Research Article
Print
Research Article
The complete mitochondrial genome sequence of Scolopendra mutilans L. Koch, 1878 (Scolopendromorpha, Scolopendridae), with a comparative analysis of other centipede genomes
expand article infoChaoyi Hu, Shuaibin Wang, Bisheng Huang, Hegang Liu, Lei Xu, Zhigang Hu§, Yifei Liu
‡ Hubei University of Chinese Medicine, Wuhan, China
§ Hubei Jingui Chinese Medicine Pieces Ltd., Wuhan, China
Open Access

Abstract

Scolopendra mutilans L. Koch, 1878 is an important Chinese animal with thousands of years of medicinal history. However, the genomic information of this species is limited, which hinders its further application. Here, the complete mitochondrial genome (mitogenome) of S. mutilans was sequenced and assembled by next-generation sequencing. The genome is 15,011 bp in length, consisting of 13 protein-coding genes (PCGs), 14 tRNA genes, and two rRNA genes. Most PCGs start with the ATN initiation codon, and all PCGs have the conventional stop codons TAA and TAG. The S. mutilans mitogenome revealed nine simple sequence repeats (SSRs), and an obviously lower GC content compared with other seven centipede mitogenomes previously sequenced. After analysis of homologous regions between the eight centipede mitogenomes, the S. mutilans mitogenome further showed clear genomic rearrangements. The phylogenetic analysis of eight centipedes using 13 conserved PCG genes was finally performed. The phylogenetic reconstructions showed Scutigeromorpha as a separate group, and Scolopendromorpha in a sister-group relationship with Lithobiomorpha and Geophilomorpha. Collectively, the S. mutilans mitogenome provided new genomic resources, which will improve its medicinal research and applications in the future.

Keywords

Chilopoda, Chinese medicinal materials, mitogenome, Scolopendra mutilans

Introduction

Animal medicine is an important part of the Chinese traditional medicine system. As a typical representative of medicinal animals, the centipede Scolopendra mutilans has been used for hundreds of years in China for treating many disorders, such as stroke-induced hemiplegia, epilepsy, apoplexy, whooping cough, tetanus, burns, tuberculosis, and myocutaneous disease (Ding et al. 2016). Moreover, centipedes have been described for the treatment of cardiovascular diseases in Korea, China, and other east Asian countries (Chen et al. 2014). Scolopendra mutilans is a venom-containing animal, which is rich in antimicrobial peptides, ion channel modulators, enzymes, and other macromolecular active substances (Yoo et al. 2014). Due to its active ingredients, it is of great interest in modern medical research. However, with the increase of medicinal applications, the wild populations of S. mutilans were over-exploited and declined greatly (Kang et al. 2017). Conservation and further artificial culture are needed, which in turn depends on the correct classification and molecular identification of the natural centipede taxa.

Centipedes (Chilopoda) are one of the oldest extant terrestrial arthropods. Approximately 3300 centipede species have been described (Chipman et al. 2014) and the majority of these taxa are distributed in tropical and subtropical regions. Six orders of centipedes are currently recognized, namely, Scolopendromorpha, Geophilomorpha, Lithobiomorpha, Scutigeromorpha, Craterostigmomorpha, and Devonobiomorpha (Bortolin et al. 2018). Devonobiomorpha is an extinct order represented by a single species (Shear and Bonamo 1988) and the Craterostigmomorpha only occur in Tasmania and New Zealand (Undheim et al. 2016). The remaining orders are distributed widely (Edgecombe et al. 2002), but their evolutionary relationships remain unclear on the basis of morphological traits. The Scutigeromorpha, with body respiratory openings on the back, was generally classified as class Notostigmophora, while the remaining orders with lateral spiracles were divided into another class, Pleurostigmophora (Giribet et al. 1999). However, both Scutigeromorpha and Lithobiomorpha have an anamorphic development in which the segment number increases during postembryonic life (Anamorpha). While Scolopendromorpha and Geophilomorpha have an epimorphic development in which the definitive number of body segments appears upon hatching (Epimorpha). The Craterostigmomorpha order is not strictly anamorphic, making its position unclear (Giribet et al. 1999; Edgecombe and Giribet 2007).

Previously, phylogenetic analysis on the basis of different molecular data provided support to these morphological classifications to some degree (Regier et al. 2008; Fernández et al. 2016). With a phylogenetic reconstruction based on a large number of protein-coding nuclear genes, the Scutigeromorpha was placed as a single evolutionary branch in Chilopoda, while the other three orders were clustered together, in which the Lithobiomorpha was a sister group of the Scolopendromorpha and the Geophilomorpha showed a distant relationship to them (Regier et al. 2008). A phylogenomic reconstruction based on transcriptomic data also suggested a similar pattern, that the Scutigeromorpha order was a sister group with the other three orders. Moreover, the Scolopendromorpha order is closer to the Geophilomorpha order than the Lithobiomorpha (Fernández et al. 2016).

The mitochondrial genome (mitogenome), including those markers derived from it as well as the whole mitogenome, is the most commonly used molecule in animal studies with relation to taxonomy, population genetics, and evolutionary biology (Wolstenholme 1992; Li et al. 2018a). Generally, an animal mitogenome is a double-stranded circular molecule, ranging from 14 to 20 kb in length and containing a typical set of 37 genes, including 13 protein-coding genes (PCGs), 22 transfer RNA (tRNA) genes, and two ribosomal RNA (rRNA) genes (Taanman 1999). Functional information on replication derived from the related genomic structures has been well investigated, but the transcription features of animal mitogenomes are still limited (Chen and Du 2017). Here, we sequenced and assembled the mitogenome of S. mutilans and compared its genome to seven other representative centipede mitogenomes derived from Scolopendromorpha, Geophilomorpha, Lithobiomorpha, and Scutigeromorpha. We obtained the phylogenetic relationship of these centipede taxa based on the 13 PCGs and our results provide new genetic information for both conservation and sustainable use of centipedes as a medicinal resource.

Materials and methods

Sample collection and DNA extraction

Scolopendra mutilans samples were collected in August 2018 from the wild in Yichang, Hubei Province, China. The specimens used in this study were preserved in 100% ethanol and stored at -20 °C. Genomic DNA was extracted from locomotory legs by Column mtDNAout kit (Tiangen Biotech Co., China) according to the instructions and stored at -20 °C until used for sequencing. The DNA quality was measured by gel electrophoresis and the concentration was estimated using the Nanodrop ND-1000.

Sequencing, assembly, and annotation of mitochondrial genomes

Whole genome sequencing was performed on an Illumina HiSeq 2500 platform (Illumina, San Diego, CA, USA). Quality control and de novo assembly of the S. mutilans mitogenome were conducted based on previously described methods (Li et al. 2018b). Briefly, raw reads were first filtered to generate clean data. De novo assembly of mitogenomes were performed using the SPAdes v3.9.0 software package (He et al. 2018), and the gaps were filled using MITObim v1.9 (Deng et al. 2018).

The mitogenomes were annotated by combining results from both MFannot and MITOS (Bernt et al. 2013), using the genetic code 4 in both programs. The PCGs, rRNA and tRNA were initially annotated at this step. The annotated PCGs were then refined using the NCBI Open Reading Frame Finder, and further annotated with BLASTp searches against the NCBI non-redundant protein sequence database (He et al. 2018; Wang et al. 2018b). The tRNA genes were also predicted using tRNAscan-SE v1.3.1 (Zhang et al. 2018). Subsequently, graphical maps of the complete mitogenomes were drawn using OGDraw v1.2 (Wang et al. 2018a).

Repetitive element analysis

In order to identify interspersed repeats or intra-genomic duplications of large fragments throughout the mitogenomes, we performed BLASTn searches of the mitogenome against itself using an E-value of 1e-10. Tandem repeats within the mitogenome were detected by MicroSAtellite (MISA) (Shaoli et al. 2018; Thiel et al. 2003), with the following thresholds: ten, six, five, five, five, and five repeat units for mono-nucleotide, di-nucleotide, tri-nucleotide, tetranucleotide, penta-nucleotide, and hexa-nucleotide SSRs. Forward (direct), reverse, complemented, and palindromic (reverse complemented) repeats were identified using the REPuter software (Kurtz et al. 2001) with default settings.

The base composition of the mitogenome was determined using the DNAStar Lasergene package v7.1 (Burland 2000). The following formulae were used to assess mitogenome strand asymmetry: AT skew = [A – T] / [A + T]; GC skew = [G – C] / [G + C]. Lastly, genomic synteny of the eight mitogenomes was analyzed with Mauve v2.4.0 (Darling et al. 2004).

Phylogenetic analysis

A maximum likelihood (ML) tree was constructed using the RAxML (Stamatakis et al. 2017) based on nucleotide sequence data of 13 PCGs derived from eight centipede species among the class Chilopoda (Table 1) and a Sphaerotheriidae sp. (NC_018361) (Dong et al. 2012) from the class Diplopoda was used as the outgroup. The nucleotide sequences of the 13 PCGs were firstly aligned with Clustal X (Larkin et al. 2007) as implemented in MEGA7 (Kumar et al. 2008) using the default settings. The best nucleotide substitution model was determined with Jmodeltest (Posada 2008) and the GTR+G+I model was predetermined for analyses. One thousand bootstrap replicates were performed and the phylogenetic tree was illustrated using the software FigTree v1.4.2 (Lemey et al. 2010).

Analysis of selective pressures was performed for 13 PCGs of eight centipedes using the codeml program in PAML (University College London, London, UK) (Yang 2007) by calculating the nonsynonymous (KA) and synonymous (KS) substitution ratio. The method reported by Yang and Nielsen (2000) was adopted to estimate the ω value (ω = KA / KS) of every gene sequence.

Results

Gene content and composition

The full circular mitogenome of S. mutilans (GenBank: MN317390) was 15,011 bp in length, which was similar to those of seven other centipede mitogenomes sequenced in the class Chilopoda (Table 1) (Robertson et al. 2015; Sun et al. 2018). The S. mutilans mitogenome contains 29 genes, including 13 PCGs, 14 tRNA genes, and two rRNA genes (Figure 1). Most PCGs, including the cox1, cox2, cox3, nad2, nad3, nad6, atp6, atp8, and cob genes, and the majority of tRNA genes (trnI, trnM, trnW, trnK, trnD, trnG, trnA, trnS1, trnV, and trnS2) are transcribed from the plus strand, while the remaining four PCGs, two ribosomal genes, and four tRNAs are transcribed from the minus strand (Table 2). The overlapping regions between genes were in relation to three neighboring gene pairs, containing a length of 27 bp in total, with each size ranging from 2 to 18 bp. We also found a total of 2111 bp of intergenic regions on the S. mutilans mitogenome, accounting for 14% of the genome size.

Figure 1. 

Mitochondrial genome map of the Scolopendra mutilans. Genes drawn inside the circle are transcribed clockwise, and those outside are counterclockwise. PCGs are shown as brown arrows, rRNA genes as green arrows, tRNA genes as pink arrows. The innermost circle shows the GC content. GC content is plotted as the deviation from the average value of the entire sequence.

The mitogenome size of eight centipedes ranged from 14,538 bp for S. dehaani Brandt, 1840 to 16,833 bp for Cermatobius longicornis Takakuwa,1939, with that of S. mutilans in the middle of the range. To identify the specific variation contributing most to the diversity of the mitogenome size in centipedes, the length variation of all PCGs, tRNA, and rRNA genes, and intergenic regions in each mitogenome was investigated. Comparatively, the length of most genes across centipede species was relatively stable except the PCGs in L. forficatus Linnaeus,1758 (AF309492.1), while the length of intergenic regions was the primary contributor to mitogenome size variation.

Table 1.

Basic information of the mitogenomes for Chilopoda used in this study.

Species Order NCBI ID Length (bp)
Scolopendra mutilans L. Koch, 1878 Scolopendromorpha MN317390 15011
Scolopendra dehaani Brandt, 1840 Scolopendromorpha KY947341.1 14538
Scolopocryptops sp. Scolopendromorpha KC200076.1 15119
Strigamia maritima (Leach, 1817) Geophilomorpha KP173664.1 14983
Cermatobius longicornis Takakuwa,1939 Lithobiomorpha NC_021403.1 16833
Bothropolys sp. Lithobiomorpha AY691655.1 15139
Lithobius forficatus (Linnaeus,1758) Lithobiomorpha AF309492.1 15695
Scutigera coleoptrata (Linnaeus, 1758) Scutigeromorpha AJ507061.2 14922
Table 2.

Organization of the Scolopendra mutilans mitogenome.

Gene Start End Strand Length Start/End codon
trnI (gat) 1 65 + 65
trnM (cat) 69 141 + 73
nad2 124 954 + 831 ATT/TAA
trnW (tca) 1071 1119 + 49
cox1 1148 2656 + 1509 ATG/TAG
cox2 2676 3341 + 666 ATG/TAA
trnK (ctt) 3355 3405 + 51
trnD (gtc) 3425 3458 + 34
atp8 3465 3614 + 150 ATA/TAA
atp6 3620 4267 + 648 ATG/TAA
cox3 4282 5049 + 768 AYG/TAA
trnG (tcc) 5084 5137 + 54
nad3 5141 5488 + 348 ATT/TAG
trnA (tgc) 5487 5542 + 56
trnS1 (gct) 5612 5662 + 51
nad1 5784 6635 - 851 ATT/TAA
rrnL 6719 7937 - 1219
rrnS 7993 8740 - 748
trnQ (ttg) 9667 9720 - 54
trnF (gaa) 9735 9791 - 57
nad5 9906 11,474 - 1569 ATT/TAA
trnH (gtg) 11,556 11,619 - 64
nad4 11,641 12,789 - 1149 ATG/TAA
nad4l 12,939 13,181 - 243 ATA/TAA
trnV (aac) 13,230 13,262 + 33
trnP (tgg) 13,272 13,315 - 44
nad6 13,373 13,759 + 387 ATT/TAA
cob 13,773 14,873 + 1101 ATG/TAA
trnS2 (tga) 14,889 14,944 + 56

Genomic repeats

The repeated DNA in animal mitogenomes can be divided into tandem repeats and interspersed repeats (Wu et al. 2017). In the S. mutilans mitogenome, 46 tandem repeats have been identified, of which the longest is 39 bp and the shortest is 9 bp. However, no interspersed repeat was found. Generally, SSRs are a group of tandem repeated sequences containing 1–6 nucleotide repeat units and are widely distributed in animal mitogenomes, and they are commonly used as molecular markers for species identification (Wang et al. 2018a). A total of nine SSRs were detected in the S. mutilans mitogenome, including three mono-nucleotides, five di-nucleotides, and one tri-nucleotide, as well as two compound SSRs (Table 3). Among these, only one mono-nucleotide SSR is distributed in the small subunit of one ribosomal RNA gene, while the other SSRs are all presented in the intergenic regions. These mitogenomic SSRs will provide additional marker information for future genetic analyses of S. mutilans samples and its related species.

Table 3.

Simple sequence repeats in Scolopendra mutilans.

Number SSR type SSR Size (bp) Start End Position
1 mono-nucleotide (A)11 11 12,790 12,800 intergenic
2 mono-nucleotide (A)12 12 8540 8551 rrnS
3 mono-nucleotide (A)20 20 12,837 12,856 intergenic
4 di-nucleotide (AT)8 16 8776 8791 intergenic
5 di-nucleotide (AT)8 17 9820 9835 intergenic
6 di-nucleotide (AT)9 19 3406 3423 intergenic
7 di-nucleotide (TA)11 22 1119 1140 intergenic
8 di-nucleotide (AT)19 39 14,968 15,005 intergenic
9 tri-nucleotide (TAA)5 17 14,954 14,968 intergenic

Protein-coding genes

For all 13 PCGs identified in the S. mutilans mitogenome, five genes (nad2, nad3, nad1, nad5 and nad6) initiated with the start codon ATT, two genes (atp8 and nad4l) started with the ATG codon, and the remaining six genes used ATA as the start codon. The most common termination codon TAA was detected in eleven PCGs (nad2, cox2, atp8, atp6, cox3, nad1, nad5, nad4, nad4l, nad6, cob). The cox1 and nad3 genes had termination codons with TAG (Table 2). We further compared the PCGs between different centipede mitogenomes (Table 1). Across the eight centipede mitogenomes investigated, we found that the length of some PCGs was variable; for instance, the NADH dehydrogenase genes in S. mutilans is a little shorter than those in other centipedes, especially for both the nad2 and nad4 genes (Figure 2A). Notably, it was found that the mean length of PCG genes in the L. forficatus (AF309492.1) mitogenome was slightly shorter; this may be caused by post-transcriptional editing that occurs in its mitochondrial tRNAs, which may play an important role in the synthesis of subunits of ATPase in PCGs according to previous reports (Lavrov et al. 2000). Moreover, the GC content of the 13 PCGs across these mitogenomes was also different. We found two subunits of both ATPase genes (atp6 and atp8) showed the lowest GC content compared with the other PCGs in the majority of all mitogenomes. The genetic relationship is usually positively correlated with the GC content of the mitogenome of a species (Bohlin 2011). Comparatively, we found that S. mutilans had the lowest GC content in all investigated species at the whole genome level, and S. dehaani, another species of the same genus, showed the second lowest GC content of all mitogenomes we investigated (Figure 2B). Interestingly, the four NADH dehydrogenase subunits (nad1, nad4, nad4l, nad5) possessed the opposite AT skew (Figure 2C) and GC skew in the S. mutilans mitogenome compared with other species (Figure 2D).

Figure 2. 

Variation in length and base composition of each of the 13 core protein coding genes (PCGs) among eight centipedes’ mitochondrial genomes A PCG length variation B GC content across PCGs C AT skew D GC skew.

Genomic arrangement analysis

By using the Mauve analysis, we identified six large genomic homologous regions (marked A–F in Figure 3). These homologous regions were commonly presented in all eight centipede mitogenomes, and their sequence lengths were variable across regions and genomes, particularly for the A and E regions, which had a relatively large fragmental size and greatly contributed to the genome size variation between centipede mitogenomes (Figure 3). Interestingly, we found the arrangement of these homologous regions was not conserved, particularly between the S. mutilans mitogenome and that of the other species (Figure 3). For example, S. mutilans contained a B-C-D-E order of four homologous regions in its mitogenome, while the majority other centipedes showed a D-E-B-C order (Figure 3). The F region was shorter and more conserved in all six homologous regions. However, there was an absence of the F region and a clearly shorter A region in the Strigamia maritima Leach, 1817 (KP173664.1) mitogenome. Alternatively, a large ratio of intergenic regions in the S. maritima mitogenome were identified, which was also previously reported (Chipman et al. 2014; Robertson et al. 2015). In the Lithobiomorpha order, the six homologous regions of Bothropolys sp. (AY691655.1) and L. forficatus (AF309492.1) were very similar for their length and the genomic location, while those in C. longicornis (NC_021403) were clearly different. Comparatively, in the Scolopendromorpha order, the lengths of these homologous regions across S. mutilans, S. dehaani, and Scolopocryptops sp. mitogenomes were conserved, though there was a clear rearrangement among them.

Figure 3. 

Mitogenome synteny among eight centipede species. Synteny analyses were generated in Mauve 2.4.0. A total of six large homologous regions were identified among the eight mitogenomes, while the sizes and relative positions of the homologous fragments varied across the mitogenomes.

Phylogenetic analysis

The constructed ML tree is presented in Figure 4. As previously expected, S. mutilans, together with S. dehaani and Scolopocryptops sp., was placed in one group belonging to the Scolopendromorpha order. Moreover, our phylogenetic analysis suggested that the Scutigeromorpha order (Scutigera coleoptrata) was a sister group with the other three centipede orders, Scolopendromorpha, Geophilomorpha (S. maritima) and Lithobiomorpha (C. longicornis, Bothropolys sp., and L. forficatus). Our analysis further showed a close relationship between the orders Geophilomorpha and Lithobiomorpha, although the traditional morphological taxonomy suggested a potentially close relationship between the Geophilomorpha and Scolopendromorpha orders due to their shared trait of a stable segment number and lateral spiracles (Fernández et al. 2014).

The ω value can be used for revealing the constraints of natural selection (Tomoko 1995). Among our calculations, the ω value of 13 PCGs were all distributed around 0.004 (Suppl. material 1: Table S1), indicated a possibly purifying selection.

Figure 4. 

A Molecular phylogeny of eight centipede species based on Maximum Likelihood inference analysis of 13 protein-coding genes (PCGs) B Traditional morphological classification based on the position of spiracles and the variation of larvae.

Discussion

We sequenced and assembled the mitogenome of S. mutilans, a representative animal widely used in Chinese traditional medicine. The mitogenome is 15,011 bp in length, which is similar to the genome size of other known centipede mitogenomes, for example, 15,119 bp in Scolopocryptops sp. and 15,139 bp in Bothropolys sp. (Table 1). The variation of the Chilopoda mitogenome size was relatively conserved, which was consistent with that reported in Diplopoda, an animal class close to Chilopoda (Dong et al. 2016). The gene distribution was mainly presented on the plus strand of the S. mutilans mitogenome, and only four PCGs and two rRNA genes were located on the minus strand (Figure 1). This was consistent with other centipede species, like S. maritima and S. dehaani reported in previous studies (Robertson et al. 2015; Sun et al. 2018). Comparatively, the 13 PCGs in the S. mutilans mitogenome revealed a relatively low GC content, which was similar to that of S. dehaani (Figure 2).

Our study predicted nine mitogenomic SSRs, which can provide additional genetic marker information in molecular identification of centipede species (Table 3). Generally, the identification and genetic evaluation of centipede taxa depend on the variation presented in the cox1 gene region (Chen et al. 2013). However, when samples were investigated within species or between the closely related taxa, it is difficult to identify variation at individual or population level by only using the cox1 gene information (Kang et al. 2017). Comparatively, due to the relatively high mutation rate and the potentially neutrally evolutionary trajectory of SSR loci, they are widely used in animal genetic research under the species level, including assessing genetic diversity of wild populations, accelerating the progress of genetic selection, and molecular assistant breeding (Zhang et al. 2014). Our nine mitogenome SSRs were valued for future genetic research of samples from both S. mutilans and its closely related taxa.

We identified six homologous regions among the eight species’ mitogenomes, which revealed obviously genomic rearrangements, in particular between S. mutilans and some other centipedes (Figure 3). Genomic rearrangement is common and potentially randomly presented in animals’ mitogenomes (Chen et al. 2016). With the increase of mitochondrial genome data of animals, it is clear that rearrangements in mitogenomes are more a matter of sampling than a product of evolution (Boore 1999). For example, Negrisolo et al. (2003) found that it is less reliable to infer phylogenetic relationships based on gene order data in Arthropoda. Genomic rearrangements also occurred randomly among different orders in Hexapoda insects, which is not directly related to the evolution of groups (Cameron et al. 2006). Nevertheless, the observed mitogenomic rearrangements of Chilopoda taxa showed information about how genes move dynamically between different mitogenomes, which may be related to each individual gene evolutionary pattern.

Previous studies revealed alternative phylogenetic relationships of different centipedes by using different molecular datasets (Regier et al. 2008; Robertson et al. 2015; Fernández et al. 2016). With the obtained whole mitogenomic information of S. mutilans and the comparative analysis with other representative centipede taxa, our phylogenetic tree revealed a close relationship between S. mutilans and S. dehaani, which commonly belongs to the Scolopendromorpha order together with Scolopocryptops sp. (Figure 4). This was consistent with previous research (Lewis et al. 2005). However, at the order level, with increased two Scolopendromorpha samples, our analysis showed a closer relationship between Geophilomorpha and Lithobiomorpha, rather than between Geophilomorpha and Scolopendromorpha, which was slightly different to previous research (Robertson et al. 2015). Given the potentially dynamic evolutionary trajectory of different genes or between nuclear and mitochondrial genomes, this discordance may reflect the complex evolutionary history of these centipedes, including the possibility of a genetic admixture or adaptive radiations of these lineages in relation to morphological or functional specification in different geographical areas.

In conclusion, we successfully sequenced the complete mitochondrial genome of S. mutilans for the first time using next-generation sequencing, which will be valued for further studies in terms of the conservation, molecular identification, and evolutionary biology of diverse centipede species, improving the medicinal applications of S. mutilans and other closely related taxa.

Acknowledgments

We acknowledge Professor Ya’hua Zhan for identifying the materials sequenced here. Special thanks to Pavel Stoev and the reviewers for constructive comments that improved this paper. The present study was supported by the central government guides local science and technology development fund in Hubei Province (2019ZYYD063), Hubei Science Foundation for Distinguished Young Scholars (2019CFA097) and Wuhan “Yellow Crane Talent Program” Talent Project (Principal: Zhigang Hu).

References

  • Bernt M, Donath A, Jühling F, Externbrink F, Florentz C, Fritzsch G, Pütz J, Middendorf M, Stadler PF (2013) MITOS: improved de novo metazoan mitochondrial genome annotation. Molecular Phylogenetics Evolution 69: 313–319. https://doi.org/10.1016/j.ympev.2012.08.023
  • Bortolin F, Fusco G, Bonato L (2018) Comparative analysis of diet in syntopic geophilomorph species (Chilopoda, Geophilomorpha) using a DNA-based approach. Soil Biology and Biochemistry 127: 223–229. https://doi.org/10.1016/j.soilbio.2018.09.021
  • Cameron SL, Beckenbach AT, Dowton M, Whiting M (2006) Evidence from mitochondrial genomics on interordinal relationships in insects. Arthropod Systematics & Phylogeny 64(1): 27–34.
  • Chen M, Li J, Zhang F, Liu Z (2014) Isolation and characterization of SsmTx-I, a Specific Kv2.1 blocker from the venom of the centipede Scolopendra Subspinipes Mutilans L. Koch. Journal of Peptide Science 20: 159–164. https://doi.org/10.1002/psc.2588
  • Chen S, Yao H, Han J (2013) Principles for molecular identification of traditional Chinese materia medica using DNA barcoding. China Journal of Chinese Materia Medica 38(2): 141. https://doi.org/10.4268/cjcmm20130201
  • Chen Z, Du Y (2016) Rearrangement of mitochondrial genome in insects. Journal of environmental insects 38(04): 843–851.
  • Chen Z, Du Y (2017) First Mitochondrial Genome from Nemouridae (Plecoptera) Reveals Novel Features of the Elongated Control Region and Phylogenetic Implications. International Journal of Molecular Sciences 18(5): 996. https://doi.org/10.3390/ijms18050996
  • Chipman AD, Ferrier DE, Brena C, Qu J, Hughes DS, Schroder R, Torres-Oliva M, Richards S (2014) The first myriapod genome sequence reveals conservative arthropod gene content and genome organisation in the centipede Strigamia maritima. PLos Biology 12(11): e1002005. https://doi.org/10.1371/journal.pbio.1002005
  • Darling C, Mau B, Blattner R, Perna T (2004) Mauve: multiple alignment of conserved genomic sequence with rearrangements. Genome research 14(7): 1394–1403. https://doi.org/10.1101/gr.2289704
  • Deng Y, Hsiang T, Li S, Lin L, Wang Q, Chen Q, Xie B, Ming R (2018) Comparison of the Mitochondrial Genome Sequences of Six Annulohypoxylon stygium Isolates suggests short fragment insertions as a potential factor leading to larger Genomic Size. Frontiers in Microbiology 9: 2079. https://doi.org/10.3389/fmicb.2018.02079
  • Ding D, Guo YR, Wu RL, Qi WY, Xu HM (2016) Two new isoquinoline alkaloids from Scolopendra subspinipes mutilans induce cell cycle arrest and apoptosis in human glioma cancer U87 cells. Fitoterapia 110: 103–109. https://doi.org/10.1016/j.fitote.2016.03.004
  • Dong Y, Xu JJ, Hao SJ, Sun HY (2012) The complete mitochondrial genome of the giant pill millipede, Sphaerotheriidae sp. (Myriapoda: Diplopoda: Sphaerotheriida). Mitochondrial DNA 23: 333–335. https://doi.org/10.3109/19401736.2012.683184
  • Dong Y, Zhu L, Bai Y (2016) Complete mitochondrial genomes of two flat-backed millipedes by next-generation sequencing (Diplopoda, Polydesmida). ZooKeys 637(637): 1–20. https://doi.org/10.3897/zookeys.637.9909
  • Edgecombe GD, Giribet G, Wheeler W (2002) Phylogeny of Henicopidae (Chilopoda: Lithobiomorpha): a combined analysis of morphology and five molecular loci. Systematic Entomology 27: 31–64. https://doi.org/10.1046/j.0307-6970.2001.00163.x
  • Fernández R, Laumer CE, Varpu V, Silvia L, Stefan K (2014) Evaluating topological conflict in centipede phylogeny using transcriptomic data sets. Molecular Biology and Evolution 31(6): 1500–1513. https://doi.org/10.1093/molbev/msu108
  • Fernández R, Edgecombe GD, Giribet G (2016) Exploring phylogenetic relationships within Myriapoda and the effects of matrix composition and occupancy on phylogenomic reconstruction. Systematic Biology 65(5): 871–889. https://doi.org/10.1093/sysbio/syw041
  • Giribet G, Carranza S, Riutort M, Baguna J, Ribera C (1999) Internal phylogeny of the Chilopoda (Myriapoda, Arthropoda) using complete 18S rDNA and partial 28S rDNA sequences. Philosophical Transactions of the Royal Society B-Biological Sciences 354: 215–222. https://doi.org/10.1098/rstb.1999.0373
  • Kang S, Liu Y, Zeng X, Deng H, Luo Y, Chen K, Chen S (2017) Taxonomy and Identification of the Genus Scolopendra in China Using Integrated Methods of External Morphology and Molecular Phylogenetics. Scientific Reports 7: 16032. https://doi.org/10.1038/s41598-017-15242-7
  • Kumar S, Nei M, Dudley J, Tamura K (2008) MEGA: a biologist-centric software for evolutionary analysis of DNA and protein sequences. Briefings in bioinformatics 9(4): 299–306. https://doi.org/10.1093/bib/bbn017
  • Kurtz S, Choudhuri JV, Ohlebusch E, Schleiermacher C, Stoye J, Giegerich R (2001) REPuter: the manifold applications of repeat analysis on a genomic scale. Nucleic Acids Research 29: 4633–4642. https://doi.org/10.1093/nar/29.22.4633
  • Larkin MA, Blackshields G, Brown NP, Chenna R, McGettigan PA, McWilliam H, Valentin F, Wallace IM, Wilm A, Lopez R (2007) Clustal W and Clustal X version 2.0. Bioinformatics 23: 2947–2948. https://doi.org/10.1093/bioinformatics/btm404
  • Lavrov DV, Brown WM, Boore JL (2000) A novel type of RNA editing occurs in the mitochondrial tRNAs of the centipede Lithobius forficatus. Proceedings of the National Academy of Sciences of the United States of America 97: 13738–13742. https://doi.org/10.1073/pnas.250402997
  • Lemey P, Rambaut A, Welch JJ, Suchard MA (2010) Phylogeography Takes a Relaxed Random Walk in Continuous Space and Time. Molecular Biology Evolution 27: 1877–1885. https://doi.org/10.1093/molbev/msq067
  • Li J, Lin RR, Zhang YY, Hu KJ, Zhao YQ, Li Y, Huang ZR, Zhang X, Geng XX, Ding JH (2018a) Characterization of the complete mitochondrial DNA of Theretra japonica and its phylogenetic position within the Sphingidae (Lepidoptera, Sphingidae). ZooKeys 754: 127–139. https://doi.org/10.3897/zookeys.754.23404
  • Li J, Zhang Y, Hu K, Zhao Y, Lin R, Li Y, Huang Z, Zhang X, Geng X, Ding J (2018b) Mitochondrial genome characteristics of two Sphingidae insects (Psilogramma increta and Macroglossum stellatarum) and implications for their phylogeny. International Journal of Biological Macromolecules 113: 592–600. https://doi.org/10.1016/j.ijbiomac.2018.02.159
  • Lewis JG, Edgecombe GD, Shelley RM (2005) A proposed standardised terminology for the external taxonomic characters of the Scolopendromorpha (Chilopoda). Fragmenta faunistica 48(1): 1–8. https://doi.org/10.3161/00159301FF2005.48.1.001
  • Regier JC, Shultz JW, Ganley AR, Hussey A, Shi D, Ball B, Cunningham CW (2008) Resolving arthropod phylogeny: exploring phylogenetic signal within 41 kb of protein-coding nuclear gene sequence. Systematic biology 57(6): 920–938. https://doi.org/10.1080/10635150802570791
  • Shaoli M, Hao Y, Chao L, Yafu Z, Fuming S, Yuchao W (2018) The complete mitochondrial genome of Xizicus (Haploxizicus) maculatus revealed by Next-Generation Sequencing and phylogenetic implication (Orthoptera, Meconematinae). ZooKeys 773: 57–67. https://doi.org/10.3897/zookeys.773.24156
  • Shear WA, Bonamo PM (1988) Devonobiomorpha, a new order of centipeds (Chilopoda) from the Middle Devonian of Gilboa, New York State, USA, and the phylogeny of centiped orders. American Museum Novitates 2927: 1–30.
  • Sun L, Yingju QI, Tian X (2018) Analysis of mitochondrial genome of Scolopendra subspinipes dehaani. Tianjin Journal of Traditional Chinese Medicine 35(03): 225–229.
  • Thiel T, Michalek W, Varshney R, Graner A (2003) Exploiting EST databases for the development and characterization of gene-derived SSR-markers in barley (Hordeum vulgare L.). Theoretical Applied Genetics 106: 411–422. https://doi.org/10.1007/s00122-002-1031-0
  • Tomoko O (1995) Synonymous and nonsynonymous substitutions in mammalian genes and the nearly neutral theory. Journal of Molecular Evolution 40: 56–63. https://doi.org/10.1007/BF00166595
  • Wang JJ, Yang MF, Dai RH, Li H, Wang XY (2018a) Characterization and phylogenetic implications of the complete mitochondrial genome of Idiocerinae (Hemiptera: Cicadellidae). International Journal of Biological Macromolecules 120: 2366–2372. https://doi.org/10.1016/j.ijbiomac.2018.08.191
  • Wang Y, Cao J, Murányi D, Li W (2018b) Comparison of two complete mitochondrial genomes from Perlodidae (Plecoptera: Perloidea) and the family-level phylogenetic implications of Perloidea. Gene 675: 254–264. https://doi.org/10.1016/j.gene.2018.06.093
  • Wu M, Li Q, Hu Z, Li X, Chen S (2017) The Complete Amomum kravanh Chloroplast Genome Sequence and Phylogenetic Analysis of the Commelinids. Molecules 22(11): 1875. https://doi.org/10.3390/molecules22111875
  • Yoo WG, Lee JH, Shin Y, Shim JY, Jung M, Kang BC, Song KD (2014) Antimicrobial peptides in the centipede Scolopendra subspinipes mutilans. Functional & Integrative Genomics 14: 275–283. https://doi.org/10.1007/s10142-014-0366-3
  • Zhang D, Li WX, Zou H, Wu SG, Li M, Jakovlic I, Zhang J, Chen R, Wang GT (2018) Mitochondrial genomes of two diplectanids (Platyhelminthes: Monogenea) expose paraphyly of the order Dactylogyridea and extensive tRNA gene rearrangements. Parasites Vectors 11: 601. https://doi.org/10.1186/s13071-018-3144-6
  • Zhang J, Ma W, Song X, Lin Q, Gui J, Mei J (2014) Characterization and development of est-ssr markers derived from transcriptome of yellow catfish. Molecules 19(10): 16402–16415. https://doi.org/10.3390/molecules191016402

1 These authors contributed equally to this work.

Supplementary material

Supplementary material 1 

Table S1

Chaoyi Hu, Shuaibin Wang, Bisheng Huang, Hegang Liu, Lei Xu, Zhigang Hu, Yifei Liu

Explanation note: The ω value of 13 PCGs.

This dataset is made available under the Open Database License (http://opendatacommons.org/licenses/odbl/1.0/). The Open Database License (ODbL) is a license agreement intended to allow users to freely share, modify, and use this Dataset while maintaining this same freedom for others, provided that the original source and author(s) are credited.
Download file (10.05 kb)
login to comment