Next Article in Journal
Evaluation of Anaesthetic Effect of Commercial Basil Ocimum basilicum on Zebrafish (Danio rerio) Embryos
Previous Article in Journal
Effects of Different Oxytocin and Temperature on Reproductive Activity in Nile tilapia (Oreochromis niloticus): Based on Sex Steroid Hormone and GtHR Gene Expression
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

The Complete Mitochondrial Genome of Hyotissa hyotis (Bivalvia: Gryphaeidae) Reveals a Unique Gene Order within Ostreoidea

1
State Key Laboratory of Marine Resource Utilization in South China Sea, Hainan University, Haikou 570228, China
2
College of Marine Science, Hainan University, Haikou 570228, China
3
Key Laboratory of Mariculture, Ministry of Education, Ocean University of China, Qingdao 266003, China
*
Authors to whom correspondence should be addressed.
Fishes 2022, 7(6), 317; https://doi.org/10.3390/fishes7060317
Submission received: 8 October 2022 / Revised: 28 October 2022 / Accepted: 29 October 2022 / Published: 31 October 2022
(This article belongs to the Section Genetics and Biotechnology)

Abstract

:
The mitochondrial (mt) genome is an important tool when studying the evolution of metazoan animals. The oyster family Gryphaeidae, together with Ostreidae, is one of the two extant taxa of superfamily Ostreoidea. Up until now, the available mitochondrial genomes of oysters were all limited to family Ostreidae. In the present study, the first complete mtDNA of family Gryphaeidae represented by Hyotissa hyotis was sequenced and compared with other available ostreoid mtDNA. The mtDNA of H. hyotis is 22,185 bp in length, encoding 13 protein-coding-genes (PCGs), two ribosomal RNA (rRNA) and 23 transfer RNA (tRNA) genes. Within all the intergenic regions that range from 2 to 1528 bp, two large non-coding regions were identified. The first large non-coding region, located between Cox1 and trnA, contains 1528 nucleotides, while the second one is 1191 bp in length and positioned between Cytb and Nad2. The nucleotide composition of the whole mtDNA is A + T biased, accounting for 59.2%, with a negative AT skew value of −0.20 and a positive GC skew value of 0.33. In contrast to the mtDNA of Ostreidae, neither the split of rrnL nor rrnS was detected in that of H. hyotis. The duplication of trnW of H. hyotis was also discovered for the first time within Ostreoidea. The gene order of H. hyotis is quite different from those of ostreids, indicating extensive rearrangements within superfamily Ostreoidea. The reconstructed phylogeny supported H. hyotis as sister to Ostreidae, with the latter clade formed by Ostrea + (Saccostrea + Crassostrea). This study could provide important information for further understanding the mitochondrial evolution of oysters.

1. Introduction

The mitochondrial (mt) genome within multicellular animals was characteristic of a closed-circular molecule that contains 37 genes, including 13 protein coding genes (PCGs), two ribosomal RNA (tRNA) genes, and the 22 transfer RNA (tRNA) genes [1].With the increase of mtDNA data collected in the last two decades, considerable diversity in the organization of animal mtDNA has been detected especially in nonbilaterian groups, including phyla Cnidaria, Ctenophora, Placozoa, and Porifera [2]. Within these groups, the differences that challenged conventional defined mtDNA characteristics mainly lie in the numbers of both linear and circular chromosomes, the presence of additional PCGs (e.g., Atp9 and TatC), large variation in the number of tRNAs (0 to 25), changes in genetic codes, presence of introns, tRNA and mRNA editing, translational frameshifting, and variable rates of sequence evolution [2,3,4,5]. The size of most bilaterian mtDNA ranges from 15 to 20 kb [1]. Even though some larger mtDNA have been reported, the differences in mitogenome size were attributed to variations in length of certain portions (usually non-coding regions) instead of changes in gene content [6,7]. The sequences and structures of animal mtDNA have been widely used for species delimitation [8,9], classification [10], and phylogenetic analyses [11,12,13] due to its advantages including high abundance, lack of recombination, a higher rate of evolution and maternal inheritance [14]. However, the following studies have indicated that molluscan mitochondrial genomes broke the rules [15] in terms of the discovery of doubly uniparental inheritance (DUI) and homologous recombination within some bivalve taxa [16,17,18,19]. In addition, mitochondrial heteroplasmy that has no correlation with DUI has also been reported in Mytilidae [20].
Compared with other metazoans, mollusks in general, and bivalves in particular, are more prone to gene rearrangements [10,21]. Some bivalves also show variations in the number of mitochondrial tRNA genes [22] and genome size, ranging from 14,622 bp (Lanternula elliptica [23]) to 56,170 bp (Scapharca kagoshimensis [6]) according to our current knowledge. Even within some closely related taxa, considerable differences in size of mtDNA can still be discovered [6,24]. A previous study indicated that size variation was mostly caused by the different lengths of the non-coding regions [25].
Oysters are marine bivalves distributed worldwide, and many of them are important fishery and aquaculture species [26]. Belonging to superfamily Ostreoidea, the extant oysters were divided into two families, Gryphaeidae and Ostreidae [27]. Since the shell morphology of oysters was proved to be highly plastic, molecular data have been applied for the revision of oyster classifications and understanding of their evolution [28]. With the development of sequencing technologies, more and more complete mtDNA of oysters are accessible (Table 1), which provides new insights into oyster phylogenetics [25,26,29,30,31]. Previous comparative mitogenomic analyses of oysters were limited to family Ostreidae, and suggested several shared characteristics, such as a split of the rrnL gene and the duplication of trnM. Within Ostreidae, although the duplications and rearrangements of tRNA genes are frequently observed, the order of PCGs is relatively conserved within the same genus in Crassostrea, Ostrea and Saccostrea. While the PCG rearrangements between different genera could be explained as the result of transversion and/or translocation of the plesiomorphic gene order belonging to Ostrea [26]. In addition to family Ostreidae, there is still no information with respect to the mtDNA of other ostreoid species, especially within Gryphaeidae.
As the representative of family Gryphaeidae, Hyotissa hyotis (Linnaeus, 1758) is distributed in the tropical regions of the Indo-Pacific Ocean, and it is an important fishery resource in coastal countries like China. Although H. hyotis has also been found in the Florida Keys of the United States, it was proven to be a recent introduction from the Indo-Pacific [41]. In the present study, the complete mtDNA of Hyotissa hyotis was sequenced and compared with those of other ostreoids available on GenBank (Table 1), with the following aims: (1) to characterize the mitogenomic features of H. hyotis; (2) to explore its gene order; and (3) to determine its phylogenetic position within Ostreoidea.

2. Materials and Methods

2.1. Samples and DNA Extraction

The specimen of H. hyotis was collected in a local market of Ximaozhou Island (18°14′22″ N; 109°22′42″ E). The soft tissue of the specimen was deposited in 95% alcohol in the Laboratory of Economic Shellfish Genetic Breeding and Culture Technology (LESGBCT), Hainan University.
Genomic DNA was extracted from small pieces of adductor muscle using TIANamp Marine Animals DNA Kit (Tiangen, Beijing, China) following the instructions. Only one specimen of H. hyotis was used for DNA extraction. The genomic DNA was visualized on 1% agarose gel for quality inspection.

2.2. Illumina Sequencing and Mitogenome Assembly

Genomic DNA of H. hyotis was sent to Novogene Company (Beijing, China) for library construction and next-generation sequencing. The DNA library was generated using NEB Next® Ultra™ DNA Library Prep Kit for Illumina (NEB, USA) according to the manufacturer’s instructions. One library with insert size of approximately 300 bp was prepared and then sequenced on the Illumina NovaSeq 6000 platform with 150 bp paired-end reads. Finally, 37,526,636 clean reads of each direction were generated for the library. Adapters and low-quality reads were filtered using Trimmomatic [42]. The generated clean data were imported in Geneious Prime 2021.0.1 [43] for mitogenome assembly, with the strategy following Irwin et al. [44]. The filtered Illumina reads and assembled mitogenome were submitted to Sequence Read Archive (SRA) and GenBank, with the accession numbers SRR21135036 (under Bioproject PRJNA871102) and OP151093, respectively.

2.3. Mitogenome Annotation and Sequence Analysis

The mitogenome of H. hyotis was annotated using MITOS Webserver [45]. The PCGs were also determined by ORF Finder (http://www.ncbi.nlm.nih.gov/orffinder, accessed on 7 October 2022) with the invertebrate mitochondrial genetic code, and their boundaries were modified by comparing them with those of other ostreoids. The identification of Atp8 was done by the comparison of the possible open reading frames predicted by ORF Finder with the Atp8 genes of Ostrea lurida (KC768038) and Crassostrea gasar (KR856227) that were available on GenBank. The secondary structure of tRNA and rRNA genes were predicted by MITOS Webserver. The tRNAs were also verified using ARWEN [46].
The nucleotide composition of the mitochondrial genome, PCGs, rRNA, tRNA genes and major non-coding region were calculated using MEGA X [47]. The base skew values for a given strand were calculated according to the following formulas: AT skew = (A − T)/(A + T) and GC skew = (G − C)/(G + C) [48], where A, T, G, and C are the occurrences of the four nucleotides. Codon usage of PCGs was also conducted by MEGA X. The mitochondrial genome map was generated using CGView [49].

2.4. Gene Order Comparisons

The mtDNA of H. hyotis, along with those of other ostreoids available on GenBank (Table 1), were used for comparative gene order analysis. Mitochondrial gene orders of all mtDNA within Ostreoidea were conducted manually. The rearrangements were analyzed directly by comparing only the 12 PCG rearrangement steps between the common gene order of each genus and visualized using Microsoft Visio 2016. Atp8 gene was not included in PCG rearrangements following Guo et al. (2018) [26].

2.5. Phylogenetic Analysis

In order to clarify the phylogenetic position of H. hyotis within Ostreoidea, a total of 18 ostreoid species (including 17 previously published mitogenomes) were used in the present study (Table 1). Pinctata maxima and P. margaritifera were selected as outgroups according to the topology by Sun et al. (2021) [7]. Phylogenetic trees were reconstructed using maximum likelihood (ML) and Bayesian inference (BI) analyses based on the nucleotide sequences of 12 PCGs. The Atp8 gene was not included since it is missing in most oyster mitogenomes, and due to its high substitution rate, it is not appropriate for phylogenetic construction. Maximum likelihood analyses were preformed using IQtree 1.6.10 [50], allowing partitions to have different evolutionary rates (-spp option) and with 10,000 ultrafast bootstrap pseudoreplications (-bb option). Bayesian inference analyses were carried out by MrBayes 3.2.6 [51], running four simultaneous Monte Carlo Markov chains (MCMC) for 10,000,000 generations, sampling every 1000 generations. Two independent Bayesian inference runs were performed to increase the chance of adequate mixing of the Markov chains and to increase the chance of detecting failure to converge, as determined using Tracer v1.6. The initial 25% of generations were discarded as burn-in according to the excellent effective sample size (ESS) values (>200) of all parameters as measured by Tracer v1.6.
The best partition schemes and best-fit substitution models for BI analyses were conducted using PartitionFinder 2 [52], under the Bayesian Information Criterion (BIC). For the nucleotide sequences of 12 PCGs dataset, the partitions tested were: all genes combined; all genes separated (except Nad4-Nad4L); and genes grouped by subunits (Atp, Cytb, Cox and Nad). Additionally, these three partition schemes were tested considering separately the three codon positions. The best partition schemes for ML analyses were also determined by PartitionFinder 2, while the best-fit substitution models for ML analyses were calculated with ModelFinder [53] as implemented in IQ-TREE 1.6.12.

3. Results and Discussion

3.1. Genome Organization of Hyotissa Hyotis

The assembled mtDNA includes 343,733 short reads, which account for 0.45% of total data.The mean coverage reaches 2302. The complete mtDNA of H. hyotis is 22,185 bp in length, containing 13 protein coding genes (PCGs), two rRNA and 23 tRNA genes (Figure 1, Table 2). This size is longer than most ostreoid mtDNA, but still falling within the size range of Ostreoidea, from 16,277 bp for O. denselamellosa [31] to 22,446 bp for C. iredalei [36]. All 38 genes are encoded on the major strand, like all marine bivalves [15]. Compared with the typical metazoan mtDNA, a duplication of trnW has been detected. The non-coding regions are distributed throughout the mtDNA, ranging from 2 to 1528 bp (Table 2), accounting for 28.9% of the whole mtDNA. Among these intergenic regions, two large non-coding regions (LNR) were identified. The first one (LNR1), which contains 1528 nucleotides, is located between Cox1 and trnA, while the second one (LNR2) is 1191 bp in length and positioned between Cytb and Nad2. The AT skew value of LNR1 (64.8%) is significantly higher than that of LNR2 (51.7%), as well as the remainder of the mtDNA (Table 3). The LNR with higher AT content values is typically used for the determination of Ostreidae mitochondrial control region and considered to contain the initiation of replication and transcription [29]. The presence of two LNRs is not common in the reported Ostreidae mtDNA, and has only been found in that of C. iredalei, within which the two regions were presumed to originate from the duplications of Nad2, followed by separate mutations and functional gene loss [36].
The molecule has an overall AT content of 59.2% (Table 3), which is lower than those of Ostreidae species (60.7% in O. denselamellosa to 65.3% in C. hongkongensis [29]). The nucleotide composition is strongly skewed away from C in favor of G (with the GC skew value of 0.33) and from A in favor of T (with the AT skew value of −0.20) (Table 3). Although the negative AT skew and positive GC skew has been detected in all Ostreoidea species, the absolute values of the two parameters in H. hyotis are relatively higher than those of other ostreoids [36].

3.2. Protein Coding Genes (PCGs)

A total of 13 PCGs were annotated in the mtDNA of H. hyotis (Figure 1, Table 2). The detection of Atp8 is quite difficult since it is a short gene that is under low selective pressure and exhibits great variability at both amino acid and nucleotide levels [54]. Previous studies indicated that Atp8 was missing in all oyster mtDNA [25,29,36,55], however, following studies suggested the identification of this short gene in some oyster species [54]. In this study, a potential Atp8 which is located between Nad1 and trnP was detected based on the slight similarity between Atp8 sequences here and those in the mtDNA of Osrea lurida (KC768038) and C. gasar (KR856227). The Atp8 of H. hyotis encodes 41 amino acids, similar to those of other oysters [54]. It is worth noting that the stop codon of Atp8 is missing. This may be caused by the trnP that breaks the transcript and makes the gene similar in length to other Atp8 genes in this superfamily.
The AT content, AT and GC skew values of the PCGs as well as of the three positions of PCGs, show the same tendency of asymmetry as the mitochondrial genome (Table 3). The 13 PCGs are 12,039 bp in length, accounting for 54.3% of whole mtDNA.
Among the 13 PCGs, seven genes start with the conventional initiation codons ATG (Cox1, Nad1, Cytb, Nad6, Atp6 and Atp8) and ATA (Cox3), while the remaining genes employ alternative start codons, including TTT (Nad3), TTG (Cox2), ATT (Nad2), AAG (Nad5), GTG (Nad4) and TAT (Nad4L) (Table 2). Eight genes stop with TAA (Atp6, Cox2, Cox3, Nad4, Nad4L, Nad5 and Nad6) and two use TAG (Cox1 and Nad2). Finally, Nad1 and Nad3 exhibit the incomplete stop codons TA and T, respectively. The truncated stop codons are common in the metazoan mtDNA [35,56], and might be modified to the TAA termini via post-transcriptional polyadenylation [57]. The relative synonymous codon usage (RSCU) values are shown in Table 4. The mtDNA encodes 3961 amino acids (excluding all stop codons), among which leucine is the most frequently chosen one (Table 4) as indicated in other mollusk taxa [58]. Considerable synonymous codon usage bias was observed in the present study (Figure 2), and these preferred codons mostly ended in T, resulting in a significant negative AT skew value at the third codon position (Table 3). Previous studies have indicated that the codon usage pattern might vary even within the same family [59].
It is also worth mentioning that the boundaries of some PCGs predicted by ORF Finder are quite long. For example, the predicted length of Nad3 in H. hyotis mtDNA reached 1000 bp, far beyond the normal length 351 bp within the superfamily. On the other hand, if the indicated boundary of Nad1 was accepted, it would overlap the Atp8 gene. Therefore, some genes were truncated to similar length according to the corresponding PCGs in this superfamily. This operation also resulted in the appearance of incomplete stop codons mentioned above. These codons are explained to be completed via post-transcriptional polyadenylation, however, no stable secondary structures that could break translation were detected after these genes. The NGS RNAseq reads are needed to finally determine the accurate boundaries of these genes in the future.

3.3. Ribosomal RNA and Transfer RNA Genes

The split of rRNAs has been considered as a common feature of Ostreidae. Firstly, the split of rrnL was found in all ostreid mtDNA [25,29,36]. Furthermore, Asian species within Crassostrea also contained a duplicated rrnS gene [25,36]. In this study, however, neither rrnL nor rrnS of H. hyotis is separated. The rrnS and rrnL genes of H. hyotis are 958 and 1379 bp, with AT content of 53.7% and 58.8% respectively. The rrnL shows a weakly negative AT skew and strongly positive GC skew compared with the rrnS that shows both positive AT and GC skew values (Table 3).
The typical metazoan mtDNA encodes 22 tRNAs, including two copies of trnL and two of trnS [1]. Nevertheless, some bivalve mtDNA have been found to show exceptions which were related to the duplication of certain tRNAs. A duplicated trnM has been detected in all Ostreidae species, as well as in other bivalve taxa, such as some species within Veneridae [60]. In addition to trnM, most Asian oysters of Crassostrea also possess two more duplications of trnK and trnQ [29]. A total of 23 tRNAs was found within the mtDNA of H. hyotis (Figure 1), but the duplication of trnW has never been reported within Ostreoidea previously. The AT and GC skew values of all tRNAs are −0.12 and 0.25 (Table 3), respectively, showing the same tendency of asymmetry as the whole mtDNA. The lengths of the tRNAs range from 63 to 91 bp (Table 2). The trnM which was predicted by different software is significantly longer than other tRNAs of H. hyotis (Figure 3). This strange secondary structure has never been found in oyster mtDNA, but reported in the trnM of the bivalve family Arcidae [61]. All tRNA genes could be folded into typical clover-leaf secondary structures except for the trnS-AGU due to the absence of dihydrouracil (DHU) arm (Figure 3), which was common in metazoan mtDNA [62].

3.4. Gene Rearrangement

Within Bivalvia, gene rearrangements are most common for tRNAs. Even within Crassostrea, variable gene orders of tRNAs could be detected [25]. The change of tRNA gene orders might be explained by the ‘duplication-random loss model’ [63], since duplicated tRNAs could often be found within ostreid mtDNA in addition to the 22 essential tRNAs. The tRNA gene order of H. hyotis shows extensive rearrangements within Ostreoidea, with little gene blocks shared with ostreids (Figure 4).
Compared with tRNAs, PCG orders are more conserved at genus level within Ostreidae. The PCG rearrangements among Crassostrea, Ostrea, and Saccostrea was explained by assuming the PCG order in Ostrea as the ancestral characteristic of Ostreidae, and those of Crassostrea and Saccostrea resulted from one transversion and one transversion plus one translocation, respectively [26]. However, the PCG order of H. hyotis shows little shared gene blocks with the pleisomorphic gene order of Ostrea (Figure 5). Instead, one identical gene block (Nad5-Nad6-Nad4-Atp6) as well as one inverted gene block (Nad1-Nad3-Cox2-Cytb) was detected between H. hyotis and Saccostrea (Figure 5). To understand how the gene order of H. hyotis evolved within superfamily Ostreoidea, more data with respect to the ostreoid mtDNA together with a robust phylogenetic framework are still needed.
Bivalves are sometimes notoriously tricky to amplify PCR products from especially typical barcoding genes often used in species identification. The potential reasons may lie in, such as the gene order rearranging and differing start codons as mentioned above. The first mtDNA of family Gryphaeidae could also help provide those working with oysters a tool for better designing relevant primers.

3.5. Phylogenetic Relationships

The identification of H. hyotis was based on morphological and molecular data. The rrnL fragment in this study is 100% matched to another individual also distributed in the South China Sea (GenBank Accession No. MT332228). However, the genetic distance of Cox1 fragment between specimens from Ximaozhou Island (this study) and Madagascar (GQ166583)/Singapore (OM946454) reaches 18%, which indicates that the H. hyotis on GenBank is comprised of at least two different species. In the future, studies including more population genetic data of Hyotissa are necessary to distinguish the existence of cryptic diversity within H. hyotis or incorrect species identification. It is also necessary to include more complete mtDNA of different species belonging to Gryphaeidae, in order to determine whether the unusual mitogenomic characteristics of H. hyotis mentioned above is species-specific or common within the whole family.
Phylogenetic relationships of Ostreoidea were reconstructed based on the concatenated nucleotide sequences of the 12 PCGs (11,118 positions in length) using probabilistic methods (Figure 6). According to the BIC, the best partition scheme was the one combining genes by subunits but analyzing each codon position separately (Table S1). Both ML (−lnL = 134,336.36) and BI (−lnL = 130,276.19 for run 1; −lnL = 130,276.78 for run 2) analyses arrived at identical topologies (Figure 6).
Within Ostreoidea, H. hyotis was recovered as sister to Ostreidae. This result is consistent with traditional classification, which was based on shell morphology and anatomic characteristics, suggesting that living oysters should be divided into two families Ostreidae and Gryphaeidae [64]. Since H. hyotis was the only representative of Gryphaeidae, it is still necessary to include more taxa of the same family (e.g., Neopycnodonte and Pycnodonte) to test the monophyly of Gryphaeidae in the future.
The phylogeny within Ostreidae was represented as Ostreinae + (Saccostreinae + Crassostreinae). In the present study, the three ostreid subfamilies were represented by three single genera (Ostrea, Saccostrea, and Crassostrea respectively). The close affinity of Saccostrea and Crassostrea here (strongly supported by BI and poorly resolved by ML, Figure 6) has been supported by morphological characteristics as they were previously classified within the same subfamily Crassostreinae [65]. However, previous molecular phylogenies mostly based on short gene fragments revealed a paraphyletic relationship between Crassostrea and Saccostrea, which did not form a monophyletic group, and supported the establishment of Saccostreinae for Saccostrea [66,67,68]. Even though the present phylogeny within Ostreidae (Figure 6) is inconsistent with previous results, the inter-relationships between Ostrea, Saccostrea and Crassostrea still remains unresolved, since different datasets or methods could result in different topologies. For example, the COI- and 16S-based NJ trees [67], as well as the ML tree of four combined sequences (16S, COI, ITS2 and 28S fragments [68]), revealed a closer relationship between Crassostrea and Ostrea, whereas Saccostrea and Ostrea were recovered more related in the BI tree of the combined dataset [68]. It is necessary to develop phylogenomic resources, such as the anchored hybrid enrichment probe set that is capable of capturing hundreds of molecular markers from diverse taxa [69], to resolve the controversial phylogeny within Ostreidae.
Within Crassostrea, the Asia-Pacific and Atlantic species formed two monophyletic lineages as previously suggested [26,68,70]. These two geographical clades could be recognized by several other molecular landmarks, such as different number of duplicated mitochondrial tRNA genes and mt gene rearrangements [25], diverse ITS2 rRNA secondary structures [68], as well as various size and shape of the rDNA-bearing chromosome [71]. Based on such molecular and biogeographical evidence, these two lineages were proposed to be assigned to different genera, and a new genus Magallana was proposed to include the Asian species of Crassostrea [68,70]. This proposal, however, has been opposed since genetic evidence is not supposed to be considered as a sole justification to define genus, without being supported by morphological and anatomical data [26,72]. In this study, we insist that Crassostrea is the valid genus name for both Asia-Pacific and Atlantic oysters and treat Magallana as a subgenus under Crassostrea.

4. Conclusions

The complete mitogenome of H. hyotis is 22,185 bp in length, including 13 PCGs, two rRNA and 23 tRNA genes as well as several non-coding regions. The AT content, AT skew and GC skew values are slightly different from those of ostreids. The gene contents, including the single rrnL and rrnS and a duplicated trnW, and gene orders of H. hyotis exhibited extensive differences with ostreid mtDNA. The phylogenetic analysis revealed that H. hyotis was sister to Ostreidae, which was grouped by Ostrea + (Saccostrea + Crassostrea) and needed further determined by phylogenomic data. In addition, this study indicates considerable Cox1 genetic divergence of H. hyotis between the sequence here and the previously published ones, suggesting either the existence of cryptic species or the possibility of incorrect species identification. However, this needs to be further determined with the inclusion of more population genetic data of Hyotissa. To finally understand the gene order evolution of H. hyotis, this study calls for a comprehensive mitogenomic information of superfamily Ostreoidea, together with a robust phylogenetic framework.

Supplementary Materials

The following are available online at https://www.mdpi.com/article/10.3390/fishes7060317/s1, Table S1: Best fit partitions and substitution models.

Author Contributions

Formal analysis, F.L., M.F., S.W., Y.Y. and S.L.; Funding acquisition, A.W. and Y.Y.; Resources, F.L., M.F., S.W., Z.G., C.L. and Y.Y.; Software, F.L., M.F., S.W., Y.Y. and S.L.; Supervision, Z.G., A.W., C.L. and Y.Y.; Writing–original draft, F.L., Y.Y. and S.L.; Writing–review & editing, Y.Y. and S.L. All authors have read and agreed to the published version of the manuscript.

Funding

This research was funded by the National Key Research and Development Program of China (2019YFD0901301), the Key Research and Development Project of Hainan Province (ZDYF2021SHFZ269), and the Hainan Provincial Natural Science Foundation of China (322QN260).

Institutional Review Board Statement

The study was conducted according to the guidelines of the Declaration of Helsinki and approved by The Committee and Laboratory Animal Department of Hainan University.

Data Availability Statement

The mitochondrial genome was deposited at NCBI, with accession number OP151093. The filtered Illumina reads were submitted to NCBI Sequence Read Archive (SRA) and are available online at https://www.ncbi.nlm.nih.gov/bioproject/ PRJNA871102 (accessed on 19 August 2022).

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Boore, J.L. Animal mitochondrial genomes. Nucleic Acids Res. 1999, 27, 1767–1780. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  2. Lavrov, D.V.; Pett, W. Animal mitochondrial DNA as we do not know it: Mt-genome organization and evolution in nonbilaterian lineages. Genome Biol. Evol. 2016, 8, 2896–2913. [Google Scholar] [CrossRef] [PubMed]
  3. Bridge, D.; Cunningham, C.W.; Schierwater, B.; Desalle, R.; Buss, L.W. Class-level relationships in the phylum Cnidaria: Evidence from mitochondrial genome structure. Proc. Natl. Acad. Sci. USA 1992, 89, 8750–8753. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  4. Kayal, E.; Bentlage, B.; Collins, A.G.; Kayal, M.; Pirro, S.; Lavrov, D.V. Evolution of linear mitochondrial genomes in medusozoan cnidarians. Genome Biol. Evol. 2012, 4, 1–12. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  5. Lavrov, D.V.; Pett, W.; Voigt, O.; Wörheide, G.; Forget, L.; Lang, B.F.; Kayal, E. Mitochondrial DNA of Clathrina clathrus (Calcarea, Calcinea): Six linear chromosomes, fragmented rRNAs, tRNA editing, and a novel genetic code. Mol. Biol. Evol. 2013, 30, 865–880. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  6. Kong, L.; Li, Y.; Kocot, K.M.; Yang, Y.; Qi, L.; Li, Q.; Halanych, K.M. Mitogenomics reveals phylogenetic relationships of Arcoida (Mollusca, Bivalvia) and multiple independent expansions and contractions in mitochondrial genome size. Mol. Phylogenet. Evol. 2020, 150, 106857. [Google Scholar] [CrossRef]
  7. Sun, S.; Li, Q.; Kong, L. Relaxation of Selective Constraint on the Ultra-Large Mitochondrial Genomes of Arcidae (Mollusca: Bivalvia). J. Ocean Univ. China 2021, 20, 1157–1166. [Google Scholar] [CrossRef]
  8. Abalde, S.; Tenorio, M.J.; Afonso, C.M.; Uribe, J.E.; Echeverry, A.M.; Zardoya, R. Phylogenetic relationships of cone snails endemic to Cabo Verde based on mitochondrial genomes. BMC Evol. Biol. 2017, 17, 231. [Google Scholar] [CrossRef] [Green Version]
  9. Abalde, S.; Tenorio, M.J.; Afonso, C.M.; Zardoya, R. Mitogenomic phylogeny of cone snails endemic to Senegal. Mol. Phylogenet. Evol. 2017, 112, 79–87. [Google Scholar] [CrossRef]
  10. Yuan, Y.; Li, Q.; Yu, H.; Kong, L. The complete mitochondrial genomes of six heterodont bivalves (Tellinoidea and Solenoidea): Variable gene arrangements and phylogenetic implications. PLoS ONE 2012, 7, e32353. [Google Scholar] [CrossRef]
  11. Boore, J.L.; Brown, W.M. Big trees from little genomes: Mitochondrial gene order as a phylogenetic tool. Curr. Opin. Genet. Dev. 1998, 8, 668–674. [Google Scholar] [CrossRef]
  12. Lee, Y.; Kwak, H.; Shin, J.; Kim, S.C.; Kim, T.; Park, J.K. A mitochondrial genome phylogeny of Mytilidae (Bivalvia: Mytilida). Mol. Phylogenet. Evol. 2019, 139, 106533. [Google Scholar] [CrossRef]
  13. Uribe, J.E.; Irisarri, I.; Templado, J.; Zardoya, R. New patellogastropod mitogenomes help counteracting long-branch attraction in the deep phylogeny of gastropod mollusks. Mol. Phylogenet. Evol. 2019, 133, 12–23. [Google Scholar] [CrossRef]
  14. Brown, W.M.; George, M., Jr.; Wilson, A.C. Rapid evolution of animal mitochondrial DNA. Proc. Natl. Acad. Sci. USA 1979, 76, 1967–1971. [Google Scholar] [CrossRef] [Green Version]
  15. Ghiselli, F.; Gomes-dos-Santos, A.; Adema, C.M.; Lopes-Lima, M.; Sharbrough, J.; Boore, J.L. Molluscan mitochondrial genomes break the rules. Phil. Trans. R. Soc. B 2021, 376, 20200159. [Google Scholar] [CrossRef]
  16. Breton, S.; Beaupre, H.D.; Stewart, D.T.; Hoeh, W.R.; Blier, P.U. The unusual system of doubly uniparental inheritance of mtDNA: Isn’t one enough? Trends Genet. 2007, 23, 465–474. [Google Scholar] [CrossRef]
  17. Burzyński, A.; Zbawicka, M.; Skibinski, D.O.; Wenne, R. Evidence for recombination of mtDNA in the marine mussel Mytilus trossulus from the Baltic. Mol. Biol. Evol. 2003, 20, 388–392. [Google Scholar] [CrossRef] [Green Version]
  18. Ladoukakis, E.D.; Zouros, E. Direct evidence for homologous recombination in mussel (Mytilus galloprovincialis) mitochondrial DNA. Mol. Biol. Evol. 2001, 18, 1168–1175. [Google Scholar] [CrossRef] [Green Version]
  19. Gusman, A.; Lecomte, S.; Stewart, D.T.; Passamonti, M.; Breton, S. Pursuing the quest for better understanding the taxonomic distribution of the system of doubly uniparental inheritance of mtDNA. PeerJ 2016, 4, e2760. [Google Scholar] [CrossRef] [Green Version]
  20. Lubośny, M.; Śmietanka, B.; Arculeo, M.; Burzyński, A. No evidence of DUI in the Mediterranean alien species Brachidontes pharaonis (P. Fisher, 1870) despite mitochondrial heteroplasmy. Sci. Rep. 2022, 12, 8569. [Google Scholar] [CrossRef]
  21. Serb, J.M.; Lydeard, C. Complete mtDNA sequence of the North American freshwater mussel, Lampsilis ornata (Unionidae): An examination of the evolution and phylogenetic utility of mitochondrial genome organization in Bivalvia (Mollusca). Mol. Biol. Evol. 2003, 20, 1854–1866. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  22. Gissi, C.; Iannelli, F.; Pesole, G. Evolution of the mitochondrial genome of Metazoa as exemplified by comparison of congeneric species. Heredity 2008, 101, 301–320. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  23. Park, H.; Ahn, D.H. Complete mitochondrial genome of the antarctic soft-shelled clam, Laternula elliptica (Bivalvia; Laternulidae). Mitochondrial DNA B 2015, 26, 2. [Google Scholar] [CrossRef]
  24. Sun, S.E.; Li, Q.; Kong, L.; Yu, H. Complete mitochondrial genomes of Trisidos kiyoni and Potiarca pilula: Varied mitochondrial genome size and highly rearranged gene order in Arcidae. Sci. Rep. 2016, 6, 33794. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  25. Ren, J.; Liu, X.; Jiang, F.; Guo, X.; Liu, B. Unusual conservation of mitochondrial gene order in Crassostrea oysters: Evidence for recent speciation in Asia. BMC Evol. Biol. 2010, 10, 394. [Google Scholar] [CrossRef] [Green Version]
  26. Guo, X.; Li, C.; Wang, H.; Xu, Z. Diversity and evolution of living oysters. J. Shellfish Res. 2018, 37, 755–771. [Google Scholar] [CrossRef]
  27. Bouchet, P.; Rocroi, J.P.; Bieler, R.; Carter, J.G.; Coan, E.V. Nomenclator of bivalve families with a classification of bivalve families. Malacologia 2010, 52, 1–184. [Google Scholar] [CrossRef]
  28. Bayne, B.L. Biology of Oysters; Academic Press: London, UK, 2017; p. 860. [Google Scholar]
  29. Danic-Tchaleu, G.; Heurtebise, S.; Morga, B.; Lapègue, S. Complete mitochondrial DNA sequence of the European flat oyster Ostrea edulis confirms Ostreidae classification. BMC Res. Notes 2011, 4, 400. [Google Scholar] [CrossRef] [Green Version]
  30. Volatiana, J.A.; Fang, S.; Kinaro, Z.O.; Liu, X. Complete mitochondrial DNA sequences of Saccostrea mordax and Saccostrea cucullata: Genome organization and phylogeny analysis. Mitochondrial DNA A 2016, 27, 3024–3025. [Google Scholar] [CrossRef]
  31. Yu, H.; Kong, L.; Li, Q. Complete mitochondrial genome of Ostrea denselamellosa (Bivalvia, Ostreidae). Mitochondrial DNA A 2016, 27, 711–712. [Google Scholar] [CrossRef]
  32. Liu, S.; Liu, Y.; He, J.; Lin, Z.; Xue, Q. The complete mitochondrial genome of Crassostrea hongkongensis from East China Sea indicates species’ range may extend northward. Mol. Biol. Rep. 2022, 49, 1631–1635. [Google Scholar] [CrossRef]
  33. Gastineau, R.; Nguyễn, Đ.H.; Lemieux, C.; Turmel, M.; Tremblay, R.; Nguyễn, V.D.; Widowati, I.; Witkowski, A.; Mouget, J.L. The complete mitochondrial DNA of the tropical oyster Crassostrea belcheri from the Cần Giò’mangrove in Vietnam. Mitochondrial DNA B 2018, 3, 462–463. [Google Scholar] [CrossRef] [Green Version]
  34. Cavaleiro, N.P.; Solé-Cava, A.M.; Melo, C.M.; de Almeida, L.G.; Lazoski, C.; Vasconcelos, A.T.R. The complete mitochondrial genome of Crassostrea gasar (Bivalvia: Ostreidae). Mitochondrial DNA A 2016, 27, 2939–2940. [Google Scholar] [CrossRef]
  35. Yu, H.; Li, Q. Complete mitochondrial DNA sequence of Crassostrea nippona: Comparative and phylogenomic studies on seven commercial Crassostrea species. Mol. Biol. Rep. 2012, 39, 999–1009. [Google Scholar] [CrossRef]
  36. Wu, X.; Xu, X.; Yu, Z.; Wei, Z.; Xia, J. Comparison of seven Crassostrea mitogenomes and phylogenetic analyses. Mol. Phylogenet. Evol. 2010, 57, 448–454. [Google Scholar] [CrossRef]
  37. Milbury, C.A.; Gaffney, P.M. Complete mitochondrial DNA sequence of the eastern oyster Crassostrea virginica. Mar. Biotechnol. 2005, 7, 697–712. [Google Scholar] [CrossRef]
  38. Yu, H.; Li, Q. Mutation and selection on the wobble nucleotide in tRNA anticodons in marine bivalve mitochondrial genomes. PLoS ONE 2011, 6, e16147. [Google Scholar] [CrossRef] [Green Version]
  39. Xiao, S.; Wu, X.; Li, L.; Yu, Z. Complete mitochondrial genome of the Olympia oyster Ostrea lurida (Bivalvia, Ostreidae). Mitochondrial DNA 2015, 26, 471–472. [Google Scholar] [CrossRef]
  40. Wu, X.; Li, X.; Li, L.; Yu, Z. A unique tRNA gene family and a novel, highly expressed ORF in the mitochondrial genome of the silver-lip pearl oyster, Pinctada maxima (Bivalvia: Pteriidae). Gene 2012, 510, 22–31. [Google Scholar] [CrossRef]
  41. Bieler, R.; Mikkelsen, P.M.; Lee, T.; Foighil, D.O’. Discovery of the Indo-Pacific oyster Hyotissa hyotis (Linnaeus, 1758) in the Florida Keys (Bivalvia: Gryphaeidae). Molluscan Res. 2004, 24, 149–159. [Google Scholar] [CrossRef]
  42. Bolger, A.M.; Lohse, M.; Usadel, B. Trimmomatic: A Flexible Trimmer for Illumina Sequence Data. Bioinformatics 2014, 30, 2114–2120. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  43. Kearse, M.; Moir, R.; Wilson, A.; Stones-Havas, S.; Cheung, M.; Sturrock, S.; Buxton, S.; Cooper, A.; Markowitz, S.; Duran, C.; et al. Geneious Basic: An Integrated and Extendable Desktop Software Platform for the Organization and Analysis of Sequence Data. Bioinformatics 2012, 28, 1647–1649. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  44. Irwin, A.R.; Strong, E.E.; Kano, Y.; Harper, E.M.; Williams, S.T. Eight New Mitogenomes Clarify the Phylogenetic Relationships of Stromboidea Within the Caenogastropod Phylogenetic Framework. Mol. Phylogenet. Evol. 2021, 158, 107081. [Google Scholar] [CrossRef] [PubMed]
  45. 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]
  46. Laslett, D.; Canbäck, B. ARWEN: A program to detect tRNA genes in metazoan mitochondrial nucleotide sequences. Bioinformatics 2008, 24, 172–175. [Google Scholar] [CrossRef] [Green Version]
  47. 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]
  48. Perna, N.T.; Kocher, T.D. Patterns of nucleotide composition at fourfold degenerate sites of animal mitochondrial genomes. J. Mol. Evol. 1995, 41, 353–358. [Google Scholar] [CrossRef]
  49. Grant, J.R.; Stothard, P. The CGView Server: A Comparative Genomics Tool for Circular Genomes. Nucleic Acids Res. 2008, 36, 181–184. [Google Scholar] [CrossRef]
  50. Nguyen, L.T.; Schmidt, H.A.; Von Haeseler, A.; Minh, B.Q. IQ-TREE: A fast and effective stochastic algorithm for estimating maximum-likelihood phylogenies. Mol. Biol. Evol. 2015, 32, 268–274. [Google Scholar] [CrossRef]
  51. Ronquist, F.; Huelsenbeck, J.P. MrBayes 3: Bayesian Phylogenetic Inference Under Mixed Models. Bioinformatics 2003, 19, 1572–1574. [Google Scholar] [CrossRef]
  52. 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. 2017, 34, 772–773. [Google Scholar] [CrossRef] [Green Version]
  53. Kalyaanamoorthy, S.; Minh, B.Q.; Wong, T.K.; Von Haeseler, A.; Jermiin, L.S. ModelFinder: Fast model selection for accurate phylogenetic estimates. Nat. Methods 2017, 14, 587–589. [Google Scholar] [CrossRef] [Green Version]
  54. Wu, X.; Li, X.; Li, L.; Xu, X.; Xia, J.; Yu, Z. New features of Asian Crassostrea oyster mitochondrial genomes: A novel alloacceptor tRNA gene recruitment and two novel ORFs. Gene 2012, 507, 112–118. [Google Scholar] [CrossRef] [PubMed]
  55. Breton, S.; Stewart, D.T.; Hoeh, W.R. Characterization of a mitochondrial ORF from the gender-associated mtDNAs of Mytilus spp. (Bivalvia: Mytilidae): Identification of the “missing” ATPase 8 gene. Mar. Genom. 2010, 3, 11–18. [Google Scholar] [CrossRef]
  56. Grande, C.; Templado, J.; Zardoya, R. Evolution of gastropod mitochondrial genome arrangements. BMC Evol. Biol. 2008, 8, 61. [Google Scholar] [CrossRef] [Green Version]
  57. Ojala, D.; Montoya, J.; Attardi, G. tRNA punctuation model of RNA processing in human mitochondria. Nature 1981, 290, 470–474. [Google Scholar] [CrossRef]
  58. Yang, Y.; Liu, H.; Qi, L.; Kong, L.; Li, Q. Complete mitochondrial genomes of two toxin-accumulated Nassariids (Neogastropoda: Nassariidae: Nassarius) and their implication for phylogeny. Int. J. Mol. Sci. 2020, 21, 3545. [Google Scholar] [CrossRef]
  59. Li, F.; Zheng, J.; Ma, Q.; Gu, Z.; Wang, A.; Yang, Y.; Liu, C. Phylogeny of Strombidae (Gastropoda) Based on Mitochondrial Genomes. Front. Mar. Sci. 2022, 9, 930910. [Google Scholar] [CrossRef]
  60. Wang, Y.; Yang, Y.; Liu, H.; Kong, L.; Yu, H.; Liu, S.; Li, Q. Phylogeny of Veneridae (Bivalvia) based on mitochondrial genomes. Zool. Scr. 2021, 50, 58–70. [Google Scholar] [CrossRef]
  61. Sun, S.; Kong, L.; Yu, H.; Li, Q. The complete mitochondrial DNA of Tegillarca granosa and comparative mitogenomic analyses of three Arcidae species. Gene 2015, 557, 61–70. [Google Scholar] [CrossRef]
  62. Wolstenholme, D.R. Animal Mitochondrial DNA: Structure and Evolution. Int. Rev. Cytol. 1992, 141, 173–216. [Google Scholar] [PubMed]
  63. Boore, J.L. The duplication/random loss model for gene rearrangement exemplified by mitochondrial genomes of deuterostome animals. In Comparative Genomics: Empirical and Analytical Approaches to Gene Order Dynamics, Map Alignment and the Evolution of Gene Families; Sankoff, D., Nadeau, J.H., Eds.; Springer: Dordrecht, The Netherlands, 2000; pp. 133–147. [Google Scholar]
  64. Stenzel, H.B. Oysters. In Treatise on Invertebrate Paleontology. Part N. Mollusca 6: Bivalvia; Moore, R.C., Ed.; Geological Society of America and University of Kansas Press: Lawrence, KS, USA, 1971; Volume 3, pp. 953–1224. [Google Scholar]
  65. Harry, H.W. Synopsis of the supraspecific classification of living oysters (Bivalvia: Gryphaeidae and Ostreidae). Veliger 1985, 28, 121–158. [Google Scholar]
  66. Li, C.; Kou, Q.; Zhang, Z.; Hu, L.; Huang, W.; Cui, Z.; Liu, Y.; Ma, P.; Wang, H. Reconstruction of the evolutionary biogeography reveal the origins and diversification of oysters (Bivalvia: Ostreidae). Mol. Phylogenet. Evol. 2021, 164, 107268. [Google Scholar] [CrossRef] [PubMed]
  67. Liu, J.; Li, Q.; Kong, L.; Yu, H.; Zheng, X. Identifying the true oysters (Bivalvia, Ostreidae) with mitochondrial phylogeny and distance-based DNA barcoding. Mol. Ecol. Resour. 2011, 11, 820–830. [Google Scholar] [CrossRef]
  68. Salvi, D.; Macali, A.; Mariottini, P. Molecular phylogenetics and systematics of the bivalve family Ostreidae based on rRNA sequence-structure models and multilocus species tree. PLoS ONE 2014, 9, e108696. [Google Scholar] [CrossRef]
  69. Pfeiffer, J.M.; Breinholt, J.W.; Page, L.M. Unioverse: A phylogenomic resource for reconstructing the evolution of freshwater mussels (Bivalvia, Unionoida). Mol. Phylogenet. Evol. 2019, 137, 114–126. [Google Scholar] [CrossRef]
  70. Salvi, D.; Mariottini, P. Molecular taxonomy in 2D: A novel ITS2 rRNA sequence-structure approach guides the description of the oysters’ subfamily Saccostreinae and the genus Magallana (Bivalvia: Ostreidae). Zool. J. Linn. Soc.-Lond. 2017, 179, 263–276. [Google Scholar] [CrossRef] [Green Version]
  71. Wang, Y.; Xu, Z.; Guo, X. Differences in the rDNA-bearing Chromosome divide the Asian-Pacific and Atlantic species of Crassostrea (Bivalvia, Mollusca). Biol. Bull. 2004, 206, 46–54. [Google Scholar] [CrossRef] [Green Version]
  72. Bayne, B.L.; Ahrens, M.; Allen, S.K.; D’auriac, M.A.; Backeljau, T.; Beninger, P.; Bohn, R.; Boudry, P.; Davis, J.; Green, T.; et al. The proposed dropping of the genus Crassostrea for all Pacific cupped oysters and its replacement by a new genus Magallana: A dissenting view. J. Shellfish Res. 2017, 36, 545–547. [Google Scholar] [CrossRef]
Figure 1. Mitochondrial genome map of Hyotissa hyotis.
Figure 1. Mitochondrial genome map of Hyotissa hyotis.
Fishes 07 00317 g001
Figure 2. Relative synonymous codon usage (RSCU) of mitochondrial genome for Hyotissa hyotis.
Figure 2. Relative synonymous codon usage (RSCU) of mitochondrial genome for Hyotissa hyotis.
Fishes 07 00317 g002
Figure 3. Inferred secondary structures of 23 transfer RNAs from Hyotissa hyotis.
Figure 3. Inferred secondary structures of 23 transfer RNAs from Hyotissa hyotis.
Fishes 07 00317 g003
Figure 4. Linearized mitochondrial gene orders of the represented oyster mitogenomes available on GenBank (Accession numbers are provided in Table 1). tRNA genes are marked by the single-letter codes of the encoded amino acids. tRNAs in the dashed box means the missing genes compared with the congeneric species.
Figure 4. Linearized mitochondrial gene orders of the represented oyster mitogenomes available on GenBank (Accession numbers are provided in Table 1). tRNA genes are marked by the single-letter codes of the encoded amino acids. tRNAs in the dashed box means the missing genes compared with the congeneric species.
Fishes 07 00317 g004
Figure 5. Linearized PCG order of Ostreidae (including Crassostrea, Saccostrea and Ostrea) and Gryphaeidae (represented by Hyotissa hyotis only) based on the phylogenetic tree (see Figure 6 for detailed relationships). By assuming the PCG order in Ostrea as pleisomorphic of Ostreidae, the transversions and translocations (happened in Crassostrea and Saccostrea) are indicated by red rectangular and oval boxs, respectively. The identical gene block (Nad5-Nad6-Nad4-Atp6) and the inverted gene block (Nad1-Nad3-Cox2-Cytb) between H. hyotis and Saccostrea are marked by solid and dotted underlines, respectively.
Figure 5. Linearized PCG order of Ostreidae (including Crassostrea, Saccostrea and Ostrea) and Gryphaeidae (represented by Hyotissa hyotis only) based on the phylogenetic tree (see Figure 6 for detailed relationships). By assuming the PCG order in Ostrea as pleisomorphic of Ostreidae, the transversions and translocations (happened in Crassostrea and Saccostrea) are indicated by red rectangular and oval boxs, respectively. The identical gene block (Nad5-Nad6-Nad4-Atp6) and the inverted gene block (Nad1-Nad3-Cox2-Cytb) between H. hyotis and Saccostrea are marked by solid and dotted underlines, respectively.
Fishes 07 00317 g005
Figure 6. Phylogenetic relationships of Ostreoidea based on the concatenated nucleotide sequences of 12 mitochondrial protein-coding genes. The reconstructed Bayesian inference (BI) phylogram using Pinctada margaritifera and P. maxima as outgroup is shown. The first number at each node is Bayesian posterior probability (PP), and the second number is bootstrap proportion (BP) of maximum likelihood (ML) analyses. Nodal with maximum statistical supports (PP = 1; BP = 100) is marked with a solid red circle. BP values under 80 are marked as dash line.
Figure 6. Phylogenetic relationships of Ostreoidea based on the concatenated nucleotide sequences of 12 mitochondrial protein-coding genes. The reconstructed Bayesian inference (BI) phylogram using Pinctada margaritifera and P. maxima as outgroup is shown. The first number at each node is Bayesian posterior probability (PP), and the second number is bootstrap proportion (BP) of maximum likelihood (ML) analyses. Nodal with maximum statistical supports (PP = 1; BP = 100) is marked with a solid red circle. BP values under 80 are marked as dash line.
Fishes 07 00317 g006
Table 1. List of mtDNA used in the present study.
Table 1. List of mtDNA used in the present study.
New Mitochondrial Genome
FamilySpeciesLength (bp)Sampling TimeAccession No.
OP151093
GryphaeidaeHyotissa hyotis22,185May, 2022
GenBank Mitochondrial Genome
FamilySpeciesLength (bp)Accession No.Reference
OstreidaeCrassostrea hongkongensis18,617MZ337404Liu et al. (2022) [32]
OstreidaeCrassostrea bilineata22,420MT985154Arshad et al. (2020)
OstreidaeCrassostrea belcheri21,020MH051332Gastineau et al. (2018) [33]
OstreidaeCrassostrea gasar17,685KR856227Cavaleiro et al. (2016) [34]
OstreidaeCrassostrea nippona20,030HM015198Yu and Li (2012) [35]
OstreidaeCrassostrea iredalei22,446FJ841967Wu et al. (2010) [36]
OstreidaeCrassostrea ariakensis18,414EU672835Ren et al. (2010) [25]
OstreidaeCrassostrea sikamea18,243EU672833Ren et al. (2010) [25]
OstreidaeCrassostrea angulata18,225EU672832Ren et al. (2010) [25]
OstreidaeCrassostrea gigas18,225EU672831Ren et al. (2010) [25]
OstreidaeCrassostrea virginica17,244AY905542Milbury and Gaffney (2005) [37]
OstreidaeOstrea denselamellosa16,277HM015199Yu and Li (2011) [38]
OstreidaeOstrea edulis16,320JF274008Danic-Tchaleu et al. (2011) [29]
OstreidaeOstrea lurida16,344KC768038Xiao et al. (2015) [39]
OstreidaeSaccostrea mordax16,532FJ841968Wu et al. (2009)
OstreidaeSaccostrea cucullata16,396KP967577Volatiana et al. (2016) [30]
OstreidaeSaccostrea kegaki16,260KX065089Hsiao (2016)
MargaritidaePinctada margaritifera15,680HM467838Wu et al. (2010)
MargaritidaePinctada maxima16,994GQ452847Wu et al. (2012) [40]
Table 2. Gene annotations of the mtDNA of Hyotissa hyotis.
Table 2. Gene annotations of the mtDNA of Hyotissa hyotis.
GeneStrandLocationSize
(bp)
Start
Codon
Stop
Codon
Intergenic
Nucleotides
Cox1H1–17251725ATGTAG1528
tRNA-AlaH3254–331966 31
tRNA-GlnH3351–341363 51
tRNA-LysH3465–353167 403
tRNA-MetH3935–402591 325
tRNA-TyrH4351–441464 36
tRNA-GluH4451–451666 8
tRNA-PheH4525–458763 0
rrnLH4588–59661379 175
rrnSH6142–7099958 93
tRNA-IleH7193–726068 22
tRNA-ThrH7283–734563 20
tRNA-Ser1H7366–743570 90
Nad1H7526–8460935ATGTA-71
Atp8H8532–8654123ATGmissing1
tRNA-ProH8656–872065 13
tRNA-AspH8734–879764 10
tRNA-Leu2H8808–887366 177
Nad3H9051–9417367TTTT--408
tRNA-AsnH9826–989267 4
tRNA-GlyH9897–996165 3
tRNA-Ser2H9965–1003470 21
tRNA-Trp1H10056–1012267 7
tRNA-Leu1H10130–1019465 20
tRNA-ValH10215–1028167 29
tRNA-HisH10311–1037565 15
Cox2H10391–11089699TTGTAA138
CytbH11228–129101683ATGTAA1191
Nad2H14102–151211020ATTTAG13
tRNA-Trp2H15135–1520167 35
Nad5H15237–169011667AAGTAA578
Nad6H17480–18034555ATGTAA64
tRNA-CysH18099–1816264 2
Nad4H18165–195081344GTGTAA66
Atp6H19575–20324750ATGTAA20
Nad4LH20345–20638294TATTAA52
tRNA-ArgH20691–2075363 252
Cox3H21006–21884879ATATAA301
Table 3. List of AT content, AT skew, and GC skew of Hyotissa hyotis.
Table 3. List of AT content, AT skew, and GC skew of Hyotissa hyotis.
Feature(A+T)%AT skewGC skew
Whole genome59.2−0.200.33
PCGs59.4−0.290.34
PCGs157.3−0.090.38
PCGs259.9−0.340.08
PCGs361.0−0.420.57
Atp659.8−0.230.40
Atp860.1−0.300.14
Cox158.6−0.280.29
Cox259.0−0.250.37
Cox359.3−0.310.30
Cytb60.5−0.200.27
Nad159.5−0.360.36
Nad259.1−0.300.39
Nad358.4−0.430.50
Nad461.6−0.330.35
Nad4L58.5−0.330.43
Nad559.1−0.290.39
Nad657.7−0.240.39
tRNAs57.9−0.120.25
rrnS53.70.110.11
rrnL58.8−0.020.20
LNR164.8−0.080.16
LNR251.7−0.210.52
Note: LNR is the abbreviation of large non-coding region.
Table 4. Codon and relative synonymous codon usage (RSCU) of 12 protein-coding genes (PCGs) in the mtDNA of Hyotissa hyotis.
Table 4. Codon and relative synonymous codon usage (RSCU) of 12 protein-coding genes (PCGs) in the mtDNA of Hyotissa hyotis.
Amino Acid CodonCount (RSCU)AminoCodonCount (RSCU)
PheTTT289.0(1.78)AlaGCT134.0(2.08)
TTC35.0(0.22) GCC24.0(0.37)
LeuTTA91.0(1.16) GCA41.0(0.64)
TTG232.0(2.96) GCG59.0(0.91)
CTT84.0(1.07)GlyGGT101.0(1.22)
CTC9.0(0.11) GGC24.0(0.29)
CTA18.0(0.23) GGA91.0(1.10)
CTG 36.0(0.46) GGG115.0(1.39)
IleATT193.0(1.74)ArgCGT40.0(1.52)
ATC29.0(0.26) CGC12.0(0.46)
MetATA47.0(0.44) CGA28.0(1.07)
ATG169.0(1.56) CGG25.0(0.95)
ValGTT155.0(1.68)TyrTAT135.0(1.56)
GTC25.0(0.27) TAC38.0(0.44)
GTA68.0(0.74)HisCAT70.0(1.65)
GTG122.0(1.32) CAC15.0(0.35)
SerTCT99.0(2.21)GlnCAA22.0(0.90)
TCC6.0(0.13) CAG27.0(1.10)
TCA25.0(0.56)AsnAAT81.0(1.59)
TCG49.0(1.09) AAC21.0(0.41)
AGT47.0(1.05)LysAAA81.0(0.98)
AGC15.0(0.34) AAG84.0(1.02)
AGA54.0(1.21)AspGAT67.0(1.52)
AGG63.0(1.41) GAC21.0(0.48)
ProCCT82.0(2.33)GluGAA65.0(0.97)
CCC11.0(0.31) GAG69.0(1.03)
CCA11.0(0.31)CysTGT77.0(1.47)
CCG37.0(1.05) TGC28.0(0.53)
ThrACT83.0(2.11)TrpTGA39.0(0.52)
ACC19.0(0.48) TGG110.0(1.48)
ACA22.0(0.56)
ACG33.0(0.84)
Publisher’s Note: MDPI stays neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Share and Cite

MDPI and ACS Style

Li, F.; Fan, M.; Wang, S.; Gu, Z.; Wang, A.; Liu, C.; Yang, Y.; Liu, S. The Complete Mitochondrial Genome of Hyotissa hyotis (Bivalvia: Gryphaeidae) Reveals a Unique Gene Order within Ostreoidea. Fishes 2022, 7, 317. https://doi.org/10.3390/fishes7060317

AMA Style

Li F, Fan M, Wang S, Gu Z, Wang A, Liu C, Yang Y, Liu S. The Complete Mitochondrial Genome of Hyotissa hyotis (Bivalvia: Gryphaeidae) Reveals a Unique Gene Order within Ostreoidea. Fishes. 2022; 7(6):317. https://doi.org/10.3390/fishes7060317

Chicago/Turabian Style

Li, Fengping, Mingfu Fan, Shunshun Wang, Zhifeng Gu, Aimin Wang, Chunsheng Liu, Yi Yang, and Shikai Liu. 2022. "The Complete Mitochondrial Genome of Hyotissa hyotis (Bivalvia: Gryphaeidae) Reveals a Unique Gene Order within Ostreoidea" Fishes 7, no. 6: 317. https://doi.org/10.3390/fishes7060317

Article Metrics

Back to TopTop