Next Article in Journal
Adults of Alderflies, Fishflies, and Dobsonflies (Megaloptera) Expel Meconial Fluid When Disturbed
Previous Article in Journal
Presence of Spodoptera frugiperda Multiple Nucleopolyhedrovirus (SfMNPV) Occlusion Bodies in Maize Field Soils of Mesoamerica
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

The Complete Mitogenomes of Three Grasshopper Species with Special Notes on the Phylogenetic Positions of Some Related Genera

1
Key Laboratory of Cultivation and Protection for Non-Wood Forest Trees (Central South University of Forestry and Technology), Ministry of Education, Changsha 410004, China
2
Guangxi Key Laboratory of Rare and Endangered Animal Ecology, Guangxi Normal University, Guilin 541004, China
3
Key Laboratory of Forest Bio-Resources and Integrated Pest Management for Higher Education in Hunan Province, Central South University of Forestry and Technology, Changsha 410004, China
4
College of Agriculture and Biological Science, Dali University, Dali 671003, China
5
Shanghai Entomological Museum, Chinese Academy of Sciences, Shanghai 200032, China
6
College of Life Sciences, Shaanxi Normal University, Xi’an 710119, China
*
Authors to whom correspondence should be addressed.
These authors contributed equally to this work.
Insects 2023, 14(1), 85; https://doi.org/10.3390/insects14010085
Submission received: 26 November 2022 / Revised: 5 January 2023 / Accepted: 10 January 2023 / Published: 13 January 2023
(This article belongs to the Section Insect Systematics, Phylogeny and Evolution)

Abstract

:

Simple Summary

The complete mitogenomes of three grasshopper species were sequenced and annotated. The phylogenetic positions of the genera Emeiacris and Choroedocus are clarified based on both complete mitogenome and morphological evidences. The results show that Emeiacris consistently has the closest relationship with the genus Paratonkinacris of the subfamily Melanoplinae, and Choroedocus has the closest relationship with the genus Shirakiacris of the subfamily Eyprepocnemidinae, respectively. In addition, the genera Conophymacris and Xiangelilacris, as well as Ranacris and Menglacris, are two pairs of the closest relatives, but their phylogenetic positions need further study to clarify.

Abstract

Clarifying phylogenetic position and reconstructing robust phylogeny of groups using various evidences are an eternal theme for taxonomy and systematics. In this study, the complete mitogenomes of Longzhouacris mirabilis, Ranacris albicornis, and Conophyma zhaosuensis were sequenced using next-generation sequencing (NGS), and the characteristics of the mitogenomes are presented briefly. The mitogenomes of the three species are all circular molecules with total lengths of 16,164 bp, 15,720 bp, and 16,190 bp, respectively. The gene structures and orders, as well as the characteristics of the mitogenomes, are similar to those of other published mitogenomes in Caelifera. The phylogeny of the main subfamilies of Acrididae with prosternal process was reconstructed using a selected dataset of mitogenome sequences under maximum likelihood (ML) and Bayesian inference (BI) frameworks. The results showed that the genus Emeiacris consistently fell into the subfamily Melanoplinae rather than Oxyinae, and the genus Choroedocus had the closest relationship with Shirackiacris of the subfamily Eyprepocnemidinae in both phylogenetic trees deduced from mitogenome protein coding genes (PCGs). This finding is entirely consistent with the morphological characters, which indicate that Emeiacris belongs to Melanoplinae and Choroedocus belongs to Eyprepocnemidinae. In addition, the genera Conophymacris and Xiangelilacris, as well as Ranacris and Menglacris, are two pairs of the closest relatives, but their phylogenetic positions need further study to clarify.

1. Introduction

The grasshopper family Acrididae is not only the largest group within Orthoptera, but also contains many species with tremendous economic importance [1]. Reconstructing a robust phylogeny will promote our understanding of the interesting biology and evolutionary patterns within this family, such as character and behaviour evolution [2,3,4,5]. In recent years, an increasing number of studies have been carried out based on molecular evidence to resolve phylogenetic problems at different taxonomic scales of Acrididae [6,7,8,9,10], and the complete mitogenome data has demonstrated great potential in this field [3,4,5,11,12,13,14,15,16,17,18,19,20,21,22,23,24,25,26,27,28,29,30]. Therefore, studies of the mitogenome for various purposes have always been a hot topic [4]. Since paraphyly is rampant across many subfamilies of Acrididae [17], as well as other groups of Orthoptera, such as Pyrgomorphidae, and so on [28], there is still a long way for us to go to resolve such problems.
The genus Longzhouacris was first erected based on the type species Longzhouacris rufipennis You and Bi, 1983 [31],and currently includes 11 species distributed across tropical to subtropical areas [32]. Originally, Longzhouacris was placed under the family Catantopidae and not assigned to a definite subfamily. Later, it was considered to be a member of the subfamily Habrocneminae [32,33,34]. The subfamily Habrocneminae contains four genera in Li and Xia’s monograph [33]: Habrocnemis, Longzhouacris, Menglacris, and Promesosternus, but the genus Promesosternus is assigned now in the subfamily Oedipodinae in the Orthoptera Species File (OSF) [32].
The genus Ranacris was established with Ranacris albicornis You and Lin, 1983, as the type species [35] and consists of three known species so far [32,34,36]. Like the genus Longzhouacris, Ranacris was also placed originally under the family Catantopidae without a definite subfamily assignment [35]. However, later, the subfamily Ranacridinae was proposed to contain the single genus Ranacris [37]. This was followed by some later scholars [33,34,38]. Storozhenko [39] then synonymized Ranacridinae with Mesambriini, and the genus was transferred to the tribe Mesambriini of Catantopinae. Although the genera Ranacris and Menglacris are placed in two different subfamilies at present, they exhibit high similarity superficially except for the difference in the presence or absence of the tegmen, i.e., the genus Ranacris is apterous, but the genus Menglacris is micropterous with narrow scaly tegmina. In addition, the genus Habrocnemis is also very similar to Ranacris and Menglacris.
The genus Conophyma is the largest one of the subfamily Conophyminae, with 102 known species distributed mainly in the mountainous and plateau areas from Central Asia to the Himalayas [32]. Most Conophyma species inhabit above the altitude of 2000 m. Conophyminae had been placed under the family Acrididae before Otte [40], but Eades [41] transferred it to the family Dericorythidae based on the comparison of the male genitalia structure. Dericorythidae lacks the arch sclerite characteristic of Acrididae. Instead, it has the distinctive pseudoarch that is easily mistaken for the true arch if dissection is not completed by opening up the spermatophore sac. However, Eades’ [41] study sampled only one species of the subfamily Conophyminae, Plotnikovia lanigera, and did not examine any materials of the genus Conophyma, the largest group of Conophyminae.
Although the three genera mentioned above have a debatable phylogenetic position and a complicated relationship with other genera, there is no complete mitogenome data available for phylogeny inference. Similar problems are also explicit in the genera Emeiacris, Choroedocus, Conophymacris, and Xiangelilacris. Emeiacris is recognized as a member of the subfamily Oxyinae in OSF [32], but was placed in the subfamily Melanoplinae by Li and Xia [33], and Mao et al. [34], and this was always supported by molecular evidences [24,25,27]. Choroedocus is currently placed in the subfamily Catantopinae in OSF [32], but was obviously regarded as a member of the subfamily Eyprepocnemidinae in nearly all published literatures. Conophymacris is presently in the subfamily Conophyminae of the family Dericorythidae in OSF [32], but belongs to the subfamily in some monographs [33,34] and always has a closest relationship with Xiangelilacris in previous molecular studies [24,25,27]. In this study, the complete mitogenome of Longzhouacris mirabilis, Ranacris albicornis, and Conophyma zhaosuensis were sequenced and annotated. Additionally, the phylogeny of the grasshoppers in Acrididae with prosternal process was reconstructed using a selected dataset of mitogenome sequences of 90 species, including the three newly sequenced mitogenomes and the ones of 87 related species downloaded from GenBank (https://www.ncbi.nlm.nih.gov/ (accessed on 12 March 2022)). The phylogenetic position of some related genera are discussed in combination with morphological characters.

2. Materials and Methods

2.1. Taxon Sampling

Three species, i.e., Longzhouacris mirabilis, Ranacris albicornis, and Conophyma zhaosuensis, were selected as representatives of the genera Longzhouacris, Ranacris, and Conophyma, respectively. Data of the materials for generating mitogenomes were: (1) Longzhouacris mirabilis (voucher number: mt1825), collected at Jiangjunzhai, Mangshan Nature Reserve, Yizhang County, Hunan Province, China; 112°55′56′′ E, 24°57′24′′ N; 22 August 2018; Jingxiao Gu leg.; (2) Ranacris albicornis (voucher number: mt1826), collected at Diding Nature Reserve, Jingxi County, Guangxi, China; 105°59′47′′ E, 23°5′44′′ N; 8 August 2010; Jianhua Huang leg. (3) Conophyma zhaosuensis (voucher number: mt1938), collected on the hill behind the stud farm, Wuzunbulake Township, Zhaosu County, Xinjiang Province, China; 81°11′56′′ E, 43°11′25′′ N, altitude 2141 m; on 15 August 2013; Jianhua Huang leg. The specimens were identified by the last author according to the keys to species in Li and Xia’s [33] and Maoet al.’s [34] monographs. They were preserved in 100% anhydrous ethanol and stored in a refrigerator(Thermo Fisher Scientific, Waltham, MA, USA) at −80 °C in the Insect Collection of the Central South University of Forestry and Technology.
All materials were collected under appropriate collection permits and approved ethics guidelines. The morphological terminology followed that of Uvarov [42] and Storozhenko et al. [43]. The terminology of male genitalia followed that of Woller and Song [44]. All photographs were taken using a Nikon D600 digital camera(Nikon Corp., Minato-ku, Tokyo, Japan) or Leica DFC 5500 system(Leica Microsystems Inc., Wetzlar, Germany), the stacking images were combined using Helicon Focus ver. 6.0, (Helicon Soft, Kharkiv, Ukraine) and the plates were edited in Photoshop CS (Adobe Inc., San Jose, CA, USA).
To clarify the phylogenetic positions of as many taxa as possible, we included in this analysis nearly all presently available mitogenome data of species with a prosternal process in Acrididae and Dericorythidae (Table 1), representing 2 families, 12 subfamilies, 50 genera and 88 species/subspecies in total. Coryphistes ruricola (MG993389, MG993390, MG993403, MG993406) in Catantopinae and Kosciuscola tristis (MG993402, MG993408, MG993414) in Oxyinae were not included in this analysis because they have only partial mitogenome sequences available. Gesonula punctifrons (MN046214) with complete mitogenome was also excluded from this analysis due to the inaccuracy of the sequence, possibly derived from the inadequate sequencing data (only 2 Gb data was generated through next-generation sequencing (NGS)) [24].

2.2. Sequencing, Assembly and Annotation

A hind femur for each sample was sent to Berry Genomics (Beijing, China) for genomic sequencing using NGS, and the remainder of the specimen was deposited as a voucher specimen at the Central South University of Forestry and Technology. Whole genomic DNA was extracted from muscle tissue of the hind femur using a modified routine phenol and chloroform method. Separate 400 bp insert libraries were created from the whole genome DNA and sequenced using the Illumina HiSeq X Ten sequencing platform. A total of 20 Gb of 150 bp paired-end (PE) reads were generated in total for each sample. Raw reads were filtered to remove reads containing adaptor contamination (>15 bp matching to the adaptor sequence), poly-Ns (>5 bp Ns), or >1% error rate (>10 bp bases with quality score < 20). The mitogenome sequence was assembled from clean reads in Mitobim (ver. 1.9.1, see https://github.com/chrishah/MITObim (accessed on 8 May 2022)) [67]. Two runs were implemented independently using the same reference with different starting points (one point is trnI and another is COXI) to improve the sequence quality of the control region. The assembled raw mitogenome sequences were primarily annotated online using the MITOS Web Server (http://mitos.bioinf.uni-leipzig.de/index.py; accessed on 20 May 2022) [68] and then checked and corrected in Geneious (ver. 8.04, see https://www.geneious.com(accessed on 5 June 2022)) [69]. The secondary structure of the RNA encoding genes predicted in MITOS were visualised and checked manually using VARNA (ver.3.93, see http://varna.lri.fr(accessed on 13 June 2022)) [70]. The three newly sequenced mitogenomes have been deposited in GenBank under accession number ON943039 for Ranacris albicornis, ON931612 for Longzhouacris mirabilis, and ON943040 for Conophyma zhaosuensis, respectively (Table 1).
Base composition, A−T- and G − C-skews, and codon usage were calculated in MEGA X [71]. The formulas used to calculate the skews of the composition were (A − T)/(A + T) for the A−T-skew and (G − C)/(G + C) for the G−C-skew.

2.3. Phylogenetic Analyses

To explore the phylogenetic position of the genera Longzhouacris, Ranacris, Conophyma and some related taxa, 96 complete mitogenome sequences in total, representing 85 species in Acrididae and 3 species in Dericorythidae, were selected as ingroups, and 2 species in Pamphagidae served as outgroups. The complete mitogenome dataset consists of the 13 protein coding genes (PCGs). The two rRNA genes were not used for the phylogeny inference due to their limited resolution above the genus level [27,29]. In order to involve more species of the genus Conophyma in our analysis, partial mitochondrial COX1 sequences of 6 Conophyma species were downloaded from NCBI (Supplementary Materials Table S1) and the corresponding fragment was extracted from the complete mitogenome to generate a new dataset of partial COX1 fragment.
The PCGs were aligned using macse_v2.03 (see https://bioweb.supagro.inra.fr/macse/index.php?menu=releases (accessed on 8 July 2022)) [72]. The alignments were manually optimized and concatenated into a single dataset in Phylosuite (see http://phylosuite.jushengwu.com/ (accessed on 9 July 2022)) [73].
The PCGs dataset was divided into 39 data blocks (13 PCGs divided into individual codon positions). Best-fit models of nucleotide evolution and best-fit partitioning schemes were selected using ModelFinder (see http://www.iqtree.org/ModelFinder/ (accessed on 9 July 2022)) [74]. The best-fitting models used for the phylogenetic analyses of the mitochondrial PCGs and partial COX1 datasets are shown in Supplementary Materials Table S2 and Supplementary Materials Table S3, respectively.
The phylogenies were reconstructed in maximum likelihood (ML) and Bayesian inference (BI) frameworks. The ML phylogenies were reconstructed using IQ-TREE (ver. 1.6.12, see http://www.iqtree.org (accessed on 13 July 2022)) [75]. The approximately unbiased branch supportvalues were calculated using UFBoot2 [76]. The analysis was performed in W-IQ-TREE (see http://iqtree.cibiv.univie.ac.at (accessed on 13 July 2022)) [77] using the default settings. Nodes with a bootstrap percentage of at least 70% were considered well supported in the ML analyses [78]. BI analyses were accomplished in MrBayes (ver. 3.2.1, see http://morphbank.Ebc.uu.SE/mrbayes/ (accessed on 15 July 2022)) [79], with two independent runs, each with four Markov Chain Monte Carlo (MCMC) chains. The analysis was run for 1 × 107 generations, sampling every 100 generations, and the first 25% of generations were discarded as burn-in, whereas the remaining samples were used to summarize the Bayesian posterior probabilities. All of the above analyses were implemented in Phylosuite (see http://phylosuite.jushengwu.com/ (accessed on 15 July 2022)) [73]. For the phylogenetic trees reconstructed from partial COXI dataset, the cutoffs of 0.95 posterior and 90 bootstrap were used to collapse the nodes below these cutoffs to a polytomy.
To overcome, at least partially, some of the issues in mtDNA, such as the generally high saturation and the among-lineages and/or among-sites compositional bias, the mitochondrial PCGs were translated into amino acids, and then the amino acids were used to run an ML and a Bayesian analysis via MtOrt [24], the taxa-specific amino acid substitution model for Orthoptera, and MtRev [74], the best-fit model chosen according to Bayesian Information Criterion (BIC), respectively.

3. Results

3.1. Characteristics of the Newly Sequenced Mitogenomes

The mitogenomes of Longzhouacris mirabilis, Ranacris albicornis, and Conophyma zhaosuensis are all circular molecules with total lengths of 16,164 bp, 15,720 bp, and 16,190 bp, respectively (Figure 1). They have the typical metazoan mitochondrial gene set consisting of 13 PCGs, 22 tRNAs, 2 rRNAs (the large and small ribosomal subunits), and a putative A + T-rich region (control region, CR). Among the 13 PCGs, 9 (ATP6, ATP8, COX1, COX II, CYTB, ND2, ND3, and ND6) are located in the J strand, and the remaining 4 (ND1, ND4, ND4L, and ND5) are located in the N strand. Among the 37 genes coded by the mitogenome, 23 genes are coded at the J strand and 14 at the N strand. The gene order of the newly sequenced mitogenomes is the same as that of other published mitogenomes in Caelifera (Figure 1). The base composition is obviously A–T-biased, with the total A + T contents ranging from 74.6% (Conophyma zhaosuensis), to 75.1% (Ranacris albicornis), to 75.2% (Longzhouacris mirabilis) (Supplementary Materials Tables S4–S6). The A–T-skews are 0.1185 (Ranacris albicornis), 0.125 (Longzhouacris mirabilis), and 0.1394 (Conophyma zhaosuensis), and the G–C-skews are −0.1498 (Ranacris albicornis), −0.1406 (Longzhouacris mirabilis), and −0.1216 (Conophyma zhaosuensis).
Most PCGs have a typical initiation codon of ATN (Table 2). However, COX1 in Longzhouacris mirabilis and Ranacris albicornis initiates from a non-standard initiation codon of ACC, COX1 in Conophyma zhaosuensis initiates from CAA, and ATP6 in Longzhouacris mirabilis initiates from GTG. Seven PCGs (ND2, COX2, COX3, ND4, ND4L, ND6, and CYTB) initiated from ATG. The initiation codon ATT has the second highest frequency of usage, followed by ACC and ATA. With respect to termination codons, the majority of PCGs have a typical termination codon of TAA in most species (Table 2). The complete termination codon TAG occurs in ND1 in all of the three species. The incomplete termination codon TA occurs only in CYTB of Longzhouacris mirabilis and ND6 of Conophyma zhaosuensis. COX1 in all of the three species, ND4 in Longzhouacris mirabilis, and ND4 and ND5 in Conophyma zhaosuensis are terminated by T.
The PCGs of the mitogenome have extremely similar codon usage pattern to other grasshoppers (Supplementary Materials Tables S7–S9). Among all codons of the PCGs, the most preferred codon with the highest average relative synonymous codon usage (RSCU) is UUA, which codes for Leucine and has an RSCU value of 3.98%. The next common codons are UCA (Serine) and CGA (Arginine), followed by UCU (Serine) and ACA (Threonine), with average RSCU values of 2.463%, 2.467%, 2.09%, and 1.993%, respectively, indicating a distinct codon usage bias in grasshoppers [29].
The sizes of the 22 tRNAs varies over a very small range in all the three newly sequenced mitogenomes (Supplementary Materials Table S10). Except for tRNASer-AGN lacking the DHU arm, all of the other 21 tRNAs can be folded into a typical clover structure (Supplementary Materials Figure S1). The numbers of base mismatches in the tRNAs varies drastically among the different mismatch types in all species, but all species have similar distribution patterns of base mismatches (Table 3). The mismatch of G–U represents the majority of the total mismatches. A–A occurs only once in trnD. U–U occurs in trnQ and some other tRNAs. A–G occurs only in trnW. For the most frequent mismatch of G–U, it does not occur in trnM, trnD, trnN,and trnW for all three species, andthe maximum mismatch number in one tRNA is five (Table 4).
The lrRNA and srRNA are located between the trnL1 and trnV, and trnV and A + T-rich regions, respectively. Their lengths vary between 1365–1389 bp (lrRNA) and 792–806 bp (srRNA). The control region is located between rrnS and trnI, and contains the highest proportion of A + T content ranging from 78.6 to 80.9%. The lengths of the control region vary between 860 and 1437 bp (Supplementary Materials Tables S4–S6).
In addition to the control region, there are also some gene intervals or base overlaps between some genes, and the maximum overlap area is between trnL1 and rrnL (Supplementary Materials Table S10). There are nine, seven, and seven tightly aligned gene pairs without overlap or interval in the mitogenome of the three species, respectively.

3.2. Phylogeny

The phylogenetic trees inferred from the dataset of the 13 mitochondrial PCGs using maximum likelihood and Bayesian inference methods have an extremely consistent topology above the genus level (Figure 2, Supplementary Materials Figure S2). At the family level, the monophylies of both Acrididae and Dericorythidae are not supported. The three species of Dericorythidae form three individual clades. Conophymacris viridis completely falls into Acrididae, having the closest relationship with Xiangelilacris zhongdianensis. Conophyma zhaosuensis and Dericorys annulata are located near the base of the trees, but do not form a single clade. Conophyma zhaosuensis forms a small clade with Leptacris sp. and Dericorys annulata forms an individual clade itself, located at the most outside of the ingroup. At the subfamily level, the monophylies of six subfamilies (Spathosterninae, Oxyinae, Cyrtacanthacridinae, Eyprepocnemidinae, Calliptaminae, and Coptacrinae) are retrieved usually with strong nodal support. However, the remaining subfamilies are not recovered as monophyletic. For the subfamily Melanoplinae, nearly all species cluster into an independent clade except for Xiangelilacris zhongdianensis, which forms a small clade with Conophymacris viridis of Dericorythidae, and has close relationships with Coptacrinae and Longzhouacris mirabilis of Habrocneminae. The two species of Habrocneminae sampled in this study, Longzhouacris mirabilis and Menglacris maculata, do not form a single clade, but fall into two distantly separated clades, one of which is Menglacris maculata + Ranacris albicornis, and the other is Longzhouacris mirabilis + Conophymacris viridis + Xiangelilacris zhongdianensis + Coptacrinae, with a bootstrap value of 100% or posterior probability of 0.96. The members of Hemiacridinae are divided into two distantly separated clades, with Hieroglyphus species having a close relationship with Spathosterninae, but Leptacris sp. forming a small clade with Conophyma zhaosuensis. The members of Catantopinae are divided into three clades: Ranacris albicornis, the genus Traulia and the typical Catantopini species. Ranacris albicornis is consistently most related to Menglacris maculata, with a bootstrap value of 100% or a posterior probability of 1. The genus Traulia has the closest relationship with the clade including Coptacrinae species. The clade of the typical Catantopini species forms the sister group of Cyrtacanthacridinae.
Although the phylogeny deduced from the mitochondrial PCGs is robust, the phylogenetic trees reconstructed from the dataset of partial COX1 fragment sequences exhibit great difference from the former (Supplementary Materials Figures S3 and S4). The monophylies of the subfamilies Calliptaminae, Coptacrinae, Eyprepocnemidinae, and Oxyinae are no longer supported in both or at least one tree from the COX1 dataset. The relationships among the key groups are also very different between the ML and BI trees, and the relationships among most clades are unsolved, forming a large polytomy at the base of the trees. Even the two Hieroglyphus species are also split into two distantly separated clades. This result indicates the extreme instability of the phylogeny reconstructed using COX1 sequence.
Despite the great difference in the topology between the trees from mitochondrial PCGs and COX1 datasets, and the instability of the COX1 trees, the small clades of Ranacris albicornis + Menglacris maculata and Conophymacris viridis + Xiangelilacris zhongdianensis are always robust in all trees (Figure 2, Supplementary Materials Figures S2–S4). The clade of the genus Conophyma is also robust in the COX1 trees, but the relationship of this clade with other groups varies (Supplementary Materials Figures S3 and S4). The position of Longzhouacris mirabilis also varies in the COX1 trees (Supplementary Materials Figures S3 and S4). It falls into the clade of the subfamily Eyprepocnemidinae in the ML tree of the COX1 dataset (Figure S3), but forms a polytomy clade with some species of the subfamilies Hemiacridinae and Oxyinae, as well as other clades, in the BI tree of the COX1 dataset (Supplementary Materials Figure S4). In addition, it is noticeable that the genus Emeiacris always falls into the clade of the subfamily Melanoplinae and has an extremely stable close relationship with Paratonkinacris in all trees (Figure 2, Supplementary Materials Figures S2–S4). The similar case also occurs in the genus Choroedocus, which always forms a stable clade with Shirakiacris species of the subfamily Eyprepocnemidinae and has a closer relationship with the subfamily Calliptaminae than Catantopinae in the trees from the dataset of mitochondrial PCGs (Figure 2, Supplementary Materials Figure S2).
With a further look at the trees reconstructed from the amino acid dataset, we find that the ML tree reconstructed using the MtOrt model (Figure 3) has an extremely high similarity in the main topology with the trees deduced from the mitochondrial PCGs (Figure 2, Supplementary Materials Figure S2), including the non-monophyly of Dericorythidae, the monophylies of the six subfamilies (Spathosterninae, Oxyinae, Cyrtacanthacridinae, Eyprepocnemidinae, Calliptaminae, and Coptacrinae), the positons of the genera Emeiacris and Choroedocus, the relationship of Menglacris with Ranacris, and that of Conophymacris with Xiangelilacris, and so on. The most important difference is that Dericorys annulata and Conophyma zhaosuensis forms an independent clade only in this tree (Figure 3). The BI tree deduced from the amino acid dataset using the MtRev model (Supplementary Materials Figure S5) is similar to the ML tree, but the monophyly of the subfamily Calliptaminae is no longer supported, with Peripolus nepalensis escaping from the clade of the genus Calliptamus.

4. Discussion

4.1. Phylogenetic Position of the Genus Emeiacris

The genus Emeiacris was established with Emeiacris maculata Zheng, 1981 as the type species [80] and three known species so far [32]. According to the original description, Emeiacris is most similar to the genus Oxyacris of the subfamily Oxyinae, and is mainly characterized by the rounded apex of the lower knee-lobes of the hind femora (Figure 4e), the widely separated metasternal lobes (Figure 4f), and the distinct process at the lateral margin of the supra-anal plate. When Emeiacris was erected, it was not definitely assigned to any subfamily. Therefore, it is possible that the authors of OSF placed Emeiacris in the subfamily Oxyinae according to the original reference, where the closest relative of Emeiacris was Oxyacris of the subfamily Oxyinae [32]. Subsequently, Emeiacris was definitely placed in the subfamily Podisminae [81], and then in Melanoplinae [33]. Morphologically, the species of Emeiacris are extremely similar to the species of the genera Ognevia and Fruhstorferiola (Figure 4a–n). The rounded apex of the lower knee-lobes of the hind femora and the widely separated metasternal lobes are typical distinguishing characters of Melanoplinae (Figure 4e,f), but not those of Oxyinae (Figure 4o–q). In addition, the absence of the ectoapical spine in the hind tibia (Figure 4g,h), and the epiphallus not divided into two separated symmetric parts (Figure 4i–k), also disagree with the diagnostic characters of Oxyinae, but match those of Melanoplinae. In the molecular study, Emeiacris consistently falls into the clade of Melanoplinae and has the robust closest relationship with Paratonkinacris in all trees (Figure 2 and Figure 3, Supplementary Materials Figures S2–S5). Therefore, the genus Emeiacris should be considered as a member of the subfamily Melanoplinae rather than Oxyinae.

4.2. Phylogenetic Position of the Genus Choroedocus

The genus Choroedocus was proposed by Bolívar [82] to replace the preoccupied genus name “Demodocus Stål, 1878” (nec Demodocus Guérin, 1843 in Coleoptera). Demodocus was proposed first as a subgenus of the genus Calliptenus, which was considered being most similar to the genus Eyprepocnemis [83], and then raised to the generic level by Brunner von Wattenwyl [84]. Kirby [85] proposed to restrict Walker’s name Heteracris to the genus because it was preoccupied in Coleoptera. Bolívar [82] proposed a new name, Choroedocus, for Demodocus. No matter Demodocus, Heteracris, or Choroedocus, they were always definitely assigned to the group Euprepocnemes [82,86], or the tribe Eyprepocnemini [87,88], or the subfamily Eyprepocnemidinae [33,34,38,85]. There are indeed a few works where Choroedocus is placed in the subfamily Catantopinae [81,89,90], but Catantopinae in this sense contains actually all Eyprepocnemidinae taxa. In other words, all Eyprepocnemidinae taxa are members of the subfamily Catantopinae, and there is no category of the subfamily Eyprepocnemidinae in that classification scheme.
We do not know why the authors of OSF finally placed the genus Choroedocus in the subfamily Catantopinae. The most probable reason may be that the genus Choroedocus was once placed by Liu [91] in the family Catantopidae without assignment of the subfamily position. However, this is not strong evidence for the decision because the truth is that Liu [91] merely listed no subfamily category in his work. After examining materials of the genus Choroedocus, we found they highly agree with the distinguishing characters of the subfamily Eyprepocnemidinae: the pronotum with distinct lateral carina and a large, black, velvety maculation, the hind tibiae with many more spines on the external margins (Figure 5a–d), and the male genitalia structure, especially the epiphallus (Figure 5e–j), which is very similar to that of Shirakiacris shirakii (Figure 5k–t). Based on molecular analysis, Choroedocus has a robust relationship with Shirakiacris of the subfamily Eyprepocnemidinae (Figure 2 and Figure 3, Supplementary Materials Figures S2 and S5). Therefore, it is more reasonable to consider Choroedocus as a member of the subfamily Eyprepocnemidinae.

4.3. Phylogenetic Relationships among Some Related Genera

4.3.1. Non-Monophyly of the Family Dericorythidae

The family name Dericorythidae was first proposed by Eades [41] according to the distinctive pseudoarch in the phallic complex. The pseudoarch found in Dericorythidae is a paired structure not connected across the midline (Figure 6j,s–t). In contrast, the arch of aedeagus rises from the median, dorsobasal region of the dorsal valves of aedeagus [44] (Figure 4m). Eades [41] thought that the presence of a well-developed arch sclerite should be treated as a crucial character in defining the family Acrididae. However, the representatives of Dericorythidae examined by Eades [41] were extremely limited, with only two species in the subfamily Dericorythinae, and one species in the subfamilies Conophyminae and Iranellinae, each, leading to a possibility that the morphological diversity was not fully represented by the limited taxon sampling.
In this study, the 3 sampled species of Dericorythidae did not cluster into a single clade in all phylogenetic trees (Figure 2 and Figure 3, Supplementary Materials Figures S2–S5). Conophymacris szechwanensis first clusters into a clade with Xiangelilacris zhongdianensis, and then with Longzhouacris mirabilis and two species of Coptacrinae in the trees from the mitogenome dataset, showing an extremely distant relationship with the two other species of Dericorythidae (Figure 2 and Figure 3, Supplementary Materials Figures S2 and S5). Furthermore, Conophymacris szechwanensis has not only a true arch rather than a pseudoarch (Figure 7j), but also an extremely different external morphology (Figure 7a–d) and geographical distribution. Dericorys annulata and Conophyma zhaosuensis are both located at the base of the trees, but do not form a single clade in most phylogenetic trees (Figure 2, Supplementary Materials Figures S2 and S5), except in the ML tree deduced from amino acid dataset with MtOrt model, where Dericorys annulata and Conophyma zhaosuensis forms an independent clade (Figure 3). Although both Dericorys annulata and Conophyma zhaosuensis have pseudoarches in the phallic complex (Figure 6j,s,u) and similar geographical distribution region, they exhibit distinct differences in external morphology, including the general appearance and the male genitalia structure, especially the epiphallus (Figure 6e–i,p–r).
Therefore, the family Dericorythidae is certainly not a monophyletic group, and the relationship among the family needs to be clarified by denser taxon and character sampling and more nuclear molecular markers, including both genome and transcriptome data [5]. After all, morphology may be heavily influenced by many abiotic and biotic factors [92]. Some morphological characters may have evolved independently multiple times [2,28] and may not be reliable for recognizing monophyletic groups within some higher categories [28,93]. Phylogenetic relationships between or within some groups may be clouded by many factors, such as gene tree discordance, introgression, and the gene tree anomaly zone [92]. Denser taxon sampling will reveal a wide range of variation and more efficiently improve an artificial classification [94,95,96].

4.3.2. Phylogenetic Relationship between the Genera Conophymacris and Xiangelilacris

The genus Conophymacris was erected by Willemse [97] with Conophymacris chinensis Willemse, 1933 as the type species, and placed originally in the subfamily Catantopinae. Later, it was respectively placed in the tribe Conophymatini of Catantopinae [43,98], the subfamily Conophyminae [40], and Podisminae [33,34,81], but not assigned to a definite tribal position within the subfamily Conophyminae in OSF [32]. The genus Xiangelilacris was established by Zheng et al. [99] with the type species, Xiangelilacris zhongdianensis, as the only known species so far. This genus is most similar to the genera Indopodisma and Pedopodisma, according to the original description. Therefore, it was undoubtedly recognized as a member of the tribe Podismini of the subfamily Melanoplinae by later acridologists [34]. However, this opinion has not been supported by mitogenome evidence, and it always has a very robust close relationship with Conophymacris in all phylogenetic trees [27] (Figure 2 and Figure 3, Supplementary Materials Figures S2–S5). The clade of Conophymacris + Xiangelilacris fell into neither the subfamily Melanoplinae nor the clades of other species of the family Dericorythidae. We examined some materials of Conophymacris szechwanensis (Figure 7a–k) as well as the types of Xiangelilacris zhongdianensis (Figure 7l–s), and found that the types of Xiangelilacris zhongdianensis are nymphs rather than adults, with the bud of hind wings distinctly covering on that of tegmina (Figure 7r,s), and most similar to Conophymacris species. In addition, Conophymacris szechwanensis has a true arch sclerite in the phallic complex (Figure 7j), indicating its possible membership in the family Acrididae. Both Conophymacris szechwanensis and Xiangelilacris zhongdianensis have distinct lateral carinae on the pronotum and ectoapical spines in the hind tibiae, which are absent in Melanoplinae. Therefore, it is undoubted that the genera Conophymacris and Xiangelilacris have a most robust close relationship with each other, and neither of them belongs to either the subfamily Melanoplinae or the family Dericorythidae, or even the Conophyminae. Their exact position needs further comprehensive research.

4.3.3. Phylogenetic Relationship between the Genera Menglacris, Ranacris and Longzhouacris

The genus Menglacris was established by Jiang and Zheng [100], with Menglacris maculata Jiang and Zheng, 1994 as the type species. Although it was not assigned to a definite subfamily position originally, its membership in Habrocneminae was recognized by Li et al. [33], and followed by Mao et al. [34]. The taxonomic history and subfamily position of the genus Ranacris have been mentioned in the introduction section. Although these two genera are placed in different subfamilies in OSF, they display a very robust close relationship with each other in all phylogenetic trees (Figure 2 and Figure 3, Supplementary Materials Figures S2–S5) and an extremely high similarity in external morphology and male genital structure (Figure 8). However, they are always far away from the genus Longzhouacris of the subfamily Habrocneminae in all phylogenetic trees (Figure 2 and Figure 3, Supplementary Materials Figures S2–S5), and show distinct differences in morphology and male genital structure (Figure 8 and Figure 9). Therefore, we have decided that the genera Menglacris and Ranacris are close relatives, but their relationship with Longzhouacris and other related groups may be resolved only in a broader context where all members of the subfamily Habrocneminae and the tribe Mesambriini, or even more related groups, could be included in the analysis.

4.4. Performance of the Mitochondrial COX1 Gene in Reconstructing Phylogeny

A previous study has suggested that COX1 barcode region may perform much better in phylogenetic reconstruction at genus and species ranks than at higher ranks [101]. In this study, the COX1 barcode region extracted from the mitogenome plus that of six additional Conophyma species [102] was used again to test: (1) the accuracy of mitogenome sequence of Conophyma zhaosuensis; (2) the phylogenetic position of the genus Conophyma using more sampled species; and (3) its performance in reconstructing phylogeny at higher categories under the phylogenetic framework derived from the complete mitogenome dataset. The result showed that Conophyma zhaosuensis always fell within the same clade together with other Conophyma species (Supplementary Materials Figures S3 and S4) [102], indicating the reliability or accuracy of the newly sequenced mitogenome of Conophyma zhaosuensis. Although the Conophyma species always formed a single clade in both ML and BI trees (Supplementary Materials Figures S3 and S4), they did not cluster into a single clade with Dericorys annulata and Conophymacris szechwanensis, just as in the trees deduced from the complete mitogenome dataset (Figure 2, Supplementary Materials Figure S2), indicating the remote relationship among these three genera. As for the performance of COX1 gene in resolving the phylogeny at higher categories, the monophylies of most well-charactered subfamilies were not supported except for Spathosterninae and Cyrtacanthacridinae. Sometimes, even the congeneric species, such as the Hieroglyphus species, were split into different clades. Therefore, the mitochondrial COX1 gene alone is not suitable for resolving phylogeny of higher categories, at least in Acridoidea, but is indeed powerful in reconstructing phylogenetic relationships among closely related species [103] despite the high error rates sometimes in individual lineages [104].

Supplementary Materials

The following supporting information can be downloaded at: https://www.mdpi.com/article/10.3390/insects14010085/s1, Figure S1. Secondary structures of 22 tRNAs of the three newly sequenced mitogenomes. Figure S2. Phylogenetic tree reconstructed from sequences of the 13 mitochondrial PCGs using Bayesian inference. The asterisk indicates the three newly sequenced species. Figure S3. Phylogenetic tree reconstructed from partial sequences of the COX1 gene using maximum likelihood. The asterisk indicates the three newly sequenced species. Figure S4. Phylogenetic tree reconstructed from partial sequences of the COX1 geneusing Bayesian inference. The asterisk indicates the three newly sequenced species. Figure S5. Phylogenetic tree reconstructed from amino acid sequences of the 13 mitochondrial PCGs gene using Bayesian inference with MtRev model. The asterisk indicates the three newly sequenced species. Table S1. Accession numbers and references of COXI sequences of Conophyma spp. sampled in this study. Table S2. The best–fitting models used for phylogenetic analyses of the mitochondrial PCGs dataset. Table S3. The best–fitting models used for the phylogenetic analyses of the partial COX1 dataset. Table S4. Nucleotide composition of the mitogenome of Longzhouacris mirabilis (mt1825). Table S5.Nucleotide composition of the mitogenome of Ranacris albicornis (mt1826). Table S6. Nucleotide composition of the mitogenome of Conophyma zhaosuensis (mt1938). Table S7. Codon usage of the PCGs of Longzhouacris mirabilis. Table S8. Codon usage of the PCGs of Ranacris albicornis. TableS9. Codon usage of the PCGs of Conophyma zhaosuensis. Table S10. Annotation of the three newly sequenced complete mitogenomes.

Author Contributions

Conceptualization, B.M., Y.H. and J.H.; data curation, C.Z.; formal analysis, C.Z. and Z.C.; funding acquisition, B.M. and J.H.; investigation, B.M., H.W., L.D. and Z.C.; methodology, Y.H.; project administration, J.H.; resources, B.M., Z.C. and J.H.; supervision, J.H.; validation, B.M. and J.H.; visualization, C.Z.; writing—original draft, C.Z. and J.H.; writing—review and editing, B.M., Y.H. and J.H. All authors have read and agreed to the published version of the manuscript.

Funding

This work was financially supported by the Open Foundation of Guangxi Key Laboratory of Rare and Endangered Animal Ecology, Guangxi Normal University (GKN22-A-02-01), the National Natural Science Foundation of China (No.31760628, 31960110, 31540055) and the Natural Science Foundation of Changsha (kq2202279).

Institutional Review Board Statement

Not applicable.

Informed Consent Statement

Not applicable.

Data Availability Statement

The mitogenome sequencesare deposited in GenBank under accession number ON943039 for Ranacris albicornis, ON931612 for Longzhouacris mirabilis and ON943040 for Conophyma zhaosuensis, respectively.

Acknowledgments

First of all, we heartily acknowledge the great contributions of our colleagues who sequenced a great quantity of grasshopper mitogenomes and submitted the data to National Center for Biotechnology Information (NCBI). It is impossible for us to discuss the related phylogenetic problems in this manuscript without the longtime accumulation of the mitogenome data by our colleagues. We would like to thank Jingxiao Gu for collecting the materials of Longzhouacris mirabilis for molecular study. We are thankful to Haiyang Xu for his help in the assembly and annotation of the mitogenome and the reconstruction of the phylogenetic trees. Yao Niu (Henan Normal University) is acknowledged for providing a lot of valuable comments on the relationship among related groups of Acrididae with prosternal process and the variations on the male genitaliae of Acrididae. Special gratitude is given to Huihui Chang for the reconstruction of phylogenetic trees using the amino acid dataset with MtOrt model, and to Dong Zhang (Lanzhou University), Fangluan Gao (Fujian Agricultural and Forestry University) and Feng Zhang (Nanjing Agricultural University) for their instructions in reconstructing phylogenetic trees using mulecular sequences.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Lecoq, M.; Zhang, L. Encyclopedia of Pest Orthoptera of the World; China Agricultural University Press: Beijing, China, 2019. [Google Scholar]
  2. Mugleston, J.; Naegle, M.; Song, H.; Bybee, S.M.; Ingley, S.; Suvorov, A.; Whiting, M.F. Reinventing the leaf: Multiple origins of leaf-like wings in katydids (Orthoptera: Tettigoniidae). Invertebr. Syst. 2016, 30, 335–352. [Google Scholar] [CrossRef]
  3. Li, X.-D.; Jiang, G.-F.; Yan, L.-Y.; Li, R.; Mu, Y.; Deng, W.-A. Positive selection drove the adaptation of mitochondrial genes to the demands of flight and high-altitude environments in grasshoppers. Front. Genet. 2018, 9, 605. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  4. Chen, L.P.; Zheng, F.Y.; Bai, J.; Wang, J.M.; Lv, C.Y.; Li, X.; Zhi, Y.C.; Li, X.J. Comparative analysis of mitogenomes among six species of grasshoppers (Orthoptera: Acridoidea: Catantopidae) and their phylogenetic implications in wing-type evolution. Int. J. Biol. Macromol. 2020, 159, 1062–1072. [Google Scholar] [CrossRef] [PubMed]
  5. Song, H.; Béthoux, O.; Shin, S.; Donath, A.; Letsch, H.; Liu, S.; McKenna, D.D.; Meng, G.; Misof, B.; Podsiadlowski, L.; et al. Phylogenomic analysis sheds light on the evolutionary pathways towards acoustic communication in Orthoptera. Nat. Commun. 2020, 11, 4939. [Google Scholar] [CrossRef]
  6. Huang, J.-P.; Hill, J.; Ortego, J.; Knowles, L.L. Paraphyletic species no more—Genomic data resolve a Pleistocene radiation and validate morphological species of the Melanoplus scudderi complex (Insecta: Orthoptera). Syst. Entomol. 2020, 45, 594–605. [Google Scholar] [CrossRef]
  7. Gu, J.X.; Jiang, B.; Wang, H.J.; Wei, T.; Lin, L.L.; Huang, Y.; Huang, J.H. Phylogeny and species delimitation of the genus Longgenacris and Fruhstorferiola viridifemorata species group (Orthoptera: Acrididae: Melanoplinae) based on molecular evidence. PLoS ONE 2020, 15, e0237882. [Google Scholar] [CrossRef]
  8. Hemp, C.; Scherer, C.; Brandl, R.; Pinkert, S. The origin of the endemic African grasshopper family Lentulidae (Orthoptera: Acridoidea) and its climate-induced diversification. J. Biogeogr. 2020, 47, 1805–1815. [Google Scholar] [CrossRef]
  9. Wang, H.; Jiang, B.; Gu, J.X.; Wei, T.; Lin, L.L.; Huang, Y.; Liang, D.; Huang, J.H. Molecular phylogeny and species delimitation of the genus Tonkinacris (Orthoptera, Acrididae, Melanoplinae) from China. PLoS ONE 2021, 16, e0249431. [Google Scholar] [CrossRef]
  10. Dey, L.-S.; Hochkirch, A.; Moussi, A.; Simões, M.V.P.; Husemann, M. Diversification in and around the Atlas Mountains: Insights into the systematics and biogeography of the genus Thalpomena (Orthoptera: Acrididae: Oedipodinae). Syst. Entomol. 2022, 47, 402–419. [Google Scholar] [CrossRef]
  11. Fenn, J.D.; Song, H.; Cameron, S.L.; Whiting, M.F. A preliminary mitochondrial genome phylogeny of Orthoptera (Insecta) and approaches to maximizing phylogenetic signal found within mitochondrial genome data. Mol. Phylogenet. Evol. 2008, 49, 59–68. [Google Scholar] [CrossRef]
  12. Erler, S.; Ferenz, H.J.; Moritz, R.F.A.; Kaatz, H.H. Analysis of the mitochondrial genome of Schistocerca gregaria gregaria (Orthoptera: Acrididae). Biol. J. Linn. Soc. 2010, 99, 296–305. [Google Scholar] [CrossRef]
  13. Sheffield, N.C.; Hiatt, K.D.; Valentine, M.C.; Song, H.; Whiting, M.F. Mitochondrial genomics in Orthoptera using MOSAS. Mitochondrial DNA 2010, 21, 87–104. [Google Scholar] [CrossRef] [PubMed]
  14. Leavitt, J.R.; Hiatt, K.D.; Whiting, M.F.; Song, H. Searching for the optimal data partitioning strategy in mitochondrial phylogenomics: A phylogeny of Acridoidea (Insecta: Orthoptera: Caelifera) as a case study. Mol. Phylogenet. Evol. 2013, 67, 494–508. [Google Scholar] [CrossRef] [PubMed]
  15. Zhang, H.L.; Huang, Y.; Lin, L.L.; Wang, X.Y.; Zheng, Z.M. The phylogeny of the Orthoptera (Insecta) as deduced from mitogenomic gene sequences. Zool. Stud. 2013, 52, 37. [Google Scholar] [CrossRef] [Green Version]
  16. Song, H.; Amédégnato, C.; Cigliano, M.M.; Desutter-Grandcolas, L.; Heads, S.W.; Huang, Y.; Otte, D.; Whiting, M.F. 300 million years of diversification: Elucidating the patterns of orthopteran evolution based on comprehensive taxon and gene sampling. Cladistics 2015, 31, 621–651. [Google Scholar] [CrossRef]
  17. Song, H.; Mariño-Pérez, R.; Woller, D.A.; Cigliano, M.M. Evolution, diversification, and biogeography of grasshoppers (Orthoptera: Acrididae). Insect Syst. Divers. 2018, 2, 3. [Google Scholar] [CrossRef] [Green Version]
  18. Liu, H.X.; Yan, L.Y.; Jiang, G.F. The complete mitochondrial genome of the jumping grasshopper Sinopodisma pieli (Orthoptera: Acrididae) and the phylogenetic analysis of Melanoplinae. Zootaxa 2017, 4363, 506–520. [Google Scholar] [CrossRef]
  19. Zhou, Z.J.; Zhao, L.; Liu, N.; Guo, H.F.; Guan, B.; Di, J.X.; Shi, F.M. Towards a higher-level Ensifera phylogeny inferred from mitogenome sequences. Mol. Phylogenet. Evol. 2017, 108, 22–33. [Google Scholar] [CrossRef]
  20. Li, R.; Shu, X.; Li, X.; Meng, L.; Li, B. Comparative mitogenome analysis of three species and monophyletic inference of Catantopinae (Orthoptera: Acridoidea). Genomics 2019, 111, 1728–1735. [Google Scholar] [CrossRef]
  21. Li, R.; Shu, X.; Deng, W.; Meng, L.; Li, B. Complete mitochondrial genome of Atractomorpha sagittaris (Orthoptera: Pyrgomorphidae) and its phylogenetic analysis for Acrididea. Biologia 2020, 75, 1571–1583. [Google Scholar] [CrossRef]
  22. Li, R.; Wang, Y.; Shu, X.; Meng, L.; Li, B. Complete mitochondrial genomes of three Oxya grasshoppers (Orthoptera) and their implications for phylogenetic reconstruction. Genomics 2020, 112, 289–296. [Google Scholar] [CrossRef]
  23. Mariño-Pérez, R.; Song, H. On the origin of the New World Pyrgomorphidae (Insecta: Orthoptera). Mol. Phylogenet. Evol. 2019, 139, 106537. [Google Scholar] [CrossRef]
  24. Chang, H.; Nie, Y.; Zhang, N.; Zhang, X.; Sun, H.; Mao, Y.; Qiu, Z.Y.; Huang, Y. MtOrt: An empirical mitochondrial amino acid substitution model for evolutionary studies of Orthoptera insects. BMC Evol. Biol. 2020, 20, 57. [Google Scholar] [CrossRef]
  25. Chang, H.; Qiu, Z.; Yuan, H.; Wang, X.; Li, X.; Sun, H.; Guo, X.; Lu, Y.; Feng, X.; Majid, M.; et al. Evolutionary rates of and selective constraints on the mitochondrial genomes of Orthoptera insects with different wingtypes. Mol. Phylogenet. Evol. 2020, 145, 106734. [Google Scholar] [CrossRef]
  26. Qiu, Z.Y.; Chang, H.H.; Yuan, H.; Huang, Y.; Lu, H.M.; Li, X.; Gou, X.C. Comparative mitochondrial genomes of four species of Sinopodisma and phylogenetic implications (Orthoptera, Melanoplinae). ZooKeys 2020, 969, 23–42. [Google Scholar] [CrossRef]
  27. Xu, H.Y.; Mao, B.Y.; Storozhenko, S.Y.; Huang, Y.; Chen, Z.L.; Huang, J.H. Phylogenetic position of the genus Alulacris (Orthoptera: Acrididae: Melanoplinae: Podismini) revealed by complete mitogenome evidence. Insects 2021, 12, 918. [Google Scholar] [CrossRef]
  28. Zahid, S.; Mariño-Pérez, R.; Song, H. Molecular phylogeny of the grasshopper family Pyrgomorphidae (Caelifera, Orthoptera) reveals rampant paraphyly and convergence of traditionally used taxonomic characters. Zootaxa 2021, 4969, 101–118. [Google Scholar] [CrossRef]
  29. Zeng, X.; Xu, H.Y.; Gu, J.X.; Mao, B.Y.; Chen, Z.L.; Huang, Y.; Huang, J.H. Phylogenetic position of the genera Caryandoides, Paratoacris, Fer and Longchuanacris (Orthoptera: Acrididae) revealed by complete mitogenome sequences. Invertebr. Syst. 2021, 35, 725–741. [Google Scholar] [CrossRef]
  30. Zheng, F.-Y.; Shi, Q.-Y.; Ling, Y.; Chen, J.-Y.; Zhang, B.-F.; Li, X.-J. Comparative analysis of mitogenomes among five species of Filchnerella (Orthoptera: Acridoidea: Pamphagidae) and their phylogenetic and taxonomic implications. Insects 2021, 12, 605. [Google Scholar] [CrossRef]
  31. You, Q.J.; Li, T.S.; Bi, D.Y. Descriptions of new genera and species of Catantopidae from Guangxi (Orthoptera: Acridoidea). Entomotaxonomia 1983, 5, 165–181. [Google Scholar]
  32. Cigliano, M.M.; Braun, H.; Eades, D.C.; Otte, D. Orthoptera Species File. Version 5.0/5.0. 2022. Available online: http://Orthoptera.SpeciesFile.org/HomePage/Orthoptera/HomePage.aspx (accessed on 10 November 2022).
  33. Li, H.C.; Xia, K.L. Fauna Sinica, Insecta, Vol. 43, Orthoptera, Acridoidea, Catantopidae; Science Press: Beijing, China, 2006. [Google Scholar]
  34. Mao, B.Y.; Ren, G.D.; Ou, X.H. Fauna, Distribution Pattern and Adaptability on Acridoidea from Yunnan; China Forestry Publishing House: Beijing, China, 2011. [Google Scholar]
  35. You, Q.J.; Lin, R.Z. Description of a new genus and new species of Catantopidae from Guangxi Province, China. Entomotaxonomia 1983, 5, 255–258. [Google Scholar]
  36. Zheng, Z.M.; Lin, L.L.; Deng, W.A.; Shi, F.M. Two new species of Catantopidae (Orthoptera: Acridoidea) from China. Entomotaxonomia 2015, 37, 177–181. [Google Scholar] [CrossRef]
  37. You, Q.J.; Li, T.S.; Zhang, Y.Q.; Lin, R.Z. Economic Insect Iconography of Guangxi, Phytophage Insects; Guangxi Press of Science & Technology: Nanning, China, 1990. [Google Scholar]
  38. Jiang, G.F.; Zheng, Z.M. Grasshoppers and Locusts from Guangxi; Guangxi Normal University Press: Guilin, China, 1998. [Google Scholar]
  39. Storozhenko, S.Y. A new species of the genus Mesambria Stål, 1878 with notes on the tribe Mesambriini (Orthoptera: Acrididae, Catantopinae). Zootaxa 2018, 4418, 55–65. [Google Scholar] [CrossRef]
  40. Otte, D. Orthoptera Species File 4. Grasshoppers (Acridomorpha) D; Orthopterists’ Society & Academy of Natural Sciences of Philadelphia: Philadelphia, PA, USA, 1995. [Google Scholar]
  41. Eades, D.C. Evolutionary relationships of phallic structures of Acridomorpha (Orthoptera). J. Orthoptera Res. 2000, 9, 181–210. [Google Scholar] [CrossRef]
  42. Uvarov, B.P. Grasshoppers and Locusts. A Handbook of General Acridology. Vol.1. Anatomy, Physiology, Development, Phase Polymorphism, IntroductiontoTaxonomy; University of Cambridge Press: Cambridge, UK, 1966. [Google Scholar]
  43. Storozhenko, S.Y.; Kim, T.W.; Jeon, M.J. Monograph of Korean Orthoptera; National Institute of Biological Resources: Incheon, Republic of Korea, 2015. [Google Scholar]
  44. Woller, D.A.; Song, H. Investigating the functional morphology of genitalia during copulation in the grasshopper Melanoplus rotundipennis (Scudder,1878) via correlative microscopy. J. Morph. 2017, 278, 334–359. [Google Scholar] [CrossRef]
  45. Han, H.; Wang, N.; Xu, L.; Gao, S.; Liu, A. The complete mitochondrial genome of Calliptamus abbreviatus Ikovnnikov (Orthoptera: Acridoidea). Mitochondrial DNA Part B 2016, 1, 770–771. [Google Scholar] [CrossRef] [Green Version]
  46. Zhi, Y.; Zhang, N.; Lu, X.; Yin, H.; Zhang, D. The complete mitochondrial genome of Peripolus nepalensis Uvarov, 1942 (Orthoptera: Acridoidea: Catantopidae). Mitochondrial DNA 2016, 27, 26–27. [Google Scholar] [CrossRef]
  47. Qiu, Z.Y.; Wang, Y.T.; Yuan, H.; Zhao, X.; Ru, N.Y.; Liu, L.; Cui, Y.Y.; Liu, Y.T. The complete mitochondrial genome of Traulia minuta Huang & Xia, 1985 (Orthoptera: Catantopinae). Mitochondrial DNA Part B 2021, 6, 2180–2181. [Google Scholar] [CrossRef]
  48. Yang, J.; Liu, Y.; Liu, N. The complete mitochondrial genome of the Xenocatantops brachycerus (Orthoptera: Catantopidae). Mitochondrial DNA Part A 2016, 27, 2844–2845. [Google Scholar] [CrossRef]
  49. Chang, H.; Huang, Y. The complete mitochondrial genome of the Hieroglyphus tonkinensis (Orthoptera: Acrididae). Mitochondrial DNA Part B 2016, 1, 534–535. [Google Scholar] [CrossRef] [Green Version]
  50. Chen, Z.N.; Xu, S.Q. The complete mitochondrial DNA genome sequence of a terrestrial grasshopper, Curvipennis wixiensis (Acrididae: Podismini). Conserv. Genet. Resour. 2017, 9, 115–118. [Google Scholar] [CrossRef]
  51. Liu, F.; Qiu, Z. The complete mitochondrial genome of Fruhstorferiola huayinensis (Orthoptera: Catantopidae). Mitochondrial DNA Part B 2016, 1, 273–274. [Google Scholar] [CrossRef] [PubMed]
  52. Yang, R.; Guan, D.L.; Xu, S.-Q. Complete mitochondrial genome of the Chinese endemic grasshopper Fruhstorferiola kulinga (Orthoptera: Acrididae: Podismini). Mitochondrial DNA 2016, 27, 3240–3241. [Google Scholar] [CrossRef] [PubMed]
  53. Zhang, X.M.; Lin, L.L. The complete mitochondrial genome of Fruhstorferiola tonkinensis (Orthoptera: Catantopidae). Mitochondrial DNA Part B 2016, 1, 434–435. [Google Scholar] [CrossRef]
  54. Zhi, Y.; Liu, B.; Han, G.; Yin, H.; Zhang, D.C. The complete mitochondrial genome of Kingdonella bicollina (Orthoptera: Acridoidea: Catantopidae). Mol. Biol. Rep. 2016, 27, 611–619. [Google Scholar] [CrossRef]
  55. Sun, H.; Zheng, Z.; Huang, Y. Sequence and phylogenetic analysis of complete mitochondrial DNA genomes of two grasshopper species Gomphocerus rufus (Linnaeus, 1758) and Primnoa arctica (Zhang and Jin, 1985) (Orthoptera: Acridoidea). Mitochondrial DNA 2010, 21, 115–131. [Google Scholar] [CrossRef] [PubMed]
  56. Li, R.; Jiang, G.F.; Liang, A.P.; Zhong, X.T.; Liu, Y. Characterization of the mitochondrial genome of the montane grasshopper, Qinlingacris elaeodes (Orthoptera: Catantopidae). Mitochondrial DNA Part A 2016, 27, 1765–1766. [Google Scholar] [CrossRef]
  57. Zhang, X.; Li, X.; Liu, F.; Yuan, H.; Huang, Y. The complete mitochondrial genome of Tonkinacris sinensis (Orthoptera: Acrididae): A tRNA-like sequence and its implications for phylogeny. Biochem. Syst. Ecol. 2017, 70, 147–154. [Google Scholar] [CrossRef]
  58. Hu, Z.; Guan, D.L.; Mao, B.Y. Characterization of the complete mitochondrial genome of the Yunnan endemic grasshopper Yunnanacris yunnaneus (Insecta: Orthoptera: Acrididae). Conserv. Genet. Resour. 2016, 8, 267–270. [Google Scholar] [CrossRef]
  59. Yuan, H.; Qiu, Z.; Yang, C.; Huang, Y. The complete mitochondrial genome sequence of Caryanda elegans (Orthoptera: Acrididae). Mitochondrial DNA Part B 2019, 4, 1580–1581. [Google Scholar] [CrossRef] [Green Version]
  60. Hu, Z.; Guan, D.-L.; Mao, B.-Y. Description of a new species, Caryanda Stål, 1878 (Acrididae, Orthoptera) from China. Orient. Insects 2017, 51, 124–134. [Google Scholar] [CrossRef]
  61. Hu, Z.; Han, Y.-P.; Guan, D.-L.; Mao, B.-Y. Characterization of the complete mitochondrial genome of the Yunnan endemic grasshopper Longchuanacris curvifurculus (Insecta: Orthoptera: Catantopidae). Mitochondrial DNA Part B 2018, 3, 670–671. [Google Scholar] [CrossRef] [PubMed]
  62. Zhang, C.Y.; Huang, Y. Complete mitochondrial genome of Oxya chinensis (Orthoptera, Acridoidea). Acta Biochim. Biophys. Sin. 2008, 40, 7–18. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  63. Song, N.; Li, H.; Song, F.; Cai, W. Molecular phylogeny of Polyneoptera (Insecta) inferred from expanded mitogenomic data. Scient. Rep. 2016, 6, 36175. [Google Scholar] [CrossRef] [Green Version]
  64. Dong, J.J.; Guan, D.L.; Xu, S.Q. Complete mitogenome of the semiaquatic grasshopper Oxya intricata (Stål.) (Insecta: Orthoptera: Catantopidae). Mitochondrial DNA Part A 2016, 27, 3233–3234. [Google Scholar] [CrossRef]
  65. Tang, M.; Tan, M.H.; Meng, G.L.; Yang, S.Z.; Su, X.; Liu, S.L.; Song, W.H.; Li, Y.Y.; Wu, Q.; Zhang, A.B.; et al. Multiplex sequencing of pooled mitochondrial genomes—A crucial step toward biodiversity analysis using mito-metagenomics. Nucleic Acids Res. 2014, 42, e166. [Google Scholar] [CrossRef] [Green Version]
  66. Zhou, F.; Huang, Y. The complete mitochondrial genome of Spathosternum prasiniferum sinense Uvarov, 1931 (Orthoptera: Acridoidea: Acrididae). Mitochondrial DNA 2016, 27, 1932–1933. [Google Scholar] [CrossRef]
  67. Hahn, C.; Bachmann, L.; Chevreux, B. Reconstructing mitochondrial genomes directly from genomic next-generation sequencing reads—A baiting and iterative mapping approach. Nucleic Acids Res. 2013, 41, e129. [Google Scholar] [CrossRef] [Green Version]
  68. Bernt, M.; Donath, A.; Jühling, F.; Externbrink, F.; Florentz, C.; Fritzsch, G.; Pütz, J.; Middendorf, M.; Stadler, P.F. MITOS: Improved de novo metazoan mitochondrial genome annotation. Mol. Phylogenet. Evol. 2013, 69, 313–319. [Google Scholar] [CrossRef]
  69. Kearse, M.; Moir, R.; Wilson, A.; Stones-Havas, S.; Cheung, M.; Sturrock, S.; Buxton, S.; Cooper, A.; Markowitz, S.; Duran, C.; et al. Geneious basic: An integrated and extendable desktop software platform for the organization and analysis of sequence data. Bioinformatics 2012, 28, 1647–1649. [Google Scholar] [CrossRef] [Green Version]
  70. Darty, K.; Denise, A.; Ponty, Y. VARNA: Interactive drawing and editing of the RNA secondary structure. Bioinformatics 2009, 25, 1974–1975. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  71. Kumar, S.; Stecher, G.; Li, M.; Knyaz, C.; Tamura, K. MEGA X: Molecular evolutionary genetics analysis across computing platforms. Mol. Biol. Evol. 2018, 35, 1547–1549. [Google Scholar] [CrossRef]
  72. Ranwez, V.; Douzery, E.J.P.; Cambon, C.; Chantret, N.; Delsuc, F. MACSE v2: Tool kit for the alignment of coding sequences accounting for frame shifts and stop codons. Mol. Biol. Evol. 2018, 35, 2582–2584. [Google Scholar] [CrossRef]
  73. Zhang, D.; Gao, F.; Jakovlić, I.; Zou, H.; Zhang, J.; Li, W.X.; Wang, G.T. PhyloSuite: An integrated and scalable desktop platform for streamlined molecular sequence data management and evolutionary phylogenetics studies. Mol. Ecol. Res. 2020, 20, 34–355. [Google Scholar] [CrossRef] [PubMed]
  74. Kalyaanamoorthy, S.; Minh, B.Q.; Wong, T.K.F.; Haeseler, A.; Jermiin, L.S. ModelFinder: Fast model selection for accurate phylogenetic estimates. Nat. Methods 2017, 14, 587–589. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  75. Nguyen, L.-T.; Schmidt, H.A.; Haeseler, A.; Minh, B.Q. IQ-TREE: A fast and effective stochastic algorithm for estimating maximum-likelihood phylogenies. Mol. Biol. Evol. 2015, 32, 268–274. [Google Scholar] [CrossRef] [PubMed]
  76. Hoang, D.T.; Chernomor, O.; Haeseler, A.; Minh, B.Q.; Vinh, L.S. UFBoot2: Improving the ultrafast bootstrap approximation. Mol. Biol. Evol. 2018, 35, 518–522. [Google Scholar] [CrossRef] [PubMed]
  77. Trifinopoulos, J.; Nguyen, L.-T.; Haeseler, A.; Minh, B.Q. WIQ-TREE: A fast online phylogenetic tool for maximum likelihood analysis. Nucleic Acids Res. 2016, 44, W232–W235. [Google Scholar] [CrossRef] [Green Version]
  78. Hillis, D.M.; Bull, J.J. An empirical test of bootstrapping as a method for assessing confidence in phylogenetic analysis. Syst. Biol. 1993, 42, 182–192. [Google Scholar] [CrossRef]
  79. Ronquist, F.; Teslenko, M.; van der Mark, P.; Ayres, D.L.; Darling, A.; Höhna, S.; Larget, B.; Liu, L.; Suchard, M.A.; Huelsenbeck, J.P. MrBayes3.2: Efficient Bayesian phylogenetic inference and model choice across a large model space. Syst. Biol. 2012, 61, 539–542. [Google Scholar] [CrossRef] [Green Version]
  80. Zheng, Z.M. New genera and new species of grasshoppers from Yunnan, Guizhou and Sichuan, China. Acta Zootax. Sin. 1981, 6, 60–68. [Google Scholar]
  81. Zheng, Z.M. Grasshoppers from the Areas of Yunnan, Guizhou, Sichuan, Shanxi and Ningxia; Science Press: Beijing, China, 1985. [Google Scholar]
  82. Bolívar, I. Estudios Entomológicos Segunda Parte. I. El grupo de los Euprepocnemes. Trab. Mus. Nac. Cienc. Nat. Zool. 1914, 20, 1–40. [Google Scholar]
  83. Stål, C. Systema Acridiodeorum. Essai d’une systématisation des Acridiodées. Bih. K. Sven. Vet.-Akad. Handl. 1878, 5, 1–100. [Google Scholar]
  84. Brunner von Wattenwyl, C. Révision du système des orthoptères et description des espèces rapportées par M. Leonardo Fea de Birmanie. Ann. Mus. Civ. Stor. Nat. Genova Ser. 2 1893, 13, 1–230. [Google Scholar]
  85. Kirby, W.F. List of a small collection of orthopterous insects formed by Sir Harry Johnston in British East Africa and Uganda in 1899 and 1900, with descriptions of five new species. P. Gener. Meet. Scient. Bus. Zool. Soc. Lond. 1902, 1, 93–101. [Google Scholar]
  86. Willemse, C. Bijdrage tot de kennis der Orthoptera s.s. van den Nederlandsch Indischen Archipel en omliggende gebieden. Zool. Meded. 1921, 6, 1–44. [Google Scholar]
  87. Uvarov, B.P. Notes on the Orthoptera in the British Museum. 1. The group of Euprepocnemini. Trans. Entomol. Soc. Lond. 1921, 69, 106–144. [Google Scholar] [CrossRef]
  88. Tinkham, E.R. Taxonomic and biological studies of the Cyrtacanthacrinae of South China. Lingnan Sci. J. 1940, 19, 269–382. [Google Scholar]
  89. Xia, K.L. Synopsis of the Classification on the Chinese Acrididae [Taxonomic Essentials (Principal classification) of Acrididae from China]; Science Press: Beijing, China, 1958. [Google Scholar]
  90. Yin, X.C. Grasshoppers and Locusts from Qinghai-Xizang Plateau of China; Science Press: Beijing, China, 1984. [Google Scholar]
  91. Liu, J.P. (Ed.) Studies on Acridoidea of Hainan Island; Tianze Publishing House: Yangling, China, 1995. [Google Scholar]
  92. Owen, C.L.; Miller, G.L. Phylogenomics of the Aphididae: Deep relationships between subfamilies clouded by gene tree discordance, introgression and the gene tree anomaly zone. Syst. Entomol. 2022, 47, 470–486. [Google Scholar] [CrossRef]
  93. Lu, L.; Dietrich, C.H.; Cao, Y.H.; Zhang, Y.L. A multi-gene phylogenetic analysis of the leafhopper subfamily Typhlocybinae (Hemiptera: Cicadellidae) challenges the traditional view of the evolution of wing venation. Mol. Phylogenet. Evol. 2021, 165, 107299. [Google Scholar] [CrossRef]
  94. Huber, B.A.; Astrin, J.J. Increased sampling blurs morphological and molecular species limits: Revision of the Hispaniolan endemic spider genus Tainonia (Araneae: Pholcidae). Invert. Syst. 2009, 23, 281–300. [Google Scholar] [CrossRef]
  95. Vahtera, V.; Edgecombe, G.D.; Giribet, G. Phylogenetics of scolopendromorph centipedes: Can denser taxon sampling improve an artificial classification? Invert. Syst. 2013, 27, 578–602. [Google Scholar] [CrossRef]
  96. Cao, Y.; Dietrich, C.H.; Zahniser, J.N.; Dmitriev, D.A. Dense sampling of taxa and characters improves phylogenetic resolution among deltocephaline leafhoppers (Hemiptera: Cicadellidae: Deltocephalinae). Syst. Entomol. 2022, 47, 430–444. [Google Scholar] [CrossRef]
  97. Willemse, C. On a small collection of Orthoptera from the Chungking District, S.E. China. Nat. Maandbl. 1933, 22, 15–18. [Google Scholar]
  98. Mistshenko, L.L. Fauna of the USSR, Orthoptera 4 Catantopinae; Bey-Bienko, G.J., Mistshenko, L.L., Eds.; Zoological Institute Akademia Nauk SSSR (NS): St. Petersburg, Russia, 1952; Volume 54, pp. 1–610. [Google Scholar]
  99. Zheng, Z.M.; Huang, Y.; Zhou, Z.J. New genus and new species of grasshoppers from Hengduanshan Region, China (Orthoptera, Acridoidea). Acta Zootax. Sin. 2008, 33, 363–367. [Google Scholar]
  100. Jiang, G.-F.; Zheng, Z.-M. A new genus and species of Catantopidae (Orthoptera: Acridoidea) from Yunnan Province, China. Acta Entomol. Sin. 1994, 37, 463–467. [Google Scholar]
  101. Huang, J.H.; Zhang, A.B.; Mao, S.L.; Huang, Y. DNA barcoding and species boundary delimitation of selected species of Chinese Acridoidea (Orthoptera: Caelifera). PLoS ONE 2013, 8, e82400. [Google Scholar] [CrossRef] [Green Version]
  102. Bugrov, A.G.; Smyshlyaev, G.A.; Blinov, A.G. Phylogeny of grasshoppers (Orthoptera, Acrididae) based on the analysis of DNA sequences in COI mitochondrial gene. Euroasian Entomol. J. 2012, 11, 493–502. [Google Scholar]
  103. Saad, Y.M.; El-Sadek, H.E.-S.A. The Efficiency of cytochrome oxidase subunit 1 gene (cox1) in reconstruction of phylogenetic relations among some crustacean species. Int. J. Anim. Vet. Sci. 2017, 11, 515–520. [Google Scholar] [CrossRef]
  104. Hendrich, L.; Pons, J.; Ribera, I.; Balke, M. Mitochondrial cox1 sequence data reliably uncover patterns of insect diversity but suffer from high lineage-idiosyncratic error rates. PLoS ONE 2010, 5, e14448. [Google Scholar] [CrossRef]
Figure 1. Linearised representation of gene arrangements in the three newly sequenced mitogenomes. mt1825: Longzhouacris mirabilis; mt1826: Ranacris albicornis; mt1938: Conophyma zhaosuensis.
Figure 1. Linearised representation of gene arrangements in the three newly sequenced mitogenomes. mt1825: Longzhouacris mirabilis; mt1826: Ranacris albicornis; mt1938: Conophyma zhaosuensis.
Insects 14 00085 g001
Figure 2. Phylogenetic tree reconstructed from the sequences of the 13 mitochondrial PCGs using maximum likelihood. The asterisk indicates the three newly sequenced species.
Figure 2. Phylogenetic tree reconstructed from the sequences of the 13 mitochondrial PCGs using maximum likelihood. The asterisk indicates the three newly sequenced species.
Insects 14 00085 g002
Figure 3. Phylogenetic tree reconstructed from amino acid sequences of the 13 mitochondrial PCGs using maximum likelihood with MtOrt model. The asterisk indicates the three newly sequenced species.
Figure 3. Phylogenetic tree reconstructed from amino acid sequences of the 13 mitochondrial PCGs using maximum likelihood with MtOrt model. The asterisk indicates the three newly sequenced species.
Insects 14 00085 g003
Figure 4. Morphological characters of Emeiacris maculata and Oxya agavisa. (an) Emeiacris maculata. (a,b) Male habitus. (c,d) Female habitus. (e) Right hind leg of male. (f) Meso- and metasterna of male. (g,h) Right hind tabia of male. (ik) Epiphallus in dorsal, frontal and dorsolateral views. (ln) Phallic complex in dorsal and lateral views. (oq) Oxya agavisa. (o) left hind femur of male. (p) left hind tibia of male. (q) Meso- and metasterna of male.
Figure 4. Morphological characters of Emeiacris maculata and Oxya agavisa. (an) Emeiacris maculata. (a,b) Male habitus. (c,d) Female habitus. (e) Right hind leg of male. (f) Meso- and metasterna of male. (g,h) Right hind tabia of male. (ik) Epiphallus in dorsal, frontal and dorsolateral views. (ln) Phallic complex in dorsal and lateral views. (oq) Oxya agavisa. (o) left hind femur of male. (p) left hind tibia of male. (q) Meso- and metasterna of male.
Insects 14 00085 g004
Figure 5. Morphological characters of Choroedocus capensis and Shirakiacris shirakii. (aj) Choroedocus capensis. (a,b) Male habitus. (c,d) Female habitus. (eh) Epiphallus in dorsal, dorsofrontal, frontal and dorsolateral views. (i,j) Phallic complex in dorsal and lateral views. (kt) Shirakiacris shirakii. (k,l) Male habitus. (m,n) Female habitus. (or) Epiphallus in dorsal, dorsofrontal, frontal and dorsolateral views. (s,t) Phallic complex in dorsal and lateral views.
Figure 5. Morphological characters of Choroedocus capensis and Shirakiacris shirakii. (aj) Choroedocus capensis. (a,b) Male habitus. (c,d) Female habitus. (eh) Epiphallus in dorsal, dorsofrontal, frontal and dorsolateral views. (i,j) Phallic complex in dorsal and lateral views. (kt) Shirakiacris shirakii. (k,l) Male habitus. (m,n) Female habitus. (or) Epiphallus in dorsal, dorsofrontal, frontal and dorsolateral views. (s,t) Phallic complex in dorsal and lateral views.
Insects 14 00085 g005
Figure 6. Morphological characters of Dericorys annulata and Conophyma zhaosuensis. (ak) Dericorys annulata. (a,b) Male habitus. (c,d) Female habitus. (ei) Epiphallus in dorsal, dorsofrontal, frontal, dorsolateral and lateral views. (j,k) Phallic complex in dorsal and lateral views. (lu) Conophyma zhaosuensis. (l,m) Male habitus. (n, o) Female habitus. (pr) Epiphallus in dorsal, frontal and dorsolateral views. (su) Phallic complex in dorsal and lateral views. The blue arrows indicate the pseudoarch.
Figure 6. Morphological characters of Dericorys annulata and Conophyma zhaosuensis. (ak) Dericorys annulata. (a,b) Male habitus. (c,d) Female habitus. (ei) Epiphallus in dorsal, dorsofrontal, frontal, dorsolateral and lateral views. (j,k) Phallic complex in dorsal and lateral views. (lu) Conophyma zhaosuensis. (l,m) Male habitus. (n, o) Female habitus. (pr) Epiphallus in dorsal, frontal and dorsolateral views. (su) Phallic complex in dorsal and lateral views. The blue arrows indicate the pseudoarch.
Insects 14 00085 g006
Figure 7. Morphological characters of Conophymacris szechwanensis and Xiangelilacris zhongdianensis. (ak) Conophymacris szechwanensis. (a,b) Male habitus. (c,d) Female habitus. (eh) Epiphallus in dorsal, dorsofrontal, frontal and dorsolateral lateral views. (ik) Phallic complex in dorsal and lateral views. (ls) Types of Xiangelilacris zhongdianensis. (l,m) Holotype male habitus. (n) Labels of the holotype male (the chinese character sting below the scientific name Xiangelilacris zhongdianensis in the left label is the chinese name for this species, and that in the upper right label is the collecting data of the holotype male: 5–6 August 2007; Xiangelila, Yunnan; Yuan Huang, Zhijun Zhou leg.). (o,p) Paratype female habitus. (q) Labels of the paratype female (the chinese character sting below the scientific name Xiangelilacris zhongdianensis in the left label is the chinese name for this species, and that in the upper right label is the collecting data of the paratype female, same as that of the holotype male). (r) Wing bud of the holotype male. (s) Wing bud of the paratype female. The blue arrows indicate the arch.
Figure 7. Morphological characters of Conophymacris szechwanensis and Xiangelilacris zhongdianensis. (ak) Conophymacris szechwanensis. (a,b) Male habitus. (c,d) Female habitus. (eh) Epiphallus in dorsal, dorsofrontal, frontal and dorsolateral lateral views. (ik) Phallic complex in dorsal and lateral views. (ls) Types of Xiangelilacris zhongdianensis. (l,m) Holotype male habitus. (n) Labels of the holotype male (the chinese character sting below the scientific name Xiangelilacris zhongdianensis in the left label is the chinese name for this species, and that in the upper right label is the collecting data of the holotype male: 5–6 August 2007; Xiangelila, Yunnan; Yuan Huang, Zhijun Zhou leg.). (o,p) Paratype female habitus. (q) Labels of the paratype female (the chinese character sting below the scientific name Xiangelilacris zhongdianensis in the left label is the chinese name for this species, and that in the upper right label is the collecting data of the paratype female, same as that of the holotype male). (r) Wing bud of the holotype male. (s) Wing bud of the paratype female. The blue arrows indicate the arch.
Insects 14 00085 g007
Figure 8. Morphological characters of Menglacris maculata and Ranacris albicornis. (ak) Menglacris maculata. (a,b) Male habitus. (c,d) Female habitus. (eh) Epiphallus in dorsal, dorsofrontal, frontal and dorsolateral views. (ik) Phallic complex in dorsal and lateral views. (lu) Ranacris albicornis. (l,m) Male habitus. (n,o) Female habitus. (ps) Epiphallus in dorsal, dorsofrontal, frontal and dorsolateral views. (tu) Phallic complex in dorsal and lateral views.
Figure 8. Morphological characters of Menglacris maculata and Ranacris albicornis. (ak) Menglacris maculata. (a,b) Male habitus. (c,d) Female habitus. (eh) Epiphallus in dorsal, dorsofrontal, frontal and dorsolateral views. (ik) Phallic complex in dorsal and lateral views. (lu) Ranacris albicornis. (l,m) Male habitus. (n,o) Female habitus. (ps) Epiphallus in dorsal, dorsofrontal, frontal and dorsolateral views. (tu) Phallic complex in dorsal and lateral views.
Insects 14 00085 g008
Figure 9. Morphological characters of Longzhouacris mirabilis. (a,b) Male habitus. (c,d) Female habitus. (eh) Epiphallus in dorsal, dorsofrontal, frontal and dorsolateral views. (i,j) Phallic complex in dorsal and lateral views.
Figure 9. Morphological characters of Longzhouacris mirabilis. (a,b) Male habitus. (c,d) Female habitus. (eh) Epiphallus in dorsal, dorsofrontal, frontal and dorsolateral views. (i,j) Phallic complex in dorsal and lateral views.
Insects 14 00085 g009
Table 1. Accession numbers and references of the mitogenomes of the species sampled in this study.
Table 1. Accession numbers and references of the mitogenomes of the species sampled in this study.
SpeciesAccession NumberReference
Caelifera, Acridoidea, Acrididae
Calliptaminae
Calliptamus abbreviatesNC_030626Han et al. (2016) [45]
Calliptamus barbarousNC_046544Chang et al. (2020) [24]
Calliptamus italicusNC_011305Fenn et al. (2008) [11]
Peripolus nepalensisNC_029135Zhi et al. (2016) [46]
Catantopinae
Diabolocatantops pinguisMT916719Zeng et al. (2021) [29]
Ranacris albicornisON943039 This study (mt1826)
Stenocatantops mistshenkoiNC_052717Chen et al. (2020) [4]
MT916714Zeng et al. (2021) [29]
Stenocatantops splendensMN083191Chang et al. (2020) [24]
MT916715Zeng et al. (2021) [29]
NC_041115Li et al. (2019) [20]
Traulia lofaoshanaNC_046551 Chang et al. (2020) [24]
Traulia minutaNC_036063Qiu et al. (2021) [47]
Traulia nigritibialisNC_041114Li et al. (2019) [20]
Traulia orchotibialisNC_046565 Chang et al. (2020) [24]
Traulia szetschuanensisNC_013826Direct Submission
Xenocatantops brachycerusMT916716Zeng et al. (2021) [29]
NC_021609Yang J. et al. (2016) [48]
Coptancrinae
Apalacris nigrogeniculataNC_046527Chang et al. (2020) [24]
Eucoptacra sp.MG993445Song et al. (2018) [17]
Cyrtacanthacridinae
Austracris guttulosaMG993415Song et al. (2018) [17]
Chondracris roseaNC_019993Direct Submission
Cyrtacanthacris tataricaMG993444Song et al. (2018) [17]
Patanga japonicaNC_036062Direct Submission
Schistocerca gregaria gregariaNC_013240Erler et al. (2010) [12]
Eyprepocnemidinae
Choroedocus capensisMN046212Chang et al. (2020) [24]
Choroedocus violaceipesMK903559Chang et al. (2020) [24]
Shirakiacris shirakiiNC_021610Direct Submission
Shirakiacris yunkweiensisNC_046531 Chang et al. (2020) [24]
Habrocneminae
Longzhouacris mirabilisON931612This study (mt1825)
Menglacris maculataMK903568Chang et al. (2020) [25]
Hemiacridinae
Hieroglyphus annulicornisMK903564Chang et al. (2020) [25]
Hieroglyphus tonkinensisNC_030587Chang and Huang (2016) [49]
Leptacris sp.MG993429Song et al. (2018) [17]
Melanoplinae
Alulacris shilinensisMW810985Xu et al. (2021) [27]
Anapodisma miramaeNC_052715Chen et al. (2020) [4]
Curvipennis wixiensisNC_031397Chen and Xu (2017) [50]
Emeiacris maculateNC_046556Chang et al. (2020) [24]
Fruhstorferiola huayinensisNC_031379Liu and Qiu (2016) [51]
Fruhstorferiola kulingaNC_026716Yang R. et al. (2016) [52]
Fruhstorferiola omeiNC_046545Chang et al. (2020) [24]
Fruhstorferiola sp.KU355786Direct submission
Fruhstorferiola tonkinensisNC_031817Zhang and Lin (2016) [53]
Indopodisma kingdoniNC_046529Chang et al. (2020) [24]
Kingdonella bicollinaNC_023920Zhi et al. (2016) [54]
Kingdonella pienbaensisMK903565Chang et al. (2020) [25]
Melanoplus bivittatusMG993426Song et al. (2018) [17]
Melanoplus differentialisNC_057646Direct Submission
Ognevia longipennisNC_013701Direct Submission
Paratonkinacris vittifemoralisNC_046530Chang et al. (2020) [24]
Pedopodisma emeiensisNC_046561Chang et al. (2020) [24]
Pedopodisma funiushanaNC_046546Chang et al. (2020) [24]
Pedopodisma tsinlingensisKX857635Qiu et al. (2020) [26]
NC_032303Direct Submission
Pedopodisma wudangshanensisNC_046547Chang et al. (2020) [24]
Prumna arcticaNC_013835Sun et al. (2010) [55]
Qinlingacris elaeodesKM363599Li et al. (2016) [56]
Qinlingacris taibaiensisNC_027187Direct Submission
Rhinopodisma eminifrontusMK903556Chang et al. (2020) [25]
Sinopodisma houshanaNC_033905Qiu et al. (2020) [26]
Sinopodisma lofaoshanaNC_046562Chang et al. (2020) [24]
Sinopodisma lushiensisNC_046549Chang et al. (2020) [24]
Sinopodisma pieliKX857633Qiu et al. (2020) [26]
NC_051867Liu et al. (2017) [18]
Sinopodisma qinlingensisNC_056238Qiu et al. (2020) [26]
Sinopodisma rostellocercaNC_052716Chen et al. (2020) [4]
Sinopodisma wulingshanensisNC_033906Qiu et al. (2020) [26]
Tonkinacris sinensisNC_032716Zhang et al. (2017) [57]
Xiangelilacris zhongdianensisNC_046533Chang et al. (2020) [24]
Yunnanacris wenshanensisKX296781Direct Submission
Yunnanacris yunnaneusNC_030586 Hu et al. (2016) [58]
Zubovskya koeppeniMK903579Chang et al. (2020) [25]
Oxyinae
Caryanda elegansNC_036750Yuan et al. (2019) [59]
Caryanda xinpingensisNC_030165Hu et al. (2017) [60]
Caryandoides hunanicaNC_053659Zeng et al. (2021) [29]
Fer nigripennisNC_053658Zeng et al. (2021) [29]
Longchuanacris curvifurculusNC_036994Hu et al. (2018) [61]
Oxya adentataMK903571Chang et al. (2020) [25]
Oxya agavisaNC_045883Li et al. (2020) [22]
Oxya chinensisNC_010219Zhang and Huang (2008) [62]
Oxya hainanensisMN083185Chang et al. (2020) [24]
NC_045928Li et al. (2020) [22]
Oxya hylaNC_032076Song et al. (2016) [63]
Oxya hyla intricateKP313875Dong et al. (2016) [64]
Oxya japonicaNC_043773Li et al. (2020) [22]
Oxytauchira brachypteraMT916721Zeng et al. (2021) [29]
NC_046570Chang et al. (2020a) [24]
Oxytauchira flangeNC_053745Zeng et al. (2021) [29]
Paratoacris reticulipennisNC_053660Zeng et al. (2021) [29]
Pseudoxya diminutaNC_025765 Tang et al. (2014) [65]
Spathosterninae
Spathosternum nigrotaeniatumMG993439Song et al. (2018) [17]
Spathosternum prasiniferum sinenseKM588074Zhou et al. (2016) [66]
Spathosternum prasiniferum prasiniferumNC_046532Chang et al. (2020) [24]
Dericorythidae
Conophyminae
Conophyma zhaosuensisON943040This study (mt1938)
Conophymacris viridisNC_046528Chang et al. (2020) [24]
Dericorythinae
Dericorys annulataNC_046555Chang et al. (2020) [24]
Pamphagidae
Pamphaginae
Filchnerella kukunorisMK903590Chang et al. (2020) [25]
Filchnerella yongdengensisMK903560Chang et al. (2020) [25]
Table 2. Initiation and termination codons of protein coding genes (PCGs) of the newly sequenced complete mitogenomes.
Table 2. Initiation and termination codons of protein coding genes (PCGs) of the newly sequenced complete mitogenomes.
PCGsInitiation CodonsTermination Codons
mt1825mt1826mt1938mt1825mt1826mt1938
ND2ATGATGATGTAATAAT
COX1ACCACCCAATTT
COX2ATGATGATGTAATAATAA
ATP8ATTATTATCTAATAATAA
APT6GTGATGATGTAATAATAA
COX3ATGATGATGTAATAATAA
ND3ATGATTATTTAATAATAA
ND5ATTATTATTTAATAAT
ND4ATGATGATGTTAAT
ND4LATGATGATGTAATAATAA
ND6ATGATGATGTAATAATA
CYTBATGATGATGTATAATAA
ND1ATAATAATGTAGTAGTAG
Note: mt1825: Longzhouacris mirabilis; mt1826: Ranacris albicornis; mt1938: Conophyma zhaosuensis.
Table 3. Total numbers of different types of base mismatches in tRNAs of the three newly sequenced mitogenomes.
Table 3. Total numbers of different types of base mismatches in tRNAs of the three newly sequenced mitogenomes.
SpeciesA–AA–GA–CG–UC–UU–U
mt18251(trnD)1(trnW)02804(trnQ,trnD,trnH,trnS2)
mt18261(trnD)1(trnW)02204(trnQ,trnE)
mt19381(trnD)1(trnW)01603(trnI, trnQ, trnH)
Table 4. Distribution of G–U base mismatches in tRNAs of the three newly sequenced mitogenome.
Table 4. Distribution of G–U base mismatches in tRNAs of the three newly sequenced mitogenome.
Transfer RNAmt1825mt1826mt1938Transfer RNAmt1825mt1826mt1938
trnI100trnR111
trnQ111trnN000
trnM000trnS1112
trnW000trnE101
trnC210trnF522
trnY222trnH130
trnL2110trnT110
trnD000trnP222
trnK110trnS2100
trnG112trnL1110
trnA322trnV221
Disclaimer/Publisher’s Note: The statements, opinions and data contained in all publications are solely those of the individual author(s) and contributor(s) and not of MDPI and/or the editor(s). MDPI and/or the editor(s) disclaim responsibility for any injury to people or property resulting from any ideas, methods, instructions or products referred to in the content.

Share and Cite

MDPI and ACS Style

Zhang, C.; Mao, B.; Wang, H.; Dai, L.; Huang, Y.; Chen, Z.; Huang, J. The Complete Mitogenomes of Three Grasshopper Species with Special Notes on the Phylogenetic Positions of Some Related Genera. Insects 2023, 14, 85. https://doi.org/10.3390/insects14010085

AMA Style

Zhang C, Mao B, Wang H, Dai L, Huang Y, Chen Z, Huang J. The Complete Mitogenomes of Three Grasshopper Species with Special Notes on the Phylogenetic Positions of Some Related Genera. Insects. 2023; 14(1):85. https://doi.org/10.3390/insects14010085

Chicago/Turabian Style

Zhang, Chulin, Benyong Mao, Hanqiang Wang, Li Dai, Yuan Huang, Zhilin Chen, and Jianhua Huang. 2023. "The Complete Mitogenomes of Three Grasshopper Species with Special Notes on the Phylogenetic Positions of Some Related Genera" Insects 14, no. 1: 85. https://doi.org/10.3390/insects14010085

Note that from the first issue of 2016, this journal uses article numbers instead of page numbers. See further details here.

Article Metrics

Back to TopTop