 2016, 9, e99. [Google Scholar] [CrossRef]
  11. Ben Abdallah, A.; Ben Souissi, J.; Méjri, H.; Capapé, C.; Golani, D. First record of Cephalopholis taeniops (Valenciennes) in Mediterranean Sea. J. Fish Biol. 2007, 71, 610–614. [Google Scholar] [CrossRef]
  12. Salameh, P.; Sonin, O.; Golani, D. A first record of the African hind (Cephalopholis taeniops) (Pisces: Serranidae) in the Levant. Ann. Ser. Hist. Nat. 2009, 190, 151–154. [Google Scholar]
  13. Guidetti, P.; Giardina, F.; Azzurro, E. A new record of Cephalopholis taeniops in the Mediterranean Sea, with considerations on the Sicily channel as a biogeographical crossroad of exotic fish. Mar. Biodivers. Rec. 2010, 3, e13. [Google Scholar] [CrossRef]
  14. Brito, A.; Clemente, S.; Herrera, R. On the occurrence of the African hind, Cephalopholis taeniops, in the Canary Islands (eastern subtropical Atlantic): Introduction of large-sized demersal littoral fishes in ballast water of oil platforms? Biol. Invasions 2011, 13, 2185–2189. [Google Scholar] [CrossRef]
  15. Engin, S.; Irmak, E.; Seyhan, D. New record of the thermophilic Cephalopholis taeniops (Osteichthyes: Serranidae) in the Aegean Sea. Zool. Middle East. 2016, 62, 184–186. [Google Scholar] [CrossRef]
  16. Canes Garcia, L.; Rangel Moreira, C.; Carvalho Filho, A. First record of African Hind, Cephalopholis taeniops (Valenciennes, 1828) (Perciformes, Epinephelidae) in the South-western Atlantic. Check List 2018, 14, 961–965. [Google Scholar] [CrossRef]
  17. Luiz, O.J.; Woods, R.M.; Madin, E.M.; Madin, J.S. Predicting IUCN extinction risk categories for the world’s data deficient groupers (Teleostei: Epinephelidae). Conserv. Lett. 2016, 9, 342–350. [Google Scholar] [CrossRef]
  18. Burton, M.L.; Potts, J.C.; Carr, D.R. Age, growth and natural mortality of coney (Cephalopholis fulva) from the southeastern United States. PeerJ 2015, 3, e825. [Google Scholar] [CrossRef] [PubMed]
  19. Schemmel, E.M.; Donovan, M.K.; Wiggins, C.; Anzivino, M.; Friedlander, A.M. Reproductive life history of the introduced peacock grouper Cephalopholis argus in Hawaii. J. Fish Biol. 2016, 89, 1271–1284. [Google Scholar] [CrossRef] [PubMed]
  20. Gaither, M.R.; Bowen, B.W.; Bordenave, T.R.; Rocha, L.A.; Newman, S.J.; Gomez, J.A.; van Herwerden, L.; Craig, M.T. Phylogeography of the reef fish Cephalopholis argus (Epinephelidae) indicates Pleistocene isolation across the Indo-Pacific Barrier with contemporary overlap in The Coral Triangle. BMC Evol. Biol. 2011, 11, 189. [Google Scholar] [CrossRef]
  21. De Souza, A.S.; Dias Júnior, E.A.; Galetti, P.M., Jr.; Machado, E.G.; Pichorim, M.; Molina, W.F. Wide-range genetic connectivity of Coney, Cephalopholis fulva (Epinephelidae), through oceanic islands and continental Brazilian coast. An. Acad. Bras. Cienc. 2015, 87, 121–136. [Google Scholar] [CrossRef]
  22. Xie, Z.; Wang, D.; Jiang, S.; Peng, C.; Wang, Q.; Huang, C.; Li, S.; Lin, H.; Zhang, Y. Chromosome-Level Genome Assembly and Transcriptome Comparison Analysis of Cephalopholis sonnerati and Its Related Grouper Species. Biology 2022, 11, 1053. [Google Scholar] [CrossRef]
  23. Pondella, D.J., 2nd; Craig, M.T.; Franck, J.P. The phylogeny of Paralabrax (Perciformes: Serranidae) and allied taxa inferred from partial 16S and 12S mitochondrial ribosomal DNA sequences. Mol. Phylogenet. Evol. 2003, 29, 176–184. [Google Scholar] [CrossRef]
  24. Craig, M.T.; Hastings, P.A. A molecular phylogeny of the groupers of the subfamily Epinephelinae (Serranidae) with a revised classification of the Epinephelini. Ichthyol. Res. 2007, 54, 1–17. [Google Scholar] [CrossRef]
  25. Gill, A.C.; Pogonoski, J.J.; Johnson, J.W.; Tea, Y.K. Three new species of Australian anthiadine fishes, with comments on the monophyly of Pseudanthias Bleeker (Teleostei: Serranidae). Zootaxa 2021, 4996, 49–82. [Google Scholar] [CrossRef] [PubMed]
  26. Williams, J.T.; Viviani, J. Pseudogramma polyacantha complex (Serranidae, tribe Grammistini): DNA barcoding results lead to the discovery of three cryptic species, including two new species from French Polynesia. Zootaxa 2016, 4111, 246–260. [Google Scholar]
  27. Harrison, H.B.; Feldheim, K.A.; Jones, G.P.; Ma, K.; Mansour, H.; Perumal, S.; Williamson, D.H.; Berumen, M.L. Validation of microsatellite multiplexes for parentage analysis and species discrimination in two hybridizing species of coral reef fish (Plectropomus spp., Serranidae). Ecol. Evol. 2014, 4, 2046–2057. [Google Scholar] [CrossRef]
  28. Craig, M.T.; Pondella, D.J.; Franck, J.P.; Hafner, J.C. On the status of the Serranid fish genus Epinephelus: Evidence for paraphyly based upon 16S rDNA sequence. Mol. Phylogenet. Evol. 2001, 19, 121–130. [Google Scholar] [CrossRef]
  29. Ding, S.; Zhuang, X.; Guo, F.; Wang, J.; Su, Y.; Zhang, Q.; Li, Q. Molecular phylogenetic relationships of China Seas groupers based on cytochrome b gene fragment sequences. Sci. China C Life Sci. 2006, 49, 235–242. [Google Scholar] [CrossRef] [PubMed]
  30. Ma, K.Y.; Craig, M.T. An Inconvenient Monophyly: An Update on the Taxonomy of the Groupers (Epinephelidae). Copeia 2018, 106, 443–456. [Google Scholar] [CrossRef]
  31. Stepien, C.A.; Rosenblatt, R.H.; Bargmeyer, B.A. Phylogeography of the spotted sand bass, Paralabrax maculatofasciatus: Divergence of Gulf of California and Pacific Coast populations. Evolution 2001, 55, 1852–1862. [Google Scholar]
  32. Liu, J.X.; Gao, T.X.; Yokogawa, K.; Zhang, Y.P. Differential population structuring and demographic history of two closely related fish species, Japanese sea bass (Lateolabrax japonicus) and spotted sea bass (Lateolabrax maculatus) in Northwestern Pacific. Mol. Phylogenet. Evol. 2006, 39, 799–811. [Google Scholar] [CrossRef]
  33. Gauthier, D.T.; Audemard, C.A.; Carlsson, J.E.; Darden, T.L.; Denson, M.R.; Reece, K.S.; Carlsson, J. Genetic population structure of US Atlantic coastal striped bass (Morone saxatilis). J. Hered. 2013, 104, 510–520. [Google Scholar] [CrossRef]
  34. Buchholz-Sørensen, M.; Vella, A. Population Structure, Genetic Diversity, Effective Population Size, Demographic History and Regional Connectivity Patterns of the Endangered Dusky Grouper, Epinephelus marginatus (Teleostei: Serranidae), within Malta’s Fisheries Management Zone. PLoS ONE 2016, 11, e0159864. [Google Scholar] [CrossRef] [PubMed]
  35. Puebla, O.; Bermingham, E.; Guichard, F. Population genetic analyses of Hypoplectrus coral reef fishes provide evidence that local processes are operating during the early stages of marine adaptive radiations. Mol. Ecol. 2008, 17, 1405–1415. [Google Scholar] [CrossRef] [PubMed]
  36. Bariche, M.; Torres, M.; Smith, C.; Sayar, N.; Azzurro, E.; Baker, R.; Bernardi, G. Red Sea fishes in the Mediterranean Sea: A preliminary investigation of a biological invasion using DNA barcoding. J. Biogeogr. 2015, 42, 2363–2373. [Google Scholar] [CrossRef]
  37. Galal-Khallaf, A.; Osman, A.G.M.; El-Ganainy, A.; Farrag, M.M.; Mohammed-AbdAllah, E.; Moustafa, M.A.; Mohammed-Geba, K. Mitochondrial genetic markers for authentication of major Red Sea grouper species (Perciformes: Serranidae) in Egypt: A tool for enhancing fisheries management and species conservation. Gene 2019, 689, 235–245. [Google Scholar] [CrossRef] [PubMed]
  38. Duranton, M.; Allal, F.; Fraïsse, C.; Bierne, N.; Bonhomme, F.; Gagnaire, P.A. The origin and remolding of genomic islands of differentiation in the European sea bass. Nat. Commun. 2018, 9, 2518. [Google Scholar] [CrossRef]
  39. Sun, C.F.; Zhang, X.H.; Dong, J.J.; You, X.X.; Tian, Y.Y.; Gao, F.Y.; Zhang, H.T.; Shi, Q.; Ye, X. Whole-genome resequencing reveals recent signatures of selection in five populations of largemouth bass (Micropterussalmoides). Zool. Res. 2023, 44, 78–89. [Google Scholar] [CrossRef] [PubMed]
  40. Coulmance, F.; Akkaynak, D.; Le Poul, Y.; Höppner, M.P.; McMillan, W.O.; Puebla, O. Phenotypic and genomic dissection of colour pattern variation in a reef fish radiation. Mol. Ecol. 2023. [Google Scholar] [CrossRef]
  41. Zhuang, X.; Qu, M.; Zhang, X.; Ding, S. A comprehensive description and evolutionary analysis of 22 Grouper (Perciformes, Epinephelidae) mitochondrial genomes with emphasis on two novel genome organizations. PLoS ONE 2013, 8, e73561. [Google Scholar] [CrossRef]
  42. Li, J.L.; Liu, M.; Wang, Y.Y. Complete mitochondrial genome of the chocolate hind Cephalopholis boenak (Pisces: Perciformes). Mitochondrial DNA 2014, 25, 167–168. [Google Scholar] [CrossRef]
  43. Hsiao, S.T.; Chen, K.S.; Tseng, C.T.; Wu, C.L. Complete mitochondrial genome of the sixblotch hind Cephalopholis sexmaculata (Pisces: Perciformes). Mitochondrial DNA Part A DNA Mapp. Seq. Anal. 2016, 27, 1018–1019. [Google Scholar] [CrossRef] [PubMed]
  44. Guo, M.; Huang, H.; Gao, Y. Complete mitochondrial genome of darkfin hind Cephalopholis urodeta (Perciformes, Epinephelidae). Mitochondrial DNA B Resour. 2016, 1, 913–916. [Google Scholar] [CrossRef] [PubMed]
  45. Meng, L.; Zhang, Y.; Gong, L.; Gao, Y. Characterization of the complete mitochondrial genome of Cephalopholis miniata (Perciformes, Serranidae) and its phylogenetic analysis. Mitochondrial DNA B Resour. 2021, 6, 1976–1978. [Google Scholar] [CrossRef] [PubMed]
  46. Wang, C.; Ye, P.; Liu, M.; Zhang, Y.; Feng, H.; Liu, J.; Zhou, H.; Wang, J.; Chen, X. Comparative Analysis of Four Complete Mitochondrial Genomes of Epinephelidae (Perciformes). Genes 2022, 13, 660. [Google Scholar] [CrossRef] [PubMed]
  47. Miya, M.; Kawaguchi, A.; Nishida, M. Mitogenomic exploration of higher teleostean phylogenies: A case study for moderate-scale evolutionary genomics with 38 newly determined complete mitochondrial DNA sequences. Mol. Biol. Evol. 2001, 18, 1993–2009. [Google Scholar] [CrossRef]
  48. Saitoh, K.; Sado, T.; Mayden, R.L.; Hanzawa, N.; Nakamura, K.; Nishida, M.; Miya, M. Mitogenomic evolution and interrelationships of the Cypriniformes (Actinopterygii: Ostariophysi): The first evidence toward resolution of higher-level relationships of the world’s largest freshwater fish clade based on 59 whole mitogenome sequences. J. Mol. Evol. 2006, 63, 826–841. [Google Scholar] [CrossRef]
  49. Yamanoue, Y.; Miya, M.; Matsuura, K.; Yagishita, N.; Mabuchi, K.; Sakai, H.; Katoh, M.; Nishida, M. Phylogenetic position of tetraodontiform fishes within the higher teleosts: Bayesian inferences based on 44 whole mitochondrial genome sequences. Mol. Phylogenet. Evol. 2007, 45, 89–101. [Google Scholar] [CrossRef]
  50. Saccone, C.; Pesole, G.; Sbisá, E. The main regulatory region of mammalian mitochondrial DNA: Structure-function model and evolutionary pattern. J. Mol. Evol. 1991, 33, 83–91. [Google Scholar] [CrossRef]
  51. Bernt, M.; Braband, A.; Schierwater, B.; Stadler, P.F. Genetic aspects of mitochondrial genome evolution. Mol. Phylogenet. Evol. 2013, 69, 328–338. [Google Scholar] [CrossRef]
  52. Sato, Y.; Miya, M.; Fukunaga, T.; Sado, T.; Iwasaki, W. MitoFish and MiFish Pipeline: A Mitochondrial Genome Database of Fish with an Analysis Pipeline for Environmental DNA Metabarcoding. Mol. Biol. Evol. 2018, 35, 1553–1555. [Google Scholar] [CrossRef]
  53. Bowen, B.W.; Rocha, L.A.; Toonen, R.J.; Karl, S.A.; ToBo Laboratory. The origins of tropical marine biodiversity. Trends Ecol. Evol. 2013, 28, 359–366. [Google Scholar] [CrossRef]
  54. Myers, N.; Mittermeier, R.A.; Mittermeier, C.G.; da Fonseca, G.A.; Kent, J. Biodiversity hotspots for conservation priorities. Nature 2000, 403, 853–858. [Google Scholar] [CrossRef]
  55. Miller, E.C.; Hayashi, K.T.; Song, D.; Wiens, J.J. Explaining the ocean’s richest biodiversity hotspot and global patterns of fish diversity. Proc. Biol. Sci. 2018, 285, 20181314. [Google Scholar] [CrossRef]
  56. Fan, H.; Huang, M.; Chen, Y.; Zhou, W.; Hu, Y.; Wei, F. Conservation priorities for global marine biodiversity across multiple dimensions. Natl. Sci. Rev. 2022, 10, nwac241. [Google Scholar] [CrossRef]
  57. Simard, F.; Laffoley, D.; Baxter, J.M. (Eds.) Marine Protected Areas and Climate Change: Adaptation and Mitigation Synergies, Opportunities and Challenges; IUCN: Gland, Switzerland, 2016; 52p. [Google Scholar]
  58. Wang, X.; Wang, Y.; Zhang, Y.; Yu, H.; Tong, J. Evolutionary analysis of cyprinid mitochondrial genomes: Remarkable variation and strong adaptive evolution. Front. Genet. 2016, 7, 156. [Google Scholar]
  59. Kosiol, C.; Vinar, T.; da Fonseca, R.R.; Hubisz, M.J.; Bustamante, C.D.; Nielsen, R.; Siepel, A. Patterns of positive selection in six Mammalian genomes. PLoS Genet. 2008, 4, e1000144. [Google Scholar] [CrossRef] [PubMed]
  60. Foote, A.D.; Morin, P.A.; Durban, J.W.; Pitman, R.L.; Wade, P.; Willerslev, E.; Gilbert, M.T.; da Fonseca, R.R. Positive selection on the killer whale mitogenome. Biol. Lett. 2011, 7, 116–118. [Google Scholar] [CrossRef] [PubMed]
  61. Hirsh, A.E.; Fraser, H.B. Protein dispensability and rate of evolution. Nature 2001, 411, 1046–1049. [Google Scholar] [CrossRef] [PubMed]
  62. Bloom, J.D.; Labthavikul, S.T.; Otey, C.R.; Arnold, F.H. Protein stability promotes evolvability. Proc. Natl. Acad. Sci. USA 2006, 103, 5869–5874. [Google Scholar] [CrossRef] [PubMed]
  63. Montoya-Burgos, J.I. Patterns of Positive Selection and Neutral Evolution in the Protein-Coding Genes of Tetraodon and Takifugu. PLoS ONE 2011, 6, e24800. [Google Scholar] [CrossRef] [PubMed]
  64. Yang, Z.; Bielawski, J.P. Statistical methods for detecting molecular adaptation. Trends Ecol. Evol. 2000, 15, 496–503. [Google Scholar] [CrossRef]
  65. Yang, Z.H.; Nielsen, R. Estimating synonymous and nonsynonymous substitution rates under realistic evolutionary models. Mol. Biol. Evol. 2000, 17, 32–43. [Google Scholar] [CrossRef]
  66. Zhu, K.C.; Liang, Y.Y.; Wu, N.; Guo, H.Y.; Zhang, N.; Jiang, S.G.; Zhang, D.C. Sequencing and characterization of the complete mitochondrial genome of Japanese Swellshark (Cephalloscyllium umbratile). Sci. Rep. 2017, 7, 15299. [Google Scholar] [CrossRef]
  67. Nei, M.; Kumar, S. Molecular Evolution and Phylogenetics; Oxford University Press: New York, NY, USA, 2000. [Google Scholar]
  68. Sato, N.S.; Hirabayashi, N.; Agmon, I.; Yonath, A.; Suzuki, T. Comprehensive genetic selection revealed essential bases in the peptidyl-transferase center. Proc. Natl. Acad. Sci. USA 2006, 103, 15386–15391. [Google Scholar] [CrossRef]
  69. Ojala, D.; Montoya, J.; Attardi, G. tRNA punctuation model of RNA processing in human mitochondria. Nature 1981, 290, 470–474. [Google Scholar] [CrossRef]
  70. Varani, G.; McClain, W.H. The G-U wobble base pair: A fundamental building block of RNA structure crucial to RNA function in diverse biological systems. EMBO Rep. 2000, 1, 18–23. [Google Scholar] [CrossRef]
  71. Cantatore, P.; Gadaleta, M.N.; Roberti, M.; Saccone, C.; Wilson, A.C. Duplication and remoulding of tRNA genes during the evolutionary rearrangement of mitochondrial genomes. Nature 1987, 329, 853–855. [Google Scholar] [CrossRef] [PubMed]
  72. Ponce, M.; Infante, C.; Jiménez-Cantizano, R.M.; Pérez, L.; Manchado, M. Complete mitochondrial genome of the blackspot seabream, Pagellus bogaraveo (Perciformes: Sparidae), with high levels of length heteroplasmy in the WANCY region. Gene 2008, 409, 44–52. [Google Scholar] [CrossRef] [PubMed]
  73. Lee, W.J.; Conroy, J.; Howell, W.H.; Kocher, T.D. Structure and evolution of teleost mitochondrial control regions. J. Mol. Evol. 1995, 41, 54–66. [Google Scholar] [CrossRef] [PubMed]
  74. San Mauro, D.; Gower, D.J.; Zardoya, R.; Wilkinson, M. A hotspot of gene order rearrangement by tandem duplication and random loss in the vertebrate mitochondrial genome. Mol. Biol. Evol. 2006, 23, 227–234. [Google Scholar] [CrossRef] [PubMed]
  75. Springer, M.S.; Douzery, E. Secondary structure and patterns of evolution among mammalian mitochondrial 12S rRNA molecules. J. Mol. Evol. 1996, 43, 357–373. [Google Scholar] [CrossRef] [PubMed]
  76. Wang, H.Y.; Lee, S.C. Secondary structure of mitochondrial 12S rRNA among fish and its phylogenetic applications. Mol. Biol. Evol. 2002, 19, 138–148. [Google Scholar] [CrossRef] [PubMed]
  77. Li, J.; Wang, X.; Kong, X.; Zhao, K.; He, S.; Mayden, R.L. Variation patterns of the mitochondrial 16S rRNA gene with secondary structure constraints and their application to phylogeny of cyprinine fishes (Teleostei: Cypriniformes). Mol. Phylogenet. Evol. 2008, 47, 472–487. [Google Scholar] [CrossRef]
  78. Kundu, S.; De Alwis, P.S.; Kim, A.R.; Lee, S.R.; Kang, H.-E.; Go, Y.; Gietbong, F.Z.; Wibowo, A.; Kim, H.-W. Mitogenomic Characterization of Cameroonian Endemic Coptodon camerunensis (Cichliformes: Cichlidae) and Matrilineal Phylogeny of Old-World Cichlids. Genes 2023, 14, 1591. [Google Scholar] [CrossRef]
  79. Kundu, S.; Palimirmo, F.S.; Kang, H.-E.; Kim, A.R.; Lee, S.R.; Gietbong, F.Z.; Song, S.H.; Kim, H.-W. Insights into the Mitochondrial Genetic Makeup and Miocene Colonization of Primitive Flatfishes (Pleuronectiformes: Psettodidae) in the East Atlantic and Indo-West Pacific Ocean. Biology 2023, 12, 1317. [Google Scholar] [CrossRef]
  80. Tine, M.; Kuhl, H.; Gagnaire, P.A.; Louro, B.; Desmarais, E.; Martins, R.S.; Hecht, J.; Knaust, F.; Belkhir, K.; Klages, S.; et al. European Sea bass genome and its variation provide insights into adaptation to euryhalinity and speciation. Nat. Commun. 2014, 5, 5770. [Google Scholar] [CrossRef]
  81. Cheung, W.W.L.; Frölicher, T.L.; Lam, V.W.Y.; Oyinlola, M.A.; Reygondeau, G.; Sumaila, U.R.; Tai, T.C.; The, L.C.L.; Wabnitz, C.C.C. Marine high temperature extremes amplify the impacts of climate change on fish and fisheries. Sci. Adv. 2021, 7, eabh0895. [Google Scholar] [CrossRef]
  82. Waterhouse, L.; Heppell, S.A.; Pattengill-Semmens, C.V.; McCoy, C.; Bush, P.; Johnson, B.C.; Semmens, B.X. Recovery of critically endangered Nassau grouper (Epinephelus striatus) in the Cayman Islands following targeted conservation actions. Proc. Natl. Acad. Sci. USA 2020, 117, 1587–1595. [Google Scholar] [CrossRef]
  83. Kumar, S.; Stecher, G.; Li, M.; Knyaz, C.; Tamura, K. MEGA X: Molecular Evolutionary Genetics Analysis across computing platforms. Mol. Biol. Evol. 2018, 35, 1547–1549. [Google Scholar] [CrossRef]
  84. Bernt, M.; Donath, A.; Jühling, F.; Externbrink, F.; Florentz, C.; Fritzsch, G.; Pütz, J.; Middendorf, M.; Stadler, P.F. MITOS: Improved de novo Metazoan Mitochondrial Genome Annotation. Mol. Phylogenet. Evol. 2013, 69, 313–319. [Google Scholar] [CrossRef]
  85. Iwasaki, W.; Fukunaga, T.; Isagozawa, R.; Yamada, K.; Maeda, Y.; Satoh, T.P.; Sado, T.; Mabuchi, K.; Takeshima, H.; Miya, M.; et al. MitoFish and MitoAnnotator: A mitochondrial genome database of fish with an accurate and automatic annotation pipeline. Mol. Biol. Evol. 2013, 30, 2531–2540. [Google Scholar] [CrossRef]
  86. Rozas, J.; Ferrer-Mata, A.; Sánchez-DelBarrio, J.C.; Guirao-Rico, S.; Librado, P.; Ramos-Onsins, S.E.; Sánchez-Gracia, A. DnaSP 6: DNA sequence polymorphism analysis of large data sets. Mol. Biol. Evol. 2017, 34, 3299–3302. [Google Scholar] [CrossRef]
  87. Perna, N.T.; Kocher, T.D. Patterns of nucleotide composition at fourfold degenerate sites of animal mitochondrial genomes. J. Mol. Evol. 1995, 41, 353–359. [Google Scholar] [CrossRef]
  88. Xia, X. DAMBE6: New tools for microbial genomics, phylogenetics and molecular evolution. J. Hered. 2017, 108, 431–437. [Google Scholar] [CrossRef]
  89. Laslett, D.; Canbäck, B. ARWEN, a program to detect tRNA genes in metazoan mitochondrial nucleotide sequences. Bioinformatics 2008, 24, 172–175. [Google Scholar] [CrossRef]
  90. Chan, P.P.; Lin, B.Y.; Mak, A.J.; Lowe, T.M. tRNAscan-SE 2.0: Improved detection and functional classification of transfer RNA genes. Nucleic Acids Res. 2021, 49, 9077–9096. [Google Scholar] [CrossRef]
  91. 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]
  92. Shen, M.; Shi, X.; Qu, M.; Chen, J. Complete mitochondrial genome of squaretail coralgrouper Plectropomus areolatus (Perciformes, Epinephelidae). Mitochondrial DNA 2013, 24, 365–367. [Google Scholar] [CrossRef]
  93. Wang, H.; Guo, L.; Ding, S.X. The complete mitochondrial genome of Diploprion bifasciatum (Perciformes, Serranidae). Mitochondrial DNA Part A DNA Mapp. Seq. Anal. 2016, 27, 3137–3138. [Google Scholar] [CrossRef]
  94. Hoban, M.L.; Whitney, J.; Collins, A.G.; Meyer, C.; Murphy, K.R.; Reft, A.J.; Bemis, K.E. Skimming for barcodes: Rapid production of mitochondrial genome and nuclear ribosomal repeat reference markers through shallow shotgun sequencing. PeerJ 2022, 10, e13790. [Google Scholar] [CrossRef]
  95. Vella, A.; Vella, N.; Acosta-Díaz, C. The first complete mitochondrial genomes for Serranus papilionaceus and Serranus scriba, and their phylogenetic position within Serranidae. Mol. Biol. Rep. 2022, 49, 6295–6302. [Google Scholar] [CrossRef]
  96. Vences, M.; Miralles, A.; Brouillet, S.; Ducasse, J.; Fedosov, A.; Kharchev, V.; Kostadinov, I.; Kumari, S.; Patmanidis, S.; Scherz, M.D.; et al. iTaxoTools 0.1: Kickstarting a specimen-based software toolkit for taxonomists. Megataxa 2021, 6, 77–92. [Google Scholar] [CrossRef]
  97. Darriba, D.; Taboada, G.L.; Doallo, R.; Posada, D. JModelTest 2: More models, new heuristics and parallel computing. Nat. Methods 2012, 9, 772. [Google Scholar] [CrossRef]
  98. Miller, M.A.; Schwartz, T.; Pickett, B.E.; He, S.; Klem, E.B.; Scheuermann, R.H.; Passarotti, M.; Kaufman, S.; O’Leary, M.A. A RESTful API for Access to Phylogenetic Tools via the CIPRES Science Gateway. Evol. Bioinform. 2015, 11, 43–48. [Google Scholar] [CrossRef]
  99. Lanfear, R.; Frandsen, P.B.; Wright, A.M.; Senfeld, T.; Calcott, B. PartitionFinder 2: New Methods for Selecting Partitioned Models of Evolution for Molecular and Morphological Phylogenetic Analyses. Mol. Biol. Evol. 2016, 34, 772–773. [Google Scholar] [CrossRef]
  100. Trifinopoulos, J.; Nguyen, L.-T.; von Haeseler, A.; Minh, B.Q. W-IQ-TREE: A fast online phylogenetic tool for maximum likelihood analysis. Nucleic Acids Res. 2016, 44, W232–W235. [Google Scholar] [CrossRef]
  101. Guindon, S.; Dufayard, J.F.; Lefort, V.; Anisimova, M.; Hordijk, W.; Gascuel, O. New Algorithms and Methods to Estimate Maximum-Likelihood Phylogenies: Assessing the Performance of PhyML 3.0. Syst. Biol. 2010, 59, 307–321. [Google Scholar] [CrossRef]
  102. Ronquist, F.; Huelsenbeck, J.P. MrBayes 3: Bayesian phylogenetic inference under mixed models. Bioinformatics 2003, 19, 1572–1574. [Google Scholar] [CrossRef]
  103. Letunic, I.; Bork, P. Interactive Tree of Life (iTOL): An online tool for phylogenetic tree display and annotation. Bioinformatics 2007, 23, 127–128. [Google Scholar] [CrossRef]
Figure 1. Global distribution pattern of African Hind, C. taeniops and other congeners in marine environment. Collection locality of C. taeniops is marked by blue pin. The circular mitochondrial genome of C. taeniops is represented and annotated using the MitoAnnotator online server. Different color arcs highlight the presence of PCGs, rRNAs, tRNAs, and CR. Species photograph was taken by Fantong Zealous Gietbong from the Ministry of Livestock, Fisheries and Animal Industries (MINEPIA), Yaounde, Cameroon. The linearized view of the complete mitochondrial genome organization reveals the duplication of tRNA-Asp (D) and CR in C. argus comparison with other Cephalopholis species.
Figure 1. Global distribution pattern of African Hind, C. taeniops and other congeners in marine environment. Collection locality of C. taeniops is marked by blue pin. The circular mitochondrial genome of C. taeniops is represented and annotated using the MitoAnnotator online server. Different color arcs highlight the presence of PCGs, rRNAs, tRNAs, and CR. Species photograph was taken by Fantong Zealous Gietbong from the Ministry of Livestock, Fisheries and Animal Industries (MINEPIA), Yaounde, Cameroon. The linearized view of the complete mitochondrial genome organization reveals the duplication of tRNA-Asp (D) and CR in C. argus comparison with other Cephalopholis species.
Ijms 25 01822 g001
Figure 2. (A) Gene genetic diversity (Pi) of mitochondrial PCGs furnishing the genetic variations of C. taeniops and other Cephalopholis congeners. (B) Graph exhibits the relationship between transitions (s) and transversions (v) concerning genetic divergence of PCGs combining with all three codon positions, utilizing the Kimura 2-parameter (TN84) distance method with crosses indicating transition events and triangles representing transversion events. The curves delineate the non-saturated trends in the variance of transitions and transversions as genetic distance increases. (C) Box plot showing the pairwise divergence of Ka/Ks ratio for each one of the mitochondrial PCGs. (D) Abundance of codon usage of Cephalopholis species mitochondrial genomes including C. taeniops.
Figure 2. (A) Gene genetic diversity (Pi) of mitochondrial PCGs furnishing the genetic variations of C. taeniops and other Cephalopholis congeners. (B) Graph exhibits the relationship between transitions (s) and transversions (v) concerning genetic divergence of PCGs combining with all three codon positions, utilizing the Kimura 2-parameter (TN84) distance method with crosses indicating transition events and triangles representing transversion events. The curves delineate the non-saturated trends in the variance of transitions and transversions as genetic distance increases. (C) Box plot showing the pairwise divergence of Ka/Ks ratio for each one of the mitochondrial PCGs. (D) Abundance of codon usage of Cephalopholis species mitochondrial genomes including C. taeniops.
Ijms 25 01822 g002
Figure 3. Schematic diagram shows the comparison of length and nucleotide composition of nine different conserved domains of C. taeniops and other six Cephalopholis congeners. The conserved and variable nucleotides are marked in black and blue circles. The gaps are denoted by dash symbol.
Figure 3. Schematic diagram shows the comparison of length and nucleotide composition of nine different conserved domains of C. taeniops and other six Cephalopholis congeners. The conserved and variable nucleotides are marked in black and blue circles. The gaps are denoted by dash symbol.
Ijms 25 01822 g003
Figure 4. The maximum likelihood (ML) phylogeny constructed by 13 concatenated PCGs clearly discriminate C. taeniops from other Cephalopholis congeners. The cladistic pattern also elucidate insights into the evolutionary relationships of different taxonomic levels (tribe and subfamily) of fishes under the family Serranidae. ML bootstrap supports (red color), NJ bootstrap support (blue color), and Bayesian posterior probability values (black color) are indicated at each node, reflecting the statistical support for each branching point in the tree.
Figure 4. The maximum likelihood (ML) phylogeny constructed by 13 concatenated PCGs clearly discriminate C. taeniops from other Cephalopholis congeners. The cladistic pattern also elucidate insights into the evolutionary relationships of different taxonomic levels (tribe and subfamily) of fishes under the family Serranidae. ML bootstrap supports (red color), NJ bootstrap support (blue color), and Bayesian posterior probability values (black color) are indicated at each node, reflecting the statistical support for each branching point in the tree.
Ijms 25 01822 g004
Table 1. List of annotated mitochondrial genes, boundaries, size, and intergenic nucleotides of Cephalopholis taeniops.
Table 1. List of annotated mitochondrial genes, boundaries, size, and intergenic nucleotides of Cephalopholis taeniops.
GenesStartStopStrandSize (bp)Intergenic NucleotideAnticodonStart CodonStop Codon
tRNA-Phe (F)170H700TTC..
12S rRNA711023H9530...
tRNA-Val (V)10241095H720GTA..
16S rRNA10962807H17120...
tRNA-Leu (L2)28082882H750TTA..
ND128833857H9755.ATGTAA
tRNA-Ile (I)38633932H70−1ATC..
tRNA-Gln (Q)39324002L710CAA..
tRNA-Met (M)40034072H700ATG..
ND240735118H10460.ATGTA-
tRNA-Trp (W)51195189H711TGA..
tRNA-Ala (A)51915259L690GCA..
tRNA-Asn (N)52605332L7337AAC..
tRNA-Cys (C)53705437L680TGC..
tRNA-Tyr (Y)54385508L711TAC..
COI55107060H15510.GTGTAA
tRNA-Ser (S2)70617131L711TCA..
tRNA-Asp (D)71337205H738GAC..
COII72147904H6910.ATGT--
tRNA-Lys (K)79057977H731AAA..
ATP879798146H168−10.ATGTAA
ATP681378819H6830.TTGTA-
COIII88209604H7850.ATGTA-
tRNA-Gly (G)96059676H720GGA..
ND3967710,025H3490.ATGT--
tRNA-Arg (R)10,02610,094H690CGA..
ND4L10,09510,391H297−7.ATGTAA
ND410,38511,765H13810.ATGT--
tRNA-His (H)11,76611,835H700CAC..
tRNA-Ser (S1)11,83611,907H726AGC..
tRNA-Leu (L1)11,91411,986H730CTA..
ND511,98713,825H1839−4.ATGTAA
ND613,82214,343L5220.ATGTAA
tRNA-Glu (E)14,34414,412L694GAA..
Cyt b14,41715,557H11410.ATGT--
tRNA-Thr (T)15,55815,630H73−1ACA..
tRNA-Pro (P)15,63015,699L700CCA..
Control region (CR)15,70016,572H873....
Disclaimer/Publisher’s Note: The statements, opinions and data contained in all publications are solely those of the individual author(s) and contributor(s) and not of MDPI and/or the editor(s). MDPI and/or the editor(s) disclaim responsibility for any injury to people or property resulting from any ideas, methods, instructions or products referred to in the content.

Share and Cite

MDPI and ACS Style

Kundu, S.; Kang, H.-E.; Kim, A.R.; Lee, S.R.; Kim, E.-B.; Amin, M.H.F.; Andriyono, S.; Kim, H.-W.; Kang, K. Mitogenomic Characterization and Phylogenetic Placement of African Hind, Cephalopholis taeniops: Shedding Light on the Evolution of Groupers (Serranidae: Epinephelinae). Int. J. Mol. Sci. 2024, 25, 1822. https://doi.org/10.3390/ijms25031822

AMA Style

Kundu S, Kang H-E, Kim AR, Lee SR, Kim E-B, Amin MHF, Andriyono S, Kim H-W, Kang K. Mitogenomic Characterization and Phylogenetic Placement of African Hind, Cephalopholis taeniops: Shedding Light on the Evolution of Groupers (Serranidae: Epinephelinae). International Journal of Molecular Sciences. 2024; 25(3):1822. https://doi.org/10.3390/ijms25031822

Chicago/Turabian Style

Kundu, Shantanu, Hye-Eun Kang, Ah Ran Kim, Soo Rin Lee, Eun-Bi Kim, Muhammad Hilman Fu’adil Amin, Sapto Andriyono, Hyun-Woo Kim, and Kyoungmi Kang. 2024. "Mitogenomic Characterization and Phylogenetic Placement of African Hind, Cephalopholis taeniops: Shedding Light on the Evolution of Groupers (Serranidae: Epinephelinae)" International Journal of Molecular Sciences 25, no. 3: 1822. https://doi.org/10.3390/ijms25031822

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