Research Article
Print
Research Article
Characterization of the complete mitochondrial genome of the longhorn beetle, Batocera horsfieldi (Coleoptera, Cerambycidae) and its phylogenetic analysis with suitable longhorn beetles
expand article infoJunhao Wu, Danping Xu, Xinju Wei, Wenkai Liao, Xiushan Li, Zhihang Zhuo
‡ China West Normal University, Nanchong, China
Open Access

Abstract

Mitochondrial genome analysis is an important tool for studying insect phylogenetics. The longhorn beetle, Batocera horsfieldi, is a significant pest in timber, economic and protection forests. This study determined the mitochondrial genome of B. horsfieldi and compared it with the mitochondrial genomes of other Cerambycidae with the aim of exploring the phylogenetic status of the pest and the evolutionary relationships among some Cerambycidae subgroups. The complete mitochondrial genome of B. horsfieldi was sequenced by the Illumina HiSeq platform. The mitochondrial genome was aligned and compared with the existing mitochondrial genomes of Batocera lineolata and B. rubus in GenBank (MF521888, MW629558, OM161963, respectively). The secondary structure of transfer RNA (tRNA) was predicted using tRNAScan-SE server v.1.21 and MITOS WebSever. Thirteen protein-coding genes (PCGs) and two ribosomal RNA gene sequences of 21 longhorn beetles, including B. horsfieldi, plus two outgroups, Dryops ernesti (Dryopidae) and Heterocerus parallelus (Heteroceridae), were analyzed. The phylogenetic tree was constructed using maximum likelihood and Bayesian inference methods. In this study, we successfully obtained the complete mitochondrial genome of B. horsfieldi for the first time, which is 15 425 bp in length. It contains 37 genes and an A + T-rich region, arranged in the same order as the recognized ancestor of longhorn beetles. The genome of B. horsfieldi is composed of 33.12% A bases, 41.64% T bases, 12.08% C bases, and 13.16% G bases. The structure, nucleotide composition, and codon usage of the new mitochondrial genome are not significantly different from other longhorn mitochondrial genomes. Phylogenetic analyses revealed that Cerambycidae formed a highly supported single clade, and Vesperidae was either clustered with Cerambycidae or formed a separate clade. Interestingly, B. horsfieldi, B. rubus and B. lineolata were clustered with Monochamus and Anoplophora species in both analyses, with high node support. Additionally, the Vesperidae Spiniphilus spinicornis and Vesperus sanzi and the 19 Cerambycidae species formed a sister clade in the Bayesian analysis. Our results have produced new complete mitogenomic data, which will provide information for future phylogenetic and taxonomic research, and provide a foundation for future relevant research.

Key words

Evolutionary relationships, mitogenome, pest species, protein-coding genes, ribosomal RNA genes, secondary structure, tRNA

Introduction

The longhorn beetle family Cerambycidae is one of the larger families in Coleoptera, with about 40 000 recorded species worldwide. As of 2005, the recorded number of longhorn beetle species in China had surpassed 3100 (Zhang 2011). Many of the Cerambycidae are notorious pests, as they damage crops, forests, orchards and garden plants. Batocera horsfieldi (Hope, 1839) is a significant pest in timber, economic and protection forests. This species is mainly distributed in Asian countries such as China, Vietnam, India, Myanmar, etc. (Xiao 1992). It has a wide range of host plants, including Populus tomentosa Carr., Juglans regia L., Eucalyptus robusta Sm. (Zhuo et al. 2022). Adults of B. horsfieldi fed on the tender bark and leaves of supplemental host plants, such as Rosa cymosa Tratt., Rosa multiflora Thunb., Fraxinus chinensis Roxb. (Xiao 1992), while larvae bore into the trunk and feed on the woody tissue of the host plants. Larvae of B. horsfieldi burrow inside tree trunks, causing the host’s xylem to decay and become hollow. The entry holes are blocked by sawdust and excrements, making it difficult to control with pesticide. Due to their biological habits, larvae are prone to causing host tree weakness, fungal infections, wind breaks and even the death of the entire host tree (Zhuo et al. 2022).

The mitochondrial genome is widely used in population genetics, phylogenetics and molecular evolution studies due to its maternal inheritance, stable genome composition, relatively conservative gene arrangement, and rare occurrence of recombination (Wolstenholme 1992; Nelson et al. 2012; Driscoll et al. 2015; Feng et al. 2015). With the development of next-generation sequencing technology in the mitochondrial genome, it is possible to obtain more molecular markers for phylogenetic analysis, resulting in more robust relationships (Li et al. 2015; Fagua et al. 2018). Tracking the expansion pathways and dynamic hybridization of arthropods is often accomplished through mitochondrial sequence analysis (Rubinoff et al. 2010; Moore et al. 2013; Wang et al. 2017). The mitochondrial genome is widely used for phylogenetic inference in animals from high to low taxonomic levels (Li et al. 2016; Zheng et al. 2018), including tribe-level (Nie et al. 2018), subfamily-level (Nie et al. 2020) and family-level (Nie et al. 2021) relationships of beetles. To date, only seven species of Batocera have had their mitochondrial genomes sequenced. Therefore, this study reports the first complete mitochondrial genome sequence of B. horsfieldi using second-generation sequencing technology and compares it with the 20 existing mitochondrial genome sequences of other longhorn beetles in GenBank. In addition, we inferred the phylogenetic relationships of 21 longhorn beetle species to better study the phylogeny of Cerambycidae. The establishment of this genome resource will provide information for future phylogenetic and taxonomic research, providing a foundation for future relevant research.

Material and methods

Sample collection and DNA extraction

The specimens of B. horsfieldi, three male adults and one female adult, were collected from the host plant Populus tomentosa in Xishan Forest Park, Nanchong City, Sichuan Province, China, on June 17, 2022 (30.803038°N, 106.063072°E, alt. 480 m), and were deposited in the Laboratory of Forest Conservation, College of Life Science, China West Normal University (Voucher No. SCNC-BH-20220617.1-3 for male adult, SCNC-BH-20220617.4 for female adult). These specimens were preserved in 95% ethanol at -24 °C for long-term storage in the specimen collection room at China West Normal University. The total genomic DNA was extracted from muscle tissue of individual specimens using the Ezup Column Animal Genomic DNA Purification Kit (Shanghai, China), following the manufacturer’s instructions. For sequencing, the extracted DNA was stored at -24 °C.

DNA sequencing, mitogenome assembly and annotation

Next-generation sequencing and assembly were performed by Beijing Aoweisen Gene Technology Co. Ltd (Beijing, China) to obtain the complete mitogenome of B. horsfieldi for the first time. After quantification of the total genome DNA, a whole-genome shotgun strategy was employed for sequencing on the Illumina HiSeq platform. A total of 14 348 102 paired-end reads with a read length of 150 bp were obtained from the sequencing process. Among these, 22 966 sequences were used for mitochondrial genome assembly. Based on the invertebrate genetic code, such as Batocera lineolata Chevrolat, 1852 (MF521888), B. lineolata (MW629558) and B. rubus Linnaeus, 1758 (OM161963), the assembly of B. horsfieldi was carried out by referring to Hahn et al. (2013), and the gene annotation was performed using Geneious v.11.0.2 by referring to Greiner et al. (2019). The tRNA genes were re-verified by tRNAScan-SE (Lowe and Eddy 1997) and MITOS WebSever (Bernt et al. 2013). The base composition and relative synonymous codon usage (RSCU) values were calculated using Shengxin Cloud Online Server. Strand asymmetry was calculated in terms of formulae: AT-skew = (A - T)/(A + T) and GC-skew = (G - C)/(G + C) (Perna and Kocher 1995).

Phylogenetic analyses

A total of 21 complete mitochondrial genomes of longhorn beetles, including one newly sequenced species (B. horsfieldi), as well as the complete mitochondrial genomes of Dryops ernesti DesGozis, 1886 and Heterocerus parallelus Gebler, 1830, were used in this study (Table 1). These species belong to 13 longhorn genera, with Heterocerus parallelus (Heteroceridae) and Dryops ernesti (Dryopidae) as outgroups. The utilization of tRNA genes in phylogenetic analysis is relatively infrequent due to their short length and highly conserved characteristics (Chen 2005), thus we constructed the phylogenetic tree based on the 13 PCGs and rRNAs by using PhyloSuite v.1.4.4, and analyzed the tree by maximum likelihood (ML) and Bayesian inference (BI) methods with different best-fit substitution models. The sequences were aligned using MAFFT(Multiple Alignment using Fast Fourier Transform)and trimmed by trimAl. The best-fit substitution models used in the ML and BI analyses were calculated using ModelFinder v.2.3. Maximum likelihood analyses were run with 1000 ultrafast bootstrap and 1000 SH-aLRT replicates to estimate node reliability. Bayesian analyses were run with two independent chains spanning 1 million generations, four Markov chains, sampling at every 100 generations, and with a burn-in of 25%. The phylogenetic tree was visualized and edited using the iTOL online server (https://itol.embl.de/).

Table 1.

GenBank accession numbers of species used in this study (accession number of the newly sequenced species in bold).

Family Genus Species GenBank Accession number
Cerambycidae Monochamus Monochamus sparsutus NC053906.1
Monochamus alternatus NC050066.1
Monochamus sartor urussovii OP856519.1
Monochamus saltuarius OP169419.1
Agapanthia Agapanthia amurensis MW617354.1
Agapanthia daurica MN473114.1
Anoplophora Anoplophora glabripennis NC008221.1
Anoplophora horsfieldi MW364565.1
Anoplophora chinensis NC029230.1
Batocera Batocera lineolata MW629558.1
Batocera rubus OM161963.1
Batocera horsfieldi OQ785650
Xylotrechus Xylotrechus grayii NC030782.1
Turanoclytus Turanoclytus namaganensis NC060874.1
Demonax Demonax pseudonotabilis OP096419.1
Allotraeus Allotraeus orientalis NC061181.1
Arhopalus Arhopalus unicolor NC053904.1
Aromia Aromia bungii NC053714.1
Cephalallus Cephalallus oberthueri NC062854.1
Vesperidae Vesperus Vesperus sanzi MN473093.1
Spiniphilus Spiniphilus spinicornis NC029515.1
Out groups Dryops Dryops ernesti KX035147.1
Heterocerus Heterocerus parallelus KX087297.1

Results

Mitogenome organization and composition

The complete mitochondrial genome of B. horsfieldi is 15 425 bp in length and is a closed circular molecule (Fig. 1). It contains 37 genes, consisting of 22 transfer RNAs, two ribosomal RNAs, 13 protein-coding genes (PCGs), and an A+T-rich region, which is a feature shared with other typical Cerambycidae mitogenomes. Of the 13 PCGs, 9 (ND2, ND3, ND6, COI, COII, COIII, ATP8, ATP6, and CYTB) are encoded on the J strand, along with 14 tRNAs (Ile, Met, Trp, Leu2, Lys, Asp, Gly, Ala, Arg, Asn, Ser1, Glu, Thr, and Ser2), and a control region. The remaining four PCGs (ND5, ND4, ND4L, and ND1), eight tRNAs (Gln, Cys, Tyr, Phe, His, Pro, Leu1, and Val), and two rRNAs (12S rRNA and 16S rRNA) are encoded on the N strand (Fig. 1). The gene order of the new mitogenome follows the ancestral arrangement proposed for all arthropods (Braband et al. 2010). The overall mitogenome is composed of 33.12% A, 41.64% T, 12.08% C and 13.16% G, with a highly biased A+T content of 74.76% (Table 2). The calculation of asymmetrically distributed nucleotide content was determined by the formulas AT-skew = (A - T)/(A + T) and GC-skew = (G - C)/(G + C). The AT-skew and GC-skew of the B. horsfieldi mitogenome are -0.11 and 0.04, respectively (Table 2), indicating a negative AT-skew and a positive GC-skew. Overall, the nucleotide composition and genomic structure of B. horsfieldi exhibit typical features of the mitogenomes of insects in the order Coleoptera.

Table 2.

Base composition in different regions of the mitochondrial genome of B. horsfieldi.

Region Length (bp) A% T% C% G% A+T% AT-skew GC-skew
ND1 951 27.44 44.58 8.31 19.66 72.02 -0.24 0.41
ND2 1011 33.14 40.36 18.30 8.21 73.50 -0.10 -0.38
ND3 354 31.36 43.22 16.38 9.04 74.58 -0.16 -0.29
ND4 1333 29.78 48.16 6.98 15.08 77.84 -0.24 0.37
ND4L 288 29.17 49.13 7.64 13.89 78.48 -0.26 0.29
ND5 1717 31.16 46.88 7.75 14.21 78.04 -0.20 0.29
ND6 504 37.7 42.86 13.29 6.15 80.56 -0.06 -0.37
ATP6 675 32.74 41.33 16.74 9.19 74.07 -0.12 -0.29
ATP8 156 42.95 44.23 8.97 3.85 87.18 -0.01 -0.40
COI 1543 30.46 35.84 18.79 14.91 66.30 -0.08 -0.12
COII 688 32.27 38.52 18.17 11.05 70.79 -0.09 -0.24
COIII 787 30.50 38.25 17.03 14.23 68.75 -0.11 -0.09
CYTB 1143 32.37 38.58 17.76 11.29 70.95 -0.09 -0.22
Rrnl 1283 34.53 39.91 7.17 18.39 74.44 -0.07 0.44
Rrns 751 33.95 43.14 7.46 15.45 77.09 -0.12 0.35
rRANs 2034 34.32 41.10 7.28 17.31 75.42 -0.09 0.41
tRNAs 1450 38.55 37.79 10.07 13.59 76.34 0.01 0.15
13PCGs 11150 31.42 42.13 13.60 12.85 73.55 -0.15 -0.03
A+T-rich region 791 43.99 43.11 6.83 6.07 87.10 0.01 -0.06
Whole Genome 15425 33.12 41.64 12.08 13.16 74.76 -0.11 0.04
Figure 1. 

Mitogenome map of B. horsfieldi. ND1, ND2, ND3, ND4, ND4L, ND5 and ND6 are in yellow. COI, COII and COIII are in magenta. ATP6 and ATP8 are in green. CYTB is in pink. rrnl and rns are in red. all tRNAs are in dark blue and the control region is in white. Figure 1 was generated by Geneious v.11.0.2 by referring to Greiner et al. (2019).

PCGs and codon usage

The total length of the 13 PCGs in the mitochondrial genome of B. horsfieldi is 11 150 bp, accounting for 72.29% of the entire genome. Among the 13 PCGs, four genes (ND4, ND4L, ND5, and ND1) are encoded on the N strand, while the other nine genes (COI, COII, COIII, ATP8, ATP6, NAD2, NAD3, NAD6, and CYTB) are encoded on the J strand. Among the 13 PCGs, the longest gene is COX1, but its A+T content is the lowest at 66.30%, while the shortest gene is ATP8, but its A+T content is the highest at 87.18%. The start codons for COI, COII, ATP8, ND5, and ND6 are ATT, while the start codon for ND1 is TTG. The start codons for COIII, ATP6, ND3, ND4, ND4L, and CYTB are ATN (N represents G or C) (Table 3). Seven of the 13 PCGs share the typical termination codons TAA and TAG, while CYTB uses CTC as the termination codon, and the remaining four PCGs use a single T residue as the termination codon (Table 3). It is widely believed that the incomplete codon structure signifies the stop of protein translation in insects and other invertebrates (Cheng et al. 2016). Besides, the calculation of relative synonymous codon usage (RSCU) values and the analysis of base composition bias are important references for studying the replication and transcription mechanisms of mitochondrial genomes (Wei et al. 2010). We summarized the usage of relative synonymous codons. The results showed that AGA, UUA, UAA, and UCU are the four most commonly used codons (Fig. 2).

Table 3.

Characteristics of B. horsfieldi mitochondrial genome. J and N represent the positive and negative chains of mitochondrial genome, respectively.

Gene Strand Region Length (bp) Start codon Stop codon Anticodon
tRNA-Ile J 1–68 68 GAU
tRNA-Gln N 70–138 69 UUG
tRNA-Met J 138–206 69 CAU
ND2 J 207–1217 1011 ATC TAA
tRNA-Trp J 1216–1281 66 UCA
tRNA-Cys N 1274–1336 63 GCA
tRNA-Tyr N 1337–1402 66 GUA
COI J 1395–2937 1543 ATT T
tRNA-Leu2 J 2938–3002 65 UAA
COII J 3003–3690 688 ATT T
tRNA-Lys J 3691–3759 69 UUU
tRNA-Asp J 3760–3827 68 GUC
ATP8 J 3828–3983 156 ATT TAG
ATP6 J 3977–4651 675 ATG TAA
COIII J 4651–5437 787 ATG T
tRNA-Gly J 5438–5501 64 UCC
ND3 J 5502–5855 354 ATC TAG
tRNA-Ala J 5854–5917 64 UGC
tRNA-Arg J 5918–5979 62 UCG
tRNA-Asn J 5979–6045 67 GUU
tRNA-Ser1 J 6046–6112 67 UCU
tRNA-Glu J 6113–6176 64 UUC
tRNA-Phe N 6176–6239 64 GAA
ND5 N 6240–7956 1717 ATT T
tRNA-His N 7957–8019 63 GUG
ND4 N 8020–9352 1333 ATG T
ND4L N 9346–9633 288 ATG TAA
tRNA-Thr J 9636–9700 65 UGU
tRNA-Pro N 9701–9765 65 UGG
ND6 J 9768–10271 504 ATT TAA
CYTB J 10271–11413 1143 ATG CTC
tRNA-Ser2 J 11416–11483 68 UGA
ND1 N 11504–12454 951 TTG TAG
tRNA-Leu1 N 12456–12520 65 UAG
rRNl N 12521–13803 1283
tRNA-Val N 13804–13872 69 UAC
rRNs N 13873–14623 751
Control region J 14624–15414 791
Figure 2. 

Relative synonymous codon usage (RSCU) of the obtained mitogenome. RSCU value represented on the y-axis, and the B. horsfieldi codons of their respective amino acids represe on the x-axis.

tRNA genes, rRNA genes and A+T-rich region

The total length of the 22 tRNAs in B. horsfieldi is 1450 bp, ranging from 61 to 70 bp (Table 3). The nucleotide composition of the tRNAs is 38.55% A, 37.79% T, 10.07% C, and 13.59% G. The AT-skew and GC-skew are 0.01 and 0.15, respectively (Table 2). Eight of the 22 tRNA genes are located on the N strand, while the other 14 are located on the J strand. The tRNA secondary structure of B. horsfieldi was forecasted by the tRNAScan-SE server v.1.21.

The outcomes revealed that all tRNA genes conform to the standard cloverleaf structure, with the exception of tRNA-ser1, which is lacking a characteristic feature - the dihydrouridine arm (Fig. 3). The result is aligned with previous studies, such as those on stingray [Acroteriobatus annulatus, Acroteriobatus blochii (Van Staden et al. 2022)] and beetles [Coomaniella copipes, Coomaniella dentata and Dicerca corrugata (Huang et al. 2022)]. Moreover, the lengths of the 16S rRNA and 12S rRNA genes are 1283 bp and 751 bp, respectively (Table 2). The 16S rRNA is located between tRNA-Leu and tRNA-Val and is 1283 bp, constituting 34.53% A, 39.91% T, 7.17% C, and 18.39% G. The 12S rRNA is located between tRNA-Val and the A+T-rich region and is 751 bp in length. The nucleotide composition of the 12S rRNA is 33.95% A, 43.14% T, 7.46% C, and 15.45% G. The AT-skew and GC-skew of the RRNs are -0.09 and 0.41, respectively (Table 2). Compared with other longhorn beetles, the positions and characteristics of the 16S rRNA and 12S rRNA genes are similar. The A+T-rich region of B. horsfieldi is located between trnI and rns genes and is 791 bp in length, with an A+T content of 87.1%, an AT-skew of 0.01, and a GC-skew of -0.06 under normal conditions.

Figure 3. 

tRNAs secondary structure of B. horsfieldi.

Phylogenetic analysis

Thirteen protein-coding genes (PCGs) and two ribosomal RNA (rRNA) sequences were utilized to construct the phylogenetic trees of B. horsfieldi via maximum likelihood (ML) and Bayesian inference (BI) methods (Figs 4, 5). Based on the results obtained from BI, the monophyly of Cerambycidae was once again confirmed, as all the species belonging to this family formed a highly supported single clade. However, the results obtained from ML showed that Cerambycidae and Vesperidae formed a highly supported single clade. In both BI and ML, the two outgroups were clearly separated from the Cerambycidae and Vesperidae clades. Interestingly, BI showed that B. horsfieldi, B. rubus and B. lineolata, along with four species of Monochamus and three species of Anoplophora, were clustered into a branch with high node support values. On the other hand, ML showed that B. horsfieldi, B. rubus and B. lineolata were only clustered with Monochamus sparsutus Fairmaire, 1889 (M. sartor urussovii Fischer-Waldheim, 1806 +M. saltuarius Gebler, 1830) +M. alternatus alternatus Hope, 1842), and (Anoplophora chinensis Forster, 1771 +A. glabripennis Motschulsky, 1853) +A. horsfieldi Hope, 1842), all of which had high node support values. Additionally, ML showed that Vesperidae was only clustered with seven genera of Cerambycidae (Cephalallus, Arhopalus, Agapanthia, Batocera, Monochamus, and Anoplophora), all with high node support values, whereas BI showed that Vesperidae was clustered with 13 genera of Cerambycidae used in this study, all with high node support values. Furthermore, BI showed that Spiniphilus spinicornis Lin & Bi, 2011 (Vesperidae), Vesperus sanzi Reitter, 1895 (Vesperidae)and the 19 Cerambycidae species used in this study formed a sister clade.

Figure 4. 

Phylogenetic relationships of 21 selected longhorn beetles and outgroups (Heterocerus parallelus and Dryops ernesti) inferred based on maximum likelihood of 13 protein-coding genes and two rRNA genes. Values less than 70 are not displayed.

Figure 5. 

Phylogenetic relationships of 21 selected longhorn beetles and outgroups (Heterocerus parallelus and Dryops ernesti) inferred based on Bayesian analyses of 13 protein-coding genes and two rRNA genes. Numbers on branches are posterior probabilities.

Discussion

The complete mitochondrial genome sequence of B. horsfieldi obtained in this work enriches the species data of Batocera, Cerambycidae. Comparative analysis and phylogenetic analysis of the sequence features of Cerambycidae mitochondrial genomes were carried out by combining 13 protein-coding genes (PCGs) and two ribosomal RNA (rRNA) data from 21 species of longhorn beetles publicly available on NCBI. The results of this study will lay the foundation for population genetics research on B. horsfieldi and phylogenetic reconstruction of the Cerambycidae.

The secondary structures of tRNAs in Cerambycidae mitochondrial genomes have a typical cloverleaf structure, except for the absence of the D-arm in tRNA-ser1 (AGN) (Yang et al. 2017). The absence of the D-arm in trnS1 (AGN) is commonly found in other animal and insect orders, such as Lepidoptera, Diptera and Hemiptera (Wolstenholme 1992; Liao et al. 2010). Based on the start and stop codons of the protein-coding regions, most of the mitochondrial protein-coding genes in Cerambycidae begin with the typical ATN start codon, with only a few species of ND1 having a special start codon TTG, and the stop codons are generally the common TAN and incomplete T codons (Yang et al. 2017). Incomplete stop codons (T or TA) are common in animal mitochondrial genomes (Boore 1999) and Coates et al. (2005) suggest that they can be completed to TAA by the A at the 3’ end of the transcript. There has been ongoing controversy regarding the start codon for COI, and the use of different start or stop codons can result in variations in the number of base pairs between genes (Nie and Yang 2014). According to Sheffield et al. (2008), the start codon for COI in the suborder Polyphaga is either AAT or AAC; however, a reannotation was conducted in this study, which revealed that ATT is the start codon for COI in B. horsfieldi. Interestingly, in a recent study (Mei et al. 2022), the start codon for COI was reported to be ATG, which is similar to our findings. The A+T-rich region is the most important non-coding region in mitochondrial genomes, with extremely high A+T content. Length variation in this region can be very large, even among closely related species (Zhang et al. 1995; Zhang and Hewitt 1997). Batocera horsfieldi has an A+T-rich region of 791 bp, which results in a significantly longer mitochondrial genome compared to other genomes.

Currently, traditional taxonomic studies of longhorn beetles are mainly based on morphological features; however, there are still many controversies due to the complexity and instability of the morphological features. In this study, a phylogenetic tree was constructed using BI and ML methods based on 13 PCGs and 2 rRNAs, exhibiting differences with previous related studies. For instance, our results suggest that A. glabripennis and A. chinensis belong to the same branch, whereas in the study by Li et al. (2011), they were not grouped together. Furthermore, our findings indicate a closer relationship between M. sartor urussovii and M. alternatus alternatus, which contrast with the results of the study conducted by Yang et al. (2017). This work aimed to analyze the phylogenetic relationships of 21 longhorn beetle species and two outgroups (H. parallelus and D. ernesti) using mitochondrial genome data. Although the ML and BI trees constructed from 13 PCGs and two rRNA sequences did not exhibit identical topologies, their expression results were quite similar. Notably, the node support values of BI trees were consistently higher than those of ML trees for the same dataset, which has been observed in many previous studies of other taxa (Yuan et al. 2016; Chen et al. 2018; Li et al. 2019). Our results indicate that mitochondrial genome sequences are useful tools for inferring relationships among Cerambycidae. Furthermore, the technological development of sequencing and assembly of mitochondrial genomes will facilitate future work in mitochondrial genome sequencing (Crampton-Platt et al. 2015; Bourguignon et al. 2017). Increasing the sampling of taxa and genome sequencing will further help resolve the classification issues of longhorn beetles. Although this study has contributed one new complete mitochondrial genome to the phylogeny of Cerambycidae, the interrelationships among subfamilies and tribes still require more data to be completely determined. These questions will be well addressed in the future when sufficient numbers of complete mitochondrial genomes of longhorn beetles are accumulated. Although our study data cannot fully resolve the phylogenetic relationships within Cerambycidae, it provides a reference for further research on the phylogeny of Cerambycidae.

Conclusions

In this study, we successfully obtained the complete mitochondrial genome of Batocera horsfieldi for the first time, which is 15 425 bp in length. The mitochondrial genome of B. horsfieldi exhibits a molecular pattern similar to that of other Cerambycidae species, comprising 37 gene segments, including 22 transfer RNAs, two ribosomal RNAs, 13 protein-coding genes (PCGs), and an A + T-rich region typical of other Cerambycidae mitogenomes. Four of the 13 PCGs are encoded on the N strand, while the remaining nine genes are encoded on the J strand. The start codons of COI, COII, ATP8, ND5, and ND6 are ATT, while that of ND1 is TTG, and the start codons of COIII, ATP6, ND3, ND4, ND4L, and CYTB are ATN (N represents G or C). Seven of the 13 PCGs share typical stop codons (TAA and TAG), while CYTB uses CTC as a stop codon and the remaining four PCGs use a single T residue as a stop codon. All tRNAs exhibit typical cloverleaf structures except for tRNASer1. Phylogenetic analyses revealed that Cerambycidae formed a highly supported single clade, and Vesperidae was either clustered with Cerambycidae or formed a separate clade. Interestingly, B. horsfieldi, B. rubus and B. lineolata were clustered with Monochamus and Anoplophora species in both analyses, with high node support values. Additionally, Spiniphilus spinicornis and the 20 longhorn beetles formed a sister clade in the Bayesian analysis.

Acknowledgements

We thank associate professor Bo Luo for sample collection (College of Life Science, China West Normal University). We also wish to express our deep appreciation to the anonymous reviewers for their comments that helped to improve the quality of this manuscript.

Additional information

Conflict of interest

The authors have declared that no competing interests exist.

Ethical statement

No ethical statement was reported.

Funding

This work was funded by the Ministry of Science and Technology of the People’s Republic of China Support Program (2022YFE0115200), Sichuan Province Science and Technology Support Program (2022NSFSC0986) and China West Normal University Support Program (20A007, 20E051, 21E040, and 22kA011).

Author contributions

Conceptualization, Zhihang Zhuo; methodology, Junhao wu and Danping Xu; software, xinju wei; formal analysis, Junhao wu and Zhihang Zhuo; investigation, Wenkai Liao; data curation, Junhao wu; writing-original draft preparation, Junhao wu; writing-review and editing, Danping Xu and Zhihang Zhuo; supervision, Zhihang Zhuo.

Author ORCIDs

Zhihang Zhuo https://orcid.org/0000-0002-2566-3172

Data availability

The mitogenomes were deposited at Figshare. https://doi.org/10.6084/m9.figshare.22666192.v1 and https://doi.org/10.6084/m9.figshare.22673647.v1.

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 and Evolution 69(2): 313–319. https://doi.org/10.1016/j.ympev.2012.08.023
  • Bourguignon T, Dahlsjö CAL, Jacquemin J, Gang L, Wijedasa LS, Evans TA (2017) Ant and termite communities in isolated and continuous forest fragments in Singapore. Insectes Sociaux 64(4): 505–514. https://doi.org/10.1007/s00040-017-0570-y
  • Braband A, Cameron SL, Podsiadlowski L, Daniels SR, Mayer G (2010) The mitochondrial genome of the onychophoran Opisthopatus cinctipes (Peripatopsidae) reflects the ancestral mitochondrial gene arrangement of Panarthropoda and Ecdysozoa. Molecular Phylogenetics and Evolution 57(1): 285–292. https://doi.org/10.1016/j.ympev.2010.05.011
  • Chen RB (2005) Molecular Evolution and Phylogenetic Analysis of Some Species of Catantopidae (Orthoptera: Caelifera: Acridoidea) in China Based on Complete Mitochondrial ND2 gene and adjacent tRNAs gene sequence Rubing CHEN, China. Master Thesis, Shaanxi Normal University, Xian, China.
  • Chen ZT, Zhao MY, Xu C, Du YZ (2018) Molecular phylogeny of Systellognatha (Plecoptera: Arctoperlaria) inferred from mitochondrial genome sequences. International Journal of Biological Macromolecules 111: 542–547. https://doi.org/10.1016/j.ijbiomac.2018.01.065
  • Cheng XF, Zhang LP, Yu DN, Storey KB, Zhang JY (2016) The complete mitochondrial genomes offour cockroaches (Insecta: Blattodea) and phylogenetic analyses within cockroaches. Gene 586(1): 115–122. https://doi.org/10.1016/j.gene.2016.03.057
  • Coates BS, Sumerford DV, Hellmich RL, Lewis LC (2005) Partial mitochondrial genome sequences of Ostrinia nubilalis and Ostrinia furnicalis. International Journal of Biological Sciences 1(1): 13–18. https://doi.org/10.7150/ijbs.1.13
  • Crampton-Platt A, Timmermans MJTN, Gimmel ML, Kutty SN, Cockerill TD, Vun Khen C, Vogler AP (2015) Soup to tree: The phylogeny of beetles inferred by mitochondrial metagenomics of a Bornean rainforest sample. Molecular Biology and Evolution 32(9): 2302–2316. https://doi.org/10.1093/molbev/msv111
  • Driscoll CC, Driscoll JG, Hazekamp C, Mitton JB, Wehausen JD (2015) A tale of two markers: Population genetics of colorado rocky mountain bighorn sheep estimated from microsatellite and mitochondrial data. The Journal of Wildlife Management 79(5): 819–831. https://doi.org/10.1002/jwmg.895
  • Fagua G, Condamine FL, Brunet BMT, Clamens AL, Laroche J, Levesque RC, Cusson M, Sperling FAH (2018) Convergent herbivory on conifers by Choristoneura moths after boreal forest formation. Molecular Phylogenetics and Evolution 123: 35–43. https://doi.org/10.1016/j.ympev.2018.01.013
  • Greiner S, Lehwark P, Bock R (2019) OrganellarGenomeDRAW (OGDRAW) version 1.3.1: Expanded toolkit for the graphical visualization of organellar genomes. Nucleic Acids Research 47(W1): W59–W64. https://doi.org/10.1093/nar/gkz238
  • Hahn C, Bachmann L, Chevreux B (2013) Reconstructing mitochondrial genomes directly from genomic next-generation sequencing reads—A baiting and iterative mapping approach. Nucleic Acids Research 41(13): e129. https://doi.org/10.1093/nar/gkt371
  • Huang X, Chen B, Wei Z, Shi A (2022) First report of complete mitochondrial genome in the tribes Coomaniellini and Dicercini (Coleoptera: Buprestidae) and phylogenetic implications. Genes 13(6): 1074. https://doi.org/10.3390/genes13061074
  • Li J, An SH, Yin XM (2011) Phylogenetic Study of 12 Species of Longhorn Beetles Based on COI Gene Sequences. In: Yuan GH, Guo XR, Wang GP, Li WZ (Eds) Insect Research in Central China (Volume 7). 2011 Academic Conference of the Entomological Society of Central China (Hubei, Hunan, and Henan Provinces), Xinxiang (China), November 2011. China Agricultural Science and Technology Press, Beijing, 251.
  • Li H, Shao R, Song N, Song F, Jiang P, Li Z, Cai W (2015) Higher-level phylogeny of paraneopteran insects inferred from mitochondrial genome sequences. Scientific Reports 5(1): 1–10. https://doi.org/10.1038/srep08527
  • Li Q, Wei S-J, Tang P, Wu Q, Shi M, Sharkey MJ, Chen X-X (2016) Multiple lines of evidence from mitochondrial genomes resolve phylogenetic relationships of parasitic wasps in Braconidae. Genome Biology and Evolution 8(9): 2651–2662. https://doi.org/10.1093/gbe/evw184
  • Li R, Shu XH, Li XD, Meng L, Li BP (2019) Comparative mitogenome analysis of three species and monophyletic inference of Catantopinae (Orthoptera: Acridoidea). Genomics 111(6): 1728–1735. https://doi.org/10.1016/j.ygeno.2018.11.027
  • Liao F, Wang L, Wu S, Li YP, Zhao L, Huang GM, Niu CJ, Liu YQ, Li MG (2010) The complete mitochondrial genome of the Fall webworm, Hyphantria cunea (Lepidoptera: Arctiidae). International Journal of Biological Sciences 6(2): 172–186. https://doi.org/10.7150/ijbs.6.172
  • Lowe T, Eddy S (1997) tRNAscan-SE: A program for improved detection of transfer RNA genes in genomic sequence. Nucleic Acids Research 25(5): 955–964. https://doi.org/10.1093/nar/25.5.955
  • Mei ZX, Li JQ, Shang SA (2022) Sequence comparison of mitochondrial COI gene in different generations of Batocera lineolata attacking different hosts. Xibu Linye Kexue 51(01): 16–21.
  • Moore M, Sylla M, Goss L, Burugu MW, Sang R, Kamau LW, Kenya EU, Bosio C, Munoz Mde L, Sharakova M, Black WC, Kading RC (2013) Dual African origins of Globalaedes aegypti s.l. populations revealed by mitochondrial DNA. PLoS Neglected Tropical Diseases 7(4): e2175. https://doi.org/10.1371/journal.pntd.0002175
  • Nelson LA, Lambkin CL, Batterham P, Wallman JF, Dowton M, Whiting MF, Yeates DK, Cameron SL (2012) Beyond barcoding: A mitochondrial genomics approach to molecular phylogenetics and diagnostics of blowflies (Diptera: Calliphoridae). Gene 511(2): 131–142. https://doi.org/10.1016/j.gene.2012.09.103
  • Nie RE, Yang XK (2014) Research progress in mitochondrial genomes of Coleoptera. Acta Entomologica Sinica 57(7): 860–868.
  • Nie RE, Breeschoten T, Timmermans MJTN, Nadein K, Xue HJ, Bai M, Huang Y, Yang X-K, Vogler AP (2018) The phylogeny of Galerucinae (Coleoptera: Chrysomelidae) and the performance of mitochondrial genomes in phylogenetic inference compared to nuclear rRNA genes. Cladistics 34(2): 113–130. https://doi.org/10.1111/cla.12196
  • Nie RE, Andújar C, Gómez‐Rodríguez C, Bai M, Xue HJ, Tang M, Yang C-T, Tang P, Yang X-K, Vogler AP (2020) The phylogeny of leaf Beetles (Chrysomelidae) inferred from mitochondrial genomes. Systematic Entomology 45(1): 188–204. https://doi.org/10.1111/syen.12387
  • Nie RE, Vogler AP, Yang XK, Lin MY (2021) Higher‐level phylogeny of longhorn Beetles (Coleoptera:Chrysomeloidea) inferred from mitochondrial genomes. Systematic Entomology 46(1): 56–70. https://doi.org/10.1111/syen.12447
  • Perna NT, Kocher TD (1995) Patterns of nucleotide composition at fourfold degenerate sites of animal mitochondrial genomes. Journal of Molecular Evolution 41(3): 353–358. https://doi.org/10.1007/BF01215182
  • Rubinoff D, Holland BS, Shibata A, Messing RH, Wright MG (2010) Rapid invasion despite lackofgenetic variation in the erythrina gall wasp (Quadrastichus erythrinae Kim). Pacific Science 64(1): 23–31. https://doi.org/10.2984/64.1.023
  • Sheffield NC, Song H, Cameron SL, Whiting MF (2008) A comparative analysis of mitochondrial genomes in Coleoptera (Arthropoda: Insecta) and genome descriptions of six new beetles. Molecular Biology and Evolution 25(11): 2499–2509. https://doi.org/10.1093/molbev/msn198
  • Van Staden M, Ebert DA, da Silva C, Bester-van der Merwe AE (2022) Comparative analyses of the complete mitochondrial genomes of two southern African endemic guitarfish, Acroteriobatus annulatus and A. blochii. International Journal of Biological Macromolecules 223: 1094–1106. https://doi.org/10.1016/j.ijbiomac.2022.10.285
  • Wang XY, Yang XM, Lu B, Zhou LH, Wu KM (2017) Genetic variation and phylogeographicstructure ofthecotton aphid, aphis gossypii, based on mitochondrial DNA and microsatellitemarkers. Scientific Reports 7(1): 1920. https://doi.org/10.1038/s41598-017-02105-4
  • Xiao GR (1992) Forest Insects of China. China Forestry Publishing House, Beijing, 472–473.
  • Yang J (2017) The mitochondrial genome sequencing and phylogenetic analysisn of Cerambycidae species (Insecta:Coleoptera), China. Master Thesis, Beijing Forestry University, Beijing, China.
  • Yang J, Cao L, Geng Y (2017) Determination and comparative analysis of mitochondrial genomes of Anoplophora glabripennis and other Cerambycidae species. Chinese Journal of Applied Entomology 54(05): 755–766.
  • Yuan ML, Zhang QL, Zhang L, Guo ZL, Liu YJ, Shen YY, Shao R (2016) High-level phylogeny of the Coleoptera inferred with mitochondrial genome sequences. Molecular Phylogenetics and Evolution 104: 99–111. https://doi.org/10.1016/j.ympev.2016.08.002
  • Zhang J (2011) Taxonomic Study on Cerambycidae Insects in Jilin Province, China. PhD dissertation, Northeast Normal University, Changchun, China.
  • Zhang DX, Hewitt GM (1997) Insect mitochondrial control region: A review of its structure,evolution and usefulness in evolutionary studies. Biochemical Systematics and Ecology 25(2): 99–120. https://doi.org/10.1016/S0305-1978(96)00042-7
  • Zhang DX, Szymura JM, Hewitt GM (1995) Evolution and structural conservation of the controlregion of insect mitochondrial DNA. Journal of Molecular Evolution 40(4): 382–391. https://doi.org/10.1007/BF00164024
  • Zheng BY, Cao LJ, Tang P, Van Achterberg K, Hoffmann AA, Chen HY, Chen XX, Wei SJ (2018) Gene arrangement and sequence of mitochondrial genomes yield insights into the phylogeny and evolution of bees and sphecid wasps (Hymenoptera: Apoidea). Molecular Phylogenetics and Evolution 124: 1–9. https://doi.org/10.1016/j.ympev.2018.02.028
  • Zhuo ZH, Jin YW, Xu DP, Liao WK (2022) Electroantennogram responses of Batocera horsfieldi (Hope) to the selected volatile components of host plants, Rosa cymosa Tratt. and Rosa multiflora Thunb. Global Ecology and Conservation 33: e01986. https://doi.org/10.1016/j.gecco.2021.e01986

Supplementary material

Supplementary material 1 

Batocera horsfieldi and 22 species

Junhao Wu,Zhihang Zhuo

Data type: Phylogenetic (docx. file)

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 (1.50 MB)
login to comment