Research Article
Print
Research Article
Complete mitochondrial genomes of Sinonovacula rivularis and Novaculina chinensis and their phylogenetic relationships within family Pharidae
expand article infoYiping Meng§, Liyuan Lv|, Zhihua Lin§, Demin Zhang, Yinghui Dong§|
‡ Ningbo University, Ningbo, China
§ Zhejiang Wanli University, Ningbo, China
| Zhejiang Wanli University, Ninghai, China
Open Access

Abstract

Pharidae is one of the most ecologically and commercially significant families of marine Bivalvia; however, the taxonomy and phylogeny of Pharidae has been ongoing for quite some time and remains a contentious issue. Here, to resolve some problematical relationships among this family, the complete mitochondrial genomes (mitogenomes) of Sinonovacula rivularis (17,159 bp) and Novaculina chinensis (15,957 bp) were assembled, and a comparative mitochondrial genomic analysis was conducted. Both mitogenomes contain 12 protein-coding genes, 22 transfer RNA genes, and two ribosomal RNA genes. Among the published Pharidae mitogenomes, N. chinensis exhibited the smallest genome size but the highest AT content. The results of the phylogenetic trees confirmed the monophyly of the family Solenoidea, and indicated that N. chinensis and Sinonovacula (S. constricta and S. rivularis) were closely related in the family Pharidae. From the CREx analysis, we found that transposition and tandem duplication random losses (TDRLs) might have occurred between Pharidae and Solenidae. Moreover, positive selection was detected in nad5 of the foreground N. chinensis, and divergent evolution occurred at site 144 in the freshwater and marine lineages. Overall, our findings provide new molecular data on the phylogenetic and evolutionary relationships of Pharidae, and contribute to unraveling the salinity adaptations of Pharidae.

Key words:

Gene arrangement, mitogenome, Pharidae, phylogeny, positive selection

Introduction

Pharidae belongs to Solenoidea which is one of the most ecologically and commercially significant superfamilies of marine Bivalvia, and the North-West and Indo-West Pacific regions exhibit the highest levels of species diversity, encompassing approximately 85% of all species, predominantly distributed in intertidal zones (Lin 2009; Saeedi et al. 2016; Costello and Saeedi 2019). According to the China Fisheries Statistics Yearbook (2024), the annual output of razor clams is 850,000 tons, accounting for 5.16% of the total output of mollusks. This Pharidae family has an extensive fossil record, dating back to approximately 103 million years ago (Mya) in the middle Cretaceous (Bolotov et al. 2018b). Although Pharidae is well-established as a clade, the internal taxonomic research has been ongoing and remains a contentious problem. Cosel (1993) promoted Solenidae to Solenoidea in 1993 and divided the superfamily into Solenidae and Pharidae according to the number of main teeth. Among them, the genus Sinonovacula was once classified by Graham into the family Solecurtidae, which belongs to the superfamily Tellinoidea (Graham 1935). However, an increasing number research findings contradict this, where the genus Sinonovacula should be categorized into the family Pharidae (Taylor et al. 2007; Guoquan et al. 2010; Yuan et al. 2012c; Yu et al. 2016). For example, the comparison of mitogenomes of six heterodont bivalves indicated that S. constricta (Lamarck, 1818) was more closely related to Solen grandis (Dunker, 1862), which belonged to Solenidae (Yuan et al. 2012c). The phylogenetic tree and molecular clock of tandem mitochondrial gene and nuclear gene (COI, 16S, 28S) revealed that Siliqua, Sinonovacula, Cultellus, and Novaculina belonged to Pharellinae, and Pharella javanica (Lamarck, 1818) was classified under the Sinonovacula subclade (Bolotov et al. 2018b). Moreover, Pharidae were divided into four subfamilies which were composed of 14 existing genera, including Pharinae (Nasopharus, Pharus, Sinupharus), Cultellinae (Afrophaxas, Cultellus, Ensis, Ensiculus, Phaxas, Sinucultellus), Siliquinae (Siliqua), and Pharellinae (Novaculina, Orbicularia, Pharella, Sinonovacula), and Bolotov et al. (2018b) argued that Novaculininae was considered to be a junior synonym of Pharellinae (Appeltans et al. 2012; Signorelli et al. 2021). Nevertheless, since the above studies are only based on a limited number of taxa, the phylogenetic relationship of Pharidae has not been fully studied.

Mitochondrial DNA (mtDNA) is a genetic material independent of the nucleus DNA. Owing to their small size, rapid evolution, maternal inheritance, and simple structure, mitogenomes have become an attractive candidate tool for resolving phylogenetic relationships across a wide spectrum of metazoans (Boore 1999; Curole and Kocher 1999; Saccone et al. 1999; Miya et al. 2001; Gissi et al. 2008; Osigus et al. 2013; Cameron 2014). Mitogenomes of metazoan are usually circular double-stranded molecules, and range in size from 14 kb to 42 kb (Okimoto et al. 1992; Wolstenholme 1992; Smith and Snyder 2007). The typical mitogenome is composed of 37 genes compactly organized in a near-invariant arrangement, including 13 protein-coding genes of the oxidative phosphorylation (OXPHOS) system (cox1–3, cob, nad1–6, nad4L, atp6, atp8), 22 transfer RNAs (tRNAs) and two ribosomal RNAs homologous to the 16s and 23s of Escherichia coli (rrnS and rrnL) (Wolstenholme 1992; Shadel and Clayton 1997; Andrews et al. 1999; Boore 1999). In general, metazoan mtDNA molecules have few or no nucleotides between genes except for a single non-coding region that contains signals for regulating replication and transcription (designated as the control region) (Clayton 1984; Wolstenholme 1992; Shadel and Clayton 1997). However, the phylum Mollusca has generated a vast array of unexpected deviations from the ‘textbook description’, including exceptional variation in size, frequent genome rearrangements, the integration of novel genes, and a complex inheritance system dubbed ‘doubly uniparental inheritance’ (Wu et al. 2012; Williams et al. 2017; Wu et al. 2019; Malkócs et al. 2022).

In mollusks, with the development of DNA sequencing technology, a large number of mitogenomes have been determined during the last thirty years (Yokobori et al. 2004; Yuan et al. 2012d; Kong et al. 2020; Ma et al. 2023; Taite et al. 2023). For instance, through comparing the complete mtDNA sequences of three scallop species from the subfamily Chlamydinae, it was found that the three genomes exhibited high variation in non-coding regions and different tRNA gene sets (Wu et al. 2009). Besides, the results of the phylogenetic analysis based on concatenated 12 protein-coding genes (PCGs) and two rRNA genes validated the monophyly of the family Mactridae and indicated that genera Mactrinula, Spisula, Rangia, and Mulinia should not be placed under subfamily Mactrinae (Ma et al. 2023). Nevertheless, to date, only four mitogenomes of Pharidae, which are ecologically and economically important deep-burrowing bivalves, are available (Zheng et al. 2010; Feng et al. 2021; Li et al. 2022).

Sinonovacula rivularis (R. Huang & Y.-F. Zhang, 2007), the member of the genus Sinonovacula, is similar to S. constricta in reproduction and morphology (Huang and Zhang 2007). In contrast to S. constricta, which exhibits tolerance to wide salinity (5–40 ppt), S. rivularis is capable of thriving in low salt aquatic environments (4–20 ppt), and can even endure in freshwater conditions for over four days (Huang and Zhang 2007; Wang et al. 2009; Peng et al. 2019; Wang et al. 2024). In addition, a typical freshwater genus Novaculina is found in the family Pharidae (Schram 2010; Bolotov et al. 2018b). As a species of Novaculina, N. chinensis (Y.-Y. Liu & W.-Z. Zhang, 1979) was first discovered in Taihu Lake and Gaoyou Lake in China (Liu 1979). However, due to the pollution of water and the lack of protection awareness, they have been in danger of extinction (Liu 1979; Rao et al. 2003). In this study, we assembled the complete mitogenome of S. rivularis and N. chinensis, and analyzed their basic genome characteristics, nucleotide composition and relative synonymous codon usage (RSCU). The phylogenetic tree of Solenoidea was constructed and gene arrangement events between Pharidae and Solenidae were predicted. Furthermore, selective pressure analysis was conducted to explore the evolutionary adaptation of freshwater and marine species. Briefly, our findings will enrich the basis for the taxonomic study of Pharidae and contribute to deepening the understanding of the phylogenetic relationship between Solenoidea and its related groups.

Materials and methods

Sample collection

The samples for whole-genome sequencing of S. rivularis and N. chinensis were collected from the coastal area of Quanzhou in Fujian Province and the Qiantang River in Zhejiang Province, respectively, following the relevant guidelines and regulations. A total of ten individuals each of S. rivularis and N. chinensis were sampled, with average shell length of 55.98 ± 3.47 mm and 45.41 ± 2.74 mm, respectively. All specimens were preserved in 85% ethanol as voucher specimens. These specimens were deposited at Zhejiang Key Laboratory of Aquatic Germplasm Resource, Zhejiang Wanli University, Ningbo, China.

Mitogenome assembly and annotation

Raw genome reads were acquired through both Illumina HiSeq sequencing and PacBio Sequel IIe third-generation sequencing (unpublished), and assembled for the mitogenomes of these two species. Initially, a de novo mitogenome assembly was carried out with SPAdes v3.9.0 after filtering the unqualified reads by Trimmomatic v. 0.39 (Bankevich et al. 2012; Bolger et al. 2014). The scaffold sequences were then obtained by extending the contigs using SSPACE. The assembly quality was evaluated by GetOrganelle software (Jin et al. 2020). Finally, the MitoZ program was used to annotate the protein-coding genes (PCGs), two ribosomal RNAs (rRNAs) and transfer RNAs (tRNAs) (Meng et al. 2019).

Mitogenome characteristics analysis

The content and proportion of nucleotide bases were analyzed by MEGA 11. The base skew values were calculated according to the formulae: AT-skew = (A − T) / (A + T) and GC-skew = (G − C) / (G + C). The RSCU of the two mitogenomes was counted using PhyloSuite v1.2.3.

Phylogenetic analysis and gene arrangement analysis

To explore the evolutionary relationship of S. rivularis and N. chinensis, the published mitogenome sequences of Solenoidea and Hiatelloidea were retrieved from GenBank, and Solecurtus divaricatus was selected as the outgroup (Table 1). The phylogenetic analysis was performed using PhyloSuite software (Zhang et al. 2020). First, using an invertebrate mitochondrial code table, MAFFT was used to independently align 12 protein-coding genes. The ATP8 gene was excluded due to its deletion in the majority of mollusks. Poorly aligned regions of the sequences were pruned by Gblocks under default parameters. The resulting alignments were then concatenated and transferred to ModelFinder for the best model prediction. Phylogenetic trees were estimated through maximum likelihood (ML) and Bayesian inference (BI) methods. The ML phylogenetic tree was generated using IQ-Tree with 1000 bootstrap replicates. The BI analyses were performed by MrBayes 3.2.6 with Markov Chain Monte Carlo (MCMC) for 5000,000 generations. The first 25% of trees were discarded as burn-in and the sampling was terminated when the convergence value was less than 0.01. The iTOL tool was exploited to visualize the phylogenetic tree (https://itol.embl.de/).

Table 1.

List of species used for phylogenetic analysis in this study and their GenBank accession numbers.

Order Superfamily Family Species Length (bp) Accession number Percent of AT (%)
Adapedonta Solenoidea Pharidae Novaculina chinensis 15,957 PP874232 71.50
Sinonovacula rivularis 17,159 PP874231 66.80
Sinonovacula constricta 17,224 JN398366.1 67.00
Ensis leei 16,926 MW727513.1 65.50
Cultellus attenuatus 16,888 MW653805.1 66.46
Siliqua minima 17,064 MT375556.1 66.41
Solenidae Solen strictus 16,535 NC_017616.1 62.70
Solen grandis 16,784 NC_016665.1 64.84
Hiatelloidea Hiatellidae Panopea abrupta 15,381 NC_033538.1 64.40
Panopea globosa 15,469 NC_025636.1 63.70
Panopea generosa 15,585 NC_025635.1 63.70
Panopea japonica 16,128 NC_072278.1 63.80
Hiatella sp. 19,507 OR420023.1 64.00
Hiatella arctica 18,244 DQ632742.1 66.40
Cardiida Tellinoidea Solecurtidae Solecurtus divaricatus 16,749 JN398367.1 60.10

In addition, the most plausible gene order rearrangement events that might have occurred between Pharidae and Solenidae were reconstructed by pairwise comparisons of mitogenomes through the Common Interval Rearrangement Explorer (CREx) (Bernt et al. 2007).

Selective pressure analysis

The branch-site model was used to analyze the selection pressure on 12 PCGs of razor clams in the PAML package. In this model, N. chinensis was marked as the foreground branch to investigate the evolutionary adaptation between freshwater and marine species. The null model (model = 2, Nssites = 2, fix_omega = 1, omega = 1) and alternative model (model = 2, Nssites = 2, fix_omega = 0, omega = 2) were compared by likelihood ratio test (LRT). Subsequently, P-values were calculated through the chi-square distribution. Then, the posterior probability of the amino acid sites under positive selection was calculated according to the Bayesian empirical Bayes (BEB) method. The inference of positively selected sites was based on a posterior probability of greater than 95%.

Results

General features of S. rivularis and N. chinensis mitogenomes

The lengths of S. rivularis and N. chinensis mitogenomes were 17,159 bp and 15,957 bp, respectively (Fig. 1A). Both mitogenomes contain 12 PCGs, 22 tRNAs, and 2 rRNAs, all of which were located on the heavy chain. The ATP8 gene was missing in this two mitogenomes. Their composition was similar to that of other species in Pharidae, indicating a certain degree of conservation in this family. The detailed genes information was shown in Table 2. The base composition of S. rivularis and N. chinensis mitogenomes was displayed in Table 3 with AT contents of 66.80% and 71.50%, respectively, both of which exhibited an obvious AT bias. The AT content of N. chinensis was the highest among the published Adapedonta mitogenomes. In addition, the two mitogenomes all exhibited negative AT-skew and positive GC-skew, reflecting that the base composition ratios were A biased to T, and G biased to C. There were some differences in the types of start and termination codons of 12 PCGs between the two species (Table 2). Specifically, the start codons of 12 genes in S. rivularis were found to be ATN, TTG and GTG types, whereas in N. chinensis, all genes began with the codon ATN, with the exception of the ND4 gene, which used TTG as the start codon. Concerning termination codons, six genes in S. rivularis (cytb, atp6, cox3, nad4, nad3, nad1) and seven genes in N. chinensis (cytb, nad6, atp6, cox3, nad4l, nad3, nad1) were detected TAA or TAG at the sequence end. The remaining genes featured an incomplete termination codon consisting of a T that might be complemented into a complete stop codon by polyadenylation following transcription to the resultant mRNA (Ojala et al. 1981). Furthermore, the non-coding regions of the mitochondrial genomes of N. chinensis and S. rivularis account for 11.92% and 19.33%, respectively. The longest non-coding region (NCR) of N. chinensis and S. rivularis was both located between nad2 and trnK, with lengths of 443 bp and 1,639 bp respectively, which was identified as a putative control region (CR).

Table 2.

Mitochondrial genome organization of Sinonovacula rivularis and Novaculina chinensis.

Gene Sinonovacula rivularis Novaculina chinensis
Size (bp) Start End Codon start/stop Intergenic nucleotide (bp) Size (bp) Start End Codon start/stop Intergenic nucleotide (bp)
CYTB 1120 13 1132 TTG/TAG 36 1146 9247 10392 ATG/TAA 12
ND6 227 1169 1395 TTG/T-- 265 531 10405 10935 ATG/TAG -30
l-rRNA 1661 2957 -35 10906 12201
ATP6 700 2923 3622 ATG/TAA 23 699 12164 12862 ATG/TAA 15
trnM 3646 3713 76 12878 12943 77
s-rRNA 3790 4637 -2 13021 13869 -2
COX3 790 4636 5425 ATG/TAG -2 789 13868 14656 ATG/TAG -1
trnS 5424 5491 5 14656 14722 6
ND2 899 5497 6395 GTG/T-- 1639 898 14729 15626 ATT/T-- 443
trnK 8035 8102 48 113 179 45
COX2 725 8151 8875 ATG/T-- 132 726 225 950 ATG/T-- 256
trnY 9008 9072 -20 1207 1270 7
ND4L 287 9053 9339 ATT/T-- 33 288 1278 1565 ATG/TAA 1
trnG 9373 9439 20 1567 1632 10
trnP 9460 9525 122 1643 1707 122
ND4 1354 9648 11001 TTG/TAG 8 1254 1830 3083 TTG/T-- 103
trnH 11010 11076 -1 3187 3250 2
trnW 11076 11144 2 3253 3319 3
trnR 11147 11213 13 3323 3387 18
trnE 11227 11294 -7 3406 3472 -6
trnS 11288 11351 30 3467 3529 33
ND3 337 11382 11718 ATA/TAA 15 333 3563 3895 ATT/TAG -1
trnT 11734 11800 9 3895 3960 3
trnI 11810 11876 8 3964 4029 15
trnD 11885 11951 -1 4045 4110 0
trnQ 11951 12018 5 4111 4178 10
trnC 12024 12092 42 4189 4253 2
trnA 12135 12200 23 4256 4320 5
trnF 12224 12288 223 4326 4389 200
COX1 1488 12512 13999 CGA/T-- 142 1512 4590 6101 ATT/T-- 154
trnL 14142 14209 8 6256 6320 0
ND1 919 14218 15136 GTG/TAA 2 927 6321 7247 ATG/TAA 4
trnL 15139 15207 1 7252 7319 0
trnV 15209 15274 2 7320 7383 0
trnN 15277 15343 35 7384 7449 36
ND5 1443 15379 16821 ATT/T-- 350 1441 7486 8926 ATT/T-- 320
Table 3.

Nucleotide composition and skewness of the mitogenomes of S. constricta, S. rivularis, and N. chinensis.

Species AT (%) GC (%) AT skew GC skew
S. constricta 67.00% 32.90 -0.22687 0.367781
S. rivularis 66.80% 28.50 -0.21958 0.319298
N. chinensis 71.50% 33.20 -0.23653 0.379518
Figure 1. 

Maps of A the mitogenomes of S. rivularis and N. chinensis and their B RSCU.

As illustrated in Fig. 1B, the preferred codons for 22 amino acids of two species ended in A or U, consisting with the result of AT bias of the mitogenome sequence. As a consequence of the duplication of tRNA-Leu and tRNA-Ser, Leu and Ser were each encoded by six and eight codons, respectively. The most frequently used codons were UUA (Leu2), UCU (Ser2), GCU (Ala) and CCU (Pro). Compared to S. rivularis, CUG (Leu1), AUC (Ile), AAC(Asn) were utilized to a lesser extent in N. chinensis.

Phylogenetic analysis

The 12 protein-coding genes from 15 taxa were concatenated to generate a sequence matrix of 10,806 bp. The tree topologies derived from the ML and BI analyses were largely congruent exhibiting high posterior probabilities (PP) and bootstrap support values (BS) in most nodes (Fig. 2). Phylogenetic analyses revealed that the genus Hiatella from Hiatelloidea was closely related to the superfamily Solenoidea, indicating a close evolutionary relationship between them. Additionally, both analyses strongly confirmed the monophyly of Solenoidea, which was divided into two major branches, Solenidae and Pharidae. In the family Pharidae, the genus Sinonovacula (including S. rivularis and S. constricta) was clustered alongside N. chinensis, with Cultellus attenuatus emerging as a sister group. Siliqua minima and Ensis leei were clustered in a separate cluster.

Figure 2. 

The phylogenetic trees based on concatenated 12 mitochondrial PCGs, and the gene orders of Adapedonta species. Values shown next to nodes are posterior probabilities (left) and ML bootstrap support values (right). Newly assembly mitogenomes are marked with triangles. Except for Panopea abrupta (https://inverts.wallawalla.edu) and Panopea globosa (Góngora-Gómez et al. 2016), the images of the other species are all from https://www.inaturalist.org.

Gene arrangement

The mitogenomes of Solenoidea all exhibited the identical composition of 12 PCGs, 22 tRNAs, and 2 rRNAs, except for Ensis leei, which contained an additional ATP8 gene (Fig. 3). The gene arrangement was consistent within each family, and there was a certain level of conservation in gene arrangement between Solenidae and Pharidae. A large block, rrnL-ATP6-M-rrnS-cox3, and five small blocks, L2-nad1-L1, S-nad2, nad5-cytb, I-D, Q-C were shared by both families, providing further evidence of the close lineage relationship observed in the phylogenetic analysis of this study. The CREx analysis suggested that three transposition and four tandem duplication random losses (TDRLs) might have occurred between Pharidae and Solenidae.

Figure 3. 

Putative gene rearrangement events between Pharidae and Solenidae. Green and red lines represent transposition and TDRL events, respectively, which were step by step identified by CREx.

Select pressure analysis

The species of Solenoidea were selected for molecular evolution analysis, with N. chinensis designated as the foreground branch (Fig. 4). The branch-site model (BSM) in the PAML package was employed to detect positively selected genes (PSGs). As illustrated in Table 4, the substitution model A was significantly better than the neutral selection model null in nad5, indicating that this gene underwent positive selection in the foreground branch (P < 0.05). According to the BEB analysis, there were five positive selection sites in the nad5 amino acid sequences (140 A 0.509, 143 F 0.547, 144 L 0.865, 442 A 0.700, 446 F 0.620). Moreover, discrepancies were observed in the 144th site between freshwater N. chinensis (Ala) and marine razor clams (Leu) (Fig. 5). However, the evidence for each site was somewhat inconclusive. These findings suggest that the nad5 gene may have played a pivotal role in the adaptive evolution of freshwater environments.

Table 4.

The results of positively selected gene sites for 12 PCGs.

Gene lnL0 lnL1 Np0 Np1 Omega P value Positively selected sites (PSGs)
nad3 -1552.22 -1552.22 17 18 2.52856 1
nad1 -4134.56 -4134.56 17 18 2.35774 1
cytb -4826.79 -4826.79 17 18 2.62875 1
nad4L -1236.77 -1236.78 17 18 3.31711 0.895254
nad5 -6913.41 -6842.58 17 18 3.34388 0 140 A 0.509, 143 F 0.547, 144 L 0.865, 442 A 0.700, 446 F 0.620
cox1 -5205.94 -5205.94 17 18 2.64645 0.998872
nad2 -4693.2 -4693.2 17 18 3.07094 1
nad6 -877.007 -877.007 17 18 2.68959 1
nad4 -6088.71 -6088.71 17 18 3.25046 1
cox3 -3173.94 -3173.94 17 18 2.39696 1
cox2 -2763.71 -2763.71 17 18 1.96195 1
atp6 -2882.5 -2884.28 17 18 3.30394 0.058789
Figure 4. 

Phylogenetic tree of Solenoidea for selective stress analysis. The branch marked in red is the foreground branch.

Figure 5. 

The difference of the 144th positive selected amino acid site in NAD5 of eight Solenoidea species. The 144th site is indicated by a red frame.

Discussion

General features of Pharidae mitogenomes

The mitogenomes of S. rivularis and N. chinensis were newly assembled, with lengths of 17,159 and 15,957 bp, respectively. In compared with the previously sequenced Adapedonta mtDNA size (ranged from 15,381 bp to 19,507 bp), their mitogenome sizes were within the normal range (Zheng et al. 2010; Yuan et al. 2012b; Feng et al. 2021; Li et al. 2022). Notably, the genome size of N. chinensis was the smallest in the family Pharidae, which was associated with the variation in length of the control region. The CR is the region with the largest sequence and length variation in the mitogenome, and has the fastest evolution, which is crucial for the regulation of mitochondrial DNA replication and transcription (Wolstenholme 1992; Boore 1999). The substantial differences in the content and structure of the control region within the mollusk lineage provide valuable insights for population genetic analysis (Sasuga et al. 1999; Tomita et al. 2002; Kawashima et al. 2013). Among the published mitogenomes of Pharidae, there is a large control region between nad2 and trnK, such as S. constricta (1,492 bp), S. minima (1,371 bp), C. attenuatus (1,173 bp) and E. leei (1,101 bp) (Zheng et al. 2010; Feng et al. 2021; Li et al. 2022). In this study, S. rivularis displayed a moderately larger control region size of 1,639 bp, whereas it was only 441 bp in N. chinensis, making it a different mitogenome size in the family Pharidae. Intriguingly, a similar control region was not observed in the species of Solenidae (Yuan et al. 2012a, b). This distinction provides evidence for the taxonomic division of the subfamily of Solenoidea.

Molecular phylogeny and gene arrangement of the family Pharidae

The topological tree constructed from the 12 mitochondrial PCGs sequence based on the BI and ML methods yielded consistent results, demonstrating that Solenoidea is clearly divided into Solenidae and Pharidae, which is consistent with the prior research results (Yuan et al. 2012d; Feng et al. 2021). Previously, S. rivularis was identified as a new species of Sinonovacula distinct from S. constricta based on morphological studies and a comparative analysis of COI and 16SrRNA fragments (Huang and Zhang 2007; Weng et al. 2013). In this research, this classification view was supported at the level of mitogenomes, and Sinonovacula belonged to the family Pharidae (Adapedonta: Solenoidea). In addition, N. chinensis was previously classified into Solecurtidae, whereas the results of this study demonstrated that N. chinensis and Sinonovacula are clustered together, forming a novel branch in the family Pharidae, which was consistent with the taxa in WoRMS (Liu 1979; Appeltans et al. 2012). Recently, the phylogenetic tree and molecular clock of tandem mitochondrial gene and nuclear gene (COI, 16S, 28S) revealed that Siliqua, Sinonovacula, Cultellus, and Novaculina belonged to Pharellinae (Bolotov et al. 2018b). However, Cultellus and Siliqua were categorized into the subfamily Cultellinae and Siliquinae, respectively, by Ahyong (Appeltans et al. 2012). Previously, Pharidae were divided into four subfamilies: Pharinae (Nasopharus, Pharus, Sinupharus), Cultellinae (Afrophaxas, Cultellus, Ensis, Ensiculus, Phaxas, Sinucultellus), Siliquinae (Siliqua), and Pharellinae (Novaculina, Orbicularia, Pharella, Sinonovacula). However, the present results indicate that Pharidae are divided into two clades, in which Cultellus is clustered alongside the Sinonovacula and Novaculina, while Siliqua and Ensis clustered together. These observations reflect that the current categorization of the subfamily Pharidae requires further research and refinement, particularly in combination with more species information.

Unlike stable gene arrangements of Vertebrata and Arthropoda, the gene orders of all genes within mtDNA exhibit considerable variability in every major molluscan lineage, including Cephalopoda, Bivalvia, Scaphopoda, and Monoplacophora (Rawlings et al. 2001; Dreyer and Steiner 2004; Yuan et al. 2012c; Stöger et al. 2016; Ma et al. 2023). Gene rearrangements may be caused by reverse transpositions, transpositions, inversions, and TDRL, which can provide important clues about the evolutionary history of species (Boore and Brown 1998; Serb and Lydeard 2003; Wang et al. 2021). In this paper, CREx analysis predicted that three transpositions and four TDRLs might have occurred between Pharidae and Solenidae, implying that dramatic mitogenome changes occurred during species differentiation. Moreover, the gene order illustration of Adapedonta revealed that species with a closer genetic relationship tended to share a similar gene arrangement, indicating that there is a potential relationship between evolution and gene rearrangement (Fig. 2). However, in this study, three distinct gene arrangement types were observed in the family Hiatellidae, especially in the genus Hiatella with nad3 and nad1 transpositions in terms of 12 PCGs arrangement (Fig. 2). The similar case that different gene arrangements in the same genus has also been reported in the genera Dendropoma and Crassostrea (Rawlings et al. 2010; Ren et al. 2010). Therefore, the taxonomic evolution of species cannot be substantiated exclusively through the examination of gene sequences; it also necessitates the integration of phylogenetic reconstruction.

Adaptive evolution of Pharidae mitochondrial genes to freshwater environment

Pharidae is a major marine family, with the exception of Novaculina, that is a relict marine-derived freshwater lineage (Annandale 1922; Bolotov et al. 2018a). The branch-site model study was used to determine whether positive selection occurs at a few places in freshwater razor clam. The results suggested that the nad5 gene underwent positive selection. NADH dehydrogenase is the initial and most substantial enzyme complex in the respiratory chain, functioning as a proton pump (da Fonseca et al. 2008). Nad2, nad4, and nad5 are considered to be the actual proton pumping devices because of their sequence homology with a class of Na+/H+ antiporters (Brandt 2006). The efficiency of the proton transfer process may be interfered by the mutation of the complex, which could be a crucial factor in adaptive evolution (Hassanin et al. 2009; Yu et al. 2011). For instance, the outcomes of positive selection sites in mussels from disparate habitats reflected that the p-value of nad4 was significant in freshwater branches and six sites were identified as positive sites with BEB analysis (> 95%), which implies that nad4 may contribute to the adaptation of Limnoperna fortunei in freshwater (Zhao et al. 2022). Significant non-synonymous changes were detected in the cytb and nad5 genes by comparing mitogenomes of panpulmonate gastropods that are distributed from marine to intertidal and terrestrial habitats (Romero et al. 2016). Therefore, the positive selection of nad5 gene in N. chinensis may be the result of the adaptive evolution of freshwater environment. Moreover, divergent selection occurred at site 144 of nad5, where the amino acids Ala and Leu were identified in the freshwater Novaculina chinensis and seven marine lineages, respectively, indicating divergent evolution exists the family Pharidae. Divergent evolution is the process by which separate species with common ancestors evolve distinct features to adapt to their unique living environment, which is one of the important mechanisms for the formation of biodiversity (Gautam 2020). However, the evidence supporting the positive selection of individual nad5 sites is insufficient. To provide more robust statistical support for the differences in evolutionary adaptation between freshwater and seawater species, it is necessary to include more freshwater razor clam sequences.

Conclusions

In summary, the mitogenomes of S. rivularis and N. chinensis were assembled using next-generation sequencing data, with the genomes measuring 17,159 bp and 15,957 bp, respectively. Both genomes consist of 12 protein-coding genes, 22 transfer RNA genes, and two ribosomal RNA genes. Among the published Pharidae mitogenomes, N. chinensis exhibits the smallest genome size but the highest AT content. The results of the phylogenetic analysis showed that N. chinensis and Sinonovacula (S. constricta + S. rivularis) were closely related and belonged to the family Pharidae. The gene order rearrangements in Solenoidea can be attributed to transposition and TDRL events. Moreover, the nad5 genes carry a signal of positive selections in the foreground N. chinensis, which promotes the adaptation to freshwater environments. We also show that divergent evolution occurred at site 144 in the freshwater and marine lineages. Overall, this study provides further theoretical support for the phylogenetic relationship of Pharidae, and contributes to deepening the understanding of the mitogenomic adaptations of Pharidae.

Acknowledgements

This work was supported financially by Zhejiang Major Program of Science and Technology (2021C02069-7), Ningbo Major Project of Science and Technology (2021Z114) and National Marine Genetic Resource Center Program.

Additional information

Conflict of interest

The authors have declared that no competing interests exist.

Ethical statement

No ethical statement was reported.

Funding

This work was supported by Zhejiang Major Program of Science and Technology (2021C02069-7), Ningbo Major Project of Science and Technology (2021Z114) and National Marine Genetic Resource Center Program.

Author contributions

Yinghui Dong and Demin Zhang conceived and designed the experiments. Yiping Meng performed the experiments and drafted the manuscript. All authors reviewed the paper.

Data availability

All of the data that support the findings of this study are available in the main text.

References

  • Andrews RM, Kubacka I, Chinnery PF, Lightowlers RN, Turnbull DM, Howell N (1999) Reanalysis and revision of the Cambridge reference sequence for human mitochondrial DNA. Nature Genetics 23(2): 147–147. https://doi.org/10.1038/13779
  • Appeltans W, Ahyong S, Anderson G, Angel M, Artois T, Bailly N, Bamber R, Barber A, Bartsch I, Berta A, Błażewicz-Paszkowycz M, Bock P, Boxshall G, Boyko C, Brandão S, Bray R, Bruce N, Cairns S, Chan TY, Cheng L, Collins A, Cribb T, Curini-Galletti M, Dahdouh-Guebas F, Davie P, Dawson M, De Clerck O, Decock W, De Grave S, De Voogd N, Domning D, Emig C, Erséus C, Eschmeyer W, Fauchald K, Fautin D, Feist S, Fransen C, Furuya H, Garcia-Alvarez O, Gerken S, Gibson D, Gittenberger A, Gofas S, Gómez-Daglio L, Gordon D, Guiry M, Hernandez F, Hoeksema B, Hopcroft R, Jaume D, Kirk P, Koedam N, Koenemann S, Kolb J, Kristensen R, Kroh A, Lambert G, Lazarus D, Lemaitre R, Longshaw M, Lowry J, Macpherson E, Madin L, Mah C, Mapstone G, McLaughlin P, Mees J, Meland K, Messing C, Mills C, Molodtsova T, Mooi R, Neuhaus B, Ng P, Nielsen C, Norenburg J, Opresko D, Osawa M, Paulay G, Perrin W, Pilger J, Poore G, Pugh P, Read G, Reimer J, Rius M, Rocha R, Saiz-Salinas J, Scarabino V, Schierwater B, Schmidt-Rhaesa A, Schnabel K, Schotte M, Schuchert P, Schwabe E, Segers H, Self-Sullivan C, Shenkar N, Siegel V, Sterrer W, Stöhr S, Swalla B, Tasker M, Thuesen E, Timm T, Todaro M, Turon X, Tyler S, Uetz P, Van der Land J, Vanhoorne B, Van Ofwegen L, Van Soest R, Vanaverbeke J, Walker-Smith G, Walter T, Warren A, Williams G, Wilson S, Costello M (2012) The magnitude of global marine species diversity. Current Biology 22(23): 2189–2202. https://doi.org/10.1016/j.cub.2012.09.036
  • Bankevich A, Nurk S, Antipov D, Gurevich AA, Dvorkin M, Kulikov AS, Lesin VM, Nikolenko SI, Son P, Prjibelski AD, Pyshkin AV, Sirotkin AV, Vyahhi N, Tesler G, Alekseyev MA, Pevzner PA (2012) SPAdes: a new genome assembly algorithm and its applications to single-cell sequencing. Journal of Computational Biology 19(5): 455–477. https://doi.org/10.1089/cmb.2012.0021
  • Bernt M, Merkle D, Ramsch K, Fritzsch G, Perseke M, Bernhard D, Schlegel M, Stadler PF, Middendorf M (2007) CREx: inferring genomic rearrangements based on common intervals. Bioinformatics 23(21): 2957–2958. https://doi.org/10.1093/bioinformatics/btm468
  • Bolotov IN, Aksenova OV, Bakken T, Glasby CJ, Gofarov MY, Kondakov AV, Konopleva ES, Lopes-Lima M, Lyubas AA, Wang Y, Bychkov AY, Sokolova AM, Tanmuangpak K, Tumpeesuwan S, Vikhrev IV, Shyu JBH, Win T, Pokrovsky OS (2018a) Discovery of a silicate rock-boring organism and macrobioerosion in fresh water. Nature Communications 9(1): 2882–2892. https://doi.org/10.1038/s41467-018-05133-4
  • Bolotov IN, Vikhrev IV, Lopes-Lima M, Lunn Z, Chan N, Win T, Aksenova OV, Gofarov MY, Kondakov AV, Konopleva ES, Tumpeesuwan S (2018b) Discovery of Novaculina myanmarensis sp nov (Bivalvia: Pharidae: Pharellinae) closes the freshwater razor clams range disjunction in Southeast Asia. Scientific Reports 8(1): 16325–16336. https://doi.org/10.1038/s41598-018-34491-8
  • da Fonseca RR, Johnson WE, O’Brien SJ, Ramos MJ, Antunes A (2008) The adaptive evolution of the mammalian mitochondrial genome. BMC Genomics 9: 119–140. https://doi.org/10.1186/1471-2164-9-119
  • Dreyer H, Steiner G (2004) The complete sequence and gene organization of the mitochondrial genome of the gadilid scaphopod Siphonondentalium lobatum (Mollusca). Molecular Phylogenetics and Evolution 31(2): 605–617. https://doi.org/10.1016/j.ympev.2003.08.007
  • Feng J, Guo Y, Yan C, Ye Y, Yan X, Li J, Xu K, Guo B, Lu Z (2021) Novel gene rearrangement in the mitochondrial genome of Siliqua minima (Bivalvia, Adapedonta) and phylogenetic implications for Imparidentia. PLoS ONE 16(4): 1–21. https://doi.org/10.1371/journal.pone.0249446
  • Gissi C, Iannelli F, Pesole G (2008) Evolution of the mitochondrial genome of Metazoa as exemplified by comparison of congeneric species. Heredity 101(4): 301–320. https://doi.org/10.1038/hdy.2008.62
  • Góngora-Gómez AM, Sotelo-Gonzalez MI, Hernández-Sepúlveda JA, Domínguez-Orozco AL, García-Ulloa Gómez M (2016) Nuevo registro de la almeja generosa Panopea globosa (Dall, 1898) (Bivalvia: Hiatellidae) en el estado de Sinaloa, México. Latin American Journal of Aquatic Research 44: 411–415. https://doi.org/10.3856/vol44-issue2-fulltext-22
  • Guoquan Z, Jun F, Shouju J, Yongpu Z, Chen C, Yaoyao Z, Junqi Y (2010) Biochemical genetic analysis of eight isozymes in intra-populations of razor clam Cultellus attenuatus. Fisheries Science 29: 669–673. https://doi.org/10.0000/1003-1111-15350
  • Hassanin A, Ropiquet A, Couloux A, Cruaud C (2009) Evolution of the mitochondrial genome in mammals living at high altitude: new insights from a study of the tribe Caprini (Bovidae, Antilopinae). Journal of Molecular Evolution 68(4): 293–310. https://doi.org/10.1007/s00239-009-9208-7
  • Huang R, Zhang Y (2007) A new species of the genus Sinonovacula. Journal of oceanography In Taiwan Strait 26(1): 115–120.
  • Jin J, Yu W, Yang J, Song Y, dePamphilis CW, Yi T, Li D (2020) GetOrganelle: a fast and versatile toolkit for accurate de novo assembly of organelle genomes. Genome Biology 21(1): 1–31. https://doi.org/10.1186/s13059-020-02154-5
  • Kawashima Y, Nishihara H, Akasaki T, Nikaido M, Tsuchiya K, Segawa S, Okada N (2013) The complete mitochondrial genomes of deep-sea squid (Bathyteuthis abyssicola), bob-tail squid (Semirossia patagonica) and four giant cuttlefish (Sepia apama, S. latimanus, S. lycidas and S. pharaonis), and their application to the phylogenetic analysis of Decapodiformes. Molecular Phylogenetics and Evolution 69(3): 980–993. https://doi.org/10.1016/j.ympev.2013.06.007
  • Kong L, Li Y, Kocot KM, Yang Y, Qi L, Li Q, Halanych KM (2020) Mitogenomics reveals phylogenetic relationships of Arcoida (Mollusca, Bivalvia) and multiple independent expansions and contractions in mitochondrial genome size. Molecular Phylogenetics and Evolution 150: 1–11. https://doi.org/10.1016/j.ympev.2020.106857
  • Lin Z (2009) The genetic structure and diversity analysis of three species of razor clam using AFLP markers. Marine Sciences 33(10): 26–30.
  • Liu Y (1979) A new species of freshwater razor clam, Novaculina chinensis, from jiangsu province, China. Acta Zootaxonomica Sinica 4: 356–358.
  • Ma P, Liu Y, Wang J, Chen Y, Zhang Z, Zhang T, Wang H (2023) Comparative analysis of the mitochondrial genomes of the family Mactridae (Mollusca: Venerida) and their phylogenetic implications. International Journal of Biological Macromolecules 249: 1–9. https://doi.org/10.1016/j.ijbiomac.2023.126081
  • Malkócs T, Viricel A, Becquet V, Evin L, Dubillot E, Pante E (2022) Complex mitogenomic rearrangements within the Pectinidae (Mollusca: Bivalvia). BMC Ecology and Evolution 22(1): 29–49. https://doi.org/10.1186/s12862-022-01976-0
  • Meng G, Li Y, Yang C, Liu S (2019) MitoZ: a toolkit for animal mitochondrial genome assembly, annotation and visualization. Nucleic Acids Research 47(11): 63–69. https://doi.org/10.1093/nar/gkz173
  • Ministry of Agriculture and Rural Affairs of the People’s Republic of China (2024) China Fisheries Statistical Yearbook. China Agriculture Press, Beijing, 23.
  • Miya M, Kawaguchi A, Nishida M (2001) Mitogenomic exploration of higher teleostean phylogenies: a case study for moderate-scale evolutionary genomics with 38 newly determined complete mitochondrial DNA sequences. Molecular Biology and Evolution 18(11): 1993–2009. https://doi.org/10.1093/oxfordjournals.molbev.a003741
  • Ojala D, Montoya J, Attardi G (1981) tRNA punctuation model of RNA processing in human mitochondria. Nature 290(5806): 470–474. https://doi.org/10.1038/290470a0
  • Okimoto R, Macfarlane JL, Clary DO, Wolstenholme DR (1992) The mitochondrial genomes of two nematodes, Caenorhabditis elegans and Ascaris suum. Genetics 130(3): 471–498. https://doi.org/10.1093/genetics/130.3.471
  • Peng M, Liu X, Niu D, Ye B, Lan T, Dong Z, Li J (2019) Survival, growth and physiology of marine bivalve (Sinonovacula constricta) in long-term low-salt culture. Scientific Reports 9(1): 1–9. https://doi.org/10.1038/s41598-019-39205-2
  • Rao X, Xu Y, Chen Y, Lin G (2003) A study on induced spawning of Novaculina chinensis. Journal of Fujian Normal University 19(3): 78–81.
  • Rawlings TA, MacInnis MJ, Bieler R, Boore JL, Collins TM (2010) Sessile snails, dynamic genomes: gene rearrangements within the mitochondrial genome of a family of caenogastropod molluscs. BMC Genomics 11(1): 440–463. https://doi.org/10.1186/1471-2164-11-440
  • Ren J, Liu X, Jiang F, Guo X, Liu B (2010) Unusual conservation of mitochondrial gene order in Crassostrea oysters: evidence for recent speciation in Asia. BMC Evolutionary Biology 10(1): 394–407. https://doi.org/10.1186/1471-2148-10-394
  • Romero PE, Weigand AM, Pfenninger M (2016) Positive selection on panpulmonate mitogenomes provide new clues on adaptations to terrestrial life. BMC Evolutionary Biology 16(1): 164–176. https://doi.org/10.1186/s12862-016-0735-8
  • Saeedi H, Basher Z, Costello MJ (2016) Modelling present and future global distributions of razor clams (Bivalvia: Solenidae). Helgoland marine research 70(1): 23–34. https://doi.org/10.1186/s10152-016-0477-4
  • Sasuga J, Yokobori S, Kaifu M, Ueda T, Nishikawa K, Watanabe K (1999) Gene contents and organization of a mitochondrial DNA segment of the squid Loligo bleekeri. Journal of Molecular Evolution 48(6): 692–702. https://doi.org/10.1007/PL00006513
  • Serb JM, Lydeard C (2003) 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). Molecular Biology and Evolution 20(11): 1854–1866. https://doi.org/10.1093/molbev/msg218
  • Signorelli JH, Trovant B, Marquez F (2021) A cryptic species of Ensis (Bivalvia: Pharidae) from the southeastern Pacific coast revealed by geometric morphometric methods. Scientia Marina 86(2): 1–9. https://doi.org/10.3989/scimar.05241.032
  • Smith DR, Snyder M (2007) Complete mitochondrial DNA sequence of the scallop Placopecten magellanicus: evidence of transposition leading to an uncharacteristically large mitochondrial genome. Journal of Molecular Evolution 65(4): 380–391. https://doi.org/10.1007/s00239-007-9016-x
  • Stöger I, Kocot KM, Poustka AJ, Wilson NG, Ivanov D, Halanych KM, Schrödl M (2016) Monoplacophoran mitochondrial genomes: convergent gene arrangements and little phylogenetic signal. BMC Evolutionary Biology 16(1): 274–291. https://doi.org/10.1186/s12862-016-0829-3
  • Taite M, Fernandez-Alvarez FA, Braid HE, Bush SL, Bolstad K, Drewery J, Mills S, Strugnell JM, Vecchione M, Villanueva R, Voight JR, Allcock AL (2023) Genome skimming elucidates the evolutionary history of Octopoda. Molecular Phylogenetics and Evolution 182: 1–11. https://doi.org/10.1016/j.ympev.2023.107729
  • Taylor JD, Williams ST, Glover EA, Dyal P (2007) A molecular phylogeny of heterodont bivalves (Mollusca: Bivalvia: Heterodonta): new analyses of 18S and 28S rRNA genes. Zoologica Scripta 36(6): 587–606. https://doi.org/10.1111/j.1463-6409.2007.00299.x
  • Tomita K, Yokobori S, Oshima T, Ueda T, Watanabe K (2002) The cephalopod Loligo bleekeri mitochondrial genome: multiplied noncoding regions and transposition of tRNA genes. Journal of Molecular Evolution 54(4): 486–500. https://doi.org/10.1007/s00239-001-0039-4
  • Wang Y, Sun T, Wu J, Cui D, Wang A, Wang X (2009) Influence of salinity on the survival and growth of Sinonovacula rivularis. Shandong Fisheries 26(12): 6–7.
  • Wang Y, Yang Y, Liu H, Kong L, Yu H, Liu S, Li Q (2021) Phylogeny of Veneridae (Bivalvia) based on mitochondrial genomes. Zoologica Scripta 50(1): 58–70. https://doi.org/10.1111/zsc.12454
  • Wang S, Shi Y, Dong Y, Meng Y, Yao H, He L (2024) Molecular identification of Sinonovacula constricta, Sinonovacula rivularis and their interspecific hybrids using microsatellite markers. Frontiers in Marine Science 11: 1–9. https://doi.org/10.3389/fmars.2024.1360596
  • Weng Z, Xie Y, Xiao Z, Ren P, Wang Z, Jianfang G (2013) Molecular identification of the taxonomic status of Sinonovacula rivularis and genus Sinonovacula using mitochondrial COI and 16S rRNA fragments. Acta Hydrobiologica Sinica 37(4): 684–690.
  • Williams ST, Foster PG, Hughes C, Harper EM, Taylor JD, Littlewood DTJ, Dyal P, Hopkins KP, Briscoe AG (2017) Curious bivalves: Systematic utility and unusual properties of anomalodesmatan mitochondrial genomes. Molecular Phylogenetics and Evolution 110: 60–72. https://doi.org/10.1016/j.ympev.2017.03.004
  • Wu X, Xu X, Yu Z, Kong X (2009) Comparative mitogenomic analyses of three scallops (Bivalvia: Pectinidae) reveal high level variation of genomic organization and a diversity of transfer RNA gene sets. BMC Research Notes 2: 69–69. https://doi.org/10.1186/1756-0500-2-69
  • Wu X, Li X, Li L, Xu X, Xia J, Yu Z (2012) New features of Asian Crassostrea oyster mitochondrial genomes: A novel alloacceptor tRNA gene recruitment and two novel ORFs. Gene 507(2): 112–118. https://doi.org/10.1016/j.gene.2012.07.032
  • Wu R, Liu X, Wang S, Roe KJ, Ouyang S, Wu X (2019) Analysis of mitochondrial genomes resolves the phylogenetic position of Chinese freshwater mussels (Bivalvia, Unionidae). ZooKeys 812: 23–46. https://doi.org/10.3897/zookeys.812.29908
  • Yokobori S, Fukuda N, Nakamura M, Aoyama T (2004) Long-term conservation of six duplicated structural genes in cephalopod mitochondrial genomes. Molecular Biology and Evolution 21(11): 2034–2046. https://doi.org/10.1093/molbev/msh227
  • Yu L, Wang X, Ting N, Zhang Y (2011) Mitogenomic analysis of Chinese snub-nosed monkeys: evidence of positive selection in NADH dehydrogenase genes in high-altitude adaptation. Mitochondrion 11(3): 497–503. https://doi.org/10.1016/j.mito.2011.01.004
  • Yu Z, Li Q, Kong L (2016) New insight into the phylogeny of Sinonovacula (Bivalvia: Solecurtidae) revealed by comprehensive DNA barcoding analyses of two mitochondrial genes. Mitochondrial DNA Part A 27(2): 1554–1557. https://doi.org/10.3109/19401736.2014.953135
  • Yuan Y, Li Q, Kong L, Yu H (2012b) The complete mitochondrial genome of the grand jackknife clam, Solen grandis (Bivalvia: Solenidae): a novel gene order and unusual non-coding region. Molecular Biology Reports 39(2): 1287–1292. https://doi.org/10.1007/s11033-011-0861-8
  • Yuan Y, Li Q, Yu H, Kong L (2012c) The complete mitochondrial genomes of six heterodont bivalves (Tellinoidea and Solenoidea): variable gene arrangements and phylogenetic implications. PLoS ONE 7(2): 1–12. https://doi.org/10.1371/journal.pone.0032353
  • Yuan Y, Li Q, Yu H, Kong L (2012d) The complete mitochondrial genomes of six Heterodont bivalves (Tellinoidea and Solenoidea): variable gene arrangements and phylogenetic implications. PLoS ONE 7(2): 1–9. https://doi.org/10.1371/journal.pone.0032353
  • Zhang D, Gao F, Jakovlic I, Zou H, Zhang J, Li WX, Wang GT (2020) PhyloSuite: an integrated and scalable desktop platform for streamlined molecular sequence data management and evolutionary phylogenetics studies. Molecular Ecology Resources 20(1): 348–355. https://doi.org/10.1111/1755-0998.13096
  • Zhao B, Gao S, Zhao M, Lv H, Song J, Wang H, Zeng Q, Liu J (2022) Mitochondrial genomic analyses provide new insights into the “missing” atp8 and adaptive evolution of Mytilidae. BMC Genomics 23(1): 738–752. https://doi.org/10.1186/s12864-022-08940-8
  • Zheng R, Li J, Niu D (2010) The complete DNA sequence of the mitochondrial genome of Sinonovacula constricta (Bivalvia: Solecurtidae). Acta Oceanologica Sinica 29(2): 88–92. https://doi.org/10.1007/s13131-010-0026-y
login to comment