Zum Hauptinhalt springen

Mitogenome of the leaf-footed bug Notobitus montanus (Hemiptera: Coreidae) and a phylogenetic analysis of Coreoidea.

Tian, Xiaoke ; Li, Yongqin ; et al.
In: PLoS ONE, Jg. 18 (2023), Heft 2, S. e0281597
Online academicJournal

Mitogenome of the leaf-footed bug Notobitus montanus (Hemiptera: Coreidae) and a phylogenetic analysis of Coreoidea  Introduction

Notobitus montanus Hsiao, 1963 is a major pest of bamboos. The mitogenome of N. montanus (ON052831) was decoded using next-generation sequencing. The mitogenome, with 42.26% A, 30.54% T, 16.54% C, and 10.65% G, is 16,209 bp in size. Codon usage analysis indicated that high frequently used codons used either A or T at the third position of the codon. Amino acid usage analysis showed that leucine 2, phenylalanine, isoleucine and tyrosine were the most abundant in 31 Coreoidea species. Thirteen protein-coding genes (PCGs) were evolving under purifying selection, nad5 and cox1 had the lowest and strongest purifying selection stress, respectively. Correlation analysis showed that evolutionary rate had positive correlation with A+T content. No tandem repeat was detected in the non-coding region of N. montanus. The phylogenetic tree showed that Alydidae and Coreidae were not monophyletic. However, the topology of phylogenetic trees, based on 13 PCGs, was in accordance with that of tree based on both mitochondrial and nuclear genes but not ultraconserved element loci or combination of 13 PCGs and two rRNAs. It seems that their relationships are complex, which need revaluation and revision. The mitogenomic information of N. montanus could shed light on the evolution of Coreoidea.

Leaf-footed bugs, hemipteran superfamily Coreoidea, are phytophagous insects which consist of several forest and agricultural pests [[1]]. For example, Notobitus montanus Hsiao, 1963 is a major pest of bamboos, which not only supply human with foods, building materials, crafts, and high-quality paper, but also are involved in landscaping and soil conservation [[3]]. It distributes in many southern provinces of China, such as Zhejiang, Sichuan, Yunnan, Guizhou, which serves as the pest of serval species of bamboo, including, Phyllostachys sulphurea var. Viridis, Phyllostachys glauca, Phyllostachys heteroclada, Phyllostachys bambusoides f. shouzhu, Phyllostachys heteroclada var. pubescens, Pleioblastus amzrus, Phyllostachys praecox f. provernalis [[4]]. The adults and nymphs of N. montanus feed on the sap of bamboo shoot and young bamboo, making bamboo falls into a decline and even dead. A female can lay 20–70 eggs, and the egg hatchability is high. Eggs take up 15–20 days to develop as nymphs, and then nymphs take up 30–50 days to develop as adult, which can live for 330–350 days [[4]]. Due to its high reproduction, strong ability of flying and gregariousness, N. montanus can cause serious economic loss of bamboo [[5]]. For example, in 1973, 90% of bamboo shoots were attacked by it in Liangping count, Sichuan province, leading 16.4% bamboo shoots dead [[4]]. Currently, approximately 30 Coreoidea species are recorded as the pest of bamboo.

In addition to two extinct families (Trisegmentatidae and Yruipopovinidae), Coreoidea consists of five families: Coreidae (2571 species), Alydidae (282 species), Rhopalidae (224 species), Stenocephalidae (30 species) and Hyocephalidae (three species) [[1], [6]]. The Coreidae is divided into four subfamilies: Coreinae, Hydarinae, Meropachyinae, and Pseudophloeinae [[6]]. Coreinae, with 2,320 (90%) species (372 genera), is the largest subfamily of Coreidae. In the past, the taxonomic classification of Coreidae was mainly based on morphological traits such as apical spine or tooth on the hind tibiae, hind femora, and metathoracic scent gland orifices [[8]]. The phylogenetic relationships among the tribal rank taxa of Coreidae have not been investigated comprehensively since 1997 [[8]]. However, some studies indicated that these morphological traits exhibited homoplasy [[1], [9]], which might explain contradictory results from different studies [[1]]. Phylogenomic trees constructed with ultraconserved element loci and mitogenome showed that both Coreidae and Alydidae were not monophyly [[1], [6]]. Furthermore, the phylogenomic analysis showed that several genera, and subfamilies of Coreidae were para- and polyphyly, which suggested that the taxonomic classification of Coreidae need revaluation and revision [[9]]. In short, leaf-footed bugs' phylogenetic relationships have remained far from solved [[1]].

Animal mitochondrial genome (mitogenome) is a circular double-stranded DNA molecule. Generally, it encodes 13 protein-coding genes (PCGs), two ribosomal RNA (rRNAs), and 22 transfer RNA (tRNAs) [[10]]. In addition, it contains a non-coding control region (A+T-rich region, D-loop), which contains essential regulatory elements for replication and transcription [[11]]. Owing to the lack of protection of histones, mitogenomic DNA has a much higher mutation frequency than nuclear DNA. As a result, mitogenomes are widely used in phylogenetics, phylogeography, population genetics, and evolutionary biology [[12]]. With the development of next-generation sequencing (high-throughput sequencing), many mitogenomes have been decoded using this technique.

In this work, the mitogenome of N. montanus has been decoded. The characteristics of the mitogenome, including nucleotide composition, codon usage, tRNA secondary structure, and the evolutionary pattern of 13 PCGs, were systematically analyzed. The phylogenetic tree of Coreoidea was constructed with the new mitogenome and 30 mitogenomes extracted from NCBI (S1 Table). The mitogenome could shed light on the evolution of Coreoidea.

Materials and methods

Samples and identification

Notobitus montanus is a common pest of bamboos in China as well as not recorded in the species list-of-ethics committees for research involving animals of the Guizhou Normal University. Therefore, no ethical approval or other relevant permission can be provided for the study. Specimens of N. montanus were collected from the campus of Guizhou Normal University (26°22′50.30″N, 106°38′11.72″E) in July 2021. All specimens were identified from their morphological characteristics. Total DNA was then extracted from the muscle tissue of an adult specimen with the phenol-chloroform extracting method [[14]]. The fragment of mitochondrial cytochrome c oxidase subunit I gene (cox1) was amplificated with primers (LCO1490 and HCO2198) [[16]]. The PCR productions were checked by agarose gel electrophoresis and sequenced by sanger sequencing. The sequence was submitted to BOLD systems v4 (http://www.boldsystems.org/) as a query for species identification. The similarity between the query sequence and reference sequence of N. montanus in the BOLD was 99.77%. The specimens (specimen ID: GZNU-cqq-136) were soaked in absolute alcohol and then stored at 4°C in the Museum of Guizhou Normal University.

Next-generation sequencing, annotation, and bioinformatics analysis

Total DNA was isolated from the muscle tissue of an adult specimen with ONE-4-ALL Genomic DNA Mini-Prep Kit (BS88504, Sangon, Shanghai, China). DNA was fragmented, then ~500 bp DNA was recycled. Paired-end libraries were constructed with the Illumina platform. The DNA was sequenced using the Illumina Hiseq X Ten at the Sangon Biotechnology Company (Shanghai, China). The adapter sequences were removed, and low-quality reads were trimmed with Trimmomatic version 0.36 [[17]]. The mitogenome of Cloresmus pulchellus (NC_042806) was used as reference, and clean reads were assembled with SOAPdenovo2 (version 2.04) [[18]]. PCGs were identified by BLAST comparison with C. pulchellus mitogenome [[19]]. The secondary structures of tRNAs were predicted with MITOS2 and tRNAscan-SE 2.0 [[20]], rRNAs and non-coding control region were determined by the boundary of tRNAs. Nucleotide composition was calculated with MEGA X [[22]]. The AT-skew values were calculated by the following formula: AT-skew = (A—T)/(A + T). Similarly, GC-skew values were calculated by GC-skew = (G—C)/(G + C). Codon usage indexes, including codons counts, and relative synonymous codon usage (RSCU), were calculated with MEGA X [[22]]. The number of nonsynonymous substitutions per nonsynonymous site (Ka), and the number of synonymous substitutions per synonymous site (Ks) were calculated with DnaSP 6 [[23]]. Tandem repeats in the non-coding control region were identified using tandem repeats online finder server [[24]]. Circular map of the N. montanus mitogenome was generated with CGView server [[25]]. Heatmap of codon and amino acid usage was generate with ggplot2 as implemented in R v4.1.2. Figures were edited with Adobe Illustrator CS5.

Phylogenetic analysis

The mitogenomes of N. montanus and 30 available complete mitogenomes of Coreoidea in NCBI were used to construct the phylogenetic tree of Coreoidea (S1 Table). Malcus inconspicuus (Hemiptera: Malcidae), Physopelta gutta (Hemiptera: Largidae), and Nezara viridula (Hemiptera: Pentatomidae) were selected as representative of the outgroups. These mitogenomes were imported into PhyloSuite V1.2.2 [[26]]. Then the nucleic acid sequences of 13 PCGs were extracted from these mitogenomes. Codon-based multiple alignments were carried out with MAFFT as implemented in PhyloSuite [[26]]. Then alignments of 13 PCGs were concatenated. PartitionFinder2 was used to select the best-fit partitioning strategy and models for the concatenated sequences. Phylogenetic trees were reconstructed by the Bayesian (Mrbayes v 3.2.6) and maximum likelihood (IQ-TREE v1.6.8) method as implemented in PhyloSuite [[26]]. The number of generations was 10, 000, 000 and Partition Models were selected as models. The phylogenetic tree was visualized with Figtree v1.4.4 and Adobe Illustrator CS5.

Results

Genomic structure and nucleotide composition

The complete mitogenome of N. montanus (ON052831) was 16209 bp, which encoded 13 PCGs, 22 tRNAs, two rRNAs, and a non-coding control region (D-loop) (Fig 1 and S2 Table). Fourteen genes, including four PCGs, eight tRNAs, and two rRNAs, were encoded by the minority strand (N strand), and the remaining genes were encoded by the majority strand (J strand). The gene arrangement of N. montanus was in accordance with that of other Coreoidea species [[19], [27]]. A total of 39 bp intergenic spacers were distributed across eight locations. The shortest intergenic spacers were one bp, and the longest intergenic spacer which was located between trnS2 and nad1, was 21 bp. A total of 28 bp overlaps were distributed across seven locations. The shortest overlaps were one bp, and the longest overlap was eight bp which was located between trnW and trnC. two seven bp overlaps were located between atp8 and atp6, between nad4 and nad4L.

MAP: Fig 1 Mitogenome map of Notobitus montanus.Protein coding genes (PCGs, CDS) and ribosomal genes (rRNAs) are presented with standard abbreviations. Genes coding for transfer RNAs (tRNAs) are presented with one letter abbreviation. S1 = AGN, S2 = UCN, L1 = CUN, L2 = UUR. Wathet blue, pink, light green and orange represents PCGs, rRNAs, tRNAs and D-loop (non-coding control region), respectively. Black, green and purple represents GC content, positive GC skew (GC skew+) and negative GC skew (GC skew-), respectively. Gene orientation is indicated by arrows.

The mitogenome consisted of 42.26% A, 30.54% T, 16.54% C and 10.65% G (S3 Table). The A+T content of the whole genome, PCGs, tRNAs, rRNAs and non-coding control region was 72.81%, 72.79%, 75.35%, 75.59% and 67.17%, respectively. It seemed that the name of A-T rich region was not suitable for non-coding control region. The AT-skew value was 0.161, and the GC-skew value was -0.220, which indicated that the mitogenome had a preference to A and C. The AT-skew of Coreidae species ranged from 0.097 (Leptoglossus membranaceus) to 0.176 (Acanthocoris sp. FS-2019), and GC-skew ranged from -0.253 (Hydaropsis longirostris) to -0.187 (Clavigralla tomentosicollis), which indicated A and C were more abundant than T and G in Coreidae species, respectively.

Protein-coding genes

The total length of 13 PCGs of N. montanus was 11047 bp (S2 Table), accounting for 72.79% of the whole genome, which was medium size in Coreidae species (from 11020 bp to 11110 bp). For all Coreidae species, all AT-skew values were positive, and GC-skew values were positive except for Leptoglossus membranaceus (-0.008) and Notopteryx soror (-0.001). Four PCGs, including nad1, nad4, nad4L, and nad5, were encoded by the N strand, and the remaining nine PCGs were encoded by the J strand (Fig 1 and S2 Table). In N. montanus, cox1 used TTG as start codon, however, the remaining 12 PCGs used ATN as start codon (ATA for atp8, nad3, and nad6; ATC for cox2; ATG for atp6, cox1, cox3, nad2, nad4, nad5, and cytb; ATT for nad1, and nad4L). Five PCGs, including atp8, atp6, nad5, nad4, and nad4L, used TAA as stop codon; however, two PCGs (nad3, and nad6) and six PCGs (cox1, cox2, cox3, cytb, nad1, and nad2) used TA and T as incomplete stop codon, respectively (S2 Table).

Relative synonymous codon usage (RSCU) analysis of N. montanus showed that a total of 62 codons were used, except for two stop codons (UAA and UAG) (S4 Table). The highest frequent four codons were UUA (264), UUU (251), AAU (243), and AUA (204), which accounted for 26.13% total number of codons. These codons were composed of either A or U. On the contrary, high G+C contents codons, including GCG (2), CGC (3), and CGG (4), were the lowest frequent codons. For codons with RSCU values more than one, they were more likely to use either A or T rather than either G or C at the third position of the codon (Fig 2A and 2B). The heatmap of codon usage for 31 Coreoidea species showed that codons used either A or T at the third position of the codon which were more frequently used than these used either G or C.

Graph: Fig 2 Codon and amino acid usage.(A) Relative synonymous codon usage (RSCU) of N. montanus. (B) Heatmap indicates codon usage in 31 Coreoidea species. (C) Heatmap indicates amino acid usage in 31 Coreoidea species.

For amino acids, leucine 2, phenylalanine, isoleucine, methionine, tyrosine, and asparagine, were the most abundant in N. montanus, which accounted for 9.92%, 8.92%, 8.86%, 7.52%, 6.96% and 6.15% the total number of amino acids, respectively (Fig 2C). The heatmap of amino acid usage for 31 Coreoidea species showed that leucine 2, phenylalanine, isoleucine and tyrosine were the most abundant, and the abundance variation was small. However, the abundance of methionine and asparagine had larger variation among species than that of leucine 2, phenylalanine, isoleucine and tyrosine. Furtherly, there was a linear correlation between the effective number of codons (ENc) and the G+C content of 13 PCGs (S1 Fig). Specifically, ENc had a better linear correlation with the G+C content of the third position of the codon than the remaining positions.

Nad5 and cytb had the lowest and highest Ks values, respectively (Fig 3). Cox1 had the lowest value of Ka and Ka/Ks, atp8 had the highest value of Ka, and nad5 the highest value of Ka/Ks (Fig 3). The value of Ka/Ks for 13 PCGs was less than 1, which indicated that all PCGs were evolving under purifying selection. There were negative correlations between the value of Ka/Ks and G+C contents (S2 Fig).

Graph: Fig 3 Evolutionary rates of 13 protein-coding genes in the mitogenome of 31 Coreoidea.Red represents nonsynonymous nucleotide substitutions per nonsynonymous site (Ka); black represents synonymous nucleotide substitutions per synonymous site (Ks); blue represents the ratio of Ka to Ks (Ka/Ks). Error bars presented the standard error of mean.

Transfer RNA and ribosomal RNA genes

The total length of 22 tRNAs of N. montanus was 1444 bp. The shortest tRNAs were 63 bp (trnI, trnC, trnY, trnD, trnG, trnA, trnT and trnP), while the longest tRNA was 75 bp (trnK). The total A+T content of tRNAs was 75.35%. The AT-skew value was 0.079, and the GC-skew value was -0.096. Eight tRNAs (trnQ, trnC, trnY, trnF, trnH, trnP, trnL, and trnV) were encoded by the N strand, and the remaining tRNAs were encoded by the J strand (Fig 1 and S2 Table). The distribution pattern of tRNAs in the mitogenome was in accordance with that of other Coreidae species.

For secondary structure, trnS1 lacked a dihydrouridine (DHU) arm (Fig 4). The remaining tRNAs had a typical cloverleaf secondary structure. Non-Watson-Crick base pairing (G-U pairing) appeared in the acceptor stem of trnA, trnH and trnV. G-U pairing also appeared in the anticodon arm of trnV, and the TΨC arm of trnS1 and trnV. Four tRNAs (trnC, trnH, trnT, and trnS2) used U as the discriminator nucleotide, and the remaining 18 tRNAs used A as the discriminator nucleotide.

Graph: Fig 4 Predicted secondary structure of 22 tRNAs of N. montanus.

The mitogenome of N. montanus encoded two rRNAs. RrnL (16s rRNA) was 1281 bp which was located between trnL1 and trnV, and rrnS (12s rRNA) was 784 bp which was located between trnV and the non-coding control region. Both were encoded by the N strand (S2 Table). The A+T content of rRNAs was 75.59%, the AT-skew value was 0.213, and the GC-skew value was -0.310 (S3 Table).

Non-coding control region

The non-coding control region (D-loop) of N. montanus, located between 12s rRNA and trnI, was 1642 bp (Fig 5). Its position in the mitogenome was accordance with that of other Coreidae species [[6]]. The A+T content of D-loop was 67.17%. AT-skew value was 0.081, and GC-skew value was -0.328 (S3 Table). Among the 15 Coreidae species, the length of D-loop ranged from 794 bp (Cletus rubidiventris) to 3441 bp (C. pulchellus), which made significant contribution to the size variation of mitogenome (Fig 5) [[28]]. No tandem repeat was detected in the non-coding region of N. montanus (Fig 5). Similar phenomenon was occurred in Hydaropsis longirostris, Pseudomictis tenebrosa, Manocoreus sp. However, most Coreidae species had tandem repeats in their non-coding control region. For Coreidae species, the shortest tandem repeat unit in D-loop was seven bp (Molipteryx lunata), and the longest tandem repeat unit was 369 bp (C. pulchellus) (Fig 5).

Graph: Fig 5 Tandem repeat sequences in the non-coding control region.The copy number of tandem repeats are shown by colored (blue, red, purple) oval with Arabic numerals. Non-repeat regions are indicated by green box with sequence size inside.

Phylogenetic relationship

Phylogenetic analyses were carried out with nucleotide sequences of 13 PCGs extracted from 31 mitogenomes of Coreoidea (S1 Table). The topologies of Bayesian inference (BI) and maximum likelihood (ML) phylogenetic trees, with high support scores at most nodes, were identical except for Alydinae (Fig 6 and S3 Fig). Bayesian tree showed that five Alydinae species' relationship was (((Megalotomus costalis + Riptortus pedestris) + Camptopus lateralis) + (Daclera levana + Melanacanthus marginatus)). However, maximum likelihood tree showed that their relationship was ((((Camptopus lateralis + Riptortus pedestris) + Melanacanthus marginatus) + Daclera levana) + Megalotomus costalis). Compared with maximum likelihood tree, Bayesian tree had higher support scores.

Graph: Fig 6 Bayesian phylogenetic tree of 31 Coreoidea species.The tree is constructed with nucleotide sequences of 13 mitochondrial PCGs. For family, Black, red, orange, purple and dark blue represents outgroup, Stenocephalidae, Rhopalidae, Alydidae and Coreidae, respectively. For subfamily, orange, bright red, wathet blue, green, bright green, and bright blue represents Rhopalinae, Micrelytrinae, Hydarinae, Pseudophloeinae, Alydinae, and Coreinae, respectively. The posterior probabilities were labeled at each node.

Stenocephalidae had a sister group relationship with the remaining families of Coreoidea (PP = 1, BS = 98.60). Stenocephalidae and Rhopalidae, were monophyly; however, Alydidae and Coreidae were not monophyly (Fig 6). At subfamily level, Alydinae and Coreinae showed monophyly with high support (PP = 1, BS>97). Their relationship was (((((Alydinae + Pseudophloeinae) + Hydarinae) + Micrelytrinae) + Coreinae) + Rhopalinae). At present, Coreinae have 13 genera with complete mitogenomic information, their relationship was (Manocoreus + (Enoplops + Cletus)) + (((Acanthocoris + (Notobitus + Cloresmus)) + (Leptoglossus + (Mictis + (Pseudomictis + (Anoplocnemis + (Molipteryx + (Notopteryx + Cletomorpha))))))).

Discussion

Two seven bp overlaps were detected in PCGs, including between atp8 and atp6, between nad4 and nad4L. These two overlaps are popular in arthropods [[14], [29]]. The A+T content of PCGs was 75.79%, the AT-skew value was -0.116, and the GC-skew value was -0.004 (S3 Table). It seemed that PCGs had a preference to T than A, and no obvious preference between G and C. In N. montanus, only cox1 used TTG as start codon, and the remaining 12 PCGs used ATN as start codon. Our results are in accordance with that most PCGs use ATN as start codon in metazoan [[10]]. TTG is usually used as start codon for cox1 in animals [[30]]. Eight PCGs used TA (nad3 and nad6) or T (cox1, cox2, cox3, cytb, nad1, and nad2) as incomplete stop codon. Incomplete stop codons are usually used in metazoan, which can be added with either A or AA by post-transcriptional polyadenylation [[10], [13]]. Relative synonymous codon usage (RSCU) analysis of N. montanus showed that they were more likely to use either A or T rather than either G or C at the third position of the codon. For 31 Coreoidea species, they preferred to use codons which used either A or T at the third position of the codon than codons which used either G or C at the third position of the codon. Furtherly, there was a positive correlation between the effective number of codons (ENc) and the G+C content of 13 PCGs (S1 Fig). Taken together, G+C content could influence codon usage [[32]]. The value of Ka/Ks for 13 PCGs was less than 1, which indicated that all PCGs were evolving under purifying selection. As a result, all PCGs could be used to construct phylogenetic tree. Nad5 had the highest value of Ka/Ks, which suggested that nad5 had the fastest evolutionary rate in all PCGs. Cox1, with the lowest value of Ka/Ks, had the strongest purifying selection stress. Cox1 is suitable for barcoding marker. Correlation analysis indicated that the value of Ka/Ks had a negative correlation with G+C content (S2 Fig). In short, G+C content not only influences codon usage but also evolutionary rate.

The AT-skew value was 0.079, and the GC-skew value was -0.096, which indicated that tRNAs had a weak preference to A and C. Most tRNAs had the typical cloverleaf secondary structure, but trnS1 lacked a dihydrouridine (DHU) arm. Previous studies reveal that the phenomenon of trnS1 lacking DHU arm is popular in insects [[34]]. Four tRNAs (trnC, trnH, trnT, and trnS2) used U as the discriminator nucleotide, and the remaining 18 tRNAs used A as the discriminator nucleotide. This phenomenon occurs in other insects, such as Dinorhynchus dybowskyi (NC_037724) [[29]]. Non-Watson-Crick base pairing (G-U pairing) appeared in the acceptor stem, anticodon arm and TΨC arm of tRNAs. Non-Watson-Crick base pairing in tRNA could increase the stability of tRNAs, which is popular in insects [[13]]. The A+T content of rRNAs was 75.59%, the AT-skew value was 0.213, and the GC-skew value was -0.310 (S3 Table). It seemed that rRNAs had similar A+T content with tRNAs, however, rRNAs had stronger preference to A and C than tRNAs.

The position of non-coding control region (D-loop) in N. montanus was accordance with that of other Coreidae species [[6]]. The A+T content of D-loop was 67.17%, which was lower than PCGs, tRNAs and rRNAs. It seems that the name of A-T rich region is not suitable for non-coding control region. AT-skew value was 0.081, and GC-skew value was -0.328 (S3 Table), which suggested that non-coding control region had a weak preference to A but had a strong preference to C. The length of non-coding control region ranged from 794 bp (Cletus rubidiventris) to 3441 bp (C. pulchellus) in 15 Coreidae species, which made significant contribution to the size variation of mitogenome (Fig 5) [[28]]. No tandem repeat was detected in the non-coding region of N. montanus (Fig 5). Similar phenomenon was occurred in Hydaropsis longirostris, Pseudomictis tenebrosa, Manocoreus sp [[34]]. Most Coreidae species had tandem repeats in their non-coding control region. For Coreidae species, the shortest tandem repeat unit was seven bp (Molipteryx lunata), and the longest tandem repeat unit was 369 bp (C. pulchellus) (Fig 5), which might influence mitochondrial replication and post-transcriptional modifications [[10], [13]]. High frequent mutations were occurred in the D-loop, making region was variable in both size and nucleotide sequence [[28]].

Coreoidea consists of five extant families: Coreidae, Alydidae, Rhopalidae, Stenocephalidae and Hyocephalidae [[1], [6]]. Currently, no mitogenome of Hyocephalidae species is available in NCBI, as a result, only four families were analyzed in the study. Stenocephalidae had a sister group relationship with the remaining families of Coreoidea (PP = 1, BS = 98.60), which was consistent with the previous study [[6]]. Stenocephalidae and Rhopalidae, were monophyly; however, Alydidae and Coreidae were not monophyly (Fig 6), which was in accordance with previous studies [[1], [6]].

At subfamily level, Alydinae and Coreinae showed monophyly with high support (PP = 1, BS>97). Our phylogenetic tree, based on 13 PCGs, showed that their relationship was (((((Alydinae + Pseudophloeinae) + Hydarinae) + Micrelytrinae) + Coreinae) + Rhopalinae). One study, based on 13 PCGs and 2 rRNAs showed their relationship was ((((Coreinae + Hydarinae) + Micrelytrinae) + (Alydinae + Pseudophloeinae)) + Rhopalinae) [[6]]. Another study, based on ultraconserved element loci showed their relationship was ((((Micrelytrinae + Hydarinae) + (Alydinae + Pseudophloeinae)) + Coreinae) + Rhopalinae) [[1]]. Rhopalinae was a sister group of the remaining five subfamilies in common. In addition, Alydinae had a closer relationship with Pseudophloeinae than the remaining subfamilies. Our tree showed that Coreinae was a sister group of the remaining four subfamilies, including Alydinae, Pseudophloeinae, Hydarinae, Micrelytrinae, which was consistent with the tree based on ultraconserved element loci but not the tree based on 13 PCGs and 2 rRNAs. Previous study indicated that the sequence of rRNAs was not suitable for phylogenetic tree reconstruction [[34]]. However, the topology of our tree was in accordance with that of tree based on both mitochondrial and nuclear genes [[35]] but not ultraconserved element loci [[1]]. It seems that their relationships are complex, and their relationships need revaluation and revision [[9]].

At present, Coreinae have 13 genera with complete mitogenomic information, their relationship was (Manocoreus + (Enoplops + Cletus)) + (((Acanthocoris + (Notobitus + Cloresmus)) + (Leptoglossus + (Mictis + (Pseudomictis + (Anoplocnemis + (Molipteryx + (Notopteryx + Cletomorpha))))))). Coreinae, with 372 genera, is the largest subfamily of Coreidae. The mitogenome of most genera remains unknown. Owing to a large number of species belonging to Coreinae (2320 species), phylogenetic analysis with ultraconserved elements showed that some genera were not monophyletic [[9]], which suggested that the relationships among Coreinae species were complex, and need further study. The mitogenomic information of N. montanus could shed light on the evolution of Coreoidea.

Conclusions

This study has decoded the complete mitogenome of N. montanus. Codon usage analysis indicated that high frequently used codons used either A or T at the third position of the codon. Amino acid usage analysis showed that leucine 2, phenylalanine, isoleucine and tyrosine were the most abundant in 31 Coreoidea species. The value of Ka/Ks for 13 PCGs was less than one, which suggested that all PCGs were evolving under purifying selection, Nad5 and Cox1 had the lowest and strongest purifying selection stress, respectively. Correlation analysis showed that A+T content could influence codon usage and evolutionary rate. The phylogenetic tree showed that Alydidae and Coreidae were not monophyletic. However, the topology of phylogenetic trees, based on 13 PCGs, was in accordance that of tree based on both mitochondrial and nuclear genes but not ultraconserved element loci or combination of 13 PCGs and two rRNAs. It seems that their relationships are complex, and their relationships need revaluation and revision. The mitogenomic information of N. montanus could shed light on the evolution of Coreoidea.

Supporting information

S1 Fig. Effective numbers of codons have a positive correlation with the G+C content of codons.

(A) total GC content of codons; (B) GC content of the first position of the codon; (C) GC content of the second position of the codon; (D) GC content of the third position of the codon.

(EPS)

S2 Fig. The Ka/Ks has a negative correlation with the G+C content of codons.

(A) total GC content of codons; (B) GC content of the first position of the codon; (C) GC content of the second position of the codon; (D) GC content of the third position of the codon.

(EPS)

S3 Fig. Maximum likelihood phylogenetic tree of 31 Coreoidea species.

The tree is constructed with nucleotide sequences of 13 mitochondrial PCGs. For family, Black, red, orange, purple and dark blue represents outgroup, Stenocephalidae, Rhopalidae, Alydidae and Coreidae, respectively. For subfamily, orange, bright red, wathet blue, green, bright green, and bright blue represents Rhopalinae, Micrelytrinae, Hydarinae, Pseudophloeinae, Alydinae, and Coreinae, respectively. The bootstrap values were labeled at each node.

(EPS)

S1 Table. List of species for phylogenetic analysis.

(DOCX)

S2 Table. Annotation of the Notobitus montanus mitogenome.

(DOCX)

S3 Table. Nucleotide composition of Notobitus montanus (%).

(DOCX)

S4 Table. Codon usage in the mitochondrial genome of Notobitus montanus.

(DOCX)

Footnotes 1 The authors have declared that no competing interests exist. References Forthman M, Miller CW, Kimball RT. Phylogenomic analysis suggests Coreidae and Alydidae (Hemiptera: Heteroptera) are not monophyletic. Zoologica Scripta. 2019; 48(4):520–34. doi: 10.1111/zsc.12353 WOS:000473237900009. 2 Fernandes JAM, Mitchell PL, Livermore L, Nikunlassi M. Leaf-Footed Bugs (Coreidae). In: Panizzi AR, Grazia J, editors. True Bugs (Heteroptera) of the Neotropics.Dordrecht: Springer Netherlands; 2015. p. 549–605. 3 Wang H, Varma Xu, editors. Insect Pests of Bamboos in Asia: An Illustrated Manual2002. 4 Zhanget al. Chinese economic entomography: Hemiptera (one): Science Press; 1985. 5 Miyatake T.Territorial mating aggregation in the bamboo bug, Notobitus meleagris, Fabricius (Heteroptera: Coreidae).Journal of Ethology. 1995; 13(2):185–9. 6 Dong X, Wang K, Tang Z, Zhang Y, Yi W, Xue H, et al. Phylogeny of Coreoidea based on mitochondrial genomes show the paraphyly of Coreidae and Alydidae. Archives of Insect Biochemistry and Physiology. 2022; 110(1):e21878. https://doi.org/10.1002/arch.21878. doi: 10.1002/arch.21878, 35181948 7 Team C.Coreoidea species file online. Version 5.0/5.0. Retreived from http://CoreoideaSpeciesFileorg. 2019. 8 Li X.Cladistic analysis of the phylogenetic relationships among the tribal rank taxa of Coreidae (Hemiptera-Heteroptera: Coreoidea).Acta Zootaxonomica Sinica. 1997; 22:60–8. 9 Forthman M, Miller CW, Kimball RT. Phylogenomics of the Leaf-Footed Bug Subfamily Coreinae (Hemiptera: Coreidae).Insect Systematics and Diversity. 2020; 4(4). doi: 10.1093/isd/ixaa009 Boore JL. Animal mitochondrial genomes. Nucleic acids research. 1999; 27(8):1767–80. Epub 1999/04/02. doi: 10.1093/nar/27.8.1767, 10101183; PubMed Central PMCID: PMC148383. Wolstenholme DR. Animal mitochondrial DNA: structure and evolution. International review of cytology. 1992; 141:173–216. Epub 1992/01/01. doi: 10.1016/s0074-7696(08)62066-5, 1452431. Wang IJ. Recognizing the temporal distinctions between landscape genetics and phylogeography. Molecular ecology. 2010; 19(13):2605–8. Epub 2010/06/22. doi: 10.1111/j.1365-294X.2010.04715.x, 20561197. Cameron SL. Insect mitochondrial genomics: implications for evolution and phylogeny. Annu Rev Entomol. 2014; 59:95–117. Epub 2013/10/29. doi: 10.1146/annurev-ento-011613-162007, 24160435. Chen QQ, Niu XJ, Fang Z, Weng QB. The complete mitochondrial genome of Melanostoma orientale (Diptera: Syrphidae).Mitochondrial DNA Part B-Resour.2020; 5(1):554–5. doi: 10.1080/23802359.2019.1710285 WOS:000507583000001., 33366643 Chen QQ, Niu XJ, Fang Z, Weng QB. The complete mitochondrial genome of Eysarcoris guttigerus (Hemiptera: Pentatomidae).Mitochondrial DNA Part B-Resour. 2020; 5(1):687–8. doi: 10.1080/23802359.2020.1714498 WOS:000507589900001., 33366704 Folmer O, Black M, Hoeh W, Lutz R, Vrijenhoek R. DNA primers for amplification of mitochondrial cytochrome c oxidase subunit I from diverse metazoan invertebrates. Molecular marine biology and biotechnology. 1994; 3(5):294–9. Epub 1994/10/01., 7881515. Bolger AM, Lohse M, Usadel B. Trimmomatic: a flexible trimmer for Illumina sequence data. Bioinformatics (Oxford, England).2014; 30(15):2114–20. Epub 2014/04/04. doi: 10.1093/bioinformatics/btu170, 24695404; PubMed Central PMCID: PMC4103590. Luo R, Liu B, Xie Y, Li Z, Huang W, Yuan J, et al. SOAPdenovo2: an empirically improved memory-efficient short-read de novo assembler. GigaScience. 2012; 1(1):18. Epub 2012/01/01. doi: 10.1186/2047-217X-1-18, 23587118; PubMed Central PMCID: PMC3626529. Liu Y, Li H, Song F, Zhao Y, Wilson J-J, Cai W. Higher-level phylogeny and evolutionary history of Pentatomomorpha (Hemiptera: Heteroptera) inferred from mitochondrial genome sequences. Systematic Entomology. 2019; 44(4):810–9. doi: 10.1111/syen.12357 Bernt M, Donath A, Juhling F, Externbrink F, Florentz C, Fritzsch G, et al. MITOS: improved de novo metazoan mitochondrial genome annotation. Molecular phylogenetics and evolution. 2013; 69(2):313–9. Epub 2012/09/18. doi: 10.1016/j.ympev.2012.08.023, 22982435. Chan PP, Lowe TM. tRNAscan-SE: Searching for tRNA Genes in Genomic Sequences.Methods in molecular biology (Clifton, NJ).2019; 1962:1–14. Epub 2019/04/26. doi: 10.1007/978-1-4939-9173-0_1, 31020551; PubMed Central PMCID: PMC6768409. Kumar S, Stecher G, Li M, Knyaz C, Tamura K. MEGA X: Molecular Evolutionary Genetics Analysis across Computing Platforms. Molecular biology and evolution. 2018; 35(6):1547–9. doi: 10.1093/molbev/msy096 WOS:000441140000027., 29722887 Rozas J, Ferrer-Mata A, Sánchez-DelBarrio JC, Guirao-Rico S, Librado P, Ramos-Onsins SE, et al. DnaSP 6: DNA Sequence Polymorphism Analysis of Large Data Sets. Molecular biology and evolution. 2017; 34(12):3299–302. Epub 2017/10/14. doi: 10.1093/molbev/msx248, 29029172. Benson G.Tandem repeats finder: a program to analyze DNA sequences. Nucleic acids research. 1999; 27(2):573–80. doi: 10.1093/nar/27.2.573, 9862982 Stothard P, Wishart DS. Circular genome visualization and exploration using CGView. Bioinformatics (Oxford, England).2005; 21(4):537–9. Epub 2004/10/14. doi: 10.1093/bioinformatics/bti054, 15479716. Zhang D, Gao FL, Jakovlic I, Zou H, Zhang J, Li WX, et al. PhyloSuite: An integrated and scalable desktop platform for streamlined molecular sequence data management and evolutionary phylogenetics studies. Molecular ecology resources. 2020; 20(1):348–55. doi: 10.1111/1755-0998.13096 WOS:000494565200001., 31599058 Chen Q, Tian X, Li Y, Chen Q. Complete mitochondrial genome of Liorhyssus hyalinus (Hemiptera: Rhopalidae).Mitochondrial DNA Part B. 2022; 7(8):1438–40. doi: 10.1080/23802359.2022.2107456, 35965648 Zhang DX, Hewitt GM. Insect mitochondrial control region: A review of its structure, evolution and usefulness in evolutionary studies. Biochem Syst Ecol. 1997; 25(2):99–120. doi: 10.1016/S0305-1978(96)00042-7 Zhao Q, Wang J, Wang MQ, Cai B, Zhang HF, Wei JF. Complete Mitochondrial Genome of Dinorhynchus dybowskyi (Hemiptera: Pentatomidae: Asopinae) and Phylogenetic Analysis of Pentatomomorpha Species.Journal of insect science (Online).2018; 18(2). Epub 2018/05/03. doi: 10.1093/jisesa/iey031, 29718506; PubMed Central PMCID: PMC5905379. Guzman LB, Vogler RE, Beltramino AA. The mitochondrial genome of the semi-slug Omalonyx unguis (Gastropoda: Succineidae) and the phylogenetic relationships within Stylommatophora.PloS one.2021; 16(6). doi: 10.1371/journal.pone.0253724 WOS:000671694400066., 34170937 Zhu LQ, Zhu ZH, Zhu LY, Wang DQ, Wang JX, Lin Q. The complete mitogenome of Lysmata vittata (Crustacea: Decapoda: Hippolytidae) with implication of phylogenomics and population genetics. PloS one. 2021; 16(11). doi: 10.1371/journal.pone.0255547 WOS:000755077100009., 34735446 Dey P, Sharma SK, Sarkar I, Ray SD, Pramod P, Kochiganti VHS, et al. Complete mitogenome of endemic plum-headed parakeet Psittacula cyanocephala—characterization and phylogenetic analysis. PloS one.2021; 16(4). doi: 10.1371/journal.pone.0241098 WOS:000639361000018., 33836001 Hershberg R, Petrov DA. Selection on codon bias. Annual review of genetics. 2008; 42:287–99. Epub 2008/11/06. doi: 10.1146/annurev.genet.42.110807.091442, 18983258. Yuan ML, Zhang QL, Guo ZL, Wang J, Shen YY. Comparative mitogenomic analysis of the superfamily Pentatomoidea (Insecta: Hemiptera: Heteroptera) and phylogenetic implications. Bmc Genomics. 2015;16. doi: 10.1186/s12864-015-1679-x WOS:000356405700002., 26076960 Ye F, Kment P, Rédei D, Luo J-Y, Wang Y-H, Kuechler SM, et al. Diversification of the phytophagous lineages of true bugs (Insecta: Hemiptera: Heteroptera) shortly after that of the flowering plants. Cladistics. 2022; 38(4):403–28. doi: 10.1111/cla.12501, 35349192

By Xiaoke Tian; Yongqin Li; Qin Chen and Qianquan Chen

Reported by Author; Author; Author; Author

Titel:
Mitogenome of the leaf-footed bug Notobitus montanus (Hemiptera: Coreidae) and a phylogenetic analysis of Coreoidea.
Autor/in / Beteiligte Person: Tian, Xiaoke ; Li, Yongqin ; Chen, Qin ; Chen, Qianquan
Link:
Zeitschrift: PLoS ONE, Jg. 18 (2023), Heft 2, S. e0281597
Veröffentlichung: Public Library of Science (PLoS), 2023
Medientyp: academicJournal
ISSN: 1932-6203 (print)
DOI: 10.1371/journal.pone.0281597
Schlagwort:
  • Medicine
  • Science
Sonstiges:
  • Nachgewiesen in: Directory of Open Access Journals
  • Sprachen: English
  • Collection: LCC:Medicine ; LCC:Science
  • Document Type: article
  • File Description: electronic resource
  • Language: English

Klicken Sie ein Format an und speichern Sie dann die Daten oder geben Sie eine Empfänger-Adresse ein und lassen Sie sich per Email zusenden.

oder
oder

Wählen Sie das für Sie passende Zitationsformat und kopieren Sie es dann in die Zwischenablage, lassen es sich per Mail zusenden oder speichern es als PDF-Datei.

oder
oder

Bitte prüfen Sie, ob die Zitation formal korrekt ist, bevor Sie sie in einer Arbeit verwenden. Benutzen Sie gegebenenfalls den "Exportieren"-Dialog, wenn Sie ein Literaturverwaltungsprogramm verwenden und die Zitat-Angaben selbst formatieren wollen.

xs 0 - 576
sm 576 - 768
md 768 - 992
lg 992 - 1200
xl 1200 - 1366
xxl 1366 -