Next Article in Journal
The MKK-Dependent Phosphorylation of p38α Is Augmented by Arginine Methylation on Arg49/Arg149 during Erythroid Differentiation
Previous Article in Journal
The Role of Lipid Metabolism in COVID-19 Virus Infection and as a Drug Target
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Complete Mitochondrial Genomes of Two Toxin-Accumulated Nassariids (Neogastropoda: Nassariidae: Nassarius) and Their Implication for Phylogeny

1
Key Laboratory of Mariculture, Ministry of Education, Ocean University of China, Qingdao 266003, China
2
Laboratory for Marine Fisheries Science and Food Production Processes, Qingdao National Laboratory for Marine Science and Technology, 1 Wenhai Road, Aoshanwei Town, Qingdao 266237, China
*
Author to whom correspondence should be addressed.
Int. J. Mol. Sci. 2020, 21(10), 3545; https://doi.org/10.3390/ijms21103545
Submission received: 30 March 2020 / Revised: 8 May 2020 / Accepted: 12 May 2020 / Published: 17 May 2020
(This article belongs to the Section Molecular Genetics and Genomics)

Abstract

:
The Indo-Pacific nassariids (genus Nassarius) possesses the highest diversity within the family Nassariidae. However, the previous shell or radula-based classification of Nassarius is quite confusing due to the homoplasy of certain morphological characteristics. The toxin accumulators Nassarius glans and Nassarius siquijorensis are widely distributed in the subtidal regions of the Indo-Pacific Ocean. In spite of their biological significance, the phylogenetic positions of N. glans and N. siquijorensis are still undetermined. In the present study, the complete mitochondrial genomes of N. glans and N. siquijorensis were sequenced. The present mitochondrial genomes were 15,296 and 15,337 bp in length, respectively, showing negative AT skews and positive GC skews as well as a bias of AT rich on the heavy strand. They contained 13 protein coding genes, 22 transfer RNA genes, two ribosomal RNA genes, and several noncoding regions, and their gene order was identical to most caenogastropods. Based on the nucleotide sequences combining 13 protein coding genes and two rRNA genes, a well-supported phylogeny of Nassarius was reconstructed, and several morphological synapomorphies were observed corresponding to the phylogenetic framework. In addition, the sister group relationship between N. variciferus and the remaining toxin-accumulated nassariids was determined, suggesting that the phylogeny might be related to their diet. The divergence time estimation analysis revealed a correlation between speciation events of nassariids and glacial cycles during the Pliocene-Pleistocene epoch.

1. Introduction

Containing more than 600 extant species [1], the worldwide distributed mudsnail nassariids (Family Nassariidae Iredale, 1916) are an ecologically important group within Neogastropoda. As scavengers, nassariids are important in maintaining the balance of benthic ecosystem. Some nassariids are also useful for monitoring Tributyltin (TBT) pollution in marine environment [2]. Within Nassariidae, the cosmopolitan subfamily Nassariinae consists of five genera that correspond to five geographical clades. One of the five genera, called Nassarius (type genus of Nassariidae), is endemic to the Indo-Pacific Ocean [3] and possesses the highest diversity (about 300 species as valid [4]) within Nassariidae.
Traditional classification of Nassarius was mainly based on morphological characteristics [5]. However, the shell morphology-based taxonomy was revealed to be subjective and often contradicted molecular phylogenies [3,6,7]. On the other hand, previous Nassarius phylogenies, which were mainly based on short gene fragments (especially COI and/or 16S genes [8,9]), were poorly supported and often contradicted each other, suggesting the inadequate capacity of partial mitochondrial gene data for resolving the phylogeny of Nassarius. Using complete mitochondrial (mt) genomic sequences, Yang et al. [6] conducted the phylogeny of 10 common nassariids along the China Seas. The result revealed the advantages of utilizing complete mt genomes to address phylogenetic relationships within Nassariidae, as indicated in other caenogastropod groups [10,11,12,13]. Furthermore, the study called for the necessity of including more taxa into the reconstruction of the Nassarius phylogeny, which could provide a framework to understand the evolution of morphological and ecological traits.
In Asia, nassariids are consumed as seafood [7,14]. However, due to their ability to accumulate toxins (e.g., algal toxin or tetrodotoxin) [15,16], several poisoning incidents caused by the consumption of certain nassariids have been reported occasionally in some Asian countries [17,18,19]. Recent studies found that the toxicity varied in species [20,21]. For example, most marine nassariids are toxin accumulators in addition to Nassarius variciferus and Reticunassa festiva which could not accumulate toxins. Among those toxin accumulators, N. siquijorensis and N. glans are widely distributed in the subtidal regions of the Indo-Pacific Ocean. The ecology and toxicology of N. siquijorensis and N. glans have been the subjects of numerous studies [20,22]. In spite of the biological significance, the classification and phylogenetic positions of N. siquijorensis and N. glans within Nassarius are still controversial. For example, the sister group of N. siquijorensis and N. nodifer revealed by Chen and Zhang [8] was not recovered by Pu et al. [23], although both studies were poorly supported.
In the present study, the complete mitochondrial genomes of N. siquijorensis and N. glans were sequenced and analyzed together with those of other nassariids publishes before (Table 1). Our aims were (1) to confirm the phylogenetic positions of N. siquijorensis and N. glans within Nassarius; (2) to test the contribution of different data sets of mitochondrial genomes for assessing phylogenetic relationships in Nassarius; and (3) to date major cladogenetic events within Nassarius.

2. Results and Discussion

2.1. Genome Structure, Organization, and Composition

The characteristics of the two mt genomes in the present study were similar to those of the other nassariids reported before [6,24,25,26] in terms of genome organization and nucleotide composition (Figure 1). The mt genomes of N. siquijorensis and N. glans are 15,337 and 15,296 bp in length, respectively. Both of them encode for 13 protein-coding genes (PCGs), 22 transfer RNA (tRNA) genes. and two ribosomal RNA (rRNA) genes, with 8 tRNA genes encoding in the minor strand while the others encode in the major strand (Table 2).
The AT contents and skew statistics are shown in Table 3, indicating a high A + T bias. The result also showed that the nucleotide compositions were skewed from A towards T but insignificantly skewed from C towards G. This strand asymmetry has been observed in other gastropod taxa, such as Vetigastropoda [27], Heterostropha [28], Neritimorpha [29], and Patellogastropoda [30], where T was more than A and C was less than G on the heavy strands. In addition, the heavy strands of mammalian mt genomes were also found to be GT rich [31] whereas both strands of arthropod mt genomes were AT rich [32,33]. The mechanism of asymmetry of the heavy strand was attributed to the hydrolytic deaminations of adenine and cytosine (leading to the mutations from adenine to guanine and from cytosine to thymine) during replication and transcription as well as during transcription but to a lesser extent [32].

2.2. PCGs, rRNA, and tRNA Genes

The AT content, AT skew, and GC skew of the PCGs were also similar in the mt genomes of N. siquijorensis and N. glans (Table 3). In the PCGs, all AT skews were negative while most GC skews were positive. It was notable that a significantly negative GC skew was found in the nad5 genes of both species. This reversed bias might be attributed to the effect of the control region [32], which was located between tRNA-Phe and cox3 genes. The average A + T content values of PCGs were 69.73% and 69.78% in N. glans and N. siquijorensis, respectively, and those proportions at the first, second, and third codon positions were 61.28% and 61.17%, 62.88% and 62.85%, and 85.03% and 85.33%, respectively, indicating a strong A + T bias at the third codon position as reported in other invertebrate mt genomes (e.g., Cnidaria [34] and Arthropoda [35]). Almost all PCGs of the 2 mt genomes started with the conventional initiation codon ATG, except for the nad4 of N. glans which began with ATA (Table 2).
As for the termination codons, all PCGs ended with the complete stop codons (TAA, n = 17; TAG, n = 9). Codon usage of PCGs was presented in Table 4. Both 2 mt genomes had 3729 PCG codons (excluding the stop codons), among which the most frequently used one was UUA (leucine) while the least chosen codon was CGC (arginine) (Table 4 and Figure 2). On the other hand, the most encoded amino acid was leucine while the least encoded one was cysteine. Considerable synonymous codon usage bias was observed in both mt genomes, with a total number of 20 codons being used more frequently than others (Table 4). Those preferred codons were detected all ending in A or U and therefore resulted in a strong A + T bias at the third codon position. Previous studies revealed that the synonymous codon usage bias might be caused by mutational bias alone or by both mutation bias and natural selection [36]. The four most used codons UUA (leucine), AUU (isoleucine), UUU (phenylalanine) and AUA (Methionine) observed here (Figure 2 and Table 4) also fit with some other metazoan taxa, such as in Annelida [37] and Nematoda [38].
The lengths of the tRNA genes between N. glans and N. siquijorensis were almost identical, ranging from 57 (trnS-UCN of N. siquijorensis) to 71 bp (trnL-CUN of N. glans). All of the tRNA genes could be folded into typical clover-leaf secondary structures except for the trnS-AGN in both mt genomes due to the missing of dihydrouracil (DHU) arms (Figure 3). The lack of DHU arm in trnS-AGN was quite common in metazoan mt genomes [39]. The average AT contents of tRNA genes were 68.90% and 68.73% in N. glans and N. siquijorensis, respectively, and the tRNA genes were neither substantially AT nor GC skewed (Table 3).
The 12S rRNA genes of N. glans and N. siquijorensis were 964 and 962 bp in length, with AT contents of 68.80% and 67.64%, respectively, while the 16S rRNA genes were 1361 and 1359 bp, with AT contents of 73.82% and 72.84% in N. glans and N. siquijorensis, respectively. Unlike the tRNA genes, the rRNA genes showed a weakly positive AT skew and strongly positive GC skew (Table 3).

2.3. Phylogenetic Relationship

Phylogenetic analysis was conducted using two data sets. Based on the Bayesian information criterion (BIC) [40], the best partition scheme for PCGs at nucleotide level was the one combining all these genes but analyzing the three codons separately. The best substitution models were GTR + I + G for all three codon positions. At the amino acid level, the best partition scheme for PCGs was the one combining genes by subunits. The best substitution models were MTART + I for atp, cob, and cox, and MTMAM + I + G + F for nad. For the rRNA genes, the best partition scheme was combining 12S and 16S rRNA genes together, with GTR + I + G as the best substitution model.
The final matrix of the first data set (nucleotide sequences of 13 PCGs and rRNA genes) was 13,277 bp in length. Both maximum likelihood (ML) [41] and Bayesian inference (BI) [42] arrived at almost identical topologies (Figure 4 and Figure S1), and most nodes received high support values, except for one internal node that connected N. foveolatus and N. javanus + (N. sinarus + N. succinctus). The second data set (amino acid sequences of PCGs plus nucleotide sequences of rRNA genes) was 5821 positions in length. However, the phylogenetic analysis based on the second data set was not well resolved due either to the relatively lower statistical support values or to the contradictory topologies reconstructed using ML and BI methods (Figure S1). Compared with the first data set, the second one had fewer genetic sites and the amino acid sequences appeared to be more conserved than the nucleotide sequences, as indicated by the short branch lengths of the phylogenetic trees (Figure S2). As a result, the present study shows that the combining the nucleotide sequences of 13 PCGs and two rRNA genes could provide enough variances for the divergence within genus Nassarius.
Except for N. variciferus, N. foveolatus, and N. sinarus, all the remaining species analyzed in the present study were also included in the phylogeny conducted by Galindo et al. [3] (Note that N. nodifer here is synonymized as N. hepaticus in Galindo et al. [3]). Compared with the most comprehensive phylogenetic study on Nassariidae published thus far [3], the present phylogeny showed different topologies. In the present tree, N. siquijorensis was revealed as the sister group of N. nodifer + N. conoidalis while N. glans clustered with N. foveolatus + (N. javanus + (N. succinctus + N. sinarus)). On contrast, N. glans showed a closer relationship with N. succinctus while N. siquijorensis was more related to N. conoidalis + (N. javanus + N. nodifer) in the nassariid tree of Galindo et al. [3]. Compared with previous phylogenies which were not well supported, the present study could improve the resolution of phylogenetic relationship of these targeted nassariids. However, it is also shown in the tree (Figure 4) that some internal branches connecting the nodes with relatively lower statistical support are rather short, which indicates a rapid radiation [43]. Consequently, achieving high statistical support of these phylogenetic relationships will be challenging [44]. In the future, it is necessary to sequence more mt genomes of taxa within this clade as well as to develop new molecular markers (e.g., exon-capture based on transcriptomic data [45,46]) in order to resolve the rapid radiation.
In the present phylogeny (Figure 4), N. variciferus was the first species branching off. The distant phylogenetic relationship between N. variciferus and other nassariids has also been revealed in previous studies [6,7]. Based on the morphology of shell characteristic, N. variciferus was previously classified into subgenus Varicinassa, which was defined as a single-species group with prominent angulate varices on whorls [5]. However, the presence of varices appears to be a morphological phenotype because individuals of N. variciferus without varices have also been found [7]. According to previous records, almost all Nassarius species were toxin accumulators except for N. variciferus (Table 1). It is inferred that nassariid toxins, which were recognised as algal toxin or tetrodotoxin, could not be produced by nassariids themselves, but might be from the environment (probably via the food chains [47]). Therefore, the differences on toxicity may reflect different food preferences. However, further studies concerning the components and sources of nassariid toxins as well as the anatomical characteristics of nassariid digestive systems are needed to support the present hypothesis.
In addition to N. variciferus, the remaining nassariids form into three lineages, including the single-species N. pullus; the clade grouped by N. nodifer, N. conoidalis, and N. siquijorensis; and the group formed by all the others (Figure 4). The distant phylogenetic relationship between N. pullus and other nassariids is also supported by the morphological characteristics. The most significant trait of N. pullus is the spreading columellar callus which was used to define subgenus Plicarcularia [5]. However, it is insufficient to determine whether this characteristic could unite other members in phylogeny since only one species of Plicarcularia is included here. In the second lineage, one unique trait that connects N. nodifer, N. conoidalis, and N. siquijorensis has also been observed in the present study. Morphologically, all these three species possess strong nodules formed by the axial and spiral ribs. In previous classification [5], most members in subgenus Niotha have these morphological traits. Even though N. siquijorensis possesses this characteristic, it was incorrectly classified into subgenus Zeuxis, which was a not particularly well-defined group, and its value from a taxonomic viewpoint is doubtful. Based on the present phylogeny, N. siquijorensis should be reassigned from Zeuxis to Niotha, and the existence of nodules on shell surface should be considered as one of the most important synapomorphies (Figure 5) to define this group. According to a previous study [23], a total of 23 individuals of N. siquijorensis collected in five sites along Guangdong Province, China were recovered to form two different clades in the cox1-based phylogeny, indicating that N. siquijorensis possessed cryptic species. Geographically, the sampling site of N. siquijorensis in our study did not correspond to any of the five sites in Pu et al. [23]. Although the phylogenetic position of N. siquijorensis was determined here, the circumscription of N. siquijorensis still remains unclear. Further studies with a high sampling coverage along its distribution range are needed to determine its diversity and to define the status of the N. siquijorensis complex. In the third clade, the other five species were previously assigned into three subgenera (Zeuxis, Telaseo, and Alectrion). Nevertheless, no significant morphological characteristic was found corresponding to the current phylogeny due to the limited species number. Consequently, the present study calls for the inclusion of more mitogenomic data into the reconstruction of Nassarius phylogeny, which is useful for the reestablishment of Nassarius taxonomy.

2.4. Divergence Times

Major cladogenetic events within Nassarius were dated using an uncorrelated relaxed molecular clock model, which was calibrated using two fossils. The first event of diversification within the crown group of Nassarius was estimated at a mean of 10.6 (13.7–8.2, credible interval) million years ago (Mya), separating N. variciferus and the remaining nassariids which have the ability to accumulate toxins. The branching of N. pullus and the other two major lineages was estimated to have occurred at 6.3 (8.2–4.8) Mya. Finally, the speciation events leading to the other extant Nassarius were estimated to have occurred between 4.3–2.8 (5.3–2.2) Mya (Figure 6). However, the estimated divergence times of previous studies were much earlier. For example, the divergence time between R. festiva A and R. festiva B was estimated to 5.05 Mya [6]; the speciation between N. hepaticus (synonymized as N. nodifer) and N. conoidalis was dated to about 12 Mya [3]. These results (earlier estimated divergence times) might be attributed to the employment of older fossils, such as using the fossil time of oldest Nassarius from Europe (28–23 Mya) as the divergence between Nassarius and Tritia [3,6]. It is shown here that most speciation events happened during Pliocene and Pleistocene, corresponding with the fossil records of most extant Nassarius in the Paleobiology Database [48]. In addition, previous studies have revealed that R. festiva A and R. festiva B were closely related species [26] and that their diversification time might date back to Pleistocene [49]. The estimated divergence time between R. festiva A and R. festiva B (1.42 Mya) shown here also falls into Pleistocene, supporting the hypothesis that the speciation might be caused by the third cooling event during the early Pleistocene (2.2–1.0 Mya) [50]. Furthermore, the divergence time between Nassarius and Tritia revealed here (12.15 Mya) is also supported by the fossil time of the oldest Tritia (about 13 Mya) recorded in the Paleobiology Database [48].The ancestral character state reconstruction analysis (Figure 5) suggests that Nassarius may have originated from an ancestor without toxin-accumulating ability. The divergence time between N. variciferus and the remaining toxin accumulators is dated to Miocene, corresponding to the period when nassariids diversified dramatically [51]. The divergence time of N. pullus also falls in this epoch. In the last 5.3 Mya, there were three major cooling periods discovered in the Western Pacific: the early Pliocene (4.7–3.5 Mya), the late Pliocene (3.1–2.7 Mya), and the early Pleistocene (2.2–1.0 Mya) [50]. The divergence times of the remaining nassariids broadly fall in the three cooling periods, suggesting a correlation between speciation events and glacial cycles. The rapid radiation of certain nassariids may be attributed to the first cooling event.

3. Materials and Method

3.1. Samples and DNA Extraction

Specimens of N. siquijorensis and N. glans were collected in the same local market in Wenchang, Hainan Province, China (19°32′6″N; 110°49′13″E). Samples were then deposited in 95% ethanol.
Genomic DNA was extracted from small pieces of foot tissue using TIANamp Marine Animals DNA Kit (TIANGEN Biotech Beijing Co. Ltd.), according to the instructions, and visualized on 1.0% agarose gel. Only one specimen of each species was used for DNA extraction. The DNA concentration was measured on the Nanodrop 2000.

3.2. DNA Sequencing and Genome Assembly

Genomic DNA of N. siquijorensis and N. glans were submitted to Novogene Company (Beijing, China) for library construction and high-throughput sequencing. Two sequencing libraries with average insert sizes of approximately 300 bp were prepared and then sequenced as 150 bp paired-end runs on the Illumina HiSeq 2000 platform. Finally, about 8 Gb of raw data were generated for each library.
After removing the contaminated reads, low-quality sequences, and adapters (following Reference [52]), the clean reads of 2 species were de novo assembled separately using SPAdes 3.12.0 [53].

3.3. Gene Annotation and Sequence Analysis

The newly determined mt genomes were annotated with Geneious Prime 2019.0.3 [54], using the previously published Nassarius mt genomes [6,26] as references. Annotations of 13 PCGs were defined by setting a limit of 75% nucleotide identity to previously published nassariid mt genomes (Table 1) in Geneious Prime 2019.0.3 and corroborated using MITOS Webserver [55] by setting the Invertebrate Mitochondrial genetic code. The secondary structure of tRNA genes was generated by MITOS Webserver and modified using Microsoft Visio 2010 according to the tRNA structure identified by tRNA scan-SE1.21 [56] and ARWEN [57]. The rRNA genes were identified by comparing with other nassariid mt genomes, and their boundaries were assumed to be between the adjacent genes.
Codon usage of PCGs; the A + T content values; and nucleotide composition of the mt genomes, PCGs, rRNA, and tRNA genes were estimated using MEGA 7 [58]. The base skew values for a given strand was calculated as AT skew = (A − T)/(A + T) and GC skew = (G − C)/(G + C) [59]. The mt gene map was generated using CGView [60].
A total of 19 taxa was included in the phylogenetic analysis. Seven species, including five Reticunassa, “Nassariusacuticostus, and “Nassarius” jacksonianus, were selected as outgroup taxa [6]. The 13 PCGs were aligned separately using Translator X [61], according to the Invertebrate Mitochondrial genetic code. Nucleotide sequences of the rRNA genes were aligned separately using MAFFT v7 [62] with default parameters. Ambiguously aligned positions were removed using Gblocks v.0.91b [63]. Finally, the different single alignments were concatenated into the two data sets in Geneious Prime 2019.0.3. Two data sets were constructed: 1. the nucleotide sequences of 13 PCGs and 2 rRNA genes and 2. the amino acid sequences of 13 PCGs and the nucleotide sequences of 2 rRNA genes.

3.4. Phylogenetic Analysis

Phylogenetic analyses were inferred using ML and BI. ML analyses were carried out using software RAxML v. 8.2.1 [64] with the rapid hill-climbing algorithm and 10,000 bootstrap pseudoreplicates (BP). BI analyses were performed with MrBayes v.3.1.2 [65], running four simultaneous Monte Carlo Markov chains (MCMC) for 10,000,000 generations, sampling every 1000 generations, and discarding the first 25% generations as burn-in. Parameter convergence was achieved within 10 million generations, and the standard deviation of split frequencies was less than 0.01. All parameters were checked with Tracer v. 1.6 [66], and the effective sample size (ESS) was more than 200. The resulting phylogenetic trees were visualized in FigTree v1.4.2 [67]. BI analyses based on nucleotide sequences of PCGs only with site-heterogeneous mixture models were conducted in PhyloBayes MPI v.1.7b [68], using the CAT-GTR model and discarding constant sites (“-dc” option) on the CIPRES webserver [69]. Parameters to convergence were set according to Uribe et al. [70].
The best partition schemes and best-fit substitution models for the two data sets were determined by PartitionFinder 2 [71] under the BIC. For the PCGs analyzed at both nucleotide and amino acid levels, the partitions tested were all genes combined; all genes separated (except atp6-atp8 and nad4-nad4L); and genes grouped by subunits (atp, cob, cox, and nad). At the nucleotide level, these three partition schemes were tested considering separately the three codon positions. The rRNA genes were analysed with two different schemes (genes grouped or separated).

3.5. Divergence Time Estimation

The divergence times among Nassarius species were estimated based on the protein-coding genes (at nucleotide level) only, using an uncorrelated, lognormal relaxed molecular clock model in BEAST v.1.10.4 [72]. For the tree prior, a Yule process of speciation was employed. The partitions selected by PartitionFinder 2 (see above) were applied. The final Markov chain was run twice for 100 million generations, sampling every 10,000 generations, and the first 10% samples was discarded as burn-in according to the convergence of chains checked with Tracer v. 1.6. The effective sample size of all the parameters was above 200.
The posterior distribution of the estimated divergence times was obtained by specifying two calibration points as priors for divergence times of the corresponding splits. The first calibration point was set for the divergence of Reticunassa (16.0 Mya, [73]), with a minimum of 16 Mya and a 95% upper limit of 17.5 Mya (exponential distribution, offset: 16.0; mean: 0.5). A second calibration point was set based on the fossil time of N. dorsatus (3.6–2.6 Mya, [74]) that likely corresponded to the divergence time between N. javanus and N. succinctus + N. sinarus [75].The minimum and 95% upper limits were set to 2.6 and 3.6 Mya, respectively (lognormal distribution, offset: 2.6; mean: 0.4; standard deviation: 0.4).

3.6. Ancestral Character State Reconstruction

The evolution of three traits (Figure 5) was analyzed by reconstruction of the ancestral character states. These characteristics were assessed with Mesquite v3.6.1 [76] using Tracing Character History option and ML approach and mapped over the phylogenetic tree generated by ML and BI analyses.

4. Conclusions

The complete mitochondrial genomes of N. glans and N. siquijorensis were sequenced in the present study. Containing 13 PCGs, 22 tRNA genes, two rRNA genes, and several noncoding regions, the two mt genomes showed a similar pattern with respect to genome size, gene order, and nucleotide composition compared with those of other nassariids reported before. The data set combining nucleotide sequences of 13 PCGs and two rRNA genes was revealed to be most useful for phylogenetic reconstruction of Nassarius. The phylogenetic positions of N. glans and N. siquijorensis were well determined, and several morphological synapomorphies were observed corresponding to the present phylogeny. The present study also proposed a hypothesis that the phylogeny of Nassarius was related to their diet, based on the sister group relationship between the N. variciferus and the remaining toxin accumulators. A correlation between nassariid speciation events and Pliocene-Pleistocene glacial cycles was presumed. Our results indicate that complete mt genomes would be a promising tool to reconstruct a robust phylogeny of Nassarius with the inclusion of more taxa in the future. This approach could be complemented with the development of nuclear markers, which could be useful for eliminating the influence of rapid radiations.

Supplementary Materials

Supplementary materials can be found at https://www.mdpi.com/1422-0067/21/10/3545/s1. Table S1: List of nucleotide composition of N. glans (Ngla) and N. siquijorensis (Nsiq), with respect to whole genome, PCGs (considering three different codon positions), tRNA and rRNA. Figure S1: Phylogenetic relationship of Nassarius based on amino acid sequences of 13 mitochondrial PCGs and nucleotide sequences of 2 rRNA genes. Figure S2: Phylogenetic relationship of Nassarius based on the nucleotide sequences of 13 PCGs only using PhyloBayes program.

Author Contributions

Y.Y. performed the experiments, analyzed the data, prepared the figures and tables, and wrote the paper. H.L. performed the experiments and analyzed the data. L.Q. performed the experiments and analyzed the data. L.K. collected the samples and polished the paper. Q.L. conceived and designed the experiments, supervised the work, and polished the paper. All authors have read and agreed to the published version of the manuscript.

Funding

This work was supported by the National Natural Science Foundation of China (31772414) and Research Project of the Ocean University of China-Auburn University Joint Research Center for Aquaculture and Environmental Science, and Shandong Province (2017LZGC009).

Acknowledgments

We are grateful to Ke Ran and Yaping Lin at Fisheries College, Ocean University of China, for their assistance with collection of nassariid samples in the field. We also appreciate the three anonymous reviewers for their careful work and valuable suggestions that have helped improve this paper substantially.

Conflicts of Interest

The authors declare no conflict of interest. The funders had no role in the design of the study; in the collection, analyses, or interpretation of data; in the writing of the manuscript; or in the decision to publish the results.

References

  1. MolluscaBase. Nassariidae Iredale, 1916 (1835). World Register of Marine Species. Available online: http://www.marinespecies.org/aphia.php?p=taxdetails&id=151 (accessed on 21 January 2020).
  2. Bryan, G.W.; Burt, G.R.; Gibbs, P.E.; Pascoe, P.L. Nassarius reticulatus (Nassariidae: Gastropoda) as an indicator of tributyltin pollution before and after TBT restrictions. J. Mar. Biolog. Assoc. U.K. 1993, 73, 913–929. [Google Scholar] [CrossRef]
  3. Galindo, L.A.; Puillandre, N.; Utge, J.; Lozouet, P.; Bouchet, P. The phylogeny and systematics of the Nassariidae revisited (Gastropoda, Buccinoidea). Mol. Phylogenet. Evol. 2016, 99, 337–353. [Google Scholar] [CrossRef] [PubMed]
  4. MolluscaBase. Nassarius Duméril, 1805. World Register of Marine Species. Available online: http://www.marinespecies.org/aphia.php?p=taxdetails&id=138235 (accessed on 1 May 2020).
  5. Cernohorsky, W.O. Systematics of the family Nassariidae (Mollusca:Gastropoda). Bull. Auckl. Inst. Mus. 1984, 14, 1–356. [Google Scholar]
  6. Yang, Y.; Li, Q.; Kong, L.; Yu, H. 2019. Mitogenomic phylogeny of Nassarius (Gastropoda: Neogastropoda). Zool. Scr. 2019, 48, 302–312. [Google Scholar] [CrossRef]
  7. Zou, S.M.; Li, Q.; Kong, L.F. Monophyly, distance and character-based multigene barcoding reveal extraordinary cryptic diversity in Nassarius: a complex and dangerous community. PLoS ONE 2012, 7, e47276. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  8. Chen, Z.Y.; Zhang, S.P. Molecular phylogeny of Nassarius (Gastropoda, Nassariidae) based on 16s rRNA gene sequences. Acta Zootaxonomica Sin. 2012, 37, 467–472. [Google Scholar]
  9. Li, H.; Lin, D.; Fang, H.; Zhu, A.; Gao, Y. Species identification and phylogenetic analysis of genus Nassarius (Nassariidae) based on mitochondrial genes. Chin. J. Oceanol. Limnol. 2010, 28, 565–572. [Google Scholar] [CrossRef]
  10. 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]
  11. Harasewych, M.G.; Sei, M.; Wirshing, H.H.; González, V.L.; Uribe, J.E. The complete mitochondrial genome of Neptuneopsis gilchristi G.B. Sowerby III, 1898 (Neogastropoda: Volutidae: Calliotectinae). Nautilus 2019, 133, 67–73. [Google Scholar]
  12. Uribe, J.E.; Zardoya, R.; Puillandre, N. Phylogenetic relationships of the conoidean snails (Gastropoda: Caenogastropoda) based on mitochondrial genomes. Mol. Phylogenet. Evol. 2018, 127, 898–906. [Google Scholar] [CrossRef]
  13. Vaux, F.; Hills, S.F.; Marshall, B.A.; Trewick, S.A.; Morgan-Richards, M. A phylogeny of Southern Hemisphere whelks (Gastropoda: Buccinulidae) and concordance with the fossil record. Mol. Phylogenet. Evol. 2017, 114, 367–381. [Google Scholar] [CrossRef] [PubMed]
  14. Yang, Y.; Li, Q.; Kong, L. Identification of Six Nassarid Snails Using COI-Based Restriction Fragment Length Polymorphism. J. Shellfish Res. 2018, 37, 239–244. [Google Scholar] [CrossRef]
  15. Huang, H.N.; Lin, J.; Lin, H.L. 2008. Identification and quantification of tetrodotoxin in the marine gastropod Nassarius by LC-MS. Toxicon 2008, 51, 774–779. [Google Scholar] [CrossRef]
  16. Lin, Y.; Jia, X.; Yang, M.; Zhong, Y.; Quan, G.; Fan, P. Paralytic shellfish poison in contaminated shellfish along coast of China. Trophic. Oceanol. 1999, 18, 90–96. [Google Scholar]
  17. Dao, H.V.; Pham, K.X.; Hoang, B.X.; Tanioka, M.; Watanabe, R.; Suzuki, T. Occurrence of tetrodotoxin in three Nassarius gastropod species in Khanh Hoa Province, Vietnam. Fish. Sci. 2020, 86, 181–186. [Google Scholar] [CrossRef]
  18. Qin, P.Z.; Yu, M. Research on Nassidae food poisoning in Ningbo City from 1985 to 2000. Chin. J. Prev. Med. 2003, 37, 226. [Google Scholar]
  19. Shui, L.M.; Chen, K.; Wang, J.Y.; Mei, H.Z.; Wang, A.Z.; Lu, Y.H.; Hwang, D.F. Tetrodotoxin-associated snail poisoning in Zhoushan: a 25-year retrospective analysis. J. Food Protect. 2003, 66, 110–114. [Google Scholar] [CrossRef]
  20. Hwang, P.A.; Tsai, Y.H.; Deng, J.F.; Cheng, C.A.; Ho, P.H.; Hwang, D.F. Identification of tetrodotoxin in a marine gastropod (Nassarius glans) responsible for human morbidity and mortality in Taiwan. J. Food Protect. 2005, 68, 1696–1701. [Google Scholar] [CrossRef]
  21. Zhang, N.; Su, J.; Liu, H.; Ye, S.; Li, L.; Cai, L.Z. The species and toxicities of Nassariidae collected from the coast of Southeast China Sea. Asian J. Ecotoxicol. 2009, 4, 289–294. [Google Scholar]
  22. Morton, B.; Chan, K. Hunger rapidly overrides the risk of predation in the subtidal scavenger Nassarius siquijorensis (Gastropoda: Nassariidae): an energy budget and a comparison with the intertidal Nassarius festivus in Hong Kong. J. Exp. Mar. Biol. Ecol. 1999, 240, 213–228. [Google Scholar] [CrossRef]
  23. Pu, C.; Li, H.; Zhu, A.; Chen, Y.; Zhao, Y.; Zhan, A. Phylogeography in Nassarius mud snails: Complex patterns in congeneric species. PloS ONE 2017, 12, e0180728. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  24. Simison, W.B.; Lindberg, D.R.; Boore, J.L. Rolling circle amplification of metazoan mitochondrial genomes. Mol. Phylogenet. Evol. 2006, 39, 562–567. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  25. Cunha, R.L.; Grande, C.; Zardoya, R. Neogastropod phylogenetic relationships based on entire mitochondrial genomes. BMC Evol. Biol. 2009, 9, 210. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  26. Yang, Y.; Li, Q.; Kong, L.; Yu, H. Comparative mitogenomic analysis reveals cryptic species in Reticunassa festiva (Neogastropoda: Nassariidae). Gene. 2018, 662, 88–96. [Google Scholar] [CrossRef] [PubMed]
  27. Williams, S.T.; Foster, P.G.; Littlewood, D.T.J. 2014. The complete mitochondrial genome of a turbinid vetigastropod from MiSeq Illumina sequencing of genomic DNA and steps towards a resolved gastropod phylogeny. Gene. 2014, 533, 38–47. [Google Scholar] [CrossRef]
  28. Grande, C.; Templado, J.; Zardoya, R. Evolution of gastropod mitochondrial genome arrangements. BMC Evol. Biol. 2008, 8, 61. [Google Scholar] [CrossRef] [Green Version]
  29. Arquez, M.; Colgan, D.; Castro, L.R. Sequence and comparison of mitochondrial genomes in the genus Nerita (Gastropoda: Neritimorpha: Neritidae) and phylogenetic considerations among gastropods. Mar. Genom. 2014, 15, 45–54. [Google Scholar] [CrossRef]
  30. Gaitán-Espitia, J.D.; González-Wevar, C.A.; Poulin, E.; Cardenas, L. Antarctic and sub-Antarctic Nacella limpets reveal novel evolutionary characteristics of mitochondrial genomes in Patellogastropoda. Mol. Phylogenet. Evol. 2019, 131, 1–7. [Google Scholar] [CrossRef]
  31. Reyes, A.; Gissi, C.; Pesole, G.; Saccone, C. Asymmetrical directional mutation pressure in the mitochondrial genome of mammals. Mol. Biol. Evol. 1998, 15, 957–966. [Google Scholar] [CrossRef] [Green Version]
  32. Bernt, M.; Braband, A.; Schierwater, B.; Stadler, P.F. Genetic aspects of mitochondrial genome evolution. Mol. Phylogenet. Evol. 2013, 69, 328–338. [Google Scholar] [CrossRef] [Green Version]
  33. Crease, T.J. The complete sequence of the mitochondrial genome of Daphnia pulex (Cladocera: Crustacea). Gene 1999, 233, 89–99. [Google Scholar] [CrossRef]
  34. Lisenkova, A.A.; Grigorenko, A.P.; Tyazhelova, T.V.; Andreeva, T.V.; Gusev, F.E.; Manakhov, A.D.; Goltsov, A.Y.; Piraino, S.; Miglietta, M.P.; Rogaev, E.I. Complete mitochondrial genome and evolutionary analysis of Turritopsis dohrnii, the “immortal” jellyfish with a reversible life-cycle. Mol. Phylogenet. Evol. 2017, 107, 232–238. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  35. Hu, P.; Wang, R. The complete mitochondrial genome of Parantica sita sita (Lepidoptera: Nymphalidae: Danainae) revealing substantial genetic divergence from its sibling subspecies P. s. niphonica. Gene 2019, 686, 76–84. [Google Scholar] [CrossRef]
  36. Wei, L.; He, J.; Jia, X.; Qi, Q.; Liang, Z.; Zheng, H.; Ping, Y.; Liu, S.; Sun, J. Analysis of codon usage bias of mitochondrial genome in Bombyx mori and its relation to evolution. BMC Evol. Biol. 2014, 14, 262. [Google Scholar] [CrossRef] [Green Version]
  37. Jennings, R.M.; Halanych, K.M. Mitochondrial genomes of Clymenella torquata (Maldanidae) and Riftia pachyptila (Siboglinidae): evidence for conserved gene order in Annelida. Mol. Biol. Evol. 2005, 22, 210–222. [Google Scholar] [CrossRef]
  38. Palomares-Rius, J.E.; Cantalapiedra-Navarrete, C.; Archidona-Yuste, A.; Blok, V.C.; Castillo, P. Mitochondrial genome diversity in dagger and needle nematodes (Nematoda: Longidoridae). Sci. Rep. 2017, 7, 41813. [Google Scholar] [CrossRef] [Green Version]
  39. Wolstenholme, D.R. Animal mitochondrial DNA: structure and evolution. Int. Rev. Cytol. 1992, 141, 173–216. [Google Scholar]
  40. Schwarz, G. Estimating the dimension of a model. Ann. Statist. 1978, 6, 461–464. [Google Scholar] [CrossRef]
  41. Felsenstein, J. Evolutionary trees from DNA sequences: a maximum likelihood approach. J. Mol. Evol. 1981, 17, 368–376. [Google Scholar] [CrossRef]
  42. Huelsenbeck, J.; Ronquist, F. MrBayes: Bayesian inference of phylogenetic trees. Bioinformatics 2001, 17, 754–755. [Google Scholar] [CrossRef] [Green Version]
  43. Whitfield, J.B.; Lockhart, P.J. Deciphering ancient rapid radiations. Trends Ecol. Evol. 2007, 22, 258–265. [Google Scholar] [CrossRef] [PubMed]
  44. Uribe, J.E.; Puillandre, N.; Zardoya, R. Beyond Conus: phylogenetic relationships of Conidae based on complete mitochondrial genomes. Mol. Phylogenet. Evol. 2017, 107, 142–151. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  45. Abalde, S.; Tenorio, M.J.; Uribe, J.E.; Zardoya, R. Conidae phylogenomics and evolution. Zool. Scr. 2019, 48, 194–214. [Google Scholar] [CrossRef]
  46. Abdelkrim, J.; Aznar-Cormano, L.; Fedosov, A.E.; Kantor, Y.I.; Lozouet, P.; Phuong, M.A.; Zaharias, P.; Puillandre, N. Exon-capture-based phylogeny and diversification of the venomous gastropods (Neogastropoda, Conoidea). Mol. Biol. Evol. 2018, 35, 2355–2374. [Google Scholar] [CrossRef] [PubMed]
  47. Zhang, N.; Su, J.; Liu, H.; Li, L. Study on source of toxion in Nassarius semiplicatus. Mar. Environ. Sci. 2010, 29, 705–707. [Google Scholar]
  48. The Paleobiology Database. Available online: https://paleobiodb.org/ (accessed on 1 May 2020).
  49. Yang, Y.; Qi, L.; Kong, L.; Li, Q. Phylogeography of marine gastropod Reticunassa festiva complex (Nassariidae) in the coast of China. J. Shellfish. Res. 2020. (Accepted). [Google Scholar]
  50. Wang, P. Response of Western Pacific marginal seas to glacial cycles: paleoceanographic and sedimentological features. Mar. Geol. 1999, 156, 5–39. [Google Scholar] [CrossRef]
  51. Lozouet, P.; Galindo, L.A. Resolution of the confused classification of some Miocene Nassariidae (Mollusca: Gastropoda) and reappraisal of their paleobiodiversity on the French Atlantic seaboard. Arch. Für Molluskenkd. 2015, 144, 31–50. [Google Scholar]
  52. Liu, H.; Yang, Y.; Sun, S.E.; Kong, L.; Li, Q. Mitogenomic phylogeny of the Naticidae (Gastropoda: Littorinimorpha) reveals monophyly of the Polinicinae. Zool. Scr. 2020, 49, 295–306. [Google Scholar] [CrossRef]
  53. Bankevich, A.; Nurk, S.; Antipov, D.; Gurevich, A.A.; Dvorkin, M.; Kulikov, A.S.; Lesin, V.M.; Nikolenko, S.I.; Pham, S.; Prjibelski, A.D.; et al. SPAdes: a new genome assembly algorithm and its applications to single-cell sequencing. J. Comput. Biol. 2012, 19, 455–477. [Google Scholar] [CrossRef] [Green Version]
  54. 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]
  55. 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] [PubMed]
  56. Lowe, T.M.; Eddy, S.R. TRNAscan-SE: a program for improved detection of transfer RNA genes in genomic sequence. Nucleic Acids Res. 1997, 25, 955–964. [Google Scholar] [CrossRef]
  57. 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]
  58. Kumar, S.; Stecher, G.; Tamura, K. MEGA7: Molecular Evolutionary Genetics Analysis Version 7.0 for Bigger Datasets. Mol. Biol. Evol. 2016, 33, 1870–1874. [Google Scholar] [CrossRef] [Green Version]
  59. 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]
  60. 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]
  61. Abascal, F.; Zardoya, R.; Telford, M.J. TranslatorX: multiple alignment of nucleotide sequences guided by amino acid translations. Nucleic Acids Res. 2010, 38, W7–W13. [Google Scholar] [CrossRef] [Green Version]
  62. Katoh, K.; Standley, D.M. MAFFT Multiple sequence alignment software version 7: improvements in performance and usability. Mol. Biol. Evol. 2013, 30, 772–780. [Google Scholar] [CrossRef] [Green Version]
  63. Castresana, J. Selection of conserved blocks from multiple alignments for their use in phylogenetic analysis. Mol. Biol. Evol. 2000, 17, 540–552. [Google Scholar] [CrossRef] [Green Version]
  64. Stamatakis, A. RAxML-VI-HPC: maximum likelihood-based phylogenetic analyses with thousands of taxa and mixed models. Bioinformatics 2006, 22, 2688–2690. [Google Scholar] [CrossRef] [PubMed]
  65. Ronquist, F.; Huelsenbeck, J.P. MrBayes 3: Bayesian phylogenetic inference under mixed models. Bioinformatics 2003, 19, 1572–1574. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  66. Tracer v1.6. Available online: http://beast.bio.ed.ac.uk.
  67. FigTree 1.4.2. Available online: http://tree.bio.ed.ac.uk/software/figtree/.
  68. Lartillot, N.; Rodrigue, N.; Stubbs, D.; Richer, J. PhyloBayes MPI: phylogenetic reconstruction with infinite mixtures of profiles in a parallel environment. Syst. Biol. 2013, 62, 611–615. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  69. Miller, M.A.; Pfeiffer, W.; Schwartz, T. Creating the CIPRES Science Gateway for inference of large phylogenetic trees. In Proceedings of the Gateway Computing Environments Workshop (GCE), New Orleans, LA, USA, 14–14 November 2010; pp. 1–8. [Google Scholar]
  70. 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. Biol. Evol. 2019, 133, 12–23. [Google Scholar] [CrossRef]
  71. Lanfear, R.; Frandsen, P.B.; Wright, A.M.; Senfeld, T.; Calcott, B. PartitionFinder 2: new methods for selecting partitioned models of evolution for molecular and morphological phylogenetic analyses. Mol. Biol. Evol. 2016, 34, 772–773. [Google Scholar] [CrossRef] [Green Version]
  72. Drummond, A.; Rambaut, A. BEAST: Bayesian evolutionary analysis by sampling trees. BMC Evol. Biol. 2007, 7, 214. [Google Scholar] [CrossRef] [Green Version]
  73. Tate, R. The lamellibranchs of the Older Tertiary of Australia (part 1). T. Roy. Soc. South. Aust. 1886, 8, 96–158. [Google Scholar]
  74. Foster, H.L. Geology of Ishigaki-shima Ryukyu-retto. U. S. Geol. Surv. Prof. Pap. 1965, 399, 1–119. [Google Scholar]
  75. Zhang, S.P. Comparative study on two confused species of Nassarius. Acta Zootaxonomica Sinica 2010, 35, 868–870. [Google Scholar]
  76. Maddison, W.P.; Maddison, D.R. Mesquite: A Modular System for Evolutionary Analysis. 2018. Version 3.6. Available online: http://www.mesquiteproject.org.
Figure 1. Gene map of the mt genomes of N. glans and N. siquijorensis.
Figure 1. Gene map of the mt genomes of N. glans and N. siquijorensis.
Ijms 21 03545 g001
Figure 2. Relative Synonymous Codon Usage (RSCU) of mitochondrial genomes for N. glans and N. siquijorensis.
Figure 2. Relative Synonymous Codon Usage (RSCU) of mitochondrial genomes for N. glans and N. siquijorensis.
Ijms 21 03545 g002
Figure 3. Inferred secondary structures of 22 tRNAs from N. glans: The tRNAs are labelled with their corresponding amino acids. Structural elements in tRNA arms and loops are illustrated as for trnF.
Figure 3. Inferred secondary structures of 22 tRNAs from N. glans: The tRNAs are labelled with their corresponding amino acids. Structural elements in tRNA arms and loops are illustrated as for trnF.
Ijms 21 03545 g003
Figure 4. Phylogenetic relationship of Nassarius based on nucleotide sequences of 13 mitochondrial PCGs and 2 rRNA genes: The reconstructed Bayesian inference (BI) phylogram (generated by Mybayes program) using “Nassariusacuticostus, “Nassariusjacksonianus, and Reticunassa as outgroup is shown. The first number at each node is bootstrap proportion (BP) of maximum likelihood (ML) analyses, and the second number is Bayesian posterior probability (PP). Nodal with maximum statistical supports (BP = 100; PP = 1) is marked with a solid red circle.
Figure 4. Phylogenetic relationship of Nassarius based on nucleotide sequences of 13 mitochondrial PCGs and 2 rRNA genes: The reconstructed Bayesian inference (BI) phylogram (generated by Mybayes program) using “Nassariusacuticostus, “Nassariusjacksonianus, and Reticunassa as outgroup is shown. The first number at each node is bootstrap proportion (BP) of maximum likelihood (ML) analyses, and the second number is Bayesian posterior probability (PP). Nodal with maximum statistical supports (BP = 100; PP = 1) is marked with a solid red circle.
Ijms 21 03545 g004
Figure 5. Ancestral character state reconstruction of three characteristics using the maximum likelihood model in Mesquite: Pie charts represent the degree of support at every node.
Figure 5. Ancestral character state reconstruction of three characteristics using the maximum likelihood model in Mesquite: Pie charts represent the degree of support at every node.
Ijms 21 03545 g005
Figure 6. Divergences time estimations for the Nassariidae using Bayesian relaxed dating methods (BEAST) based on the nucleotide sequences of 13 PCGs: Dates (and credibility intervals) are in millions of years, and horizontal bars represent 95% credibility intervals of relevant nodes. Calibration constraints are indicated with an asterisk on the corresponding nodes.
Figure 6. Divergences time estimations for the Nassariidae using Bayesian relaxed dating methods (BEAST) based on the nucleotide sequences of 13 PCGs: Dates (and credibility intervals) are in millions of years, and horizontal bars represent 95% credibility intervals of relevant nodes. Calibration constraints are indicated with an asterisk on the corresponding nodes.
Ijms 21 03545 g006
Table 1. List of the mt genomes analysed in the present study.
Table 1. List of the mt genomes analysed in the present study.
New mt genomes
SpeciesLength (bp)Sampling timeAccession no.Toxicity Record
Nassarius siquijorensis15,337April, 2017MN983149Toxin accumulators [19]
Nassarius glans15,296October, 2018MN983150Toxin accumulators [20]
Genbank mt genome
SpeciesLength (bp)Accession no.ReferenceToxicity Record
Tritia obsoleta15,263DQ238598[24]Unknown
Tritia reticulata15,271EU827201[25]Unknown
Nassarius variciferus15,269KM603509[26]Not toxin accumulators [21]
Nassarius succinctus15,329KT768016[6]Toxin accumulators [19]
Nassarius nodifer15,337KT818617[6]Toxin accumulators [21]
Nassarius conoidalis15,332KT826694[6]Toxin accumulators [21]
Nassarius pullus15,278KT900947[6]Toxin accumulators [17]
Nassarius sinarus15,325MH346208[6]Toxin accumulators [19]
Nassarius foveolatus15,343MH346209[6]Toxin accumulators [21]
Nassarius javanus15,325MH346210[6]Toxin accumulators [21]
Reticunassa fratercula15,174KT826695[26]Unknown
Reticunassa hiradoensis15,194MG744569[26]Unknown
Reticunassa festiva-A15,195KT735055[26]Not toxin accumulators [21]
Reticunassa festiva-B15,194MF148855[26]Not toxin accumulators [21]
Reticunassa festiva-C15,172MG744570[26]Not toxin accumulators [21]
Nassarius jacksonianus15,234MH346212[6]Unknown
Nassarius acuticostus15,240MH346211[6]Unknown
Table 2. Gene annotations of the complete mt genome of N. glans and N. siquijorensis: Identical values are shown only once.
Table 2. Gene annotations of the complete mt genome of N. glans and N. siquijorensis: Identical values are shown only once.
N. glans/N. siquijorensis
GeneStrandLocation (bp)Size
(bp)
Start
Codon
Stop
Codon
Intergenic
Nucleotides (bp)
cox3H1–780/1–780780ATGTAA10/9
tRNA-LysH791–855/790–85665/67 4/3
tRNA-AlaH860–927/860–92868/69 9
tRNA-ArgH937–1005/938–100669 11/17
tRNA-AsnH1017–1082/1024–109266/69 20/18
tRNA-IleH1103–1171/1111–117969 3
nad3H1175–1528/1183–1536354ATGTAA/TAG0
tRNA-SerH1529–1596/1537–160468 0
nad2H1597–2652/1605–26601056ATGTAA3/0
cox1H2656–4191/2661–41961536ATGTAA/TAG21/23
cox2H4213–4899/4220–4906687ATGTAA−2
tRNA-AspH4898–4965/4905–497268 1
atp8
atp6
tRNA-Met
H
H
L
4967–5125/4974–5132
5136–5831/5141–5836
5865–5931/5870–5935
159
696
67/66
ATG
ATG
TAA
TAG/TAA
10/8
33
4/5
tRNA-TyrL5936–6001/5941–600766/67 1
tRNA-CysL6003–6069/6009–607267/64 0
tRNA-TrpL6070–6136/6073–613967 −2
tRNA-GlnL6135–6201/6138–620467 0
tRNA-GlyL6202–6269/6205–627168/67 25/20
tRNA-GluL6295–6361/6292–635867 0
12s
tRNA-Val
16s
H
H
H
6362–7325/6359–7320
7326–7393/7321–7389
7394–8754/7390–8748
964/962
68/69
1361/1359
0
0
0
tRNA-LeuH8755–8825/8749–881771/69 11/7
tRNA-LeuH8837–8905/8825–889369 0
nad1H8906–9847/8894–9835942ATGTAA15/14
tRNA-ProH9863–9931/9850–991869 1
nad6H9933–10433/9920–10420501ATGTAA8
cytbH10442–11581/10429–115681140ATGTAA1/16
tRNA-SerH11583–11643/11585–1164161/57 17/8
tRNA-ThrL11661–11728/11650–1171868/69 26/35
nad4LH11755–12051/11754–12050297ATGTAG11
nad4H12063–13418/12062–134171356ATA/ATGTAG−1
tRNA-HisH13418–13485/13417–1348468 1
nad5H13487–15208/13486–152071722ATGTAG−1
tRNA-PheH15208–15274/15207–1527467/68 22/63
Table 3. List of AT content, AT skew, and GC skew of N. glans (Ngla) and N. siquijorensis (Nsiq).
Table 3. List of AT content, AT skew, and GC skew of N. glans (Ngla) and N. siquijorensis (Nsiq).
(A + T)%AT SkewGC Skew
NglaNsiqNglaNsiqNglaNsiq
Genome70.4070.20−0.12−0.120.070.06
PCGs69.7369.78−0.18−0.180.050.04
PCGs 1th61.2861.17−0.10−0.100.250.25
PCGs 2th62.8862.85−0.41−0.41−0.15−0.14
PCGs 3th85.0385.33−0.07−0.070.02−0.02
rrnS68.8067.640.090.090.170.15
rrnL73.8272.840.030.040.150.15
tRNAs68.9068.73−0.03−0.020.070.06
atp671.1471.72−0.26−0.23−0.04−0.05
atp875.6474.36−0.08−0.070.050
cob68.3468.43−0.20−0.22−0.040
cox166.6066.60−0.19−0.190.080.08
cox267.9869.01−0.12−0.130.130.15
cox364.2264.86−0.25−0.240.220.16
nad168.1668.48−0.22−0.220.100.07
nad272.0872.27−0.20−0.200.230.25
nad369.8070.37−0.27−0.270.230.19
nad471.9971.40−0.16−0.13−0.04−0.06
nad4L72.1173.13−0.14−0.150.100.11
nad570.8070.80−0.11−0.11−0.14−0.14
nad674.9072.89−0.22−0.220.040.08
Table 4. Codon and relative synonymous codon usage (RSCU) of 13 PCGs in the mt genomes of N. glans (Ngla) and N. siquijorensis (Nsiq).
Table 4. Codon and relative synonymous codon usage (RSCU) of 13 PCGs in the mt genomes of N. glans (Ngla) and N. siquijorensis (Nsiq).
Amino CodonCount (RSCU)AminoCodonCount (RSCU)
Acid NglaNsiqAcid NglaNsiq
PheUUU263 (1.66) a249 (1.59)AlaGCU130 (2.15)140 (2.27)
UUC53 (0.34)65 (0.41) GCC23 (0.38)23 (0.37)
LeuUUA345 (3.53)359 (3.64) GCA77 (1.27)75 (1.21)
UUG56 (0.57)49 (0.5) GCG12 (0.2)9 (0.15)
CUU96 (0.98)95 (0.96)GlyGGU76 (1.23)73 (1.19)
CUC10 (0.1)11 (0.11) GGC4 (0.06)5 (0.08)
CUA68 (0.7)67 (0.68) GGA126 (2.03)131 (2.14)
CUG12 (0.12)11 (0.11) GGG42 (0.68)36 (0.59)
IleAUU286 (1.83)278 (1.78)ArgCGU21 (1.42)24 (1.63)
AUC26 (0.17)35 (0.22) CGC1 (0.07)1 (0.07)
MetAUA151 (1.56)152 (1.6) CGA32 (2.17)32 (2.17)
AUG42 (0.44)38 (0.4) CGG5 (0.34)2 (0.14)
ValGUU118 (1.9)118 (1.9)TyrUAU116 (1.63)114 (1.61)
GUC13 (0.21)9 (0.15) UAC26 (0.37)28 (0.39)
GUA93 (1.49)100 (1.61)HisCAU62 (1.57)63 (1.56)
GUG25 (0.4)21 (0.34) CAC17 (0.43)18 (0.44)
SerUCU125 (2.67)118 (2.52)GlnCAA63 (1.7)57 (1.58)
UCC26 (0.55)19 (0.41) CAG11 (0.3)15 (0.42)
UCA80 (1.71)88 (1.88)AsnAAU107 (1.67)103 (1.65)
UCG7 (0.15)9 (0.19) AAC21 (0.33)22 (0.35)
AGU47 (1)54 (1.15)LysAAA80 (1.67)80 (1.67)
AGC7 (0.15)8 (0.17) AAG16 (0.33)16 (0.33)
AGA75 (1.6)67 (1.43)AspGAU63 (1.62)65 (1.65)
AGG8 (0.17)12 (0.26) GAC15 (0.38)14 (0.35)
ProCCU80 (2.16)96 (2.54)GluGAA74 (1.72)70 (1.61)
CCC17 (0.46)5 (0.13) GAG12 (0.28)17 (0.39)
CCA42 (1.14)38 (1.01)CysUGU33 (1.69)34 (1.7)
CCG9 (0.24)12 (0.32) UGC6 (0.31)6 (0.3)
ThrACU82 (1.95)79 (1.94)TrpUGA90 (1.64)100 (1.82)
ACC9 (0.21)10 (0.25) UGG20 (0.36)10 (0.18)
ACA72 (1.71)66 (1.62)*UAA9 (1.38)8 (1.23)
ACG5 (0.12)8 (0.2) UAG4 (0.62)5 (0.77)
a The higher values of preferentially used codons are in bold.

Share and Cite

MDPI and ACS Style

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. https://doi.org/10.3390/ijms21103545

AMA Style

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. International Journal of Molecular Sciences. 2020; 21(10):3545. https://doi.org/10.3390/ijms21103545

Chicago/Turabian Style

Yang, Yi, Hongyue Liu, Lu Qi, Lingfeng Kong, and Qi Li. 2020. "Complete Mitochondrial Genomes of Two Toxin-Accumulated Nassariids (Neogastropoda: Nassariidae: Nassarius) and Their Implication for Phylogeny" International Journal of Molecular Sciences 21, no. 10: 3545. https://doi.org/10.3390/ijms21103545

Note that from the first issue of 2016, this journal uses article numbers instead of page numbers. See further details here.

Article Metrics

Back to TopTop