Next Article in Journal
Genome-Wide Association Analysis Reveals Genetic Architecture and Candidate Genes Associated with Grain Yield and Other Traits under Low Soil Nitrogen in Early-Maturing White Quality Protein Maize Inbred Lines
Previous Article in Journal
Exploring Relevant mRNAs and miRNAs in Injured Urethral Tissues of Rats with High-Throughput Sequencing
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

The Complete Mitochondrial Genome of Stichopus naso (Aspidochirotida: Stichopodidae: Stichopus) and Its Phylogenetic Position

1
CAS Key Laboratory of Tropical Marine Bio-Resources and Ecology (LMB), Guangdong Provincial Key Laboratory of Applied Marine Biology (LAMB), South China Sea Institute of Oceanology Chinese Academy of Sciences, Guangzhou 510301, China
2
University of Chinese Academy of Sciences, Beijing 100049, China
3
Marin College, Beibu Gulf University, Qinzhou 535011, China
4
Southern Marine Science and Engineering Guangdong Laboratory (Guangzhou), Guangzhou 510301, China
*
Author to whom correspondence should be addressed.
Genes 2022, 13(5), 825; https://doi.org/10.3390/genes13050825
Submission received: 7 April 2022 / Revised: 2 May 2022 / Accepted: 3 May 2022 / Published: 5 May 2022
(This article belongs to the Section Animal Genetics and Genomics)

Abstract

:
The mitochondrial genome is widely used to study the molecular evolution of and perform phylogenetic analyses on animals. In this study, the complete mitochondrial genome (mitogenome) of Stichopus naso was sequenced. The mitogenome was 16,239 bp in length and contained 13 protein-coding genes (PCGs), 22 transfer RNA genes (tRNAs), and 2 ribosomal RNA genes (rRNAs). The genome composition showed positive AT-skew (0.023) and negative GC-skew (−0.158). The order of the mitochondrial genes was consistent with those from the Stichopus and Isostichopus species, whereas it was different from those of other species of Aspidochirotida. The phylogenetic analysis, based on the nucleotide sequences of 13 PCGs through the methods of Bayesian inference (BI) and maximum likelihood (ML), indicated that S. naso has close relationships with S. horrens and S. monotuberculatus, and belongs to a member of Stichopodidae. Our study provides a reference mitogenome for further molecular evolution studies and phylogenetic research on sea cucumbers.

1. Introduction

Mitochondrial DNA (mtDNA) is a type of extrachromosomal genome with unique characteristics, such as simple structure, low-rate recombination, maternal inheritance, and rapid evolution [1,2], and it is widely used as an effective molecular marker in diverse evolutionary research. Additionally, complete mitogenomes provide powerful tools for analyzing gene rearrangements and genomic evolution [3]. Therefore, mtDNA has been widely used in the study of genetic relationships and molecular identification, both within and among species of sea cucumbers [4,5]. The mitogenomes of most animals are typically closed circular molecules with a size of about 14-20 kb, including 13 protein-coding genes (PCGs), 22 transfer RNA genes (tRNAs), and 2 ribosomal RNA (12S rRNA and 16S rRNA) genes [2].
Sea cucumbers are an abundant and commercially important group of echinoderms. They inhabit a variety of marine environments, from intertidal zones to abyssal trenches [6,7]. Aspidochirotida is one of the most diverse orders in the Holothuroidea, including holothurians of various sizes [8], and the sea cucumbers in Aspidochirotida primarily occur in tropical waters featuring soft sediment and reefs [9]. Sea cucumbers play an important role in the benthic food web due to their deposit-feeding habit, acting as "vacuum cleaners" for coral reefs [10,11]. They are also available to humans as nutritious seafood with high economic value [12]. Although sea cucumbers have specific diversity, ecological importance, and high economic value, our understanding of their phylogenetic relationships and their evolution remains limited due to poor genomic and mtDNA data. Meanwhile, easily confused species (e.g. the members in the family Stichopodidae) cannot be accurately distinguished using traditional morphology, which has long been a taxonomic challenge [13]. Complete mitochondrial genomes provide abundant sequence information covering 13 PCGs, and they can be used for investigating phylogenetic relationships among species [14,15]. Among the 13 PCGs, 16S RNA and COX1 gene sequences are generally used for distinguishing confused species [16,17,18,19]. However, until now, only ~30 complete mitochondrial sequences of sea cucumbers have been released, which has seriously hampered the exploration of the phylogenetic relationships among sea cucumbers and other applications based on mitochondrial sequences.
Stichopus naso is a sea cucumber that is widely distributed in the Indo-Pacific region [20]. The sea cucumber is typically found in shallow water (1–19 m) on sandy muddy bottoms with or without sparse sea-grass beds [21]. The boundaries between species are not comprehensive or clear enough; therefore, S. naso is frequently misidentified as S. horrens [13]. In the present study, the complete mitogenome of S. naso was initially sequenced and compared with other mitogenomes of sea cucumbers. The genomic rearrangement and phylogeny were analyzed to deepen our understanding of the evolutionary relationships within the family Stichopodidae. The results of this research may provide new data for phylogenetic studies of the Aspidochirotida species and expand our knowledge about the mitochondrial genomic characteristics of sea cucumbers and the taxonomy within the family Stichopodidae.

2. Materials and Methods

2.1. Sample Collection and DNA Extraction

Stichopus naso was collected from Qionghai, Hainan Province, China. The body wall tissue was excised and preserved in 95% ethanol and stored at −20 °C until DNA extraction. Total genomic DNA was extracted using TIANamp Marine Animals DNA Kit (TianGen, Beijing, China).

2.2. Mitogenome Sequencing

The S. naso DNA library was sequenced by BIOZERON Co., Ltd. (Shanghai, China) using an Illumina NovaSeq 6000 with an average insert size of 450 bp, which was constructed using the NEBNext Ultra DNA Library Prep Kit. Approximately 5.11 GB of raw data from S. naso were generated with 150-base-pair paired-end read lengths. Adapters were removed and quality was trimmed with Trimmomatic v 0.39 [22]. Clean data was assembled de novo using SPAdes v3.14.1 (http://bioinf.spbau.ru/spades (accessed on 1 August 2021)) software with default parameters. Sequences with high coverage depth and long assembly length were selected as candidates and compared to NT libraries to confirm mitochondrial scaffold sequences, and sequences were ligated according to overlap. The starting position and orientation of the mitochondrial assembly sequences were determined from the reference genome of Holothuria pervicax (GenBank: MK328500.1) to obtain complete mitochondrial genome sequence of S. naso.

2.3. Sequence Annotation and Analysis

The complete mitogenome of S. naso was annotated using the online MITOS tool with the selection of echinoderm mitochondrial genetic code and BLAST searches in NCBI databases (https://blast.ncbi.nlm.nih.gov/Blast.cgi (accessed on 1 August 2021)) [23]. The base compositions and relative synonymous codon usage (RSCU) were conducted with MEGA 7.0 based on the echinoderm mitochondrial codon system [24]. Composition skew values were calculated according to the following formulas: AT- skew = (A − T)/(A + T); GC-skew = (G − C)/(G + C) [25]. The graphical map of the S. naso mitogenome was drawn using the CGview tool [26].

2.4. Phylogenetic Analysis

Phylogenetic analyses were performed using the sequences of 13 PCGs in mitochondrial genomes from 35 species, including sea cucumbers, sea stars, and sea urchins. In addition, three species of Echinoidea and three species of Asteroidea were included in the analyses as outgroups. Thirty-five mitogenomes were downloaded from GenBank (Supplementary Table S1) and the nucleotide sequences of their 13 PCGs were used to construct phylogenetic trees through maximum likelihood (ML) and Bayesian inference (BI) methods. Nucleotide sequences of the 13 PCGs were aligned with MUSCLE 3.8 [27]. The alignment results were further trimmed to eliminate the ambiguous positions using Gblocks v0.91 [28]. The trimmed sequences were concatenated into a supermatrix with FASconCAT [29]. Substitution saturation analysis was performed to detect the suitability of the dataset with 13 PCGs from all analyzed species for constructing a phylogenetic tree with DAMBE v. 5.3.15 [30]. The maximum likelihood (ML) and Bayesian inference (BI) were determined by using PhyML v3.0 [31] and MrBayes v3.2.6 [32], respectively. The best model was selected with jModeltest v2.1.10 [33] and the model of GTR + I + G was optimal for analysis with nucleotide alignment. Bayesian analysis was performed using four simultaneous Markov Chain Monte Carlo chains for 2,000,000 generations and sampled every 1000 generations, using a burn-in of 25% generations. The average standard deviation of split frequencies was <0.01. The resulting phylogenetic trees were visualized in FigTree v1.4.4 (http://tree.bio.ed.ac.uk/software/figtree/ (accessed on 22 November 2021)).

3. Results and Discussion

3.1. Genome Organization and Base Composition

The complete mitogenome of S. naso was acquired and deposited in GenBank under accession number MZ469138. The raw reads of the sequencing were deposited in the NCBI Sequence Read Archive (SRA Accession number: SRR19020101). The mtDNA of S. naso comprises 16,239 bp and includes 13 PCGs (cox1–3, nad1–6, nad4L, cob, atp6, and atp8), 22 tRNA genes, and 2 rRNA genes (rrnS and rrnL) (Figure 1) (Table 1). The order and orientation of the genes in the mitogenome are identical with those of other Stichopus species [34]. One of the thirteen PCGs (nad6) and five tRNAs (trnS2, trnQ, trnA, trnV, and trnD) are encoded on the light (−) strand, while the remaining 31 genes are located on the heavy (+) strand (Table 1). The base composition of the S. naso mitogenome is as follows: A = 4883 (30.07%), T = 4659 (28.69%), G = 2819 (17.36%), C = 3878 (23.88%) (Table 2). The A + T content of the mitogenome is 58.76%, which is lower than those of most of the sea cucumbers analyzed, but higher than those of S. chloronotus (58.55%), Holothuria polii (58.33%), and H. leucospilota (57.55%) (Supplementary Table S1) [35,36,37]. The bias toward high AT content is consistent with the general findings on the mitogenomes in metazoans [38]. The overall AT-skew and GC-skew in the entire S. naso mitogenome were 0.023 and −0.158, respectively. The AT-skew for the S. naso mitogenome was weakly positive, indicating a higher occurrence of A than T. By contrast, the GC-skew was slightly negative, indicating a higher incidence of C than G in the mitogenome. The potential mechanism leading to base composition bias in animal mitochondrial DNA may be related to transcription or replication [39]. The mutational stresses primarily associated with mitochondrial replication can have effects on the base composition of the mitochondrial genomes, which in turn creates base composition skews [40,41]. Thus, it suggests that these species may have been subjected to similar mutational pressures associated with mitochondrial replication. Both the base skew patterns of the S. naso were similar to those of other species of Aspidochirotida (Supplementary Table S1). The S. naso mitogenome possessed a 51-base-pair overlap between the genes in 7 locations ranging in size from 1 bp to 25 bp, and the longest overlap occurred between nad2 and rrnL (Table 1). Three overlapping regions (atp8/atp6, cox3/trnS2, and nad4/trnH) were found in a variety of sea cucumbers [34,42,43,44,45].

3.2. Protein-Coding Genes and Codon Usage

The total length of the 13 PCGs in the S. naso mitogenome is 11,133 bp, which accounts for ~70.16% of the entire mitogenome. The average content of AT in the 13 PCGs is 58.07%. Moreover, the average AT-skew and GC-skew of the 13 PCGs are negative (Table 2). Most of the PCGs in S. naso begin with the standard ATG, except nad3 and nad2, which initiate with the codon GTG. The start codon GTG is frequently found in genes encoding NADH dehydrogenase subunits (nad1, nad4l and nad5) and has also been observed in several sea cucumbers [43,44,45,46]. On the other hand, in the S. naso mitogenome, 10 PCGs (cox1, nad4l, cox2, atp8, atp6, cox3, nad5, nad6, cob, and nad2) end with TAA, while the remaining three PCGs (nad3, nad4, and nad1) terminate with TAG (Table 1). The usage of termination codons in the S. naso mitogenome is similar to those of other echinoderms [34,47,48].
The relative synonymous codon usage (RSCU) values for the 13 PCGs are summarized in Table 3 and Figure 2. Excluding stop codons, there are 5252 codons in the mitogenome of S. naso. The codons encoding Leu (CUN), Asn, Ile, and Phe are the most frequent, while those encoding Cys and Met are scarce (Figure 2). These results were similar to those on the codon usage in the S. horrens mitogenome [34]. Moreover, AAA-Asn (RSCU = 1.74), is the most commonly used codon in the mitogenome of S. naso. The RSCU reflects the effect of biased codon usage [49]. In the mitogenome of S. naso, the bias toward greater representation of nucleotides A and T led to a corresponding bias in the encoding amino acids, and it is also a common feature of the mitochondrial genomes in most metazoans [15,50].

3.3. Transfer RNA and Ribosomal RNA Genes

As with most Aspidochirotida mtDNAs, the S. naso mitogenome contains a set of 22 tRNA genes, ranging from 64 bp to 72 bp. The total length of the tRNAs in the S. naso mitogenome is 1518 bp, whereas the overall contents of AT and GC are 61.99% and 38.01%, respectively (Table 2). The tRNA genes of the mitogenome have a weakly positive AT skew (0.063) and a positive GC skew (0.043). The secondary cloverleaf structure of the 22 tRNAs was predicted (Figure 3). The trnS (AGN) lacked a dihydrouridine (DHU) arm, which is considered a common phenomenon in metazoan mitogenomes [51,52,53,54,55]. Except for trnS, the other tRNA genes exhibit typical cloverleaf structures. Previous studies on metazoans have shown that truncated tRNAs lack the D-arm without losing their function [56,57]. Except for trnR, 32 non-Watson–Crick G-U were recognized in the other tRNAs of the S. naso mitogenome. G-U base pairs were also found in Benthodytes marianensis and Phyllophorus liuwutiensis [42,44]. Previous studies have shown that mismatches can be corrected by the editing process without impacting the tRNA transport function [58].
As in other sea cucumbers, the rrnS and rrnL genes generally locate between trnF and trnE, and between nad2 and cox1, respectively (Table 1). The lengths of the rrnS and rrnL genes are 827 bp and 1496 bp, respectively. The A + T content of the rRNAs is 58.37%, which is lower than that of other mitogenomes from the Stichopus species (Supplementary Table S1). Nevertheless, the average skewed values of AT and GC for the two rRNAs are positive (0.203) and negative (−0.036), respectively, which is similar to other Stichopus species (Supplementary Table S1).

3.4. Gene Rearrangement

Mitochondrial gene rearrangement can be an effective tool to study phylogenetic relationships [59,60]. In order to detect the occurrence of rearrangements in S. naso, the gene orders of S. naso and other Aspidochirotida species were compared (Figure 4). A transposition event was observed in the S. naso mtDNA, and it shifted trnM from the site between trnG and trnL2 to the site between trnV and trnD (Figure 4). In a manner that was consistent with previous reports, the tRNA genes appeared to be among the most active elements in the mitochondrial genomes [34,57], and the rearrangement of tRNA genes may be achieved by replication or intramolecular translocation [57]. The tandem duplication/random-loss model (TDRL) is one of the most widely adopted mechanisms for mitochondrial gene rearrangement, which assumes that some genes undergo rearrangement by tandem duplication followed by the random deletion of duplicated genes [58]. The translocation of tRNA genes in the S. naso mitogenome may be explained by TDRL. The gene block (D-Y-G-M) is repeated in tandem and thus produces two clusters of identical genes (D-Y-G-M)-(D′-Y′-G′-M′). Subsequently, redundant genes are likely randomly deleted. For example, D-Y-G-M-D′-Y′-G′-M′ (in which the underlined letters represent deleted genes) probably results in a new M-D′-Y′-G′ gene sequence.
Furthermore, S. horrens, S. chloronotus, S. monotuberculatus, and S. naso were found to have the same mitochondrial gene order, which was preliminarily identified as the ancestral gene order of Stichopus (Figure 4). It is possible that the ancestral gene order of Stichopus could be finally verified as more complete mitochondrial sequences in Stichopus are acquired. In addition, Holothuria, Actinopyga, Parastichopus, and Apostichopus species have the same gene order, while Isostichopus species have the same gene order as Stichopus species, which suggests a closer evolutionary relationship between Isostichopus and Stichopus.

3.5. Phylogenetic Analysis

The phylogenetic relationships were investigated on the basis of 13 PCGs from 28 Holothuroidea species, three Echinoidea species, and three Asteroidea species through ML and BI methods. The topologies of the phylogenetic trees determined by the BI and ML analyses are very similar on the whole (Figure 5). In the BI analysis, the genus Holothuria formed several clades, which were embedded by the genus Actinopyga. However, in the ML analysis, each family analyzed in Holothuroidea formed a separate clade, which was more consistent with their morphological classification. Nevertheless, the topology of Stichopodidae was consistent in both analyses.
In both phylogenetic trees, S. horrens and S. monotuberculatus formed a well-supported clade (BI posterior probabilities (PP) = 1; ML bootstrap (BP) = 100), which was consistent with the result of the phylogenetic tree constructed using the COI gene [43]. Stichopus naso was recovered as a sister to the S. horrens/S. monotuberculatus clade. Additionally, S. chloronotus and (S. naso + (S. horrens + S. monotuberculatus)) formed a monophyletic clade. This finding was consistent with previous phylogenetic studies, in which the phylogenetic trees were constructed based on the COI gene and the 16S gene, respectively [13,43,59]. Meanwhile, I. badionotus clustered with Stichopus with strong nodal support in the phylogenetic tree (Figure 5), which indicated a close relationship between I. badionotus and Stichopus; this cluster result was in line with those of previous studies [13,43,59,60]. In addition, both I. badionotus and Stichopus shared the identical rearrangement pattern of the genes in present study.
Holothuroids are a diverse group in echinoderms [60]. A broader analysis of the mitochondrial genomes in sea cucumbers would help to provide a more comprehensive understanding of their phylogenetic relationships. However, the available data on the mitochondrial genomes of sea cucumbers are limited. The complete mitochondrial genomes of more species need to be obtained and comprehensively studied to further elucidate the phylogenetic relationships of sea cucumbers and to identify confused species in sea cucumbers on a molecular level.

4. Conclusions

In this study, the complete mitochondrial genome of S. naso was characterized for the first time. The S. naso mitogenome is 16,239 bp in length and encodes 13 PCGs, 22 tRNA genes, and two rRNAs. The mitogenome composition exhibited positive AT-skews and negative GC-skews, which were consistent with most of the sequenced mtDNAs in Aspidochirotida. The gene order of S. naso is the same as those of other Stichopus species and Isostichopus, whereas it is different from those of other species of Aspidochirotida. The phylogenetic analysis based on the mitochondrial genomes of sea cucumbers demonstrated that S. naso belongs to a member of the Stichopodidae.

Supplementary Materials

The following are available online at https://www.mdpi.com/article/10.3390/genes13050825/s1. Table S1: GenBank accession numbers, base compositions, and skew values of the species analyzed in this paper.

Author Contributions

Conceptualization, P.L. and Z.L.; investigation, Z.L., B.M., X.L., Y.L., X.J. and C.R.; data curation, Z.L. and B.M.; writing—original draft preparation, Z.L.; writing—review and editing, Z.L., C.H. and P.L.; supervision, P.L.; project administration, C.H. and P.L.; funding acquisition, C.H. and P.L. All authors have read and agreed to the published version of the manuscript.

Funding

This study was supported by the National Key Research and Development Program: Blue Granary Scientific and Technological Innovation (2020YFD0901104 and 2018YFD0901605), and Key Special Project for Introduced Talents Team of Southern Marine Science and Engineering Guangdong Laboratory (GML2019ZD0402).

Institutional Review Board Statement

Not applicable.

Informed Consent Statement

Not applicable.

Data Availability Statement

The mitochondrial genome was deposited at NCBI, with accession number MZ469138.

Acknowledgments

We would like to thank the two anonymous reviewers for providing valuable comments to improve the manuscript.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Caccone, A.; Gentile, G.; Burns, C.E.; Sezzi, E.; Bergman, W.; Ruelle, M.; Saltonstall, K.; Powell, J.R. Extreme difference in rate of mitochondrial and nuclear DNA evolution in a large ectotherm, Galapagos tortoises. Mol. Phylogenet. Evol. 2004, 31, 794–798. [Google Scholar] [CrossRef] [PubMed]
  2. Boore, J.L. Animal mitochondrial genomes. Nucleic Acids Res. 1999, 27, 1767–1780. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  3. Zhang, Y.; Gong, L.; Lu, X.; Jiang, L.; Liu, B.; Liu, L.; Lue, Z.; Li, P.; Zhang, X. Gene rearrangements in the mitochondrial genome of Chiromantes eulimene (Brachyura: Sesarmidae) and phylogenetic implications for Brachyura. Int. J. Biol. Macromol. 2020, 162, 704–714. [Google Scholar] [CrossRef] [PubMed]
  4. Rubinoff, D.; Holland, B.S. Between two extremes: Mitochondrial DNA is neither the panacea nor the nemesis of phylogenetic and taxonomic inference. Syst. Biol. 2005, 54, 952–961. [Google Scholar] [CrossRef]
  5. Wen, J.; Hu, C.; Zhang, L.; Fan, S. Genetic identification of global commercial sea cucumber species on the basis of mitochondrial DNA sequences. Food Control. 2011, 22, 72–77. [Google Scholar] [CrossRef]
  6. Kerr, A.M.; Kim, J.H. Phylogeny of Holothuroidea (Echinodermata) inferred from morphology. J. Linn. Soc. Lond. Zool. 2001, 133, 63–81. [Google Scholar] [CrossRef]
  7. Gallo, N.D.; Cameron, J.; Hardy, K.; Fryer, P.; Bartlett, D.H.; Levin, L.A. Submersible- and lander-observed community patterns in the Mariana and New Britain trenches: Influence of productivity and depth on epibenthic and scavenging communities. Deep. Sea Res. Part I 2015, 99, 119–133. [Google Scholar] [CrossRef] [Green Version]
  8. Prata, J.; Manso, C.L.C.; Christoffersen, M.L. Aspidochirotida (Echinodermata: Holothuroidea) from the northeast coast of Brazil. Zootaxa 2014, 3889, 127–150. [Google Scholar] [CrossRef]
  9. Rowe, F.W.E.; Richmond, M.D. A preliminary account of the shallow-water echinoderms of Rodrigues, Mauritius, western Indian Ocean. J. Nat. Hist. 2004, 38, 3273–3314. [Google Scholar] [CrossRef]
  10. Samyn, Y.; Tallon, I. Zoogeography of the shallow-water holothuroids of the western Indian Ocean. J. Biogeogr. 2005, 32, 1523–1538. [Google Scholar] [CrossRef]
  11. Wolfe, K.; Vidal-Ramirez, F.; Dove, S.; Deaker, D.; Byrne, M. Altered sediment biota and lagoon habitat carbonate dynamics due to sea cucumber bioturbation in a high-pCO(2) environment. GCB Bioenergy 2018, 24, 465–480. [Google Scholar] [CrossRef]
  12. Tuwo, A.; Tresnati, J. Sea Cucumber Farming in Southeast Asia (Malaysia, Philippines, Indonesia, Vietnam); Wiley: Hoboken, NJ, USA, 2015; pp. 331–352. [Google Scholar]
  13. Byrne, M.; Rowe, F.; Uthicke, S. Molecular taxonomy, phylogeny and evolution in the family Stichopodidae (Aspidochirotida: Holothuroidea) based on COI and 16S mitochondrial DNA. Mol. Phylogenet. Evol. 2010, 56, 1068–1081. [Google Scholar] [CrossRef] [PubMed]
  14. Ma, H.; Ma, C.; Ma, L. Molecular identification of genus Scylla (Decapoda: Portunidae) based on DNA barcoding and polymerase chain reaction. Biochem. Syst. Ecol. 2012, 41, 41–47. [Google Scholar] [CrossRef]
  15. Nishihama, S.; Yamana, Y.; Yoshimura, T. First record of the tropical holothurian Stichopus naso Semper, 1867 (Echinodermata: Holothuroidea: Synallactida) from the temperate coast of Kyushu mainland, Japan, in relation to ocean warming. Plankton Benthos. Res. 2020, 15, 66–71. [Google Scholar] [CrossRef] [Green Version]
  16. Massin, C. Redescription of Stichopus naso Semper, 1868 (Echinodermata, Holothuroidea, Stichopodidae). Bull. L’institut R. Sci. Nat. Belg. 2007, 77, 123–130. [Google Scholar]
  17. Bolger, A.M.; Lohse, M.; Usadel, B. Trimmomatic: A flexible trimmer for Illumina sequence data. Bioinformatics 2014, 30, 2114–2120. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  18. Bernt, M.; Donath, A.; Juehling, F.; Externbrink, F.; Florentz, C.; Fritzsch, G.; Puetz, J.; Middendorf, M.; Stadler, P.F. MITOS: Improved de novo metazoan mitochondrial genome annotation. Mol. Phylogenet. Evol. 2013, 69, 313–319. [Google Scholar] [CrossRef]
  19. 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]
  20. 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]
  21. Grant, J.R.; Stothard, P. The CGView Server: A comparative genomics tool for circular genomes. Nucleic Acids Res. 2008, 36, W181–W184. [Google Scholar] [CrossRef]
  22. Edgar, R.C. Muscle: Multiple sequence alignment with high accuracy and high throughput. Nucleic Acids Res. 2004, 32, 1792–1797. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  23. 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] [PubMed] [Green Version]
  24. Kueck, P.; Meusemann, K. FASconCAT: Convenient handling of data matrices. Mol. Phylogenet. Evol. 2010, 56, 1115–1118. [Google Scholar] [CrossRef] [PubMed]
  25. Xia, X.; Xie, Z. DAMBE: Software package for data analysis in molecular biology and evolution. J. Hered. 2001, 92, 371–373. [Google Scholar] [CrossRef] [Green Version]
  26. Guindon, S.; Dufayard, J.-F.; Lefort, V.; Anisimova, M.; Hordijk, W.; Gascuel, O. New Algorithms and Methods to Estimate Maximum-Likelihood Phylogenies: Assessing the Performance of PhyML 3.0. Syst. Biol. 2010, 59, 307–321. [Google Scholar] [CrossRef] [Green Version]
  27. Huelsenbeck, J.P.; Ronquist, F. MRBAYES: Bayesian inference of phylogenetic trees. Bioinformatics 2001, 17, 754–755. [Google Scholar] [CrossRef] [Green Version]
  28. Darriba, D.; Taboada, G.L.; Doallo, R.; Posada, D. jModelTest 2: More models, new heuristics and parallel computing. Nat. Methods 2012, 9, 772. [Google Scholar] [CrossRef] [Green Version]
  29. Fan, S.; Hu, C.; Wen, J.; Zhang, L. Characterization of mitochondrial genome of sea cucumber Stichopus horrens: A novel gene arrangement in Holothuroidea. Sci. China-Life Sci. 2011, 54, 434–441. [Google Scholar] [CrossRef] [Green Version]
  30. Chen, X.; Sun, Y.; Zhao, H.; Hu, J.; Chen, B.; Li, H.; Huang, W. Complete mitochondrial genome of a tropical sea cucumber, Stichopus chloronotus. Mitochondrial DNA B 2021, 6, 2788–2790. [Google Scholar] [CrossRef]
  31. Utzeri, V.J.; Ribani, A.; Bovo, S.; Taurisano, V.; Calassanzio, M.; Baldo, D.; Fontanesi, L. Microscopic ossicle analyses and the complete mitochondrial genome sequence of Holothuria (Roweothuria) polii (Echinodermata; Holothuroidea) provide new information to support the phylogenetic positioning of this sea cucumber species. Mar. Genom. 2020, 51, 100735. [Google Scholar] [CrossRef]
  32. Yang, Q.; Lin, Q.; Wu, J.; Ngoc Tuan, T.; Huang, R.; Sun, Z.; Zhu, Z.; Lu, Z.; Li, S.; Zhou, C. Complete mitochondrial genome of Holothuria leucospilata (Holothuroidea, Holothuriidae) and phylogenetic analysis. Mitochondrial DNA B 2019, 4, 2751–2752. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  33. Wang, Q.; Tang, D.; Guo, H.; Wang, J.; Xu, X.; Wang, Z. Comparative mitochondrial genomic analysis of Macrophthalmus pacificus and insights into the phylogeny of the Ocypodoidea & Grapsoidea. Genomics 2020, 112, 82–91. [Google Scholar] [CrossRef] [PubMed]
  34. Francino, M.P.; Ochman, H. Strand asymmetries in DNA evolution. Trends Genet. 1997, 13, 240–245. [Google Scholar] [CrossRef]
  35. 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] [PubMed] [Green Version]
  36. Xia, X. DNA Replication and Strand Asymmetry in Prokaryotic and Mitochondrial Genomes. Curr. Genom. 2012, 13, 16–27. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  37. Mu, W.; Liu, J.; Zhang, H. Complete mitochondrial genome of Benthodytes marianensis (Holothuroidea: Elasipodida: Psychropotidae): Insight into deep sea adaptation in the sea cucumber. PLoS ONE 2018, 13, e0208051. [Google Scholar] [CrossRef]
  38. Fan, S.; Hu, C.; Zhang, L.; Sun, H.; Wen, J.; Luo, P. Complete mitochondrial genome of the sea cucumber Stichopus sp and its application in the identification of this species. Aquacult. Res. 2012, 43, 1306–1316. [Google Scholar] [CrossRef]
  39. Yang, F.; Zhou, C.; Tran, N.T.; Sun, Z.; Wu, J.; Ge, H.; Lu, Z.; Zhong, C.; Zhu, Z.; Yang, Q.; et al. Comparison of the complete mitochondrial genome of Phyllophorus liuwutiensis (Echinodermata: Holothuroidea: Phyllophoridae) to that of other sea cucumbers. FEBS Open Bio. 2020, 10, 1587–1600. [Google Scholar] [CrossRef]
  40. Shen, X.; Tian, M.; Liu, Z.; Cheng, H.; Tan, J.; Meng, X.; Ren, J. Complete mitochondrial genome of the sea cucumber Apostichopus japonicus (Echinodermata: Holothuroidea): The first representative from the subclass Aspidochirotacea with the echinoderm ground pattern. Gene 2009, 439, 79–86. [Google Scholar] [CrossRef]
  41. Sun, S.; Sha, Z.; Xiao, N. The first two complete mitogenomes of the order Apodida from deep-sea chemoautotrophic environments: New insights into the gene rearrangement, origin and evolution of the deep-sea sea cucumbers. Comp. Biochem. Physiol. Part D Genom. Proteom. 2021, 39, 100839. [Google Scholar] [CrossRef]
  42. Lin, J.-P.; Tsai, M.-H.; Kroh, A.; Trautman, A.; Machado, D.J.; Chang, L.-Y.; Reid, R.; Lin, K.-T.; Bronstein, O.; Lee, S.-J.; et al. The first complete mitochondrial genome of the sand dollar Sinaechinocyamus mai (Echinoidea: Clypeasteroida). Genomics 2020, 112, 1686–1693. [Google Scholar] [CrossRef] [PubMed]
  43. Lee, T.; Bae, Y.J.; Shin, S. Mitochondrial gene rearrangement and phylogenetic relationships in the Amphilepidida and Ophiacanthida (Echinodermata, Ophiuroidea). Mar. Biol. Res. 2019, 15, 26–35. [Google Scholar] [CrossRef]
  44. Sharp, P.M.; Li, W.H. An evolutionary perspective on synonymous codon usage in unicellular organisms. J. Mol. Evol. 1986, 24, 28–38. [Google Scholar] [CrossRef] [PubMed]
  45. Shen, X.; Li, X.; Sha, Z.; Yan, B.; Xu, Q. Complete mitochondrial genome of the Japanese snapping shrimp Alpheus japonicus (Crustacea: Decapoda: Caridea): Gene rearrangement and phylogeny within Caridea. Sci. China-Life Sci. 2012, 55, 591–598. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  46. Wang, Z.; Shi, X.; Guo, H.; Tang, D.; Bai, Y.; Wang, Z. Characterization of the complete mitochondrial genome of Uca lacteus and comparison with other Brachyuran crabs. Genomics 2020, 112, 10–19. [Google Scholar] [CrossRef]
  47. Luis Teixeira, D.C.; Powell, C.; van Noort, S.; Costa, C.; Sinno, M.; Caleca, V.; Rhode, C.; Kennedy, R.J.; van Staden, M.; van Asch, B. The complete mitochondrial genome of Bactrocera biguttula (Bezzi) (Diptera: Tephritidae) and phylogenetic relationships with other Dacini. Int. J. Biol. Macromol. 2019, 126, 130–140. [Google Scholar] [CrossRef]
  48. Su, T.; Liang, A. Comparative analysis of seven mitochondrial genomes of Phymatostetha (Hemiptera: Cercopidae) and phylogenetic implications. Int. J. Biol. Macromol. 2019, 125, 1112–1117. [Google Scholar] [CrossRef]
  49. Kundu, S.; Kumar, V.; Tyagi, K.; Chakraborty, R.; Singha, D.; Rahaman, I.; Pakrashi, A.; Chandra, K. Complete mitochondrial genome of Black Soft-shell Turtle (Nilssonia nigricans) and comparative analysis with other Trionychidae. Sci. Rep. 2018, 8, 1–11. [Google Scholar] [CrossRef]
  50. Krebes, L.; Bastrop, R. The mitogenome of Gammarus duebeni (Crustacea Amphipoda): A new gene order and non-neutral sequence evolution of tandem repeats in the control region. Comp. Biochem. Physiol. Part D Genom. Proteom. 2012, 7, 201–211. [Google Scholar] [CrossRef]
  51. Liu, Y.; Wu, P.-D.; Zhang, D.-Z.; Zhang, H.-B.; Tang, B.-P.; Liu, Q.-N.; Dai, L.-S. Mitochondrial genome of the yellow catfish Pelteobagrus fulvidraco and insights into Bagridae phylogenetics. Genomics 2019, 111, 1258–1265. [Google Scholar] [CrossRef]
  52. Juehling, T.; Duchardt-Ferner, E.; Bonin, S.; Woehnert, J.; Puetz, J.; Florentz, C.; Betat, H.; Sauter, C.; Moerl, M. Small but large enough: Structural properties of armless mitochondrial tRNAs from the nematode Romanomermis culicivorax. Nucleic Acids Res. 2018, 46, 9170–9180. [Google Scholar] [CrossRef] [PubMed]
  53. Watanabe, Y.-I.; Suematsu, T.; Ohtsuki, T. Losing the stem-loop structure from metazoan mitochondrial tRNAs and co-evolution of interacting factors. Front. Genet. 2014, 5, 109. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  54. Montelli, S.; Peruffo, A.; Patarnello, T.; Cozzi, B.; Negrisolo, E. Back to Water: Signature of Adaptive Evolution in Cetacean Mitochondrial tRNAs. PLoS ONE 2016, 11, e0158129. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  55. Chen, C.; Li, Q.; Fu, R.; Wang, J.; Deng, G.; Chen, X.; Lu, D. Comparative mitochondrial genome analysis reveals intron dynamics and gene rearrangements in two Trametes species. Sci. Rep. 2021, 11, 1–12. [Google Scholar] [CrossRef] [PubMed]
  56. Su, T.; He, B.; Li, K.; Liang, A. Comparative analysis of the mitochondrial genomes of oriental spittlebug trible Cosmoscartini: Insights into the relationships among closely related taxa. BMC Genom. 2018, 19, 961. [Google Scholar] [CrossRef]
  57. Arndt, A.; Smith, M.J. Mitochondrial gene rearrangement in the sea cucumber genus Cucumaria. Mol. Biol. Evol. 1998, 15, 1009–1016. [Google Scholar] [CrossRef] [Green Version]
  58. Zhang, K.; Zhu, K.; Liu, Y.; Zhang, H.; Gong, L.; Jiang, L.; Liu, L.; Lu, Z.; Liu, B. Novel gene rearrangement in the mitochondrial genome of Muraenesox cinereus and the phylogenetic relationship of Anguilliformes. Sci. Rep. 2021, 11, 1–14. [Google Scholar] [CrossRef]
  59. Uthicke, S.; Byrne, M.; Conand, C. Genetic barcoding of commercial Beche-de-mer species (Echinodermata: Holothuroidea). Mol. Ecol. Resour. 2010, 10, 634–646. [Google Scholar] [CrossRef]
  60. Miller, A.K.; Kerr, A.M.; Paulay, G.; Reich, M.; Wilson, N.G.; Carvajal, J.I.; Rouse, G.W. Molecular phylogeny of extant Holothuroidea (Echinodermata). Mol. Phylogenet. Evol. 2017, 111, 110–131. [Google Scholar] [CrossRef] [Green Version]
Figure 1. Gene maps of the Stichopus naso mitogenome. Blue arrows represent CDS, brown arrows represent tRNA, and lavender arrows represent rRNA, black, GC content, green, GC skew +, and dark purple, GC skew −.
Figure 1. Gene maps of the Stichopus naso mitogenome. Blue arrows represent CDS, brown arrows represent tRNA, and lavender arrows represent rRNA, black, GC content, green, GC skew +, and dark purple, GC skew −.
Genes 13 00825 g001
Figure 2. Relative synonymous codon usage in the Stichopus naso mitogenome. Codon families are indicated below the X-axis.
Figure 2. Relative synonymous codon usage in the Stichopus naso mitogenome. Codon families are indicated below the X-axis.
Genes 13 00825 g002
Figure 3. Putative secondary structures of tRNAs from the Stichopus naso mitogenome. The tRNAs are labeled with the abbreviations of their corresponding amino acids.
Figure 3. Putative secondary structures of tRNAs from the Stichopus naso mitogenome. The tRNAs are labeled with the abbreviations of their corresponding amino acids.
Genes 13 00825 g003
Figure 4. Linear representation of gene rearrangements of Stichopus naso, S. horrens, S. chloronotus, S. monotuberculatus, Isostichopus badionotus, Holothuria, Actinopyga, Parastichopus, and Apostichopus. All the genes are transcribed from left to right. The tRNA genes are represented by the corresponding single-letter amino acid code. S1 (AGN), S2 (UCN), L1 (CUN), and L2 (UUR). The rrnL and rrnS are the large and small ribosomal RNA subunits, respectively. The red arrow represents the transposition event, which shifts trnM from the position between trnG and trnL2 to the position between trnV and trnD.
Figure 4. Linear representation of gene rearrangements of Stichopus naso, S. horrens, S. chloronotus, S. monotuberculatus, Isostichopus badionotus, Holothuria, Actinopyga, Parastichopus, and Apostichopus. All the genes are transcribed from left to right. The tRNA genes are represented by the corresponding single-letter amino acid code. S1 (AGN), S2 (UCN), L1 (CUN), and L2 (UUR). The rrnL and rrnS are the large and small ribosomal RNA subunits, respectively. The red arrow represents the transposition event, which shifts trnM from the position between trnG and trnL2 to the position between trnV and trnD.
Genes 13 00825 g004
Figure 5. Phylogenetic trees constructed using the nucleotide sequences of the 13 PCGs in the Stichopus naso mitogenome. (a) BI method. (b) ML method. The numbers beside the nodes are posterior probabilities (BI) and bootstrap (ML), respectively.
Figure 5. Phylogenetic trees constructed using the nucleotide sequences of the 13 PCGs in the Stichopus naso mitogenome. (a) BI method. (b) ML method. The numbers beside the nodes are posterior probabilities (BI) and bootstrap (ML), respectively.
Genes 13 00825 g005
Table 1. Summary of the genes in the Stichopus naso mitogenome.
Table 1. Summary of the genes in the Stichopus naso mitogenome.
GeneStrandLocationSizeAnticodonCodonContinuity
StartStop
cox1+1-15541554 ATGTAA167
trnR+1560-162768TCG 5
nad4l+1628-1924297 ATGTAA0
cox2+1925-2614690 ATGTAA0
trnK+2635-269864CTT 20
atp8+2699-2875177 ATGTAA0
atp6+2869-3552684 ATGTAA−7
cox3+3555-4337783 ATGTAA2
trnS24336-440671TGA −2
nad3+4438-4782345 GTGTAG31
nad4+4790-61571368 ATGTAG7
trnH+6148-621568GTG −10
trnS1+6218-628568GCT 2
nad5+6286-81301845 ATGTAA0
nad68143-8631489 ATGTAA12
cob+8640-97821143 ATGTAA8
trnF+9783-985371GAA 0
rrnS+9852-10678827 −2
trnE+10678-1074669TTC −1
trnT+10749-1081870TGT 2
trnP+11483-1155068TGG 664
trnQ11547-1161670TTG −4
trnN+11626-1169469GTT 9
trnL1+11703-1177472TAG 8
trnA11791-1185868TGC 16
trnW+11859-1192870TCA 0
trnC+11937-1200266GCA 8
trnV12004-1207572TAC 1
trnD12089-1215668GTC 13
trnY+12163-1223068GTA 6
trnG+12283-1235169TCC 52
trnM+12354-1242269CAT 2
trnL2+12436-1250772TAA 13
nad1+12508-13479972 ATGTAG0
trnI+13487-1355468GAT 7
nad2+13555-146011047 GTGTAA0
rrnL+14577-160721496 −25
Table 2. Nucleotide composition of the whole Stichopus naso mitogenome and the genes.
Table 2. Nucleotide composition of the whole Stichopus naso mitogenome and the genes.
RegionA%T%G%C%A + T (%)G + C (%)AT-SkewGC-Skew
mitogenome30.0728.6917.3623.8858.7641.240.023−0.158
PCGs27.4130.6717.3224.6058.0741.93−0.056−0.174
cox126.4532.1118.6022.8458.5641.44−0.097−0.102
nad4l26.2637.7112.7923.2363.9736.03−0.179−0.290
cox228.8430.0017.1024.0658.8441.16−0.020−0.169
atp833.9029.3815.2521.4763.2836.720.071−0.169
atp628.5130.7015.2025.5859.2140.79−0.037−0.255
cox325.8031.1619.1623.8856.9643.04−0.094−0.110
nad322.6132.1715.6529.5754.7845.22−0.175−0.308
nad430.2630.2615.7923.6860.5339.470.000−0.200
nad531.6526.6716.4225.2658.3241.680.085−0.212
nad614.7242.7426.3816.1657.4642.54−0.4880.240
cob28.6128.1716.5426.6856.7843.220.008−0.235
nad123.5630.8619.3426.2354.4245.58−0.134−0.151
nad226.1730.7516.1426.9356.9243.08−0.080−0.251
tRNAs32.9429.0519.8318.1861.9938.010.0630.043
rRNAs35.1323.2520.0621.5758.3741.630.203−0.036
Table 3. The codon numbers and relative synonymous codon usages in 13 PCGs of Stichopus naso mitogenome.
Table 3. The codon numbers and relative synonymous codon usages in 13 PCGs of Stichopus naso mitogenome.
CodonCountRSCUCodonCountRSCUCodonCountRSCUCodonCountRSCU
UUU(F)1891.01UCU(S)1381.65UAU(Y)780.98UGU(C)350.99
UUC(F)1860.99UCC(S)1121.34UAC(Y)821.02UGC(C)361.01
UUA(L)1301.13UCA(S)800.95UAA(*)851.06UGA(W)751.20
UUG(L)520.45UCG(S)400.48UAG(*)760.94UGG(W)500.80
CUU(L)1581.37CCU(P)1071.37CAU(H)791.06CGU(R)260.89
CUC(L)1311.14CCC(P)881.13CAC(H)700.94CGC(R)190.65
CUA(L)1471.28CCA(P)871.12CAA(Q)1051.13CGA(R)401.37
CUG(L)730.63CCG(P)300.38CAG(Q)810.87CGG(R)321.09
AUU(I)1301.03ACU(T)1001.14AAU(N)800.57AGU(S)480.57
AUC(I)910.72ACC(T)1021.17AAC(N)960.69AGC(S)670.80
AUA(I)1571.25ACA(T)971.11AAA(N)2421.74AGA(S)1241.48
AUG(M)951.00ACG(T)510.58AAG(K)1281.00AGG(S)620.74
GUU(V)691.19GCU(A)851.31GAU(D)600.94GGU(G)631.09
GUC(V)510.88GCC(A)841.29GAC(D)671.06GGC(G)310.53
GUA(V)731.26GCA(A)600.92GAA(E)1171.33GGA(G)891.53
GUG(V)380.66GCG(A)310.48GAG(E)590.67GGG(G)490.84
Publisher’s Note: MDPI stays neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Share and Cite

MDPI and ACS Style

Li, Z.; Ma, B.; Li, X.; Lv, Y.; Jiang, X.; Ren, C.; Hu, C.; Luo, P. The Complete Mitochondrial Genome of Stichopus naso (Aspidochirotida: Stichopodidae: Stichopus) and Its Phylogenetic Position. Genes 2022, 13, 825. https://doi.org/10.3390/genes13050825

AMA Style

Li Z, Ma B, Li X, Lv Y, Jiang X, Ren C, Hu C, Luo P. The Complete Mitochondrial Genome of Stichopus naso (Aspidochirotida: Stichopodidae: Stichopus) and Its Phylogenetic Position. Genes. 2022; 13(5):825. https://doi.org/10.3390/genes13050825

Chicago/Turabian Style

Li, Zhuobo, Bo Ma, Xiaomin Li, Ying Lv, Xiao Jiang, Chunhua Ren, Chaoqun Hu, and Peng Luo. 2022. "The Complete Mitochondrial Genome of Stichopus naso (Aspidochirotida: Stichopodidae: Stichopus) and Its Phylogenetic Position" Genes 13, no. 5: 825. https://doi.org/10.3390/genes13050825

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