The complete mitogenome of Arion vulgaris Moquin-Tandon, 1855 (Gastropoda: Stylommatophora): mitochondrial genome architecture, evolution and phylogenetic considerations within Stylommatophora

View article
Bioinformatics and Genomics

Introduction

Gastropoda is the most speciose class of Mollusca, including snails and slugs with very diverse feeding habits and a wide range of habitats (Barker, 2009). The about 63,000 gastropod species represent 476 families (Bouchet et al., 2017) and radiated in marine, freshwater and terrestrial ecosystems with detritivorous, herbivorous, carnivorous, predatory or parasitic life styles (Ponder & Lindberg, 1997). Most of the terrestrial gastropods are stylommatophoran pulmonates, with approximately 30,000 species distributed from polar to tropical regions (Mordan & Wade, 2008). Stylommatophorans are economically important as human food and because of their status of being major agricultural pests and/or vectors of parasites and invasiveness (Barker, 2009). The origin of Stylommatophora is within panpulmonate heterobranchs (Jörger et al., 2010) and the monophyly of the order is undisputed. Internal phylogenetic relationships of stylommatophorans were poorly resolved based on morphology but then investigated molecularly in different sampling sets of taxa with various methods and basically relatively short sequences. Tillier, Masselmot & Tillirt (1996) used the D2 region of 28S rRNA to explore the phylogenetic relationships of pulmonates including a few stylommatophoran species; however, they reported that these short sequences would not have sufficient resolving power for investigating the relationships owing to the probable rapid radiation of pulmonate species. Wade, Mordan & Clarke (2001) and Wade, Mordan & Naggs (2006) presented more comprehensive molecular phylogenies based on relatively longer sequence information of the rRNA gene-cluster using 104 species (Wade, Mordan & Clarke, 2001) and 160 species (Wade, Mordan & Naggs, 2006) from Stylommatophora. Although these phylogenetic reconstructions accurately supported the monophyly of achatinoid and non-achatinoid clades, some clades of families that traditionally have been assumed to be monophyletic and some of the morphological groups based on excretory system; in particular, monophyly of some families and morphological groups were not supported.

The emergence and divergence time of Stylommatophora is also doubtful due to the fragmentary fossil records. The earliest land snails identified as stylommatophoran species are from upper Carboniferous and Permian but their classification has still been controversial (Solem & Yochelson, 1979; Hausdorf, 2000). Bandel (1991) and Roth et al. (1996) suggested the oldest known fossil records from late Jurassic and Early Cretaceous (Cheruscicola) and Early Cretaceous (Pupilloidea). Tillier, Masselmot & Tillirt (1996) inferred that the Stylommatophora emerged in the transition between late Cretaceous and Paleocene (65–55 Ma) congruent with fossil records, based on the molecular data. However, all of the previous molecular dating analyses on Stylommatophora have been performed either with limited numbers of taxa or molecular markers (Tillier, Masselmot & Tillirt, 1996; Jörger et al., 2010; Dinapoli & Klussmann-Kolb, 2010; Zapata et al., 2014), therefore there is a need for further investigations in a more comprehensive sampling using more markers for better understanding of the phylogeny and timing of evolution of Stylommatophora.

In recent years, there has been a rapid increase in the number of sequenced mitochondrial genomes (mitogenomes) in parallel to revolution on high throughput DNA sequencing technology and data mining, providing a powerful tool for phylogenetic analysis (Moritz, Dowling & Brown, 1987; Boore, 1999; Bernt et al., 2013a). Animal mitogenomes are double-stranded circular molecules which are ∼16 kb in length and contain 13 protein coding genes (PCGs) forming the respiratory chain complexes: Complex I or NADH: ubiquinone oxidoreductase contains seven subunits of NADH dehydrogenase (ND1–6 and ND4L), complex III or ubiquinol: cytochrome c oxidoreductase consists of cytochrome b (CYTB), complex IV or cytochrome c oxidase comprises three subunits of cytochrome c oxidase (COX1COX3) and complex V or ATP synthase includes two subunits of the ATPase (ATP6 and ATP8). The mitochondrial PCGs have generally been supposed to be evolving under neutral or nearly neutral selection (Ballard & Kreitman, 1995). Although it has been suggested that these genes are likely to be under strong purifying selection considering their functional importance, the selective pressures might vary even among closely related species and be influenced by environmental conditions (Meiklejohn, Montooth & Rand, 2007). The mitogenomes also encode the small and large subunit rRNAs (rrnL and rrnS) and twenty-two tRNA genes for the translation process of PCGs. In general, they harbour a single large non-coding region containing control elements necessary for replication and transcription (Boore, 1999). Mitogenomes have become widely used tools in recent phylogeny, phylogeography and molecular dating analyses in various taxa, because of their (1) relatively small size, (2) the high copy number, (3) maternal inheritance type and (4) relatively rapid rate of evolutionary change (Moritz, Dowling & Brown, 1987; Gray, 1989). The sequence information of mitogenomes has also been used in reconstructing phylogenies of several taxonomic groups within/including Gastropoda (White et al., 2011; Stöger & Schrödl, 2013; Sevigny et al., 2015; Uribe et al., 2016a; Uribe et al., 2016b; Romero, Weigand & Pfenninger, 2016; Yang et al., 2019). Although there have been some criticisms about the usage of mitogenomes in construction of gastropod phylogeny because of long branch attraction, substitution saturation and strand-specific skew bias (Stöger & Schrödl, 2013), within the recently diversified lineages of gastropods, the use of mitogenomes resulted in highly resolved phylogenies (Williams, Foster & Littlewood, 2014; Osca, Templado & Zardoya, 2014). Besides the use of the mitogenome in sequence-based phylogenies, mitogenome rearrangements can also provide phylogenetic signals (Grande, Templado & Zardoya, 2008; Stöger & Schrödl, 2013; Xie et al., 2019b). Although the mitogenome is widely used in phylogeny of many gastropod groups (Arquez, Colgan & Castro, 2014; Osca, Templado & Zardoya, 2014; Sevigny et al., 2015; Uribe, Zardoya & Puillandre, 2018), there are limited numbers of reported stylommatophoran mitogenomes and phylogenetic studies in Stylommatophora in terms of usage of mitogenome sequence and rearrangement (Romero, Weigand & Pfenninger, 2016; Xie et al., 2019a; Yang et al., 2019). To date, complete or nearly complete mitogenomes have been reported for only 34 stylommatophoran species (NCBI, September, 2019).

In this study, we sequenced and annotated the complete mitogenome of Arion vulgaris Moquin-Tandon, 1855 (Gastropoda: Stylommatophora), which is considered as a serious invasive pest both in agriculture and private gardens. We compared it with the mitogenome of its congener Arion rufus (Linnaeus, 1758), and with all other previously reported stylommatophoran mitogenomes. We also reconstructed a phylogeny from stylommatophoran mitogenomes to estimate the phylogenetic position of A. vulgaris and to test the informativeness of mitogenome data in the reconstruction of Stylommatophora phylogeny. In addition, we obtained a dated phylogeny using this mitogenome dataset and fossil calibrations to estimate divergence times within Stylommatophora. Furthermore, selection analyses were performed to investigate the role of different selective forces shaping stylommatophoran mitogenomes. Finally, we compared the mitogenome organisations of stylommatophoran species using a comparative and phylogeny based method and tried to uncover the evolutionary pathways of mitogenome rearrangements.

Materials and Methods

Specimen collection and DNA extraction

The specimen of A. vulgaris was collected from the garden of the Zoologische Staatssammlung München (ZSM), Germany. Total genomic DNA was extracted from mantle tissue using CTAB method (Doyle & Doyle, 1987).

Mitogenome sequencing, annotation and analyses

The whole-genome sequencing was conducted with 150 bp pair-end reads on the Illumina Hiseq4000 Platform (Illumina, San Diego, CA) using 350 bp insert size libraries. Raw reads were processed by removing low quality reads, adapter sequences and possible contaminated reads using Fastp v0.20.0 (Chen et al., 2018) and Lighter v1.0.7 (Song, Florea & Langmead, 2014). In total, about 7.5G high quality base pairs of sequence data were obtained and the mitogenome was assembled using the MitoZ software (Meng et al., 2019), followed by manual curation using Geneious R9 (Kearse et al., 2012).

The annotation of tRNA genes of the A. vulgaris mitogenome was performed using MITOS (http://mitos.bioinf.uni-leipzig.de/index.py) (Bernt et al., 2013b) and ARWEN web servers (Laslett & Canbäck, 2008) based on their secondary structures and anticodon sequences. The locations and boundaries of PCGs and rRNA genes were identified manually by comparing with the A. rufus (KT626607) homologous gene sequences. The visualization of the secondary structure of tRNA genes was performed using VARNA v3-93 (Darty, Denise & Ponty, 2009) and RNAviz 2.0.3 (De Rijk, Wuyts & De Wachter, 2003). Intergenic spacers and overlapping regions between genes were estimated manually. The largest non-coding region was defined as control region and the Mfold server (Zuker, 2003) was used to predict the secondary structure of this region. The “palindrome” tool within the European Molecular Biology Open Software Suite (EMBOSS) (Rice, Longden & Bleasby, 2000) was used for searching the palindromic sequences in the control region. Finally, the complete mitogenome of A. vulgaris was deposited in GenBank under accession number MN607980. The mitogenome of A. vulgaris is visualized using OrganellarGenomeDRAW (OGDRAW) (Greiner, Lehwark & Bock, 2019).

The nucleotide compositions, average nucleotide and amino acid sequence divergences and the relative synonymous codon usages (RSCU) of PCGs were computed using MEGA v7.0 (Kumar, Stecher & Tamura, 2016). The strand asymmetries were calculated according to the following formulas: AT-skew = [A − T]/[A + T] and GC-skew = [G − C]/[G + C] (Perna & Kocher, 1995).

Phylogenetic and comparative analyses

Alignment and model selection

Phylogenetic and comparative analyses were performed using the mitogenome dataset of 35 stylommatophoran species representing 18 families, and using one species from Systellommatophora, one species from Hygrophila, and one species from Ellobioidea as outgroups (Table 1). Each tRNA and rRNA gene was aligned individually using MAFFT (Katoh & Standley, 2013) algorithm in Geneious R9 (Kearse et al., 2012). The alignment of nucleotide sequences of each PCG was performed using MAFFT algorithm and the “translation align” option implemented in Geneious R9. The final alignment files were then concatenated using SequenceMatrix v.1.7.8 (Vaidya, Lohman & Meier, 2011). The optimal partitioning scheme and substitution models were inferred by PartitionFinder v1.1.1 (Lanfear et al., 2012) using the Bayesian Information Criterion (BIC) and the “greedy” algorithm with the option of “unlinked” branch lengths. The best-fit partitioning scheme and nucleotide substitution models were used in phylogenetic analyses (Table S1).

Table 1:
List of stylommatophoran mitogenomes used in phylogenetic and comparative analyses.
Species Family Accession number References
STYLOMMATOPHORA Arion vulgaris Arionidae MN607980 This study
Arion rufus Arionidae KT626607 Romero, Weigand & Pfenninger (2016)
Achatinella fulgens Achatinellidae MG925058 Price et al. (2018)
Achatinella mustelina Achatinellidae NC030190 Price et al. (2016a)
Achatinella sowerbyana Achatinellidae KX356680 Price et al. (2016b)
Partulina redfieldi Achatinellidae MG925057 Price et al. (2018)
Achatina fulica Achatinidae KM114610 He et al. (2016)
Deroceras reticulatum Agriolimacidae NC035495 Ahn et al. (2017)
Aegista aubryana Bradybaenidae NC029419 Yang et al. (2016)
Aegista diversifamilia Bradybaenidae NC027584 Huang, Lin & Wu (2015)
Dolicheulota formosensis Bradybaenidae NC027493 Huang, Lin & Wu (2015)
Mastigeulota kiangsinensis Bradybaenidae NC024935 Deng et al. (2016)
Camaena cicatricosa Camaenidae NC025511 Wang et al. (2014)
Camaena poyuensis Camaenidae KT001074 Unpublished
Cerion incanum Cerionidae NC025645 González et al. (2016)
Cerion tridentatum costellata Cerionidae KY249249 Unpublished
Cerion uva Cerionidae KY124261 Harasewych et al. (2017)
Albinaria caerulea Clausiliidae NC001761 Hatzoglou, Rodakis & Lecanidou (1995)
Gastrocopta cristata Gastrocoptidae NC026043 Unpublished
Cernuella virgata Geomitridae NC030723 Lin et al. (2016)
Helicella itala Geomitridae KT696546 Romero, Weigand & Pfenninger (2016)
Cepaea nemoralis Helicidae NC001816 Yamazaki et al. (1997)
Cylindrus obtusus Helicidae NC017872 Groenenberg et al. (2012)
Cornu aspersum Helicidae NC021747 Gaitán-Espitia, Nespolo & Opazo (2013)
Helix pomatia Helicidae NC041247 Korábek, Petrusek & Rovatsos (2019)
Orcula dolium Orculidae NC034782 Groenenberg et al. (2017)
Naesiotus nux Orthalicidae NC028553 Hunter et al. (2016)
Meghimatium bilineatum Philomycidae NC035429 Xie et al. (2019a) and Xie et al. (2019b)
Philomycus bilineatus Philomycidae MG722906 Yang et al. (2019)
Polygyra cereolus Polygyridae NC032036 Unpublished
Praticolella mexicana Polygyridae KX240084 Minton et al. (2016)
Pupilla muscorum Pupillidae NC026044 Unpublished
Succinea putris Succineidae NC016190 White et al. (2011)
Microceramus pontificus Urocoptidae NC036381 Unpublished
Vertigo pusilla Vertiginidae NC026045 Unpublished
Ellobioidea Carychium tridentatum Ellobiidae KT696545 Romero, Weigand & Pfenninger (2016)
Hygrophila Galba pervia Lymnaeidae NC018536 Liu et al. (2012)
Systellommatophora Platevindex mortoni Onchidiidae GU475132 Sun et al. (2016)
DOI: 10.7717/peerj.8603/table-1

Assessing the substitution saturation level

The substitution saturation levels in different genes and codon positions were estimated comparing the uncorrected p-distances and the distances calculated by applying the GTR + G + I evolutionary model selected based on the BIC using jModelTest v2.1.7 (Darriba et al., 2012). All genetic distances were computed with PAUP v4.0 b10 (Swofford, 2002).

Phylogenetic reconstruction

Two different datasets were created for phylogenetic analyses to test the influence of saturated genes and codon positions: (1) 13 PCGs including all codon positions plus the 22 tRNAs and two rRNAs (P123RNA) and (2) PCGs excluding the five saturated genes and third codon positions, plus 22 tRNAs and two rRNAs (8P12RNA, Table S2). Maximum likelihood (ML) trees were constructed with RAxML v8.0.9 (Stamatakis, 2014) implemented in Geneious R9 applying the best-fit evolutionary model for each partition under 1,000 bootstrap replicates. For Bayesian Inference (BI) analyses, MrBayes v3.2.2 (Ronquist et al., 2012) was employed with two independent runs of 10 million generations with four Markov chains (three cold, one heated), sampling every 1,000 generations and a burn-in of 25% trees. The stationarity of the chains was assessed using the program Tracer v1.7 (Rambaut et al., 2018). The consensus phylogenetic trees were visualized using FigTree v1.4.0 (Rambaut, 2012).

Divergence time estimation

MCMCTree program implemented in the Phylogenetic Analysis by Maximum Likelihood (PAML) package v4.9 (Yang, 2007) was used for Bayesian estimation of divergence times of each species. Substitution rate per site was estimated by BASEML and was used to set the prior for the mean substitution rate in the Bayesian analysis. MCMC was run by 50 × 10,000 iterations with the REV substitution model. The soft bounds of Helix pomatia + Cornu aspersum [divergence time between 34 million years ago (Ma) and 42 Ma], Mastigeulota kiangsinensis + (Dolicheulota formosensis + (Aegista aubryana + Aegista diversifamilia)) (divergence time between 25 Ma and 51 Ma), and Camaena cicatricosa + Camaena poyuensis (divergence time between 16 Ma and 39 Ma) were used as external calibrations (Razkin et al., 2015) and the estimated nodal age of Tectipleura [244 Ma (210–279 Ma)] was used for the calibration of the root (Kano et al., 2016).

Selection analyses

The CODEML implemented in PAML was used to estimate the ratio of nonsynonymous/synonymous substitution rate (ω = dN/dS) and to explore the role of different selective constraints working on each PCG under the one-ratio model (Model A: model = 0, NSsites = 0, fix_omega = 0, omega = 1). Gaps and ambiguous sites of sequence alignments were included in the analyses. For each PCG, likelihood ratio tests (LRTs) were used to compare the null neutral model (Model B: model = 2, NSsites = 2, fix_omega = 1, omega = 1) against alternative models of branch-specific positive selection (Model C: model = 2, NSsites = 2, fix_omega = 0, omega = 1.5). The Bayes Empirical Bayes (BEB) algorithm in CODEML was used to detect the positively selected sites. Furthermore, the adaptive branch-site random effects likelihood (aBSREL) (Smith et al., 2015) implemented in DATAMONKEY webserver (Weaver et al., 2018) was used to search the signatures of episodic positive diversifying selection testing each branch. In addition, mixed effects model of evolution (MEME) (Murrell et al., 2012) was used to detect episodic or diversifying selection at individual sites and a fast, unconstrained Bayesian approximation for inferring selection (FUBAR) (Murrell et al., 2013) was used for providing additional support to the detection of sites evolving under positive or negative selection. Each PCG was also evaluated in terms of properties and magnitude of amino acid changes using TreeSAAP v3.2 (Woolley et al., 2003), which uses 31 properties of amino acids and categorizes the degree of substitutions to eight categories (1–8).

Comparison of mitogenome organizations

Mitogenome organizations and gene rearrangements of stylommatophoran species were analysed via the CREx web server (http://pacosy.informatik.uni-leipzig.de/crex) (Bernt et al., 2007). The gene orders of ancestral nodes were reconstructed using the Maximum Likelihood for Gene Order Analysis (MLGO, http://geneorder.org/) (Hu, Lin & Tang, 2014) with the input tree obtained by phylogenetic approaches, and the orders of the protein coding, rRNA and tRNA genes were compared with the inferred ancestral mitogenomes. A distance matrix was calculated based on number of common intervals, and the output diagram visually examined to identify shared and/or derived gene rearrangements as well as mechanisms of rearrangements.

Results and Discussion

Mitogenome characteristics and nucleotide composition

The complete mitogenome sequence of A. vulgaris was obtained with a length of 14,547 bp (Table 2) and its size was within the range of the those of other reported stylommatophoran mitogenomes, varying between 13,797 bp in Camaena poyuensis and 16,879 bp in Partulina redfieldi (Price et al., 2018). It includes the entire set of 37 mitochondrial genes: 13 PCGs, 22 tRNAs and two rRNAs. Twenty-four genes were located on the J strand, while the remainings were encoded by the opposite N strand (Table 2, Fig. 1).

Table 2:
Mitogenome summary of Arion vulgaris.
Gene Strand From To Size Start codon Stop codon Anticodon IGN
COX1 J 1 1,530 1,530 TTG TAG 4
tRNA-Val J 1,535 1,600 66 UAC 0
16S rRNA J 1,601 2,613 1,013 0
tRNA-Leu J 2,614 2,675 62 UAG 11
tRNA-Pro J 2,687 2,752 66 UGG 13
tRNA-Ala J 2,766 2,831 66 UGC 7
ND6 J 2,839 3,312 474 ATG TAG −41
ND5 J 3,272 4,960 1,689 ACA TAA −10
ND1 J 4,951 5,853 903 ATG TAG 15
ND4L J 5,869 6,163 295 ATA T- −15
CYTB J 6,149 7,228 1,080 ATG TAA −2
tRNA-Asp J 7,227 7,296 70 GUC 10
tRNA-Cys J 7,307 7,363 57 GCA 0
tRNA-Phe J 7,364 7,425 62 GAA 0
COX2 J 7,426 8,094 669 ATG TAG 1
tRNA-Trp J 8,096 8,160 65 UCA 91
tRNA-Tyr J 8,252 8,315 67 GUA 0
Control region J 8,316 8,685 370 0
tRNA-Gly J 8,686 8,763 78 UCC −20
tRNA-His J 8,744 8,809 66 GUG −3
tRNA-Glu N 8,807 8,873 67 UUC 5
tRNA-Gln N 8,879 8,942 64 UUG 0
12S rRNA N 8,943 9,689 747 0
tRNA-Met N 9,690 9,754 65 CAU 0
tRNA-Leu N 9,755 9,820 66 UAA −32
ATP8 N 9,789 9,971 183 GTG TAA 0
tRNA-Asn N 9,972 10,033 62 GUU −8
ATP6 N 10,026 10,688 663 ATA TAA −9
tRNA-Arg N 10,680 10,746 67 UCG 3
ND3 N 10,750 11,094 345 ATG TAA 13
tRNA-Ser2 N 11,108 11,176 69 UGA 49
tRNA-Ser1 J 11,226 11,283 58 GCU 36
ND4 J 11,320 12,633 1,314 ATA TAG −18
tRNA-Thr N 12,616 12,681 66 UGU 0
COX3 N 12,682 13,462 781 ATG T- 41
tRNA-Ile J 13,504 13,567 64 GAU 1
ND2 J 13,569 144,86 918 ATG TAA 0
tRNA-Lys J 14,487 6 67 UUU −6
DOI: 10.7717/peerj.8603/table-2

Notes:

J

major

N

minor

IGN

intergenic nucleotides

Minus indicates overlapping sequences between adjacent genes.

The nucleotide composition of A. vulgaris mitogenome was distinctly biased towards A and T, with a 70.20% A + T content, and comparable to other reported stylommatophoran mitogenomes, varying between 59.79% A + T in Cepea nemoralis (Yamazaki et al., 1997) and 80.07% A + T in Achatinella mustelina (Price et al., 2016a) (Table 3 and Table S3). A bias towards A and T nucleotides was also observed in PCGs of the A. vulgaris mitogenome with a 69.34% A + T content (Table 3). The A + T content of the 3rd codon position (79.64%) was higher than those of the 2nd (64.21%) and 1st codon positions (64.18%). Similar to other reported stylommatophoran mitogenomes (Table S3), the AT- and GC-skews were found slightly negative (−0.0756) and positive (0.0431) in the whole mitogenome of A. vulgaris, respectively. A pronounced T and G skew was also observed in all PCGs (−0.1508, 0.0472), PCGs on the majority strand (−0.1447, 0.0596), and tRNA genes (−0.0010, 0.1582) (Table 3). The T- and G-skewed mitogenome of A. vulgaris might be explained by the spontaneous deamination of cytosine during replication and transcription processes (Reyes et al., 1998). The PCGs encoded on the minority strand displayed a T- and C-skewed pattern (−0.1783 AT-skew, −0.0065 GC-skew), contrary to the expected high rates of Ts and Gs on the minority strand for most of the metazoans (Hassanin, Léger & Deutsch, 2005).

Mitogenome organization of Arion vulgaris.

Figure 1: Mitogenome organization of Arion vulgaris.

Genes transcribed from the J and N strands are shown outside and inside of the circle, respectively. PCGs coding complex I, complex III, complex IV and complex V components are marked with yellow, purple, pink and green, respectively. rRNA genes are coloured with red and the putative control region is coloured with cyan, while tRNA genes are coloured with dark blue and labelled by the single letter amino acid code.
Table 3:
Nucleotide composition of the Arion vulgaris mitogenome.
Feature T% C% A% G% A + T% AT-skew GC-skew
Whole mitogenome 37.75 14.26 32.45 15.54 70.20 −0.076 0.043
Protein coding genes 39.90 14.61 29.44 16.05 69.34 −0.151 0.047
First codon position 33.54 14.01 30.64 21.81 64.18 −0.045 0.218
Second codon position 45.65 19.42 18.55 16.38 64.21 −0.422 −0.085
Third codon position 40.50 10.39 39.14 9.97 79.64 −0.017 −0.021
Protein coding genes-J 39.78 14.34 29.72 16.16 69.50 −0.145 0.060
First codon position-J 32.94 13.93 31.23 21.90 64.17 −0.027 0.222
Second codon position-J 45.91 19.07 18.53 16.49 64.44 −0.425 −0.073
Third codon position-J 40.50 10.01 39.41 10.08 79.90 −0.014 0.003
Protein coding genes-N 40.42 15.80 28.19 15.60 68.60 −0.178 −0.006
First codon position-N 36.24 14.37 27.98 21.41 64.22 −0.129 0.197
Second codon position-N 44.50 20.95 18.65 15.90 63.15 −0.409 −0.137
Third codon position-N 40.52 12.08 37.92 9.48 78.44 −0.033 −0.121
tRNA genes 36.40 11.48 36.33 15.80 72.72 −0.001 0.158
rRNA genes 33.24 13.58 38.18 15.00 71.42 0.069 0.050
Control region 38.92 18.65 30.81 11.62 69.73 −0.116 −0.232
DOI: 10.7717/peerj.8603/table-3

Protein coding genes and codon usage

In comparison, the lengths of the PCGs of A. vulgaris mitogenome were within the range of those of other stylommatophoran mitochondrial PCGs. The ND4 gene was the most variable gene in length and has a variability of 53 codons among stylommatophorans (419 codons in Microceramus pontificus and 472 codons in Orcula dolium). The most conserved gene in length was COX1 and it exhibits variability with only 16 codons between species of Stylommatophora (501 codons in Achatinella mustelina and 517 codons in Achatina fulica). Compared with the mitogenome of A. rufus, the lengths of PCGs of A. vulgaris were distinct except for COX1, COX2, CYTB and ND1 genes. The ND6 gene was the most variable gene in length and was longer in the A. vulgaris mitogenome by 11 codons. Based on the amino acid identities, the most conserved PCG was COX1 (56.45%) whereas the least conserved was ND6 (11.92%) among the stylommatophoran mitogenomes. The most conserved PCG was COX1 (97.45%) whereas the least conserved was ATP8 (68.18%) based on the amino acid identities between the two Arion mitogenomes (Table S4).

In the A. vulgaris mitogenome, most of the PCGs initiated with typical ATN start codon, except for COX1, ND5 and ATP8 genes which use TTG, ACA and GTG triplets as start codons, respectively (Table 2). The TTG and GTG start codons are also accepted as canonical start codons for invertebrate mitogenomes (Yang et al., 2019), however, ACA as start codon for ND5 gene was reported for the first time for stylommatophoran mitogenomes. Most of the PCGs were inferred to use TAR as termination codon, except for ND4L and COX3 which have an abbreviated T-termination codon and their products are probably completed via post-transcriptional polyadenylation (Anderson et al., 1981; Ojala, Montoya & Attardi, 1981).

The most frequently used amino acids by the PCGs of the mitogenomes of A. vulgaris and A. rufus were leucine (16.71% and 15.91% respectively) and serine (10.33% and 10.18% respectively), similar to PCGs of the mitogenome of other stylommatophoran species (Leu 16,60%, Ser 10.21% on average). The codons rich in A and T, such as UUA-Leu, AUU-Ile, UUU-Phe, AUA-Met, UAU-Tyr, were the most frequently used codons in all stylommatophoran mitochondrial PCGs. The codons rich in terms of G and C content, CGC-CGG-Arg, CAG-Gln, UGC-Cys, CUC-Leu and UCG-Ser were rarely used in both Arion mitogenomes (Table S5, Fig. 2). CGN-Arg, CCS-Pro, GCS, UCG and UGC codons are seldom used or never used also in the stylommatophoran mitogenomes and reflected a significant relationship between codon usage and nucleotide content (Table S5).

Relative synonymous codon usage (RSCU) of the (A) A. vulgaris and (B) A. rufus mitogenomes.

Figure 2: Relative synonymous codon usage (RSCU) of the (A) A. vulgaris and (B) A. rufus mitogenomes.

Codon families are provided on the x axis. The stop codons are not given.

tRNA and rRNA genes

All of the tRNA genes could be folded into a usual clover-leaf secondary structure, except for trnS1 (AGN) and trnC which lacked dihydrouridine (DHU) and TΨC arms, respectively and formed simple loops (Fig. S1). Their lengths ranged between 57 bp (trnC) and 78 bp (trnG), with an average 72.72% A + T content. 26 mismatched positions were observed in stem regions and all of the mismatches were G–U pairs (Fig. S1).

The exact boundaries of rRNA genes were determined as being bounded by the adjacent tRNA genes. The rrnL gene was located between trnV and trnL1 genes, and the rrnS gene was located between trnQ and trnM genes. The length of the rrnL gene was 1,013 bp, with a 71.17% A + T content, while that of rrnS gene was 747 bp, with a 71.75% A + T content. These were comparable in ranges to homologous genes in other reported stylommatophoran species, ranging from 605 to 1215 bp in rrnL and from 564 to 857 bp in rrnS.

Non-coding and overlapping regions

The total length of intergenic regions in the A. vulgaris mitogenome was 670 bp in 16 locations ranging between 1 and 370 bp (Table 2). In general, the largest non-coding region in the animal mitogenomes is considered to contain the signals for replication and transcription, and so called as the control region (Wolstenholme, 1992). The possible candidate for the control region in A. vulgaris mitogenome was the largest non-coding region located between trnY and trnG genes with 370 bp in length. This sequence did not give BLAST hits with other putative CRs of other molluscan mitogenomes, however a part of the sequence with 67 bp in length displayed 79.11% sequence similarity with the mitochondrial control region of an amphibian species (Indotyphlus maharashtraensis, KF540157). Nucleotide composition of this region was slightly biased towards A + T with a 69.73% A + T content. The putative control region had a nine bp poly-T stretch and formed a stable secondary structure comprising seven stems and loops (Fig. 3). Furthermore, this sequence also contained a lot of palindromic sequences which are varying between 4 and 8 bp, but tandemly repeated sequences were not found.

Predicted secondary structure of putative control region of A. vulgaris mitogenome.

Figure 3: Predicted secondary structure of putative control region of A. vulgaris mitogenome.

Nucleotides are coloured as follows: Adenine is green, thymine is red, cytosine is blue and guanine is black. The poly-T stretch is labelled with purple.

The second largest non-coding region was found between trnW and trnY with a length of 91 bp (Table 2). The A + T composition of the sequence was higher than that of whole genome and putative control region with an 86.81% A + T. This non-coding region also contained a seven bp poly-A stretch and was folded into a secondary structure with two stem and loops. This secondary structure forming AT-rich sequence might function as the origin of the second strand (Wolstenholme, 1992).

Eleven overlapping regions with a total length of 164 bp were found throughout the mitogenome of A. vulgaris. The largest overlapping region was 41 bp in length and located between ND6 and ND5 genes, while the second largest was 32 bp and located between trnL2 and ATP8 (Table 2).

Phylogeny and divergence times of stylommatophoran species

Regression analyses of pairwise distances revealed that the 1st and 2nd codon positions of ATP8, ND2, ND3, ND4L and ND6 genes, as well as the 3rd codon positions of all PCGs were saturated (Table S2). Four phylogenetic reconstruction analyses were performed with combination of inference methods and different data matrices to test the influence of inference methods and saturation level of genes/codon positions on tree topology and nodal support. Three different tree topologies were obtained as the results of these analyses, and topologies were sensitive to both inference methods and exclusion of saturated genes/codon positions (Fig. 4 and Figs. S2S4). Nodal support values were always higher in BI trees than ML trees of the corresponding dataset. The usage of all mitochondrial genes and codon positions (P123RNA dataset) under both approaches resulted with identical tree topology (Figs. S3 and S4), which were similar to the results of Yang et al. (2019) obtained using only amino acid sequences of mitochondrial PCGs. The results of these analyses supported the monophyly of all included superfamilies with high nodal supports except for the superfamily Helicoidea which recovered with low nodal support [Bayesian Posterior Probability (BPP) = 0.75, Bootstrap support (BS) = 58%] and recovered Arionoidea superfamily as sister group to Urocoptoidea + (Polygyroidea + Helicoidea) clade (BPP = 1.00, BS = 100%), and Succineoidea + Orthalicoidea clade was recovered as sister group to Arionoidea + (Urocoptoidea + (Polygyroidea + Helicoidea)) (BPP = 1.00, BS = 50%). The ML and BI analyses performed using the dataset constructed with the removal of the saturated PCGs and codon positions (8P12RNA) resulted in two different tree topologies (Fig. 4 and Fig. S2). The phylogenetic tree obtained from ML analysis did not support the monophyly of the superfamily of Helicoidea and the superfamily of Polygyroidea placed within the superfamily of Helicoidea (BS = 61%, Fig. S2). A highly resolved tree with higher nodal support values was obtained from the BI approach of the dataset 8P12RNA, and hence considered as most reliable tree for discussion. The results confirmed the taxonomic position of A. vulgaris as sister species to A. rufus and recovered the monophyly of the Arionoidea superfamily (Arionidae + Philomycidae) with high support values (BPP = 1.00). A well-supported sister group relationship between Arionoidea and Orthalicoidea was recovered (BPP = 0.98) for the first time. However, previous studies using different datasets and sampling of taxa have proposed different sister groups with Arionoidea superfamily. Wade, Mordan & Naggs (2006) have found the superfamily Limacoidea as sister group to the superfamily Arionoidea using 160 stylommatophoran species, however they used only 823 nucleotides from rRNA gene-cluster. Holznagel, Colgan & Lydeard (2010) have proposed a sister group relationship between Arionoidea and Limacoidea + Zonitoidea based on the 28S rRNA sequences using seven species from Stylommatophora. The sister group relationships between Arionoidea and Urocoptoidea + Enoidea + Helicoidea (Jörger et al., 2010) or Limacoidea + (Succineoidea + Helicoidea) (Dayrat et al., 2011) have also been suggested by previous studies using relatively longer DNA sequences, however in both studies, only five stylommatophoran species were included for phylogenetic analyses. Furthermore, Xie et al. (2019b) have proposed sister group relationship between Arionoidea and Succineoidea using only amino acid dataset of mitochondrial PCGs, and stated it might be an artefact of poor taxon sampling.

Stylommatophoran phylogenetic tree constructed under BI using the dataset 8P12RNA.

Figure 4: Stylommatophoran phylogenetic tree constructed under BI using the dataset 8P12RNA.

Carychium tridentatum (Ellobioidea), Platevindex mortoni (Systellommatophora) and Galba pervia (Hygrophila) were used as outgroup. Nodes are labelled with numbers refer to hypothetical ancestral mitogenome organizations inferred by MLGO.

In the phylogenetic tree obtained from 8P12RNA under BI approach, the monophyly of all included families and superfamilies were also supported with high support values except for the superfamily Helicoidea which supported with a low nodal support (BPP = 0.82) (Fig. 4). Arionoidea + Orthalicoidea clade was recovered as sister group to Succineoidea + (Urocoptoidea + (Polygyroidea + Helicoidea)). The tree (Fig. 4) also recovered Deroceras reticulatum (Limacoidea: Agriolimacidae) at the most basal placement and did not support the monophyly of the suborder Helicina similar to the tree in Yang et al. (2019).

A chronogram for Stylommatophora divergence times based on the obtained tree topology is shown in Fig. 5. According to our divergence time analysis, the crown age of stylommatophorans was estimated as 138.55 Ma (95% CI [180.8–107.4 Ma]) corresponding to Early Cretaceous. Our estimated times for initial diversification of Stylommatophora are slightly older but broadly congruent with the fossil records and previous studies (Tillier, Masselmot & Tillirt, 1996; Jörger et al., 2010; Dinapoli & Klussmann-Kolb, 2010). Although Solem & Yochelson (1979) suggested a Paleozoic origin for Stylommatophora, the widely accepted fossil records with recognizable taxa began from Late Cretaceous (Bandel & Riedel, 1994). The Cretaceous origin of stylommatophoran species was also suggested by sequence studies of 28S rDNA fragments by Tillier, Masselmot & Tillirt (1996), of combined data of 18S, 28S, 16S rDNA and COI by Dinapoli & Klussmann-Kolb (2010) and Jörger et al. (2010). The diversification of the stylommatophoran species may have been influenced by the explosive radiation of angiosperms and speciation by host-switching during Cretaceous (Friis, Pedersen & Crane, 2010).

Dated phylogenetic tree.

Figure 5: Dated phylogenetic tree.

The axis on the bottom refers to million years. Letters in the boxes refers to external calibration points. The split between A. vulgaris and A. rufus was shown with a red bar, while the remainings were shown with blue bars.

The split time of Achatina fulica from other stylommatophoran species was inferred as 131.91 Ma in Early Cretaceous. The splits of the superfamilies Orthalicoidea and Arionoidea, of Succineoidea from Urocoptoidea + (Polygyroidea + Helicoidea), and of Clausilioidea + (Pupilloidea + Achatinelloidea) were dated to 114.18 Ma (95% CI [148.7–87.2 Ma]), 113.30 Ma (95% CI [146.1–87.9 Ma]) and 111.88 Ma (95% CI [148.1–84.2 Ma]), respectively, coinciding to the beginning of the Albian (Early Cretaceous). The crown ages of the superfamilies Arionoidea, Urocoptoidea, Helicoidea and Pupilloidea were estimated corresponding to Late Cretaceous (84.82, 75.23, 74.39 and 70.06 Ma, respectively). The split of the two Arion species and the crown age of Achatinelloidea species were dated to 15.24 Ma (95% CI [30.0–7.6 Ma]) and 13.01 Ma (95% CI [19.7–8.2 Ma]), respectively, corresponding to the Miocene. The divergence time of A. vulgaris and A. rufus corresponds to one of Earth’s most recent, global warming events, the Mid-Miocene Climatic Optimum (MMCO, 17–14.75 Ma) (Böhme, 2003). The MMCO is thought to have contributed to floristic and faunistic diversity across the world and so to animal-plant interactions, correlating with the rise in temperature (Barnosky & Carrasco, 2002; Vicentini et al., 2008; Tolley, Chase & Forest, 2008). The change of plant diversity, emergence of new host plants and the relative warm period may have triggered the diversification of the two Arion species. The divergence time of two polygyroid species was inferred as 0.45 Ma (95% CI [1.1–0.1 Ma]), in the Pleistocene.

Selective pressures on stylommatophoran mitogenomes

The ω value for each of the 13 PCGs was inferred under one-ratio model using PAML and presented in Table 4. All of the ω values were extremely low (ω < 1), ranging between 0.0129 for COX1 and 0.2198 for ATP8, reflecting that all genes were under strong purifying selection consistent with the general mitogenome evolution pattern in animals (Rand, 2001; Bazin, Glemin & Galtier, 2006). Although purifying selection is the predominant selective force shaping stylommatophoran mitogenomes, the comparison of the null neutral model and alternative branch-specific positive selection model revealed six of the PCGs (ATP6, COX2, COX3, ND2, ND4 and ND5) have variation in ω values along different branches. The variability in ω values indicated different selective forces acting on each gene as well as each branch. A more sensitive branch-site method, aBSREL, providing three states for each branch and allowing each site to evolve under any kind of the value (<1, 1 or >1) (Smith et al., 2015), was used for evaluating and confirming the selective forces across lineages determined by PAML analysis. All of the branches in the stylommatophoran phylogeny were tested with aBSREL analysis for each PCG, and the genes detected as under episodic diversifying selection were different from the results of branch-site model of PAML (Table 5) except for COX3 and ND4. The aBSREL analyses discovered episodic diversifying selection in ATP8 (at the branch leading to Microceramus pontificus), COX1 (at the branch leading to Achatinella mustelina), COX3 (at the branch leading to Arionoidea and the branch leading to Philomycus bilineatus), ND3 (at the branch leading to Helicella itala), ND4 (at the branch leading to Succinea putris) and ND6 (at the branch leading to Vertigo pusilla). Due to their important function, mitochondrial genes might have a few positively selected sites and the signatures of purifying selection likely mask those of positive selection (Meiklejohn, Montooth & Rand, 2007; Da Fonseca et al., 2008). Therefore, two different methods were used to detect positive selection in addition to BEB analysis: FUBAR which estimates the rates of nonsynonymous and synonymous substitutions at each codon in a phylogeny, and MEME which estimates the probability for a codon to have experienced episodic positive selection and allows the ω ratio to vary across branches and codons. BEB analysis identified eight positively selected codons in total in three genes (ND2, ND4 and ND5), whereas FUBAR defined six positively selected codons in five genes (ATP6, ATP8, COX2, CYTB and ND4L). The MEME analysis found the signals of episodic positive selection at 22 codons in nine genes (ATP6, ATP8, CYTB, ND2-6, and ND4L). There was not any shared codon determined by all of the three analyses (Table 6). Only four codons in three genes were shared by the results of FUBAR and MEME analyses: 44th codon in ATP8 gene, 12th codon in CYTB gene, and 13th and 57th codons in ND4L gene. Therefore, we focused only on these four codons in the TreeSAAP analyses. The positively selected substitution at codon 44 in ATP8 gene was the change of TTA (Leu) to ATT (Ile) at branches leading to M. kiangsinensis, Cerion incanum and Cerion uva. This substitution was a radical chemical change with a magnitude category of 8 and had an impact on the increment of the equilibrium constant (ionization of COOH). The change at the codon 12 in CYTB gene was a conserved change with a magnitude category of 1 and was a substitution of TTG (Leu) to ATG (Met). The positively selected substitutions in ND4L gene were the change of ATT (Ile) to ATA (Met) at branch leading to H. pomatia, to GTT (Val) at branch leading to C. nemoralis at codon 13, and the change of TTT (Phe) to AAT (Asn) at branch leading to Arionidae family at codon 57. The substitution at the 13th codon was a radical change with a magnitude category of 8 altering the equilibrium constant (ionization of COOH), while that at the 57th codon was a radical change with a magnitude category of 7 and modifying the solvent accessibility of the protein.

Table 4:
Likelihood ratios of PAML analysis showing different selective pressures on the mitochondrial PCGs in Stylommatophora.
Modelsa A B C A–B B–C
Gene ω lnLb Npc lnL Np lnL Np LRT P LRT Pd
ATP6 0.0509 −19658.6890 76 −19377.7446 78 −19374.7961 79 561.8887 0.000 −5.896982 0.015
ATP8 0.2198 −6460.1406 −6355.5745 −6355.5745 209.1322 0.000 −0.000006 1.000
COX1 0.0129 −26428.6827 −26169.9332 −26169.9332 517.4990 0.000 0.000000 1.000
COX2 0.0393 −16022.1910 −15851.6973 −15848.6152 340.9874 0.000 −6.164250 0.013
COX3 0.0317 −18617.4781 −18345.8456 −18341.7284 543.2650 0.000 −8.234404 0.004
CYTB 0.0416 −26893.7004 −26297.6496 −26297.6496 1192.1017 0.000 0.000082 0.992
ND1 0.0430 −24857.6972 −24528.7292 −24529.0150 657.9359 0.000 0.571480 0.450
ND2 0.0670 −31769.5279 −31502.5684 −31498.5036 533.9191 0.000 −8.129668 0.004
ND3 0.0607 −11403.3678 −11183.4729 −11183.4729 439.7899 0.000 0.000010 1.000
ND4 0.0511 −40498.0727 −40032.1913 −40026.0861 931.7628 0.000 −12.210500 0.000
ND4L 0.0727 −10118.1495 −10076.4001 −10075.2708 83.4989 0.000 −2.258408 0.133
ND5 0.0676 −51970.2413 −51088.0681 −51092.3346 1764.3464 0.000 8.533012 0.003
ND6 0.1009 −16916.1956 −16691.7006 −16691.7006 448.9899 0.000 0.000018 1.000
DOI: 10.7717/peerj.8603/table-4

Notes:

Degrees of freedom = 1.

A, All branches have one ω; B, All branches have same ω = 1; C, Each branch has its own ω.
The natural algorithm of the likelihood value.
Number of parameters.
Bold faced figure indicate the statistical significance (P < 0.05).
Table 5:
Genes and branches detected to be exposed episodic diversifying selection using the aBSREL approach.
Gene Number of selected branches (P < 0.05) Taxon ω Proportion of codons under selection
ATP8 1 Microceramus pontificus 288 0.460
COX1 1 Achatinella mustelina 2,180 0.086
COX3 2 Arionoidea 670 0.053
Philomycus bilineatus 119 0.092
ND3 1 Helicella itala 49.4 0.220
ND4 1 Succinea putris 4.18 0.370
ND6 1 Vertigo pusilla 15.6 0.370
DOI: 10.7717/peerj.8603/table-5
Table 6:
Genes/codons under diversifying or positive selection under codon-based models.
Gene BEB FUBAR MEME
ATP6 4 44
ATP8 44 44, 57, 64, 92, 109
COX1
COX2 32
COX3
CYTB 12 12
ND1
ND2 188 14, 16, 174
ND3 27
ND4 109, 170, 192, 301, 386, 427 9, 99
ND4L 13, 57 13, 57, 109, 111
ND5 451 260, 501
ND6 109, 179, 183
DOI: 10.7717/peerj.8603/table-6

Consequently, six positive selected genes (ATP8, COX1, COX3, ND3, ND4 and ND6) detected by branch-specific aBSREL approach and three genes (ATP8, CYTB and ND4L) detected by codon-based BEB, FUBAR and MEME approaches were exposed to diversifying selection. Four of these genes (ND3, ND4, ND4L and ND6) play an important role in oxidative phosphorylation and are subunits of NADH dehydrogenase (Complex I) which is the most complicated and largest proton pump of the respiratory chain coupling electron transfer from NADH to ubiquinone. In addition to its important role in energy production, it has been shown that complex I is implicated in the regulation of reactive oxygen species (ROS) (Sharma, Lu & Bai, 2009). Substitutions in this complex might have been favoured for increasing the efficiency of proton pumping or regulating the response to ROS depending varying amount of oxygen in the atmosphere and adaptation to conditions in new habitats (temperature, humidity, altitude) and/or hosts. CYTB gene encodes only mitogenome derived subunit of Complex III and catalyses reversible electron transfer from ubiquinol to cytochrome c (Da Fonseca et al., 2008). The positively selected sites in complexes I and III have been suggested to contribute to environmental adaptation in different groups such as mammals, birds, fishes and insects (Da Fonseca et al., 2008; Garvin, Bielawski & Gharrett, 2011; Garvin et al., 2014; Melo-Ferreira et al., 2014; Morales et al., 2015; Li et al., 2018). In the cytochrome c oxidase complex (Complex IV), COX1 protein catalyses electron transfer to the molecular oxygen; COX2 and COX3 belong to the catalytic core of the complex may act as a regulator. ATP8 gene encodes the part of ATP synthase (Complex V) regulating the assembly of complex (Da Fonseca et al., 2008). The favoured substitutions in COX3 and ATP8 gene might have an impact on assembly of the complexes IV and V. The positively selected substitutions and random accumulation of variation in mitochondrial PCGs of stylommatophoran species thus seem to be adaptive and affecting mitochondrial ATP production or protection from ROS effects, however effects of substitutions should be examined in a larger sample by considering protein folding and three-dimensional structure of complexes.

Gene rearrangements in stylommatophoran mitogenomes

The ancestral mitogenome organisation of each node in the phylogeny was inferred using the maximum likelihood approach. The organisation of the hypothetical ancestral Stylommatophora mitogenome (node: A34, Fig. 4) was identical with that of Deroceras reticulatum as well as those of Albinaria caerulea, Cernuella virgata and Helicella itala. The mitogenome of Achatina fulica has only experienced the transposition of trnP to the downstream of trnA compared to its most recent ancestral mitogenome organisation. The common ancestors of Clausilioidea + (Pupilloidea + Achatinelloidea) (node: A31, Fig. 4), Orthalicoidea + Arionoidea (node: A22, Fig. 4), and Succineoidea + (Urocoptoidea + (Polygyroidea + Helicoidea)) (node: A18, Fig. 4) maintained the same order of hypothetical ancestral stylommatophoran mitogenome. In the mitogenome of the most recent common ancestor (MRCA) of Pupilloidea + Achatinelloidea (node: A30, Fig. 4), the reversal of trnW, trnG and trnH genes occurred individually and were followed by the reversal of the cluster trnW-trnG-trnH. In the superfamily Pupilloidea, rearrangements of several tRNA genes were observed: the transposition of the cluster trnD-trnC to downstream of trnW in Pupilla muscorum, transpositions of cluster trnH-trnG to downstream of trnW and of trnT to upstream of COX3 in Orcula dolium, transposition of trnG to downstream of trnW in Vertigo pusilla and the reversal of trnQ in Gastrocopta cristata. In the mitogenome of the MRCA of the superfamily Achatinelloidea (node: A26, Fig. 4), trnF-COX2-trnY-trnH-trnG- trnW-trnQ-ATP8-trnN-ATP6-trnR-trnE-rrnS-trnM gene cluster rearranged as trnW-trnQ-ATP8-ATP6-trnR-trnE-rrnS-trnM-trnF-COX2-trnY-trnH-trnG- trnN via tandem duplication random loss (TDRL) mechanism. The organisation of the mitogenomes of achatinelloid species nearly matched with the putative ancestral order, except for Achatinella sowerbyana which has a transposed position of trnK to downstream of ATP8 and a second copy of trnL2, and for Partulina redfieldi which has the inversion of trnE and trnN genes.

The mitogenome of Naesiotus nux has almost the same organisation with its MRCA (node: A22, Fig. 4), except for the second inverted copy of ND4L located between trnL1 and trnP. The MRCA of the superfamily Arionoidea (node: A21, Fig. 4) had also identical mitogenome organisation with the ancestor of Stylommatophora, and the MRCAs of the families Arionidae (node: A19, Fig. 4) and of Philomycidae (node: A20, Fig. 4) were derived from this ancestor. The mitogenome of node A19 had shuffled positions of trnY and trnW, and also transpositions of trnE to downstream of trnQ and of rrnS-trnM to upstream of trnQ. Both Arion mitogenomes also shared this mitogenome organisation and the rearranged orders of trnW-trnY and trnE-trnQ-rrnS-trnM-trnL2-ATP8-trnN-ATP6-trnR clusters seem to be synapomorphies of this genus. The mitogenome organisation of the node A20 was quite different from those of other stylommatophoran species, which had rearranged positions of almost all genes between COX1 and trnI via two-steps TDRL, and two Philomycidae species also had identical organisation except for Philomycus bilineatus had a second copy of trnC located downstream of the original copy.

The mitogenome of Succinea putris has experienced the transpositions of trnF to upstream of trnD and of trnW to upstream of trnY, and also reverse transposition of the cluster trnW-trnY to the upstream of ND3. The MRCA of the all urocoptoid species (node: A16, Fig. 4) only had the reversal of trnQ gene from minor to major strand and M. pontificus has also maintained the identical arrangement. The four step requiring rearranged gene cluster was identified in the mitogenome of the MRCA of the genus Cerion (node: A15, Fig. 4): (i) reversal of trnV-rrnL-trnL1, (ii) reversal of trnP, (iii) reversal of trnA, and (iv) reversal of the cluster trnL1-rrnL-trnV-trnP-trnA. The mitogenome organisation remained the same in all three Cerion species and the rearranged state of trnA-trnP-trnV-rrnL-trnL1 cluster might be a synapomorphy for this genus.

The MRCAs of the polygroid species (node: A12, Fig. 4) and Camaenidae + Bradybaenidae (node: A5, Fig. 4), as well as the Polygyra cereolus and Praticolella mexicana, had the transposed position of trnG-trnH to the upstream of trnY. The rearrangement of this cluster as trnG-trnH-trnY could be suggested as a synapomorphy for Polygyroidea, but more sampling is required to confirm its status at superfamily level. In the superfamily Helicoidea, the MRCAs of Geomitridae (node: A9, Fig. 4) and Geomitridae + Helicidae (node: A10, Fig. 4) shared the identical mitogenome organisation with the MRCA of Stylommatophora. Both of the Geomitridae species had also same mitogenome organisation except for ATP8 in Cernuella virgata, in which this gene was missing, however it seems to be likely a misannotation. The mitogenome of MRCA of Helicidae species (node: A8, Fig. 4) had experienced the transpositions of trnP and cluster trnT-COX3 to downstream of ND6 and to upstream of trnS1, respectively. The mitogenome organisations of Helix pomatia, Cornu aspersum and Cepaea nemoralis have not changed and trnA-ND6-trnP and trnS2-trnT-COX3-trnS1 gene orders might be interpreted as synapomorphic for these three species. However, the individual reversals of trnA, ND6 and trnP genes followed by reversal of the cluster trnA-ND6-trnP, and reversal of trnS1 were observed in Cylindrus obtusus mitogenome. In the mitogenomes of the species of the family Camaenidae, only the transpositions of trnD and trnY to downstream of COX2 and to upstream of trnG were found, respectively. The arrangement of the trnC-trnF-COX2-trnD-trnY-trnG cluster could be considered as a synapomorphy for camaenid species, however the taxonomic level of this synapomorphy need to be evaluated in a wider taxonomic range. In the family of Bradybaenidae, the MRCA mitogenome had experienced only the reversal of trnW. In addition to this rearrangement, Aegista species also have the transposition of ND3 gene to the downstream of trnW and the rearranged position of ND3-trnW cluster appears to be a synapomorphy for the genus.

Conclusions

The sequencing and annotation of the mitogenome of A. vulgaris and its comparison with other stylommatophoran mitogenomes allow us to denote several conclusions: (i) the mitogenome characteristics of A. vulgaris are mostly consistent with the reported stylommatophoran mitogenomes; (ii) rearrangement events are detected in the trnW-trnY and trnE-trnQ-rrnS-trnM-trnL2-ATP8-trnN-ATP6-trnR gene clusters which may be apomorphic for the genus Arion, but further investigations are necessary; (iii) stylommatophoran mitogenome sequence information without the saturated positions seems to be useful for reconstructing phylogeny and estimating divergence times, and the taxon set used should be expanded; (iv) although purifying selection is the dominant force in shaping the stylommatophoran mitogenomes, in the background, several codons or different branches have experienced diversifying selection suggesting adaptation to new environmental conditions.

Supplemental Information

Summary of best fit partition schemes and nucleotide substitution models

DOI: 10.7717/peerj.8603/supp-1

Regression of the pairwise distances of the first and second codon positions of each PCG, all positions of RNA genes and the first, second, first and second, and third codon positions of the concatenated mitochondrial PCGs

DOI: 10.7717/peerj.8603/supp-2

Nucleotide compositions and skewness values of whole mitogenomes of stylommatophoran species

DOI: 10.7717/peerj.8603/supp-3

Nucleotide and amino acid identities of mitochondrial protein coding genes among stylommatophoran and between Arion species

DOI: 10.7717/peerj.8603/supp-4

Codon usages of the mitochondrial PCGs of stylommatophoran species

DOI: 10.7717/peerj.8603/supp-5

Predicted secondary structures for the 22 typical tRNA genes of A. vulgaris mitogenome. Watson–Crick pairs are indicated by lines and wobble GU pairs are indicated by dots

DOI: 10.7717/peerj.8603/supp-6

The phylogenetic tree constructed under ML using the dataset 8P12RNA

DOI: 10.7717/peerj.8603/supp-7

The phylogenetic tree constructed under BI using the dataset P123RNA

DOI: 10.7717/peerj.8603/supp-8

The phylogenetic tree constructed under ML using the dataset P123RNA

DOI: 10.7717/peerj.8603/supp-9

The complete mitogenome sequence of Arion vulgaris

DOI: 10.7717/peerj.8603/supp-10

Gene annotations of A. vulgaris mitogenome

DOI: 10.7717/peerj.8603/supp-11
14 Citations   Views   Downloads