Research Article
Print
Research Article
Characterization of the complete mitochondrial genome of Myrmus lateralis (Heteroptera, Rhopalidae) and its implication for phylogenetic analyses
expand article infoWanqing Zhao, Dajun Liu, Qian Jia, Xin Wu, Hufang Zhang
‡ Xinzhou Teachers University, Xinzhou, China
Open Access

Abstract

Mitochondrial genomes (mitogenomes) are widely used in research studies on phylogenetic relationships and evolutionary history. Here, we sequenced and analyzed the mitogenome of the scentless plant bug Myrmus lateralis Hsiao, 1964 (Heteroptera, Rhopalidae). The complete 17,309 bp genome encoded 37 genes, including 13 protein-coding genes (PCGs), 22 transfer RNA (tRNA) genes, two ribosomal RNA (rRNA) genes, and a control region. The mitogenome revealed a high A+T content (75.8%), a positive AT-skew (0.092), and a negative GC-skew (–0.165). All 13 PCGs were found to start with ATN codons, except for cox1, in which TTG was the start codon. The Ka/Ks ratios of 13 PCGs were all lower than 1, indicating that purifying selection evolved in these genes. All tRNAs could be folded into the typical cloverleaf secondary structure, except for trnS1 and trnV, which lack dihydrouridine arms. Phylogenetic trees were constructed and analyzed based on the PCG+rRNA from 38 mitogenomes, using maximum likelihood and Bayesian inference methods, showed that M. lateralis and Chorosoma macilentum Stål, 1858 grouped together in the tribe Chorosomatini. In addition, Coreoidea and Pyrrhocoroidea were sister groups among the superfamilies of Trichophora, and Rhopalidae was a sister group to Alydidae + Coreidae.

Keywords

Mitogenome, next-generation sequencing, phylogeny, scentless plant bugs

Introduction

Mitochondria are important cytoplasmic organelles in eukaryotic cells, playing a critical role in cell metabolism, disease, apoptosis, and senescence. Mitochondrial genomes are characterized by a simple structure, stable composition, conserved arrangement, maternal inheritance, and rapid evolutionary rate (Simon et al. 2006; Cameron 2014; Tyagi et al. 2020). Therefore, mitochondrial genomes (mitogenomes) have been widely used in phylogenetic analyses and studies on population genetics (Yuan et al. 2015; Song et al. 2019; Xu et al. 2021). For most insects, the mitogenome is typically a double-stranded circular DNA molecule from 15 to 20 kb in size, including 13 protein-coding genes (PCGs), 22 transfer RNA (tRNA) genes, two ribosomal RNA (rRNA) genes, as well as a long non-coding region, also known as the control or AT-rich region (Wolstenholme 1992; Boore 1999; da Fonseca et al. 2008; Yang et al. 2018). Next-generation sequencing enables large amount of sequence data to be obtained and analyzed in a more economically efficient manner. As a result of advances in this technology, it is now more feasible to obtain the complete mitogenome of large taxa; thus, molecular phylogenetic analysis has been revolutionized (Beckenbach et al. 2009; Kocher et al. 2014).

Rhopalidae, commonly known as scentless plant bugs, is a family of Pentatomomorpha insects in the superfamily Coreoidea (Schaefer and Chopra 1982; Schaefer and Panizzi 2000; Steill and Meyer 2003). The Rhopalidae includes 26 genera and 279 species distributed throughout the biogeographical areas of the world (CoreoideaSF Team 2021); many species are of critical importance as pests of pasture grass (Carroll and Loye 2012). In China, Rhopalidae includes two subfamilies, Serinethinae (containing only one genus) and Rhopalinae (containing 20 genera) (Dolling 2006). Phylogeny for the tribes of Rhopalinae has been hypothesized based on morphology, but several concurrent hypotheses still exist (Schaefer 1993; Li and Zheng 1994; Davidova-Vilimova et al. 2000; Henry 2019). Although there are several published studies examining the relationships among infrafamilial levels of the Coreoidea, a consensus regarding the relationships among different families and subfamily lineages has not been reached (Souza et al. 2009; Ghahari et al. 2012; Forthman et al. 2019). In addition, some studies support Aradoidea as sister group of Trichophora (which includes Pentatomoidea, Lygaeoidea, Pyrrhocoroidea and Coreoidea), but the relationship within Trichophora is still under debate (Li et al. 2005; Xie et al. 2005; Linnavuori 2007; Tian et al. 2011; Weirauch and Schuh 2011).

Myrmus lateralis Hsiao, 1964 is an endemic species of China (Beijing, Hebei, eastern Inner Mongolia), Korea, Mongolia, and Russia (Far East) (Liu et al. 1994; Aukema and Rieger 2006; CoreoideaSF Team 2021). Adults are present in both forests and grasslands during July and August (Nonnaizab 1986). To date, only four complete mitogenomes from the Rhopalidae have been published in GenBank, and all of them belong to Rhopalinae (Table 1). According to the latest classification, five tribes are classified in the Rhopalidae (CoreoideaSF Team 2021). The tribe Chorosomatini Fieber, 1860 contains five genera (Agraphopus Stål, 1872; Chorosoma Curtis, 1830; Ithamar Kirkaldy, 1902; Leptoceraea Jakovlev, 1873; Myrmus Hahn, 1832 and Xenogenus Berg, 1883), of which only Myrmus and Chorosoma are distributed in China.

Table 1.

List of sequences used to reconstruct the phylogenetic tree.

Family Species Accession number
Outgroup Nabidae Himacerus nodipes JF927832
Ingroup Aradidae Aradacanthia heissi HQ441233
Aradus compar NC_030362
Aneurus similis NC_030360
Cydnidae Macroscytus gibbulus NC_012457
Adrisa magna NC_042429
Scoparipes salvazai MF614955
Pentatomidae Eurydema dominulus NC_044762
Palomena viridissima NC_050166
Eysarcoris guttigerus MN831205
Arma custos MT535604
Pentatoma semiannulata MT985377
Colobathristidae Phaenacantha marcida EU427342
Rhyparochromidae Panaorus albomaculatus NC_031364
Malcidae Malcus inconspicuus EU427339
Lygaeidae Kleidocerys resedae KJ584365
Pyrrhocoridae Dindymus rubiginosus NC_042439
Euscopus rufipes NC_042436
Melamphaus faber NC_042435
Dysdercus decussatus NC_042438
Antilochus coquebertii NC_042441
Rhopalidae Stictopleurus subviridis EU826088
Myrmus lateralis MN412595
Chorosoma macilentum MN412594
Aeschyntelus notatus EU427333
Corizus sp. KM983397
Alydidae Riptortus pedestris EU427344
Neomegalotomus parvus MG253274
Coreidae Hydaropsis longirostris EU427337
Clavigralla tomentosicollis KY274846
Acanthocoris sp. MF497707
Enoplops potanini MF497720
Leptoglossus membranaceus MF497724
Anoplocnemis curvipes KY906099
Mictis tenebrosa MF497729
Pseudomictis brevicornis MF497732
Molipteryx lunata MF497721
Notopteryx soror KX505857

In this study, we sequenced and annotated the complete mitogenome of M. lateralis. We examined the genomic features, nucleotide composition, codon usage, RNA secondary structure, evolutionary pattern of 13 PCGs and characteristics of the control region. Finally, we evaluated the phylogenetic relationship of the mitochondrial sequence data at different taxonomic levels.

Material and methods

Sampling, DNA extraction and sequencing

Adult specimens of Myrmus lateralis were collected from Jincheng City, Shanxi Province, China on 20 July 2014. Specimens were preserved in 100% ethanol and stored at –20 °C. The genomic DNA was extracted from leg muscles using a ONE-4-ALL Genomic DNA Mini-Prep Kit (BS88504; Sangon, Shanghai, China) according to the manufacturer’s protocols. The mitogenome was sequenced using the whole-genome shotgun method on an Illumina Miseq platform (Personalbio, Shanghai, China). After filtering low-quality and adapter contaminated reads, A5-miseq version 20150522 (Coil et al. 2015) was used for contig assembly.

Gene annotation and sequence analysis

Sequence annotation was performed using Geneious 10.1.3 (Kearse et al. 2012) and the MITOS web server (Bernt et al. 2013). Annotations of 13 protein-coding regions were edited manually by predicting open reading frames using the invertebrate mitochondrial code. The secondary structures of the tRNA genes were predicted with tRNAscan-SE web server (http://lowelab.ucsc.edu/tRNAscan-SE/) (Tamura et al. 2013). The identification of rRNA genes was performed based on their putative secondary structures and by comparing nucleotide sequences with those of previously reported mitogenomes. The control region was identified through the boundary of neighboring genes. The number of tandem repeats of control region was investigated with Tandem Repeats Finder (http://tandem.bu.edu/trf/trf.html) (Benson 1999) and the presence of stem-loop was predicted by Mfold Web Server (http://mfold.rna.albany.edu) (Zuker 2003). The annotated mitogenome sequence of Myrmus lateralis has been submitted to GenBank under the accession number MN412595.

The base compositions, codon usages, and relative synonymous codon usage (RSCU) values were calculated by MEGA 6.0. AT and GC skew were calculated by the following formulae: AT skew = (A–T)/(A+T) and GC skew=(G–C)/(G+C). To analyse the evolutionary patterns of the 13 PCGs, the rates of nonsynonymous substitution (Ka), the rates of synonymous substitution (Ks), and the ratio of Ka/Ks for each gene were calculated by the software DnaSP 5.0 (Librado and Rozas 2009).

Phylogenetic analyses

In addition to the mitogenome newly sequenced here, 37 other mitogenomes were taken from Genbank for the phylogenetic analyses (Table 1). The nucleic acids of 13 PCGs and two rRNAs were extracted by Geneious 10.1.3 and aligned with the Muscle algorithm in MEGA 6.0. Finally, the individual alignments were concatenated to make the datasets of PCG+rRNA (nucleotide alignment including 13 protein-coding genes and two rRNA genes) with SequenceMatrix (Vaidya et al. 2011).

The dataset was used to reconstruct phylogenetic trees under Bayesian inference (BI) and maximum likelihood (ML) using MrBayes 3.2.6 (Ronquist et al. 2012) and RaxML 8.0.2 (Stamatakis 2015) respectively. BI analysis was performed using a TVM+I+G model selected by jModeltest (Darriba et al. 2012). Two independent runs of four Markov Chains Monte Carlo (MCMC) chains in parallel for 5,000,000 generations were applied and sampled every 1,000 generations. Tracer 1.7 (Rambaut et al. 2018) was used to analyze the trace files from the Bayesian MCMC runs. Stationarity was considered to be reached when ESS (estimated sample size) value above 100 as MrBayes suggested. The first 25% of sampled trees were discarded as burn-in, and the remaining trees were used to calculate a 50% majority rule consensus tree. The ML tree was constructed using the GTR+I+G model, and nodal support values were evaluated through an ultrafast bootstrap approach, with 10,000 replicates.

Results

Genome organization

The 17,309 bp mitochondrial genome of Myrmus lateralis was determined to be a typical circular nucleotide molecule, consisting of 13 PCGs, 22 tRNA genes, two rRNA genes, and a putative control region (Table 2). For the gene arrangement, 23 genes (nine PCGs and 14 tRNAs) were found on the majority strand and 14 genes (four PCGs, eight tRNAs, and two rRNAs) were found on the minority strand. In addition, the mitogenome contains 10 gene overlaps totaling 37 bp and ranging from 1 to 8 bp; the longest overlap is being found between trnW and trnC. Thirteen intergenic spacers equal to 77 bp were observed, ranging from 1 to 18 bp, the longest being found between trnS2 and nad1.

Table 2.

Organization of the mitochondrial genome of Myrmus lateralis.

Gene Strand Size(bp) Position Intergenicnucleotides(IGN) Anticodon Codons
Start Stop Start Stop
trnI J 66 1 66 0 GAU
trnQ N 69 64 132 2 UUG
trnM J 69 137 205 4 CAU
nad2 J 985 207 1191 1 ATG T
trnW J 63 1204 1266 12 UCA
trnC N 62 1259 1320 −8 GCA
trnY N 64 1321 1384 0 GUA
cox1 J 1539 1387 2925 2 TTG TAA
trnL2UUR J 66 2921 2986 −5 UAA
cox2 J 676 2987 3662 0 ATT T
trnK J 72 3663 3734 0 CUU
trnD J 65 3738 3802 3 GUC
atp8 J 162 3803 3964 0 ATA TAA
atp6 J 669 3958 4626 −7 ATG TAA
cox3 J 790 4626 5415 −1 ATG T
trnG J 63 5413 5475 −3 UCC
nad3 J 349 5476 5824 0 ATA T
trnA J 63 5825 5887 9 UGC
trnR J 64 5891 5954 3 UCG
trnN J 66 5954 6019 −1 GUU
trnS1AGN J 69 6023 6091 3 GCU
trnE J 64 6091 6154 −1 UUC
trnF N 66 6154 6219 −1 GAA
nad5 N 1702 6220 7921 0 ATA T
trnH N 63 7931 7993 9 GUG
nad4 N 1317 7994 9310 0 ATG TAA
nad4L N 288 9304 9591 −7 ATT TAA
trnT J 62 9594 9655 3 UGU
trnP N 64 9656 9719 0 UGG
nad6 J 483 9728 10210 8 ATA TAA
cytB J 1137 10210 11346 −1 ATG TAG
trnS2UCN J 69 11345 11413 −2 UGA
nad1 N 927 11432 12358 18 ATT TAA
trnL1CUN N 67 12359 12425 0 UAG
rrnL N 1261 12426 13686 0
trnV N 69 13687 13755 0 UAC
rrnS N 967 13756 14722 0
Control region 2587 14723 17309 0

Nucleotide composition and codon usage

The mitochondrial genome of M. lateralis was strongly biased toward A+T (75.8%) in nucleotide composition, with 75.5%, 75.7%, 71.3%, and 74.8% A+T content in PCGs, rRNAs, tRNAs, and control regions, respectively. The nucleotide composition and skewness of the mitogenome is shown in Table 3. The three codon positions of PCGs possessed different A+T content; the third site had the highest value (80.1%), whereas the second site had the lowest (74.3%). The mitogenome had more A and G content than T and C, with a positive AT-skew (0.092) and a negative GC-skew (–0.165). The AT-skew of PCGs and three codon positions were all negative, –0.018 in PCGs, –0.214 in the first, –0.050 in the second, and –0.099 in the third codon positions. The GC-skew of PCGs (0.033) and the second codon position (0.127) was slightly and moderately positive, respectfully, while the first and third codon positions were both negative (–0.018 and –0.020, respectively). AT-skew values for the tRNA and rRNA genes were both negative (–0.057 and –0.001, respectively), while the GC-skew was obviously positive for tRNAs (0.292) and slightly positive for rRNAs (0.019).

Table 3.

Nucleotide composition and skewness of the Myrmus lateralis mitochondrial genome.

Feature T% C% A% G% A+T% AT-skew GC-skew
Whole genome 34.4 14.1 41.4 10.1 75.8 0.092 −0.165
PCGs 42.2 11.8 33.3 12.6 75.5 −0.118 0.033
PCGs-1st 44 14.1 28.5 13.6 72.5 −0.214 −0.018
PCGs-2nd 39 11 35.3 14.2 74.3 −0.05 0.127
PCGs-3rd 44 10.4 36.1 10 80.1 −0.099 −0.02
tRNAs 40 8.6 35.7 15.7 75.7 −0.057 0.292
rRNAs 35.7 14.1 35.6 14.7 71.3 −0.001 0.019
Control region 35 15.9 39.8 9.4 74.8 0.064 −0.257

Relative synonymous codon usage (RSCU) values of the 13 PCGs were calculated based on 3633 codons (Fig. 1). UUU (F, Phe), UAU (Y, Tyr), UUA (L, Leu2), and AUU (I, Ile) were the most frequently used codons, accounting for 28.24% of all codons. All synonymous codons ending with A or U were more frequent than those ending with C or G, except in three cases: CGG (RSCU = 0.97) was used more than CGU (RSCU = 0.55) for Arg, AGG (RSCU=1.03) was used more than AGA (RSCU = 0.99) for Ser1, and GGG (RSCU=1.27) was used more than GGU (RSCU = 1.12) for Gly.

Figure 1. 

The relative synonymous codon usage (RSCU) in the Myrmus lateralis mitogenome.

Protein coding genes

In the M. lateralis mitogenome, nine PCGs were located on the J-strand (majority strand) and four PCGs were located on the N-strand (minority strand). The PCGs had a total length of 11,024 bp that accounted for 63.69% of the complete mitogenome. Among the mitochondrial proteins, Leu (16.67%), Phe (11.02%), and Tyr (9.58%) were the most frequent amino acids.

Most PCGs started with a typical ATN codon; five started with ATG (nad2, atp6, cox3, nad4, and cytb), four with ATA (atp8, nad3, nad5, and nad6), and three with ATT (cox2, nad4L, and nad1). The only unusual initiation codon was TTG in cox1. Among the 13 PCGs, seven genes ended with the complete stop codon TAA (cox1, atp8, atp6, nad1, nad4, and nad4L) or TAG (cytb), whereas five genes ended with the partial termination codon T (nad2, cox2, cox3, nad3, and nad5).

The evolutionary patterns of the 13 PCGs were analyzed and shown in Figure 2. The highest values for both Ka and Ks were observed for atp8 and cytb. atp8 exhibited the highest Ka/Ks value, whereas cox1 exhibited the lowest. The Ka/Ks values for all 13 PCGs were below 0.29.

Figure 2. 

The rates of nonsynonymous substitution (Ka), the rates of synonymous substitution (Ks), and the ratio of Ka/Ks for each PCGs of Myrmus lateralis mitogenome

Transfer RNAs and ribosomal RNAs

A total of 22 tRNA genes were encoded by the M. lateralis mitogenome, ranging from 62 bp (trnC and trnT) to 72 bp (trnK) in length (Table 2). Eight tRNA genes (trnQ, trnC, trnY, trnF, trnH, trnP, trnL1(CUN) and trnV) were encoded on the N-strand; the remaining 14 genes were encoded on the J-strand. The 22 tRNA genes had a total length of 1,445 bp, accounting for 8.34% of the complete mitogenome. The predicted secondary structures are shown in Figure 3. All tRNA genes could be folded into typical cloverleaf secondary structures, except for trnS1 and trnV, in which the necessary dihydrouridine (DHU) arms were replaced with a simple loop. A total of 16 wobbled G-U pairs were found (six in acceptor stems, eight in DHU stems, one anticodon stems, and one in TΨC stems) that formed weak bonds in the tRNAs.

Figure 3. 

Predicted secondary structure of tRNA genes in the Myrmus lateralis mitogenome.

The large 1,261 bp rrnL gene was located between trnL1 (CUN) and trnV, and the small 967 bp rrnS gene was located between trnV and the control region. The secondary structures of both the rrnL and rrnS genes are shown in Figures 4, 5. The secondary structure of the rrnL gene contained five domains (I, II, IV, V, VI); domain III was absent, whereas the rrnS gene consisted of three domains (I, II, III). Both rRNAs included many mismatched base pairs, most of which were G-U.

Figure 4. 

Predicted secondary structure of the rrnL in the Myrmus lateralis mitogenome.

Figure 5. 

Predicted secondary structure of the rrnS in the Myrmus lateralis mitogenome.

Control region

The 2587 bp mitochondrial control region of M. lateralis was located between the rrnS gene and trnI. The A+T content (74.76%) was higher than the G+C content (25.24%), with a positive AT-skew and a negative GC-skew. In the control region, tandem repeat sequences, polyT stretch, polyA stretch, stem-loop structure, tandem repeats, and G(A)n motif were commonly found. Based on these features, we identified the following important elements: a 13 bp polyT stretch, a G(A)9T sequence, an ATAGA motif, and a 9 bp polyA stretch. In addition, a stem-loop structure was observed at the end of the control region (Fig. 6). Although no longer tandem repeat sequences were identified, stretches of T(A)4 occurred many times.

Figure 6. 

The stem-loop of control region in the Myrmus lateralis mitogenome.

Phylogenetic analyses

We conducted phylogenetic analyses based on the nucleotide sequences of the PCG+rRNA from eleven families within five superfamilies; one species from the Nabidae (damsel bugs) was used as an outgroup. The dataset contained 14,128 nucleotide sites from these 38 taxa. The BI and ML analyses generated identical phylogenetic results with most posterior probabilities (PP) of one and bootstrap pseudoreplicates (BP) of 100 (Fig. 7). Pentatomoidea, Lygaeoidea, Pyrrhocoroidea, and Coreoidea, the trichophorans, were all monophyletic and highly supported in both analyses (PP > 0.85 and BP > 92). A sister relationship between Coreoidea and Pyrrhocoroidea was recovered, and Pentatomoidea formed a sister group with the other superfamilies of the remaining Trichophora. Both analyses provided robust support (PP > 0.95 and BP > 85) for three families within Coreoidea: Alydidae was closer to Coreidae than Rhopalidae. Subfamilies of Coreidae were also supported with an internal relationship of ((Hydarinae + Pseudophloeinae) + Coreinae) (PP > 0.77 and BP > 75). Among Rhopalinae species, Stictopleurus subviridis Hsiao, 1977 was identified as sister to a group of other Rhopalinae species (PP = 0.99 and BP = 100). M. lateralis was close to Chorosoma macilentum Stål, 1858 (PP = 1 and BP = 100) and they both belonged to the tribe Chorosomatini.

Figure 7. 

The phylogenetic relationships of PCG+rRNA using BI and ML methods. Numbers above each node indicate Bayesian posterior probabilities values and ML bootstrap values.

Discussion

In this study, we sequenced the mitogenome of Myrmus lateralis and built the molecular phylogenetic relationships with 37 other heteropteran taxa. The mitogenome was found to be 17,309 bp long, with 37 genes arranged consistent with other species of Hemiptera (Song et al. 2012; Li et al. 2017; Zhao et al. 2021). The control region represented the largest reported non-coding region for insect mitogenomes owing to its various motifs and tandem repeats (Cook et al. 1999; Li et al. 2015). Although the 2587 bp long control region possessed several elements, such as a polyT stretch, a G(A)nT sequence, an ATAGA motif, a stem-loop, and a polyA stretch, tandem repeats were not observed.

The present study analyzed the initiation and termination codons of the 13 PCGs. The majority of PCGs used common start codons (ATN), except for cox1, which was initiated with TTG, a frequently used start codon in heteropteran mitogenomes (Quek et al. 2021; Zhao et al. 2021). The partial termination codon T was found in nad2, cox2, cox3, nad3 and nad5. Some researchers have proposed that the incomplete termination codon could be completed via post-transcriptional polyadenylation (Ojala et al. 1981). Among the synonymous codons, codons ending with an A or U were more frequent than those ending in a G or C and were also observed in other heteropteran species (Zhao et al. 2019; Li et al. 2021). In this study, the lowest evolutionary rate was observed in cox1, indicating that this gene can be effectively used as a DNA barcoding marker. The Ka/Ks values for all 13 PCGs were far less than one, indicating that all of the PCGs evolved under purifying selection, thus, they can be used to investigate phylogenetic relationships.

Although many studies have reconstructed phylogenetic relationships among members of Pentatomomorpha, the relationships of different levels are still ambiguous (Li et al. 2005; Liu et al. 2019; Souza-Firmino et al. 2020; Kaur and Singh 2021). The results from previous research were congruent with the result that Trichophora includes four superfamilies of Pentatomomorpha except Aradoidea, and with the monophyly of four superfamilies of Trichophora (Henry 1997; Hua et al. 2008; Yuan et al. 2015; Song et al. 2019). Hua et al. (2008) and Liu et al. (2019) supported Pentatomoidea as the basal group of Trichophora and Coreoidea as a sister group to Lygaeoidea based on mitochondrial genomes. The superfamilies Coreoidea and Pyrrhocoroidea were supported as sister taxa based on cladistic analysis (Henry 1997), as well as 37 mitochondrial genes analysis (Yuan et al. 2015). Whereas our mitogenomic data showed that Coreoidea and Pyrrhocoroidea were sister groups based on ML and BI analyses using the PCG+rRNA dataset. This result from our study was consistent with that from the study of Song et al. (2019) based on relationship analysis of Pentatomomorpha PCGs.

At the family level, Rhopalidae was found to be a sister group to Alydidae + Coreidae. This relationship, based on mitogenomes, was consistent with that based on cladistic analysis by Li (1996) and Henry (1997). Despite the availability of numerous species, the classification patterns for the family Coreidae were weak in terms of both morphology and molecular data (Bressa et al. 2008; Forthman et al. 2019; Froeschner 2019). In this study, Hydarinae and Pseudophloeinae grouped the sister cluster. However, due to the limited samples, the relationships among Coreidae needs more studies in the future.

Previous studies on the classification of Rhopalinae came to a variety of conclusions and the positions of some species were not clear (Schaefer and Chopra 1982; Liu 1994; Aukema and Rieger 2006). The hypothesis that the genus Stictopleurus belongs to the Rhopalini tribe was corroborated based on morphological studies by Aukema and Rieger (2006). However, it did not group together with other Rhopalini species in this study. The monophyly of Rhopalini should be explored in the future studies with additional specimens.

Conclusions

The complete mitogenome of the scentless plant bug Myrmus lateralis was sequenced using next-generation sequencing technologies, providing the fifth mitogenome sequence from approximately 230 species of Rhopalidae. The nucleotide composition, codon usage, RNA structures, and protein-coding genes evolution were analyzed in our paper. The mitogenome of M. lateralis revealed the phylogenetic position of Myrmus. However, more mitogenomes should be sequenced to investigate the mitogenomic evolution and phylogenetic relationships of Rhopalidae.

Acknowledgments

This project was supported by the National Science Foundation Project of China (no. 31501876, 31872272) and supported by Scientific Research Project of Xinzhou Teachers University (2018KY04).

Reference

  • Dolling WR (2006) Family Rhopolidae Amyot & Serville, 1843. In: Catalog of the Heteroptera of the Palaearctic Region. The Netherlands Entomological Society, Amsterdam, 8–27.
  • Beckenbach AT, Stewart JB (2009) Insect mitochondrial genomics 3: the complete mitochondrial genome sequences of representatives from two neuropteroid orders: a dobsonfly (order Megaloptera) and a giant lacewing and an owlfly (order Neuroptera). Genome 52(1): 31–38. https://doi.org/10.1139/G08-098
  • 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 and Evolution 69(2): 313–319. https://doi.org/10.1016/j.ympev.2012.08.023
  • Bressa MJ, Franco MJ, Toscani MA, Papeschi AG (2008) Heterochromatin heteromorphism in Holhymenia rubiginosa (Heteroptera: Coreidae). European Journal of Entomology 105(1): 65–72. https://doi.org/10.14411/eje.2008.009
  • Carroll SP, Loye JE (2012) Soapberry bug (Hemiptera: Rhopalidae: Serinethinae) native and introduced host plants: biogeographic background of anthropogenic evolution. Annals of the Entomological Society of America 105(5): 671–684. https://doi.org/10.1603/AN11173
  • Cook CE, Wang Y, Sensabaugh G (1999) A mitochondrial control region and cytochromebphylogeny of sika deer (Cervus nippon) and report of tandem repeats in the control region. Molecular Phylogenetics and Evolution 12(1): 47–56. https://doi.org/10.1006/mpev.1998.0593
  • da Fonseca RR, Johnson WE, O’Brien SJ, Ramos MJ, Antunes A (2008) The adaptive evolution of the mammalian mitochondrial genome. BMC Genomics 9(1): 1–22. https://doi.org/10.1186/1471-2164-9-119
  • Darriba D, Taboada GL, Doallo R, Posada D (2012) jModelTest 2: more models, new heuristics and parallel computing. Nature methods 9(8): 772–772. https://doi.org/10.1038/nmeth.2109
  • Davidova-Vilimova J, Nejedla M, Schaefer CW (2000) Dorso-abdominal scent glands and metathoracic evaporatoria in adults of central European Rhopalidae (Hemiptera: Heteroptera), with a discussion of phylogeny and higher systematics. European Journal of Entomology 97(2): 213–222. https://doi.org/10.14411/eje.2000.039
  • Forthman M, Miller CW, Kimball RT (2019) Phylogenomic analysis suggests Coreidae and Alydidae (Hemiptera: Heteroptera) are not monophyletic. Zoologica Scripta 48(4): 520–534. https://doi.org/10.1111/zsc.12353
  • Froeschner RC (1988) Family Coreidae Leach, 1815: the coreid bugs. In: Catalog of the Heteroptera, or true bugs, of Canada and the continental United States. CRC Press, Boca Raton, 69–92. https://doi.org/10.1201/9781351070447-9
  • Ghahari H, Moulet P, Linnavuori R, Ostovan H (2012) An annotated catalog of the Iranian Coreidae, Rhopalidae, and Stenocephalidae (Hemiptera: Heteroptera: Pentatomomorpha: Coreoidea). Zootaxa 3519(1): 1–31. https://doi.org/10.11646/zootaxa.3519.1.1
  • Henry TJ (1997) Phylogenetic analysis of family groups within the infraorder Pentatomomorpha (Hemiptera: Heteroptera), with emphasis on the Lygaeoidea. Annals of the Entomological Society of America 90(3): 275–301. https://doi.org/10.1093/aesa/90.3.275
  • Henry TJ (1988) Family Rhopalidae Amyot and Serville, 1843 (= Corizidae Douglas and Scott, 1865): the scentless plant bugs. In: Catalog of the Heteroptera, or True Bugs, of Canada and the Continental United States. CRC Press, Boca Raton, 652–664. https://doi.org/10.1201/9781351070447-37
  • Hua J, Li M, Dong P, Cui Y, Xie Q, Bu W (2008) Comparative and phylogenomic studies on the mitochondrial genomes of Pentatomomorpha (Insecta: Hemiptera: Heteroptera). BMC genomics 9(1): 1–15. https://doi.org/10.1186/1471-2164-9-610
  • Kaur R, Singh D (2021) Cytochrome b sequence divergence and phylogenetic relationships among different species of family Pentatomidae (Hemiptera: Heteroptera). International Journal of Tropical Insect Science 41(2): 1177–1183. https://doi.org/10.1007/s42690-020-00303-8
  • Kearse M, Moir R, Wilson A, Stones-Havas S, Cheung M, Sturrock S, Buxton S, Cooper A, Markowitz S, Duran C (2012) Geneious Basic: an integrated and extendable desktop software platform for the organization and analysis of sequence data. Bioinformatics 28: 1647–1649. https://doi.org/10.1093/bioinformatics/bts199
  • Kocher A, Kamilari M, Lhuillier E, Coissac E, Péneau J, Chave J, Murienne J (2014) Shotgun assembly of the assassin bug Brontostoma colossus mitochondrial genome (Heteroptera, Reduviidae). Gene 552(1): 184–194. https://doi.org/10.1016/j.gene.2014.09.033
  • Linnavuori RE (2007) Studies on the Piesmatidae, Berytidae, Pyrrhocoridae, Stenocephalidae, Coreidae, Rhopalidae, Alydidae, Cydnidae, and Plataspidae (Heteroptera) of Gilan and the adjacent provinces in northern Iran. Acta Entomologica Musei Nationalis Pragae 47: 77–91. https://www.aemnp.eu/data/article-1138/1119-47_0_77.pdf
  • Li H, Leavengood JM, Chapman EG, Burkhardt D, Song F, Jiang P, Liu J, Zhou X, Cai W (2017) Mitochondrial phylogenomics of Hemiptera reveals adaptive innovations driving the diversification of true bugs. Proceedings of the Royal Society B: Biological Sciences 284(1862): e20171223. https://doi.org/10.1098/rspb.2017.1223
  • Li HM, Deng RQ, Wang JW, Chen ZY, Jia FL, Wang XZ (2005) A preliminary phylogeny of the Pentatomomorpha (Hemiptera: Heteroptera) based on nuclear 18S rDNA and mitochondrial DNA sequences. Molecular Phylogenetics and Evolution 37(2): 313–326. https://doi.org/10.1016/j.ympev.2005.07.013
  • Li R, Li M, Yan J, Bai M, Zhang H (2021) Five Mitochondrial Genomes of the Genus Eysarcoris Hahn, 1834 with Phylogenetic Implications for the Pentatominae (Hemiptera: Pentatomidae). Insects 12(7): 597. https://doi.org/10.3390/insects12070597
  • Li Y, Kocot KM, Schander C, Santos SR, Thornhill DJ, Halanych KM (2015) Mitogenomics reveals phylogeny and repeated motifs in control regions of the deep-sea family Siboglinidae (Annelida). Molecular Phylogenetics and Evolution 85: 221–229. https://doi.org/10.1016/j.ympev.2015.02.008
  • Liu Y, Li H, Song F, Zhao Y, Wilson JJ, Cai W (2019) Higher‐level phylogeny and evolutionary history of Pentatomomorpha (Hemiptera: Heteroptera) inferred from mitochondrial genome sequences. Systematic Entomology 44(4): 810–819. https://doi.org/10.1111/syen.12357
  • Liu Q, Zheng LY (1994) The classification and faunal analyses of Rhopalidae (Hemiptera) in China. Journal of Arid Land Resources and Environment 8(3): 102–115.
  • Nonnaizab (1986) Fauna of Inner Mongolia (Hemiptera: Heteroptera). Inner Mongolia People’s Publishing House, Hohhot, 298 pp.
  • Ojala D, Montoya J, Attardi G (1981) tRNA punctuation model of RNA processing in human mitochondria. Nature 290(5806): 470–474. https://doi.org/10.1038/290470a0
  • Quek ZBR, Chang JJM, Ip YCA, Chan YKS, Huang D (2021) Mitogenomes reveal alternative initiation codons and lineage-specific gene order conservation in echinoderms. Molecular Biology and Evolution 38(3): 981–985. https://doi.org/10.1093/molbev/msaa262
  • Rambaut A, Drummond AJ, Xie D, Baele G, Suchard MA (2018) Posterior summarization in Bayesian phylogenetics using Tracer 1.7. Systematic Biology 67(5): 901. https://doi.org/10.1093/sysbio/syy032
  • Ronquist F, Teslenko M, Van Der Mark P, Ayres DL, Darling A, Höhna S, Larget B, Liu L, Suchard MA, Huelsenbeck JP (2012) MrBayes 3.2: efficient Bayesian phylogenetic inference and model choice across a large model space. Systematic Biology 61(3): 539–542. https://doi.org/10.1093/sysbio/sys029
  • Schaefer CW, Chopra NP (1982) Cladistic analysis of the Rhopalidae, with a list of food plants. Annals of the Entomological Society of America 75(3): 224–233. https://doi.org/10.1093/aesa/75.3.224
  • Schaefer CW, Panizzi AR (2000) Scentless plant bugs (Rhopalidae). In: Schaefer CW, Panizzi AR (Eds) Heteroptera of Economic Importance. CRC Press, Boca Raton, 331–342. https://doi.org/10.1201/9781420041859
  • Simon C, Buckley TR, Frati F, Stewart JB, Beckenbach AT (2006) Incorporating molecular evolution into phylogenetic analysis and a new compilation of conserved polymerase chain reaction primers for animal mitochondrial DNA. Annual Review of Ecology, Evolution and Systematics 37: 545–579. https://doi.org/10.1146/annurev.ecolsys.37.091305.110018
  • Souza HV, Souza FB, Maruyama SRC, Castanhole MMU, Itoyama MM (2009) Meiosis, spermatogenesis and nucleolar behavior in the seminiferous tubules of Alydidae, Coreidae and Rhopalidae (Heteroptera) species. Genetics and Molecular Research 8(4): 1383–1396 https://doi.org/10.4238/vol8-4gmr672
  • Souza-Firmino TSD, Alevi KCC, Itoyama MM (2020) Chromosomal divergence and evolutionary inferences in Pentatomomorpha infraorder (Hemiptera, Heteroptera) based on the chromosomal location of ribosomal genes. Plos ONE 15(2): e0228631. https://doi.org/10.1590/S0074-02762013000300017
  • Tamura K, Stecher G, Peterson D, Filipski A, Kumar S (2013) MEGA6: molecular evolutionary genetics analysis version 6.0. Molecular Biology and Evolution, 30(12): 2725–2729. https://doi.org/10.1093/molbev/mst197
  • Tian XX, Xie Q, Li M, Gao CQ, Cui Y, Xi L, Bu W (2011) Phylogeny of pentatomomorphan bugs (Hemiptera-Heteroptera: Pentatomomorpha) based on six Hox gene fragments. Zootaxa 2888: 57–68. https://doi.org/10.11646/zootaxa.2888.1.5
  • Tyagi K, Chakraborty R, Cameron SL, Sweet AD, Chandra K, Kumar V (2020) Rearrangement and evolution of mitochondrial genomes in Thysanoptera (Insecta). Scientific reports 10(1): 1–16. https://doi.org/10.1038/s41598-020-57705-4
  • Vaidya G, Lohman DJ, Meier R (2011) SequenceMatrix: concatenation software for the fast assembly of multi‐gene datasets with character set and codon information. Cladistics 27(2): 171–180. https://doi.org/10.1111/j.1096-0031.2010.00329.x
  • Xu N, Ding J, Que Z, Xu W, Ye W, Liu H (2021) The mitochondrial genome and phylogenetic characteristics of the Thick-billed Green-Pigeon, Treron curvirostra: the first sequence for the genus. ZooKeys 1041: 167–182. https://doi.org/10.3897/zookeys.1041.60150
  • Xie Q, Bu W, Zheng L (2005) The Bayesian phylogenetic analysis of the 18S rRNA sequences from the main lineages of Trichophora (Insecta: Heteroptera: Pentatomomorpha). Molecular Phylogenetics and Evolution 34: 448–451. https://doi.org/10.1016/j.ympev.2004.10.015
  • Yang QQ, Liu SW, Song F, Liu GF, Yu XP (2018) Comparative mitogenome analysis on species of four apple snails (Ampullariidae: Pomacea). International Journal of Biological Macromolecules 118: 525–533. https://doi.org/10.1016/j.ijbiomac.2018.06.092
  • Yuan ML, Zhang QL, Guo ZL, Wang J, Shen YY (2015) The complete mitochondrial genome of Corizus tetraspilus (Hemiptera: Rhopalidae) and phylogenetic analysis of Pentatomomorpha. PLoS ONE 10(6): e0129003. https://doi.org/10.1371/journal.pone.0129003
  • Zhao L, Wei J, Zhao W, Chen C, Gao X, Zhao Q (2021) The complete mitochondrial genome of Pentatoma rufipes (Hemiptera, Pentatomidae) and its phylogenetic implications. ZooKeys 1042: 51–72. https://doi.org/10.3897/zookeys.1042.62302
  • Zhao W, Zhao Q, Li M, Wei J, Zhang X, Zhang H (2019) Comparative mitogenomic analysis of the Eurydema genus in the context of representative Pentatomidae (Hemiptera: Heteroptera) taxa. Journal of Insect Science 19(6): 20. https://doi.org/10.1093/jisesa/iez122
login to comment