Skip to main content
Advertisement
Browse Subject Areas
?

Click through the PLOS taxonomy to find articles in your field.

For more information about PLOS Subject Areas, click here.

  • Loading metrics

Mitochondrial genome sequencing of a vermivorous cone snail Conus quercinus supports the correlative analysis between phylogenetic relationships and dietary types of Conus species

  • Bingmiao Gao ,

    Contributed equally to this work with: Bingmiao Gao, Chao Peng

    Roles Writing – original draft

    Affiliation Hainan Provincial Key Laboratory of Research and Development of Tropical Medicinal Plants, Hainan Medical University, Haikou, China

  • Chao Peng ,

    Contributed equally to this work with: Bingmiao Gao, Chao Peng

    Roles Data curation, Software, Writing – review & editing

    Affiliation Shenzhen Key Lab of Marine Genomics, Guangdong Provincial Key Lab of Molecular Breeding in Marine Economic Animals, BGI Academy of Marine Sciences, BGI Marine, BGI, Shenzhen, China

  • Qin Chen,

    Roles Data curation, Software

    Affiliation School of Agricultural and Forestry Science and Technology, Hainan Radio & TV University, Haikou, China

  • Junqing Zhang ,

    Roles Supervision, Validation

    jqzhang2011@163.com (JZ); shiqiong@genomics.cn (QS)

    Affiliation Hainan Provincial Key Laboratory of Research and Development of Tropical Medicinal Plants, Hainan Medical University, Haikou, China

  • Qiong Shi

    Roles Writing – review & editing

    jqzhang2011@163.com (JZ); shiqiong@genomics.cn (QS)

    Affiliation Shenzhen Key Lab of Marine Genomics, Guangdong Provincial Key Lab of Molecular Breeding in Marine Economic Animals, BGI Academy of Marine Sciences, BGI Marine, BGI, Shenzhen, China

Abstract

Complete mitochondrial genome (mitogenome) sequence of a worm-hunting cone snail, Conus quercinus, was reported in this study. Its mitogenome, the longest one (16,460 bp) among reported Conus specie, is composed of 13 protein-coding genes (PCGs), 22 transfer RNA (tRNA) genes, two ribosomal RNA (rRNA) genes and one D-loop region. The mitochondrial gene arrangement is highly-conserved and identical to other reported. However, the D-loop region of C. quercinus is the longest (943 bp) with the higher A+T content (71.3%) and a long AT tandem repeat stretch (68 bp). Subsequent phylogenetic analysis demonstrated that three different dietary types (vermivorous, molluscivorous and piscivorous) of cone snails are clustered separately, suggesting that the phylogenetics of cone snails is related to their dietary types. In conclusion, our current work improves our understanding of the mitogenomic structure and evolutionary status of the vermivorous C. quercinus, which support the putative hypothesis that the Conus ancestor was vermivorous.

Introduction

Cone snails (Conus spp.), a species-rich genus of venomous marine gastropods, produce a complex of conotoxins for prey capture and defense. They are usually classified into fish-hunting (piscivorous), snail-hunting (molluscivorous) and worm-hunting (vermivorous) groups [13]. The number of piscivorous species is the least, while these snails are assessed as deadly to humans. A larger number of molluscivorous species is dangerous, although some snails have been implicated in unconfirmed fatalities. Forming the largest group, the vermivorous species account for 70% of the Conus genus, while they seem to be nonthreatening [24]. There are more than 800 Conus species, and each typically contains 100~200 venom peptides; therefore, a total of over 80,000 conotoxins have been identified from various cone snails around the world [5,6].

The colors and diets, along with the composed conotoxins, in different cone snails are abundant and complicated; hence, related taxonomy, population genetics, evolutionary biology and phylogenetics have aroused the interest of scientists [7,8]. Evolutionary relationships between feeding and conotoxins have been discussed on basis of transcriptomics, proteomics and genomics [9,10]. The diversity of peptides in the venom of cone snails confirms that most species are able to produce a variety of conotoxins, as widely reported in literatures [11,12]. There is a specific hypothesis for the shift from an ancestral worm-hunting to more recent fish-hunting [13]. However, the poor performance of venom components in predicting prey taxonomic class suggests that conotoxins (gene superfamily) and traditional means of categorizing prey types (worms, mollusa, fish) do not accurately clarify the evolutionary dynamics between venom composition and diets [1416].

Nowadays, mitochondrial genome (mitogenome) has been one of the most popular tools widely applied for gastropod mollusk taxonomy, population genetics, evolutionary biology and phylogenetics [17]. Gastropod mollusk mitogenomes usually exhibit high diversity of gene orders, and accordingly offer a suitable model system to study the patterns, rates, and mechanisms of mitogenome rearrangement as well as the phylogenetic utility of arrangement comparisons [18]. So far, the mitogenome sequences have been reported for nine cone snails, including two piscivorous (C. consors and C. striatus), three molluscivorous (C. tulipa, C. textile and C. gloriamaris), three vermivorous (C. borgesi, C. capitaneus and C. tribblei), and one broad dietary (C. californicus) species [1926].

Here, we reported the mitogenome of an additional vermivorous cone snail, C. quercinus, and described some outstanding features of its mitogenome sequence. Related mitogenomic structure and phylogentic status are going to provide more supportive evidence for the putative hypothesis about the vermivorous Conus ancestor and the correlation between traditional classification of prey types and mitogenome evolution.

Materials and methods

Ethical statement

No specific permits were required for the collection of specimens for this study. These Conus specimens are common in China, and the field collection did not involve any endangered or protected species. Our experimental procedures complied with the current laws on animal welfare and research in China, and were specifically approved by the Animal Research Ethics Committee of Hainan Medical University.

Genomic DNA extraction and sequencing

Live C. quercinus were collected in the offshore areas of Lingshui City, Hainan Province, China. Around 150 mg of foot tissue was ground to powder using mortar and pestle under liquid nitrogen. Total genomic DNA was extracted by the Column mtDNAout kit (Tianda, Beijing, China) according to the manufacturer’s instructions with minor modifications. The purified genomic DNA was quantified with a Nanodrop 2000 spectrometer (ThermoFisher Scientific, Wilmington, DE, USA).

Normalized DNA of 3 μg was employed to prepare a paired-end library using the NEB Next DNA sample libraries kit (New England Biolabs, New England) in accordance with the manufacturer’s instructions. Quantification and size estimation of the library were performed on a Bioanalyzer 2100 High Sensitivity DNA chip (Agilent, Palo Alto, CA, USA). Finally, the library was normalized to 2 nM and sequenced on the Illumina HiSeq2000 (Illumina, San Diego, CA, USA).

Sequence assembly

Illumina paired-end reads were filtered on the basis of quality values, and the low-quality bases (quality < 20, perror> 0.01) at upstream and downstream were trimmed. The remained clean data were de novo assembled by SOAPdenovo2 (http://soap.genomics.org.cn/soapdenovo.html) on the basis of overlapping and paired-end relationships. All the cleaned reads were also mapped onto the assembled contigs with Bowtie 2 (2.2.5) [27] to estimate the sequencing depth. Those contigs with sequencing depth over 30× were mapped to the Conoidea mitochondrial genomes that were downloaded from the NCBI non-redundant nucleotide database (Nt) with blastn (2.2.31+) to validate mito-contigs, and the remaining genome gaps were filled with a python script.

Genome confirmation was indispensable to perform after assembling. Finally, the paired-end clean reads were mapped onto the assembled genome with 100% coverage, and the insert-size matched the information of sequenced library. The sequencing depth, coverage and relationship of the paired-end reads were the main criteria for confirmation.

Genome annotation and analysis

Preliminary gene annotation was realized through the online program Dual Organellar GenoMe Annotator (DOGMA) [28] and ORF Finder [29] with invertebrate mitochondrial genetic codes and default parameters. To verify the exact gene and exon boundaries, putative nucleotide and protein sequences were BLAST searched in the public Nt and Nr (the NCBI Non-redundant protein) databases. All tRNA genes were further confirmed through online tRNAscan-SE and ARWEN search server [3032], in combination with the annotated results of ARAGORN. Graphical map of the circular plastome was drawn with Organellar Genome DRAW (OGDRAW v1.2) [33]. Relative synonymous codon usage (RSCU) value was employed to evaluate the synonymous codon bias in accordance with a previous report [34]. The skewing of nucleotide composition was calculated according to the following formulas: AT skew = (A–T)/(A+T) and GC skew = (G−C)/(G+C) [35, 36]. To further analyze the evolutionary adaptation in the Conus lineage, we applied DnaSP 6 [37] to estimate the ratios of non-synonymous (Ka) and synonymous (Ks) substitutions in the mitochondrial genomes among cone snails with the three different dietary types.

Phylogenetic analysis

Phylogenetic analysis was performed among ten taxa in the Conidae based on the nucleotide sequences of eleven protein-coding genes without ATP8 and ND6 from GenBank, and Oxymeris dimidiata (NC_013239.1), Fusiturris similis (NC_013242.1) and Lophiotoma cerithiformis (NC_008098.1) were employed as the out-groups. Before reconstructing phylogenetic trees, both nucleotide and protein sequences of eleven protein-coding genes were subjected to concatenated alignments using MUSCLE 3.8.31 (http://www.drive5.com/muscle/) [38]. The best-fit model GTR+G+I for nucleotide sequences was selected using the Akaike Information Criterion (AIC) with jModeltest [39]. Bayesian analyses of both nucleotide and protein alignments were carried out using PhyloBayes version 3.3f [40] under the best-fit model. Two independent Markov Chain Monte Carlo (MCMC) chains were run simultaneously to determine whether the searching reached stabilization, and were immediately stopped when all chains converged (maxdiff less than 0.1). The phylogenetic trees were eventually constructed using the Tree View program v.1.65 and Evolview (www.evolgenius.info/evolview/) [41].

Results and discussion

Genome organization and nucleotide composition

The mitogenome of C. quercinus is a closed circular molecule of 16,430 bp in length (GenBank accession No. KY609509; see more details in Fig 1). It encodes a high mutation region (D-loop), and a typical set of 37 mitochondrial genes including 13 protein-coding genes (PCGs), two ribosomal RNA (rRNA) genes (12S rRNA and 16S rRNA), and 22 transfer RNA (tRNA) genes. Eight tRNA genes are encoded on the light (L) strand, whereas the other genes are located on the heavy (H) strand (Table 1).

thumbnail
Fig 1. Mitochondrial map of Conus quercinus.

Genes outside the map are transcribed in a clockwise direction, whereas those inside the map are transcribed counterclockwise. Gene blocks are filled with different colors as the cut line shows.

https://doi.org/10.1371/journal.pone.0193053.g001

thumbnail
Table 1. The detailed mitogenome structure of Conus quercinus.

https://doi.org/10.1371/journal.pone.0193053.t001

In the C. quercinus mitogenome, gene overlapping occurred three times (the negative numbers in Table 1), spanning 1~7 nucleotides (nts), for a total of 9 nts. The intergenic spacer region occurred 20 times (the positive numbers in Table 1), spanning 1~162 bp, for a total of 421 bp. The overall base composition is estimated to be 28.15% for A, 38.32% for T, 14.82% for C and 18.71% for G, with a high A+T content at 66.47% (Table 2).

Complete mitogenomes of the ten Conus species, including C. quercinus and nine previously reported [1926], displayed moderate size variation, with the mean size of 15,755 bp (SD = 298.4, n = 10), ranging from 15,444 bp (C. californicus) to 16,430 bp (C. quercinus). AT and GC skews are measures of compositional asymmetry. In the C. quercinus mitogenome, GC-skew values are always positive, while the values of AT-skew are negative (Table 2).

The mitogenome gene arrangement is conserved and identical to those of other reported Conus species. The intergenic sequences vary between 0 and 41 nts, and one relatively large region of 162 nts happens between COX1 and COX2 (Fig 1). The gene sequences of NADH dehydrogenase subunit 4L (NAD4l) and subunit 4 (NAD4) are overlapped by 7 nts, NAD4 and tRNA-His by 1 nt, and NAD5 and tRNA-Phe by 1 nt.

Protein-coding, tRNA and rRNA genes

The 13 PCGs of C. quercinus are similar in length and arrangement to the nine previously sequenced Conus mitogenomes. All PCGs are transcribed from the H strand in C. quercinus, with initiation of the standard start codon ATG. They also display the typical TAN termination codon, in which 8 PCGs have the complete termination codon TAA and 5 PCGs have the TAG (Fig 1 and Table 1). The RSCU values of the C. quercinus mitogenome were calculated (Fig 2), indicating that TTA (Leu), TCT (Ser), GCT (Ala), ACT (Thr), and GTT (Val) are the five most frequently used codons.

thumbnail
Fig 2. RSCU values in the mitogenome of Conus quercinus.

Codon families are indicated below the X-axis.

https://doi.org/10.1371/journal.pone.0193053.g002

D-loop region of the C. quercinus mitogenome

The D-loop region between tRNA-Phe and COX3 in C. quercinus (Fig 1) is the longest (943 bp), which is much higher than those in other Conus species (97~698 bp; see more details in Table 3). Based on the sequence alignment of C. quercinus with other Conus species, we observed that the intergenic sequences of the C. quercinus mitogenome are also the longest in these Conus species. Usually, in animal mitochondrial genomes, the longest intergenic sequences were reported to play a key role in the initiation of replication and transcription [22, 42]. Interestingly, the D-loop region in C. quercinus (among the 10 Conus species except C. californicus) also presents the higher A+T content (71.3%) with a long AT tandem repeat stretch (68 bp; Fig 3).

thumbnail
Fig 3. Scheme of the D-loop region in C. quercinus compared with C. consors.

The region spans 943 bp and exhibits several outstanding motifs. The upper and lower red boxes denote the tRNA-Phe and COX3, and the blue box points to a long AT tandem repeat stretch.

https://doi.org/10.1371/journal.pone.0193053.g003

Synonymous and nonsynonymous substitutions

In genetics, the Ka/Ks ratio is of significance to estimate the balance between neutral mutations and is especially useful for understanding the evolutionary relations between homologous PCGs in closely related species [43]. To detect the influence of selection on the C. quercinus, Ka and Ks were estimated [44]. In all the 13 PCGs of three cone snails (Fig 4), the ratio of Ka/Ks is much less than 1 (between 0 and 0.16), indicating existence of a strong purifying or stabilizing selection.

thumbnail
Fig 4. Ka/Ks ratios for the 13 mitochondrial PCGs among three representative Conus species.

GenBank accession numbers: KY609509 for C. quercinus, NC_008797 for C. textile, and NC_030536 for C. striatus.

https://doi.org/10.1371/journal.pone.0193053.g004

The average Ka/Ks in ATP8 is the highest among the 13 PCGs, suggesting that this protein is under the least selective pressure among all the mitochondrial genes. Interestingly, in C. textile and C. striatus, the ratio of Ka/Ks is the least in nine protein-coding genes (except COX3, CYTB, ND4L and ND6) compared to C. quercinus, indicating that these two cone snails have a closer phylogenetic relationship than C. quercinus. These data are consistent with their dietary difference (Fig 5).

thumbnail
Fig 5. The phylogenetic tree of Conus species based on 13 complete mitogenome sequences.

GenBank accession numbers: NC_008098.1 for L. cerithiformis, NC_008797.1 for C. textile, NC_013239.1 for O. dimidiate, NC_013242.1 for F. similis, NC_013243.1 for C. borgesi, NC_023460.1 for C. consors, NC_027518.2 for C. tulipa, NC_027957.1 for C. tribblei, NC_030213.1 for C. gloriamaris, NC_030354.1 for C. capitaneus, NC_030536.1 for C. striatus, and NC_032377.1 for C. californicus. The red branch highlights our present study of C. quercinus. Please note that the outgroups L. cerithiformis, O. dimidiata and F. similis are allochthonous species. Branch lengths and topologies were obtained using the Bayesian inference analysis.

https://doi.org/10.1371/journal.pone.0193053.g005

Phylogenetic relationships of Conus species with different dietary types

Molecular phylogeny of the taxonomy is a hypothesis of its evolutionary patterns and processes. The molecular-based phylogenetic tree can estimate divergence times and ancestral distributions, and provides evidence relevant to taxonomic hypotheses [8]. To further validate the mitogenome sequence of C. quercinus and understand the evolutionary history of the Conus species with different feeding ecologies, we constructed a phylogenetic tree using Bayesian inference analysis with 13 complete mitogenomes downloaded from the NCBI (Fig 5). It is obvious that three different dietary types (vermivorous, molluscivorous and piscivorous) of cone snails, except the broad C. californicus, are clustered separately. This is consistent with previous reports [43, 45] and supports the putative hypothesis that the cone snail ancestor was vermivorous. However, the inclusion of C. californicus in the phylogenetic tree analysis may bias the results, because it is often regarded as an atypical member of Conidae due to its extremely broad diet and distant phylogenetic relationship to the rest of Conidae [14, 26].

Venom composition across Conus species has been hypothesized to be shaped by prey type and dietary breadth. Several studies [14] proved a significant positive relationship between dietary breadth and measures of conotoxin complexity by transcriptome sequencing of venom duct from 12 Conus species. However, given the high evolutionary lability of venom toxins, it is unclear that a clear relationship between dietary preference and venom composition should be expected. The prey taxonomic class of Conus can be predicted by venom components, but the performance of prey taxonomic class in predicting venom components was poor [1417, 26]. By far, we are certain that the selective pressures driven by diets play a major role in shaping evolutionary patterns in venom across the cone snails.

Conclusion

In this present study, we sequenced and annotated the complete mitogenome of C. quercinus. We used other nine publically available Conus mitogenomes to illustrate the structure of C. quercinus mitogenome and investigated the evolutionary relationships among Conus species. Interestingly, the mitochondrial gene arrangement of C. quercinus is highly conserved and identical to other Conus species. However, the D-loop region (943 bp) of C. quercinus is the longest with the higher A+T content (71.3%) and a long AT tandem repeat stretch (68 bp). The phylogenetic tree (Fig 5) revealed that three different dietary types of cone snails are clustered separately, suggesting that the phylogenetics of cone snails is related to their dietary types. Our current work improves our understanding of the mitogenomic structure and evolutionary status of the vermivorous C. quercinus, which support the putative hypothesis that the Conus ancestor was vermivorous.

Acknowledgments

This work was supported by Natural Science Foundation of Hainan Province (No. 317170), Shenzhen Dapeng Special Project for Industrial Development (No. PT20170302) and National Natural Science Foundation of China (No. 81560611).

References

  1. 1. Prashanth JR, Dutertre S, Jin AH, Lavergne V, Hamilton B, Cardoso FC, et al. The role of defensive ecological interactions in the evolution of conotoxins. Mol. Ecol. 2016; 25: 598–615. pmid:26614983
  2. 2. Himaya SW, Jin AH, Dutertre S, Giacomotto J, Mohialdeen H, Vetter I, et al. Comparative Venomics Reveals the Complex Prey Capture Strategy of the Piscivorous Cone Snail Conus catus. J. Proteom. Res. 2015; 14:4372–4381.
  3. 3. Robinson SD, Safavi-Hemami H, McIntosh LD, Purcell AW, Norton RS, Papenfuss AT. Diversity of conotoxin gene superfamilies in the venomous snail, Conus victoriae. PLoS ONE 2014; 9: e87648. pmid:24505301
  4. 4. Kumar PS, Kumar DS, Umamaheswari S. A perspective on toxicology of Conus venom peptides. Asian Pac. J. Trop. Med. 2015; 8:337–351. pmid:26003592
  5. 5. Peng C, Yao G, Gao BM, Fan CX, Bian C, Wang J, et al. High-throughput identification of novel conotoxins from the Chinese tubular cone snail (Conus betulinus) by multi-transcriptome sequencing. Gigascience. 2016; 5:17. pmid:27087938
  6. 6. Barghi N, Concepcion GP, Olivera BM, Lluisma AO. High conopeptide diversity in Conus tribblei revealed through analysis of venom duct transcriptome using two high-throughput sequencing platforms. Mar. Biotechnol. 2015; 17:81–98. pmid:25117477
  7. 7. Bouchet P, Kantor Yu I, Sysoev A, Puillandre N. A new operational classification of the Conoidea (Gastropoda). J Molluscan Stud. 2011; 77: 273–308.
  8. 8. Puillandre N, Bouchet P, Duda TF, Kauferstein S, Kohn AJ, Olivera BM, Watkins M, Meyer C. Molecular phylogeny and evolution of the cone snails (Gastropoda, Conoidea). Mol Phylogenet Evol. 2014;78: 290–303. pmid:24878223
  9. 9. Dutertre S, Jin AH, Kaas Q, Jones A, Alewood PF, Lewis RJ. Deep venomics reveals the mechanism for expanded peptide diversity in cone snail venom. Mol. Cell Proteom. 2013; 12:312–329.
  10. 10. Prashanth JR, Brust A, Jin AH, Alewood PF, Dutertre S, Lewis RJ. Cone snail venomics: From novel biology to novel therapeutics. Future Med. Chem. 2014; 6:1659–1675. pmid:25406006
  11. 11. Lavergne V, Harliwong I, Jones A, Miller D, Taft RJ, Alewood PF. Optimized deep-targeted proteotranscriptomic profiling reveals unexplored Conus toxin diversity and novel cysteine frameworks. Proc Natl Acad Sci USA. 2015; 112: E3782–3791. pmid:26150494
  12. 12. Safavi-Hemami H, Li Q, Jackson RL, Song AS, Boomsma W, Bandyopadhyay PK, et al. Rapid expansion of the protein disulfide isomerase gene family facilitates the folding of venom peptides. Proc Natl Acad Sci USA. 2016; 113:3227–3232. pmid:26957604
  13. 13. Aman JW, Imperial JS, Ueberheide B, Zhang MM, Aguilar M, Taylor D, et al. Insights into the origins of fish hunting in venomous cone snails from studies of Conus tessulatus. Proc Natl Acad Sci USA. 2016; 112: 5087–5092.
  14. 14. Phuong MA, Mahardika GN, Alfaro ME. Dietary breadth is positively correlated with venom complexity in cone snails. BMC Genomics. 2016; 17:401. pmid:27229931
  15. 15. Olivera BM, Seger J, Horvath MP, Fedosov AE. Prey-capture strategies of fish-hunting Cone snails: behavior, neurobiology and evolution. Brain Behav Evol. 2015; 86: 58–74. pmid:26397110
  16. 16. Jin AH, Dutertre S, Kaas Q, Lavergne V, Kubala P, Lewis RJ, et al. Transcriptomic messiness in the venom duct of Conus miles contributes to conotoxin diversity. Mol. Cell. Proteom. 2013;12: 3824–3833.
  17. 17. Remigio EA, Duda TF Jr. Evolution of ecological specialization and venom of a predatory marine gastropod. Molecular ecology. 2008; 17: 1156–1162. pmid:18221274
  18. 18. Modica MV, Puillandre N, Castelin M, Zhang Y, Holford M. A Good Compromise: Rapid and robust species proxies for inventorying biodiversity hotspots using the Terebridae (Gastropoda: Conoidea). PLoS One. 2014; 9: e102160. pmid:25003611
  19. 19. Brauer A, Kurz A, Stockwell T, Baden-Tillson H, Heidler J, Mebs D, et al. The mitochondrial genome of the venomous cone snail Conus consors. PLoS One. 2012; 7: e51528. pmid:23236512
  20. 20. Chen PW, Hsiao ST, Chen KS, Tseng CT, Wu WL, Hwang DF. The complete mitochondrial genome of Conus striatus (Neogastropoda: Conidae). Mitochondrial DNA Part B-Resources. 2016; 1: 493–494.
  21. 21. Chen PW, Hsiao ST, Huang CW, Chen KS, Tseng CT, Wu WL, et al. The complete mitochondrial genome of Conus tulipa (Neogastropoda: Conidae). Mitochondrial DNA A DNA Mapp Seq Anal. 2016; 27: 2738–2739. pmid:26057007
  22. 22. Bandyopadhyay PK, Stevenson BJ, Ownby JP, Cady MT, Watkins M, Olivera BM. The mitochondrial genome of Conus textile, coxI-coxII intergenic sequences and Conoidean evolution. Mol Phylogenet Evol. 2008; 46: 215–223. pmid:17936021
  23. 23. Barghi N, Concepcion GP, Olivera BM, Lluisma AO. Characterization of the complete mitochondrial genome of Conus tribblei Walls, 1977. Mitochondrial DNA A DNA Mapp Seq Anal. 2016; 27: 4451–4452. pmid:26470735
  24. 24. Cunha RL, Grande C, Zardoya R. Neogastropod phylogenetic relationships based on entire mitochondrial genomes. BMC Evol Biol. 2009; 9: 210. pmid:19698157
  25. 25. Chen PW, Hsiao ST, Chen KS, Tseng CT, Wu WL, Hwang DF. The complete mitochondrial genome of Conus capitaneus (Neogastropoda: Conidae). Mitochondrial DNA Part B Resources. 2016; 1: 520–521.
  26. 26. Uribe JE, Puillandre N, Zardoya R. Beyond Conus: Phylogenetic relationships of Conidae based on complete mitochondrial genomes. Mol Phylogenet Evol. 2017; 107: 142–151. pmid:27794464
  27. 27. Langmead B, Salzberg S. Fast gapped-read alignment with Bowtie 2. Nature Methods. 2012; 9: 357–359. pmid:22388286
  28. 28. Wyman SK, Jansen RK, Boore JL. Automatic annotation of organellar genomes with DOGMA. Bioinformatics. 2004; 20: 3252–3255. pmid:15180927
  29. 29. Cheng J, Zeng X, Ren G, Liu Z. CGAP: a new comprehensive platform for the comparative analysis of mitochondrial genomes. BMC Bioinformatics. 2013; 14:95. pmid:23496817
  30. 30. Schattner P, Brooks AN, Lowe TM. The tRNAscan-SE, snoscan and snoGPS web servers for the detection of tRNAs and snoRNAs. Nucleic Acids Res. 2005; 33: W686–689. pmid:15980563
  31. 31. Griffiths-Jones S, Bateman A, Marshall M, Khanna A, Eddy SR. Rfam: an RNA family database. Nucleic Acids Res. 2003; 31: 439–441. pmid:12520045
  32. 32. Abe T, Ikemura T, Sugahara J, Kanai A, Ohara Y, Uehara H, et al. tRNADB-CE 2011: tRNA gene database curated manually by experts. Nucleic Acids Res. 2011; 39: D210–213. pmid:21071414
  33. 33. Lohse M, Drechsel O, Bock R. Organellar Genome DRAW (OGDRAW): a tool for the easy generation of high-quality custom graphical maps of plastid and mitochondrial genomes. Curr Genet. 2007; 52: 267–274. pmid:17957369
  34. 34. Sharp PM, Li WH. The codon Adaptation Index-a measure of directional synonymous codon usage bias, and its potential applications. Nucleic Acids Res. 1987; 15: 1281–1295. pmid:3547335
  35. 35. Lobry JR. Asymmetric substitution patterns in the two DNA strands of bacteria. Mol Biol Evol. 1996; 13:660–665. pmid:8676740
  36. 36. Perna NT, Kocher TD. Patterns of nucleotide composition at fourfold degenerate sites of animal mitochondrial genomes. J Mol Evol. 1995; 41: 353–358. pmid:7563121
  37. 37. Rozas J, Ferrer-Mata A, Carlos Sánchez-DelBarrio J, Guirao-Rico S, Librado P, Ramos-Onsins SE, Sánchez-Gracia A. DnaSP 6: DNA sequence polymorphism analysis of large data sets. Mol Biol Evol. 2017; 34: 3299–3302. pmid:29029172
  38. 38. Edgar RC. MUSCLE: multiple sequence alignment with high accuracy and high throughput. Nucleic Acids Res. 2004; 32: 1792–1797. pmid:15034147
  39. 39. Posada D. jModelTest: phylogenetic model averaging. Mol Biol Evol, 2008; 25(7), 1253–1256. pmid:18397919
  40. 40. Lartillot N, Lepage T, Blanquart S. PhyloBayes 3: a Bayesian software package for phylogenetic reconstruction and molecular dating. Bioinformatics. 2009; 25: 2286–2288. pmid:19535536
  41. 41. He Z, Zhang H, Gao S, Lercher MJ, Chen WH, Hu S. Evolview v2: an online visualization and management tool for customized and annotated phylogenetic trees. Nucleic Acids Res. 2016; 44: 236–241.
  42. 42. Pastukh VM, Gorodnya OM, Gillespie MN, Ruchko MV. Regulation of mitochondrial genome replication by hypoxia: The role of DNA oxidation in D-loop region. Free Radic Biol Med. 2016; 96: 78–88. pmid:27091693
  43. 43. Xing Y, Lee C. Assessing the application of Ka/Ks ratio test to alternatively spliced exons. Bioinformatics. 2005; 21: 3701–3703. pmid:16091412
  44. 44. Sahyoun AH, Bernt M, Stadler PF, Tout K. GC skew and mitochondrial origins of replication. Mitochondrion. 2014;17: 56–66. pmid:24911382
  45. 45. Duda T. Origins of diverse feeding ecologies within Conus, a genus of venomous marine gastropods. Biological Journal of the Linnean Society. 2001; 73: 391–409.