Two new species of Erythroneurini (Hemiptera, Cicadellidae, Typhlocybinae) from southern China based on morphology and complete mitogenomes

View article
Zoological Science

Introduction

The tribe Erythroneurini (Young, 1952) is the largest tribe leafhopper subfamily Typhlocybinae, (Hemiptera, Cicadellidae) comprising 209 genera and 2,027 described species (Dmitriev et al., 2022) and is widely distributed in all major zoogeographic regions of the world (Chen et al., 2021). As plant sap sucking insects they can damage fruit trees and vegetables, and their small size makes them difficult to detect and identify (Ghauri, 1967; Ghauri, 1974; Ghauri, 1975; Knight, 1974). Damage to plants is by egg laying and as virus vectors of plant pathogens (Womack & Schuster, 1986; Bellota, 2011; Bosco et al., 2007). Moreover, erythroneurine species have adopted to various habitats and plants such as trees, rocks, grasslands, sandy substrates, and bushy areas, etc (Morris, 1971; Roddee, Kobori & Hanboonsong, 2018). The species of Erythroneurini have been divided and classified by researchers in different ways, as a result of high morphological diversity and wide geographical distributions.

China is the country with the most widely distributed, fully developed and most complete types of karst landforms in the world, which are mainly concentrated in carbonate outcropping areas, of which Guangxi, Guizhou, and eastern Yunnan account for the largest area (Xiong et al., 2008). Guizhou is an important part of the Yunnan-Guizhou Plateau and is believed to be the most well-developed representative of karst areas. The terrain is violently undulating, the types of landforms are diverse, and the composition of surface material and soil types is complex. Additionally, the climate of this area is warm and humid, with small annual temperature changes, warm in winter and cool in summer (Zhu et al., 2020), resulting in high biodiversity (plants, insects, birds, snails and bats), except that singular and goodliness natural landscape (Luo et al., 2016; Zhu, 2007). Many new species of Erythroneurini were discovered in karst regions from Guizhou (Chen et al., 2020; Song, Li & Xiong, 2011; Zhang, Song & Song, 2021).

Phylogenetic relationships of major lineages of Cicadellidae have been researched for many of years (Skinner et al., 2019; Wang et al., 2020a; Wang et al., 2020b; Lu et al., 2021; Yan et al., 2022; Hu et al., 2023; Cao et al., 2023). More recently, diverse markers have been applied to perform phylogenetic inferences of Hemiptera, which consist of shape characteristic, mitochondrial genes, nuclear genes and a combination of them, together with transcriptomes on the basis of next-generation sequencing (Almeida et al., 2009; Wang et al., 2010; Yao et al., 2021). With the purpose of confirming the results of traditional classification of Eurythroneurini we also use molecular markers. Based on morphological and molecular data, Erythroneurini has been divided into 209 genera (Dmitriev et al., 2022). However, in most instances, short time intervals between speciation events generated incongruous divergence in morphological features and molecular markers (Chen, Li & Song, 2021).

The relationships among the multiple species of Cicadellidae were established by means of morphological characters, and a few nuclear genes and mitochondrial sequences (Longo et al., 2017; Jiang et al., 2021). However, despite the tendency to expand genome coverage, the number of specimens that can be collected is relatively limited, and existing species were chosen to conduct genetic sequencing, as it is impossible to establish a relatively complete molecular identification of the family. Therefore, in our work, the mitochondrial genomes of two new species and two known species (Mitjaevia bifurcata, Mitjaevia diana) were picked via Sequencing Technology to provide a comprehensive comparative analysis of mitochondrial gene structure. We propose a hypothesis that phylogenetic trees based on mitochondrial genomes can better validate the accuracy of traditional classification. Phylogenetic trees based on the mitochondrial genomes of A. (A.) rongchangensis sp. nov., T. (T.) jiulongensis sp. nov., M. bifurcata and M. diana and another 24 species were built adopting the Bayesian inference and maximum likelihood methods. This research will enrich the mitochondrial gene bases of the erythroneurine leafhoppers and improve the accuracy of the traditional classification.

Material and Methods

Leafhopper collections and species identification based on the morphology

The species of leafhopper are collected according to Table 1. The specimens were preserved in absolute ethanol. Images of the appearance and genitalia of species were taken by a KEYENCE VHX-5000 digital microscope. Male/female specimens were identified under a stereoscope, and the whole abdomen of the specimens was separated and moistened in a hot 10% NaOH solution. Afterward, the abdomen was washed with ordinary water, blotted up with qualitative filter paper, and transferred to a clean glass slide with a drop of glycerin. Genital dissections were dissected in glycerin to inhibit parts from drying out. Then, they were viewed and plotted by way of Olympus SZX16 and BX53 microscopes. The remaining specimen was stored in 95% ethanol and put in a refrigerator at −20 °C. The analyzed specimens were examined using Olympus SZX16 dissecting microscope and Olympus BX53 stereoscopic microscopes respectively and identified by Prof. Yuehua Song. All specimens inspected are reserved in the School of Karst Science, Guizhou Normal University, China (GZNU).

Table 1:
Study sites and dates for leafhopper sample collection in this study.
Species Locality Collector Latitude Longitude Date
A. (Arboridia) rongchangensis sp. nov. Rongchang, Chongqing Guimei Luo 29°25′43″N 105°39′21″E 14 Aug 2021
T. (Thaia) jiulongensis sp. nov. Jiulongpo, Chongqing Weiwen Tan 29°28′36″N 106°25′11″E 14 Aug 2021
M. bifurcata Bijie, Guizhou Zhouwei Yuan 27°14′51″N 105°5′52″E 27 May 2019
M. diana Huajiang, Guizhou Zhouwei Yuan 25°41′36″N 105°37′46″E 29 May 2019
DOI: 10.7717/peerj.16853/table-1

DNA extraction, mitogenome sequencing and assembly

Extraction of DNA originated from the whole body removing the abdomen and wings. The bodies were incubated at 56 °C for 6 h for complete lysis and total genomic DNA was eluted in 50 µL double-distilled water (ddH2O), and the remaining other steps were performed according to the manufacturer’s protocol. Genomic DNA was stored at −20 °C. The whole mitochondrial genomes of A. (A.) rongchangensis sp. nov. and T. (T.) jiulongensis sp. nov. were sequenced at Berry Genomics (Beijing, China) by an Illumina Novaseq 6000 platform (Illumina, Alameda, CA, USA) using 150 bp paired-end reads. Firstly, the obtained sequence reads were filtered following Zhou et al. (2013), the remaining high-quality reads were assembled by an iterative De Bruijin graph de novo assembler, the IDBA-UD toolkit, with a similarity threshold of 98%, and k values of 40 and 160 bp (Peng et al., 2012). The mitogenome was initially assembled by Geneious Prime v 2021.1.1, and then manually proofread based on sequencing peak figures.

The complete mitochondrial genomes of M. Bifurcata and M. diana were sequenced at Bio-Transduction Lab Co.Ltd. (Wuhan, China) by Sanger sequencing. PCR primers were designed according to conserved region sequences and used to amplify the mitochondrial DNA sequence in PCR reactions (Tables 2 and 3). The PCR reaction was performed using the LA Taq polymerase. The thermal cycling conditions comprised an initial denaturation step at 94 °C for 2 min, then 35 cycles of denaturation at 94 °C for 30 s, 30 s for annealing at 55 °C, and elongation at 72 °C for 1 min/kb, followed by the final extension at 72 °C for 10 min. The PCR products were purified and sequenced using an ABI 3730 automatic sequencer. After quality-proofing of the obtained DNA fragments, and BLASTed were used to confirm that the amplification is the actual target sequence (Meng et al., 2013; Yu, Wu & Han, 2017). The complete mitogenome sequence was assembled manually through DNAStar v7.1 (Burland, 2000).

Genome annotation and analyses

First of all, raw mitogenomic sequences were entered into MITOS web servers (http://mitos.bioinf.uni-leipzig.de/index.py, accessed on 15 Jun 2021) in an effort to fix the rough boundaries of genes. Accurate locations of protein-coding genes (PCGs) were determined by seeking ORFs (employing genetic code 5, the invertebrate mitochondrion). All tRNAs were characteried by using tRNAscan SE v. 1.21 and ARWEN (Lowe & Eddy, 1997; Laslett & Canbck, 2008). The precise boundaries of rrnL and rrnS were defined by homologous comparison. Genomes manually annotated were parsed and extracted by means of PhyloSuite, and GenBank (NCBI) submission files and organization tables for mitogenomes were also created through the same software (Zhang et al., 2020).

Table 2:
Primers used for amplification of the mitochondrial genome of M. bifurcata.
Fragment no. Gene or region Sequence (5′–3′) Length (bp)
F1 tRNA-Met-COX1 GCTAACTTAAGCTATTAGGTTC 1,720
CGTATGTTAATTACTGTTGTG
F2 COX1 CTGGTTGAACAGTTTACCC 598
CATCTAAAAACCTTAATACC
F3 COX1-ATP6 GAGTCATTTGGTTATATTGG 1,949
GAAATTTCTCCTTGAAGAGA
F4 ATP6 CAGTTTTTGATCCTTGTACTG 473
GCCTGCAATTATGTTAGCAG
F5 ATP6-COX3 GACATTTAGTACCTGTTGGTACG 1,206
CTCAAATCCTACATGATGCC
F6 COX3-ND5 CAGGTGTTTCTATTACATGAG 2,428
CGTTTAGGGGATATTGGTCTG
F7 ND5 TGCAGTTACCAGGGTTGAAG 328
GTTAGGTTGAGATGGCTTGG
F8 ND5-ND4 CCAATATCCCCTAAACGGTTAG 1,067
GTTTACTACAAGGAGATGTA
F9 ND4 CTGAAGAACATAACCCATGAG 330
GATTACCAAAAGCGCATGTTC
F10 ND4-12S GTGAATACCAAACATAACTG 5,295
AAGCAGACATGTGTTACT
F11 12S CCAGTACAATTACTTTGTTACG 385
CTTTAACATTAATAGTTTATTTTC
F12 12S-ND2 CAATTAAGATACAGGTTCCC 3,235
GAGTGCAAAAGAGGCAGGAATG
DOI: 10.7717/peerj.16853/table-2
Table 3:
Primers used for amplification of the mitochondrial genome of M. diana.
Fragment no. Gene or region Primer name Sequence (5′–3′) Length (bp)
F1 tRNA-Met-COX1 D2F1 GCTAATTTAAGCTATTAGGTTC 2,197
D2R1 GTGACTCCATGTATTGTAGC
F2 COX1-ATP6 D2F2 GGTTTGTTGTTTGGGCTCATC 1,927
D2R2 AGTTGGATACCCCTGTAAGG
F3 ATP6 D2F3 TGTTTTCAGTATTTGACCCTTG 481
D2R3 TGCCCTGCAATTATATTAGC
F4 ATP6-COX3 D2F4 GACATTTAGTTCCGGTAGGA 957
D2R4 GAAGGTTATACATTCGAATCC
F5 COX3 D2F5 TAGCAACAGGATTTCATGGA 142
D2R5 TCTACAAAGTGTCAATACCAAG
F6 COX3-ND5 D2F6 TAGTATCTGGGATTCGAATG 2,159
D2R6 ATGTCTTTTGGTAGTTGAC
F7 ND5 D2F7 GGAAGAATGAACTAGAGATG 314
D2R7 TGCTGGGTTGAGATGGTTTAG
F8 ND5-ND4 D2F8 CCTAAACGATTAGTTAAGCAAG 1,103
D2R8 GGTATTCATTAAACTTAGTAGG
F9 ND4 D2F9 CCAGATGAACATAAACCGTGAG 326
D2R9 CCAAAAGCTCATGTTCAAGC
F10 ND4-CYTB D2F10 CAAAGATACTTATAACTCGG 2,222
D2R10 CTGTGATGTGTAGAAAGAAG
F11 CYTB D2F11 GTAATCACTAATTTACTATCTGC 383
D2R11 CATTCTGGTTGAATATGAATC
F12 CYTB-16S D2F12 GATTTACTGGGAATTGTAATTAC 1,773
D2R12 GTTACCTTAGGGATAACAGC
F13 16S D1F13 CACCGATTTGAACTCAAATC 987
D1R13 GGTTTTGTACCTTTTGTATTAGG
F14 16S-12S D1F14 GTAAAGATTATCCCTTAC 639
D1R14 GTTAGGTCAAGGTGCAGT
F15 12S D1F15 CTTTGTTACGACTTATCTC 419
D1R15 TTAGGATTAGATACCCTAT
F16 12S-ND2 D1F16 GTGGTTTATCAATTAAGAAAC 2,976
D1R16 GCTTAATTCCAAGCCACACC
DOI: 10.7717/peerj.16853/table-3

The mitogenomic circular map was generated by OrganellarGenomeDRAW (OGDRAW) version 1.3.1 (https://chlorobox.mpimp-golm.mpg.de/OGDraw.html, accessed on 3 March 2023) (Greiner, Lehwark & Bock, 2019). Intergenic spacers and overlapping regions between genes were performed manually. The nucleotide base composition, codon usage, as well as values of A + T content were calculated with MEGA 11.0 (Tamura, Stecher & Kumar, 2021). The bias of nucleotide composition was computed according to AT skew = [A − T]/[A + T] and GC skew = [G − C]/[G + C] (Perna & Kocher, 1995). Additionally, the nucleotide diversity (Pi) and nonsynonymous (Ka)/synonymous (Ks) mutation rate ratios were operated by DNAsp 6.0 (Rozas et al., 2017).

Phylogenetic analysis

A molecular phylogenetic analysis was constructed on the basis of mitogenomes of 28 species and two species regarded as outgroups (Table 4). All complete mitochondrial sequences were selected to accomplish phylogenetic analyses. The Gblocks version 0.91b was adopted to clean out the gaps and fuzzy-alignment sites, and all alignments were verified and revised in MEGA 11.0 prior to phylogenetic analysis (Tamura, Stecher & Kumar, 2021). The phylogenetic trees were constructed by introducing two methods both the maximum likelihood (ML) method and the Bayesian Inference (BI) method (Nguyen et al., 2015; Zhou et al., 2011). The ML analysis was performed with IQ-TREE under a ML + rapid bootstrap (BS) algorithm with 10,000 replicates used to calculate bootstrap scores for each node (BP). The BI analysis was carried out using MrBayes 3.2.7 elected GTR + G + I as the optimal model, running 10 million generations, sampling every 1000 trees, 25% of samples were abandoned as burn-in.

Table 4:
List of mitochondrial genomes analyzed in the present.
Subfamily/Tribe Species Length (bp) Accession number Reference
Typhlocybinae/ Typhlocybini Eurhadina acapitata 15,419 MZ457331.1 Direct submission
Eurhadina jarrayi 15,332 MZ014455.1 Lin, Huang & Zhang (2021)
Eurhadina dongwolensis 15,708 MZ457332.1 Direct submission
Eurhadina fusca 15,302 MZ983367.1 Direct submission
Agnesiella kamala 15,209 MZ457327.1 Direct submission
Agnesiella roxana 15,901 MZ457328.1 Direct submission
Eupteryx adspersa 15,178 MZ014454.1 Lin, Huang & Zhang (2021)
Eupteryx minuscula 16,944 MN910279.1 Yang et al. (2020)
Eupteryx gracilirama 17,173 MT594485.1 Yuan et al. (2021a) and Yuan et al. (2021b)
Typhlocybinae/ Erythroneurini Limassolla emmrichi 14,677 MW272458.1 Yan et al. (2022)
Limassolla lingchuanensis 15,716 NC_046037.1 Yuan, Li & Song (2020)
Limassolla sp. 17,053 MT683892.1 Zhou, Dietrich & Huang (2020)
Mitjaevia bifurcata 16,589 OK448488.1 Direct submission
Mitjaevia protuberanta 15,472 NC_047465.1 Yuan, Li & Song (2020)
Mitjaevia diana 16,183 OK448489.1 Direct submission
Mitjaevia dworakowskae 16,399 MT981880.1 Chen, Li & Song (2021)
Mitjaevia shibingensis 15,788 MT981879.1 Chen, Li & Song (2021)
Arboridia (Arboridia) rongchangensis sp. nov. 15,596 OQ404948.1 Direct submission
Thaia (Thaia) jiulongensis sp. nov. 15,676 OQ630475.1 Direct submission
Elbelus tripunctatus 15,308 MZ014452.1 Lin, Huang & Zhang (2021)
Empoascanara sipra 14,827 NC_048516.1 Tan et al. (2020)
Empoascanara wengangensis 14,830 MT445764.1 Chen et al. (2021)
Empoascanara dwalata 15,271 MT350235.1 Chen et al. (2016)
Empoascanara gracilis 14,627 MT576649.1 Chen et al. (2021)
Typhlocybinae/ Empoascini Empoasca onukii 15,167 NC_037210.1 Song, Zhang & Zhao (2019)
Empoasca vitis 15,154 NC_024838.1 Zhou et al. (2016)
Empoasca flavescens 15,152 MK211224.1 Luo et al. (2019)
Empoasca serrata 15,131 MZ014453.1 Lin, Huang & Zhang (2021)
Outgroups Bothrogonia ferruginea 15,262 KU167550.1 Yu et al. (2019)
Iassus dorsalis 15,176 NC_046066.1 Wang et al. (2020b)
DOI: 10.7717/peerj.16853/table-4

Nomenclature

The electronic version of this article in Portable Document Format (PDF) will represent a published work according to the International Commission on Zoological Nomenclature (ICZN), and hence the new names contained in the electronic version are effectively published under that Code from the electronic edition alone. This published work and the nomenclatural acts it contains have been registered in ZooBank, the online registration system for the ICZN. The ZooBank LSIDs (Life Science Identifiers) can be resolved and the associated information viewed through any standard web browser by appending the LSID to the prefix http://zoobank.org/. The LSID for this publication is: http://zoobank.org/urn:lsid:zoobank.org:pub:2E7A89DA-21F4-41D1-B9DF-A33BCD5F3086; http://zoobank.org/urn:lsid:zoobank.org:pub:7710E194-8D88-413B-B3D2-7AF0B0777BE7. The online version of this work is archived and available from the following digital repositories: PeerJ, PubMed Central SCIE and CLOCKSS.

Results and Discussion

Taxonomy based on morphology

Arboridia(Arboridia)rongchangensisZhang & Song, sp. nov. (Figs. 12)

Description.

Dorsum dark brownish (Figs. 1A and 1C). Color pattern brown. Vertex with a pair of dark preapical spots. Face yellowish white, frontoclypeus dark (Figs. 1B and 1D).

Head narrower than pronotum (Figs. 1A and 1C). Crown fore margin weakly produced medially. Face, with anteclypeus narrow and pale, and frontoclypeus dark (Figs. 1B and 1D). Pronotum wide, scutellum with lateral triangles (Figs. 1A and 1C). Forewings without spots or markings.

Male genitalia. Pygofer dorsal appendage simple, without branch, hook-like apically (Figs. 2E and 2F). Subgenital plate with two macrosetae on lateral surface, and row of peg-like setae from subbase to apex, and several microsetae scattered on apical portion (Fig. 2D). Style long and slender, with two points apically; preapical lobe obtuse and distinct (Fig. 2A). Aedeagus with a large lamellate process arising from base of aedeagal shaft ventrally; aedeagal shaft broad and flat, slightly bifurcated at apex; gonopore subapical on ventral surface; preatrium little longer than shaft (Figs. 2B and 2C). Connective V-shaped, with arms long (Fig. 2G).

Male abdominal apodemes small, not exceeding 3rd sternite (Fig. 2H).

Measurement. Male length 3.0∼3.1 mm, female length 3.1∼3.2 (including wings).

Specimen examined. Holotype: ♂, CHINA, Chongqing, Rongchang, 14 VIII 2021, coll. Guimei Luo. Paratypes: 11 ♂♂, 32 ♀♀, same data as holotype.

Remarks. This species is similar to Arboridia reniformis Song & Li (2013), but differs in having the aedeagal shaft distinctly bifurcate apically.

Etymology. The new species is named after its type locality: “Rongchang” county, Chongqing, China.

Thaia(Thaia)jiulongensisZhang & Song, sp. nov. (Figs. 34)

Description. Vertex yellow, light brownish in middle apically (Figs. 3A and 3C). Face yellowish brown (Figs. 3B and 3D). Pronotum orange brown. Scutellum with anterior margin yellow and posterior part milky yellow; lateral triangles brownish black (Figs. 3A and 3C).

Face with anteclypeus ovoid (Figs. 3B and 3D). Pronotum, with pair of large triangular impressions, posterior and anterior part lighter, yellowish (Figs. 3A and 3C).

Male abdominal apodeme small, not surpassing 3rd sternite (Fig. 4I).

Male genitalia. Pygofer lobe with scattered fine microsetae on dorsal surface, with dorso-caudal margin angulated (Fig. 4F). Anal tube with well-developed basal appendages, extending ventro-caudally (Fig. 4G). Subgenital plate broadened at subbase, provided with three macrosetae on lateral surface at midlength, numerous peg-like small setae along dorsal margin from near midlength part to apex; several small setae scattered apically (Figs. 4E and 4F). Style slender apically, preapical lobe well developed (Fig. 4A). Aedeagus expanded at base in ventral view, with pair of long basal process arising from preatrium ventrally, which slim and curved, tapering towards apex; gonopore apical on ventral surface (Figs. 4B, 4C and 4D). Connective V-shaped, without central lobe, lateral arms long and slim (Fig. 4H).

Specimen examined. Holotype: ♂, CHINA, Chongqing, Jiulongpo District, Zhongliang Yunling Forest Park, 14 VIII, 2021. coll. Weiwen Tan. Paratypes: 5 ♂♂, 15 ♀♀, same data as holotype.

Measurement. Body length ♂ 3.0∼3.2 mm; ♀ 3.0∼3.2 mm (including wings).

Remarks. This species is similar to Thaia (Thaia) barbata Dworakowska (1979), but can be distinguished from the latter species by from the shape of the adedeagus, with a small subapical protrusion and, the anal tube appendages, which are particularly wide compared to T. (T.) barbata Dworakowska (1979). Connective arms are relatively slender.

Arboridia (Arboridia) rongchangensisZhang & Song,sp. nov.

Figure 1: Arboridia (Arboridia) rongchangensisZhang & Song,sp. nov.

(A) Dorsal habitus. (B) Lateral habitus. (C) Head and thorax, dorsal view (D) Face. (E) Terminalia of female, ventral view. (F) Sternite VII of female, ventral view.
Arboridia (Arboridia) rongchangensis Zhang & Song, sp. nov.

Figure 2: Arboridia (Arboridia) rongchangensis Zhang & Song, sp. nov.

(A) Style. (B) Aedeagus, lateral view. (C) Aedeagus, ventral view. (D) Subgenital plate. (E) Pygofer lobe, lateral view. (F) Pygofer dorsal appendage, lateral view. (G) Connective. (H) Abdominal apodemes.
Thaia (Thaia) jiulongensisZhang & Song, sp. nov.

Figure 3: Thaia (Thaia) jiulongensisZhang & Song, sp. nov.

(A) Dorsal habitus. (B) Lateral habitus. (C) Head and thorax, dorsal view (D) Face. (E) Terminalia of female, ventral view. (F) Sternite VII of female, ventral view.
Thaia (Thaia) jiulongensisZhang & Song, sp. nov.

Figure 4: Thaia (Thaia) jiulongensisZhang & Song, sp. nov.

(A) Style. (B) Aedeagus, lateral view. (C) Aedeagus, lateral-ventral view. (D) Aedeagus, dorsal view. (E) Subgenital plate. (F) Pygofer lobe, lateral view. (G) Anal tube appendage, lateral view. (H) Connective. (I) Abdominal apodemes.

Etymology. The new species is named after its type locality, Jiulong, Chongqing.

Taxonomy based on molecular data

Organization and composition of the genome

The complete mitogenomes of A. (A.) rongchangensis sp. nov., T. (T.) jiulongensis sp. nov., M. bifurcata and M. diana are 15,596, 15676, 16,183 and 16,589 bp, respectively. Both species comprise 13 PCGs, 22 tRNA genes, two rRNA genes, and a control region (CR) (Fig. 5). Two strands, the majority strand (H-strand) and the minority strand (L-strand), exist in the mitochondrial genome. The H-strand consists of 23 genes (nine PCGs, 14 tRNAs) and CR, and meanwhile, the L-strand encompasses 14 genes (four PCGs, eight tRNAs, and two rRNAs).

Mitochondrial map of A. (A.) rongchangensis Zhang & Song, sp. nov., T. (T.) jiulongensis Zhang & Song, sp. nov., M. bifurcata and M. diana.

Figure 5: Mitochondrial map of A. (A.) rongchangensis Zhang & Song, sp. nov., T. (T.) jiulongensis Zhang & Song, sp. nov., M. bifurcata and M. diana.

There are 50 bp, 66 bp, and 70 bp intergenic spaces presented in total length of all the intergenic space ranging from one to 10 bp, one to 13 bp, one to nine bp in A. (A.) rongchangensis sp. nov., M. bifurcata, M. diana. However, 52 bp intergenic space existed in 11 regions from one to 15 bp in T. (T.) jiulongensis sp. nov. It can be observed that ten genes overlapped by 28 bp in A. (A.) rongchangensis sp. nov., eleven genes overlapped by 31 bp in T. (T.) jiulongensis sp. nov., ten genes overlapped by a grand total of 32 bp in M. bifurcata, nine genes overlapped by 40 bp M. diana (Table 5). The heavy AT nucleotide bias appears in the mitochondrial genomes in A. (A.) rongchangensis sp. nov., T. (T.) jiulongensis sp. nov., M. bifurcata and M. diana, the A + T contents are 80.7%, 78.0%, 78.4% and 78.5%, respectively (Table 5).

Protein-coding genes and codon usage

As with most other Typhlocybinae, the overall length of 13 PCGs of A. (A.) rongchangensis sp. nov., T. (T.) jiulongensis sp. nov., M. bifurcata and M. diana are 10,946, 10,968, 10,966 bp and 10,966 bp, 70.3%, 69.8%, 65.20% and 66.1% of the total genome of each species, respectively. In addition, nad2 and cox3 in four species have the same start codons and stop codons. The longest PCG is nad5 (1,675 bp) in A. (A.) rongchangensis sp. nov., the shortest is atp8 (144 bp) in M. bifurcata and M. diana. Only four genes (nad5, nad4, nad4L and nad1) are presented on the J-strand, and the remaining nine genes are presented on the H-strand.

The relative synonymous codon usage (RSCU) values of the 13 PCGs are generalized in Fig. 6. The codon usage analyses of A. (A.) rongchangensis sp. nov., T. (T.) jiulongensis sp. nov., M. bifurcata and M. diana revealed that codon UUA-Leu2 (214, 194, 180, 241), AUU-Ile (297, 249, 274, 210), AUA-Met (245, 202, 223, 180) AAU-Asn (273, 239, 236, 256), and AAA-Lys (290, 280, 227, 238) are the most frequently used. The highest RSCU value of each species occur in UUA-Leu2. The results showed that UUA is the most preferred codon. In addition, it can be seen from the RSCU values of the PCGs that AT is used more frequently than GC.

Transfer RNA and ribosomal RNA genes

The 22 tRNAs were deconcentrated between two regions, the rRNAs and the protein-coding region. The total tRNA lengths of A. (A.) rongchangensis sp. nov., T. (T.) jiulongensis sp. nov., M. bifurcata and M. diana are 1,434, 1,436, 1,441 and 1,455 bp, respectively, which range from 61 to 70 bp in A. (A.) rongchangensis sp. nov., 61 to 71 in T. (T.) jiulongensis sp. nov., 61 to 71 bp in M. bifurcata and 62 to 71 bp in M. diana (Table 5). The sequences of most tRNA genes demonstrated the exemplary clover-leaf secondary structure, including four structural domains and a short flexible loop: the acceptor stem, the dihydrouridine stem and loop (DHU), the anticodon stem (DHU) and loop, the thymidine stem and loop (T ψ C), and the variable (V) loop (Fig. S1), as observed in many other leafhoppers mitogenomes. However, the dihydrouridine (DHU) arm of trnS1 shapes an uncomplicated loop. Additionally, non-Waston-Crick base pairs were harbored the stems of the secondary structures, 21, 27, 21 and 18 weak G-U (or U-G) base pairs are revealed in the tRNAs of A. (A.) rongchangensis sp. nov., T. (T.) jiulongensis sp. nov., M. bifurcata and M. diana (Fig. S1). The location of mismatched base pairs in the acceptor arm, DHU arm, T ψ C arm and anticodon arm of tRNA from four species were shown in Table 6. These mismatches could be rectified by way of editing process, and the transport function is not influenced (Yuan et al., 2021a; Yuan et al., 2021b).

Table 5:
Organization of the A. (A.) rongchangensis Zhang & Song, sp. nov., T. (T.) jiulongensis Zhang & Song, sp. nov., M. bifurcata and M. diana mitochondrial genome.
Gene Position Intergenic Start codon Stop codon Strand
A. (Arboridia) rongchangensis sp. nov. T. (Thaia) jiulongensis sp. nov. M. bifurcata M. diana
trnI 1–63 1–64 1–64 1–63 0 0 0 0 H
trnQ 61–129 63–131 62–130 63–129 −3 −2 −3 −3 L
trnM 140–208 135–203 140–207 140–207 10 3 9 10 H
nad2 212–1180 204–1175 208–1179 208–1179 3 0 0 0 ATA ATA ATA ATA TAA TAA TAA TAA H
trnW 1179–1242 1174–1237 1178–1240 1178–1240 −2 −2 −2 −2 H
trnC 1235–1300 1237–1297 1233–1293 1233–1294 −8 −1 −8 −8 L
trnY 1306–1367 1298–1372 1294–1356 1295–1356 5 0 4 0 L
cox1 1369–2907 1366–2904 1361–2896 1361–2896 1 −7 0 4 ATT ATG ATG ATG TAA TAA TAA TAA H
trnL2 2911–2977 2907–2973 2898–2964 2898–2964 3 2 1 1 H
cox2 2978–3656 2974–3652 2965–3643 2965–3643 0 0 0 0 ATA TTG ATT ATT T T T T H
trnK 3657–3726 3653–3723 3644–3714 3644–3714 0 0 0 0 H
trnD 3727–3794 3724–3786 3719–3782 3723–3785 0 0 4 8 H
atp8 3794–3946 3786–3938 3792–3935 3795–3938 −1 −1 9 9 TTG TTG ATG ATG TAA TAA TAA TAA H
atp6 3943–4593 3932–4585 3929–4582 3932–4585 −4 −7 −7 −7 ATA ATG ATG ATG TAA TAA TAA TAA H
cox3 4594–5373 4586–5365 4583–5362 4586–5365 0 0 0 0 ATG ATG ATG ATG TAA TAA TAA TAA H
trnG 5380–5441 5366–5428 5367–5428 5371–5432 6 0 4 5 H
nad3 5442–5795 5429–5782 5429–5782 5433–5786 0 0 0 0 ATT ATT ATA ATA TAA TAA TAA TAA H
trnA 5797–5857 5797–5859 5796–5856 5792–5853 1 14 13 5 H
trnR 5857–5919 5864–5924 5856–5916 5853–5917 −1 4 −1 −1 H
trnN 5919–5983 5924–5990 5916–5979 5917–5982 −1 −1 −1 −1 H
trnS1 5983–6049 5990–6056 5979–6046 5982–6049 −1 −1 −1 −1 H
trnE 6052–6118 6060–6123 6054–6116 6059–6122 2 3 7 9 H
trnF 6120–6183 6139–6203 6123–6190 6128–6195 1 15 6 5 L
nad5 6184–7858 6203–7876 6193–7866 6197–7870 0 −1 2 1 ATT TTG TTG TTG T TAA TAA TAA L
trnH 7856–7917 7877–7945 7867–7931 7871–7935 −3 0 0 0 L
nad4 7918–9241 7949–9248 7931–9259 7935–9263 0 3 −1 −1 ATA ATG ATG ATG T T TAA TAA L
nad4L 9238–9516 9242–9517 9253–9531 9257–9535 −4 −7 −6 −6 ATG ATT ATG ATG TAG TAA TAA TAA L
trnT 9519–9584 9520–9583 9534–9597 9538–9601 2 2 2 2 H
trnP 9585–9649 9584–9651 9598–9661 9602–9665 0 0 0 0 L
nad6 9652–10137 9654–10139 9664–10149 9668–10153 2 2 2 2 ATG ATT ATT ATT TAA TAA TAA TAA H
cytb 10138–11274 10142–11278 10153–11289 10161–11297 0 2 3 7 ATG ATG ATG ATG TAA TAG TAG TAA H
trnS2 11279–11342 11278–11341 11288–11343 11300–11365 4 −1 −2 2 H
nad1 11333–12274 11344–12274 11344–12285 11356–12297 10 2 0 −10 ATT ATT ATT ATT TAA T TAA TAA L
trnL1 12275–12339 12275–12342 12286–12352 12298–12362 0 0 0 0 L
rrnL 12340–13586 12343–13537 12353–13538 12363–13548 0 0 0 0 L
trnV 13522–13586 13538–13599 13539–13605 13549–13618 0 0 0 0 L
rrnS 13587–14320 13600–14331 13606–14341 13619–14356 0 0 0 0 L
D-loop 14321–15596 14332–15676 14342–16813 14357–16589
DOI: 10.7717/peerj.16853/table-5
Relative synonymous codon usage (RSCU) in the mitogenomes of A. (A.) rongchangensis Zhang & Song, sp. nov., T. (T.) jiulongensis Zhang & Song, sp. nov., M. bifurcata and M. diana.

Figure 6: Relative synonymous codon usage (RSCU) in the mitogenomes of A. (A.) rongchangensis Zhang & Song, sp. nov., T. (T.) jiulongensis Zhang & Song, sp. nov., M. bifurcata and M. diana.

Table 6:
The location of mismatched base pairs (G-U or U-G) in tRNA from four species.
A. (A.) rongchangensis sp. nov. T. (T.) jiulongensis sp. nov. M. bifurcata M. diana
Acceptor arm trnY, trnR, trnP trnC, trnG, trnN, trnF, trnY, trnR, trnP trnA, trnC, trnP, trnV, trnY
DHU arm trnQ, trnC, trnY, trnG, trnR, trnF, trnH, trnS2, trnL1, trnV trnQ, trnC, trnY, trnG, trnF, trnH, trnP, trnV trnQ, trnC, trnY, trnG, trnR, trnF, trnH, trnS2, trnL1, trnV trnC, trnE, trnF, trnG, trnH, trnP, trnQ, trnS1, trnV
T ψ C arm trnW, trnA, trnR, trnS1, trnS2 trnR, trnS1, trnT, trnP, trnS2 trnW, trnA, trnR, trnS1, trnS2 trnA, trnP
Anticodon arm trnQ, trnL2, trnH trnC, trnL2, trnH, trnS2 trnQ, trnL2, trnH trnH, trnL2
DOI: 10.7717/peerj.16853/table-6

Control region

The control region, also known as the A + T region, acts a crucial part in the size variation of mitogenomes. The largest non-coding regions of the two species, putative control regions, were placed between rrnS and trnI. The control region in length of A. (A.) rongchangensis sp. nov., T. (T.) jiulongensis sp. nov., M. bifurcata and M. diana are 1,276 bp, 1,345 bp, 2,472 bp and 2,233 bp, the AT contents are 99.0%, 97.9%, 89.9% and 92.0%, respectively.

Phylogenetic analysis

In this study, complete mitochondrial genomes from 28 Typhlocybine species were collected as a dataset to establish phylogenetic trees by BI and ML methods, Bothrogonia ferruginea and Iassus dorsalis were regarded as outgroups. The GenBank accession numbers of all selected species used in this study were listed in Table 4. The phylogenetic topologies constructed by the two methods were completely consistent (Fig. 7). The monophyly of each tribe was generally well supported in the subfamily Typhlocybinae, which is consistent with the findings of some previous molecular phylogenetic studies (Chen et al., 2021; Chen, Li & Song, 2021). Twelve species of Typhlocybini, twelve species of Erythroneurini, and four species of Empoascini are clustered together, respectively, and all phylogenetic relationships demonstrated higher nodal support in both ML and BI analyses. All species from Typhlocybinae (inner group) are clustered together, all Mitjaevia species are gathered together. Our results further confirmed that the genus Arboridia has a closer relationship with Mitjaevia. Among them, M. bifurcata, M. protuberanta and M. diana are gathered into one clade, while M. bifurcata and M. protuberanta are sister groups of each other in ML tree and BI tree. In addition, Zyginellini is a junior synonym of Typhlocybini, our result supports recent author’s viewpoint (Dietrich, 2013; Zhou, Dietrich & Huang, 2020; Yan et al., 2022). The previous primary diagnosis of Typhlocybinae was made by morphological features, meanwhile, our phylogenetic tree based on molecular data is in agreement with morphological taxonomy. Because the external appearance of Mitjaevia species is very similar, the only difference lies in male genitalia including the pygofer, subgenital plate and aedeagus, so, molecular technologies have become particularly important as a supplement to identification of Mitjaevia species. This study indicated that mitochondrial genome sequences are the most popularly adopted genomic markers in leafhoppers and becoming increasingly important toward studies in the insect molecular field, involving molecular evolution, phylogeny and phylogeography.

Phylogenetic tree of Typhlocybinae produced from maximum-likelihood (ML) and Bayesian inference (BI) analyses based on complete mitochondrial gene.

Figure 7: Phylogenetic tree of Typhlocybinae produced from maximum-likelihood (ML) and Bayesian inference (BI) analyses based on complete mitochondrial gene.

Discussion

The traditional classification of leafhoppers mainly relies on the morphology of their appearance and male genitalia (Song & Li, 2013; Ramaiah, Meshram & Dey, 2023; Xu & Zhang, 2023). However, due to the large number of leafhoppers and the small size of the Erythroneurini leafhoppers, generally about 2∼4 mm (Dietrich & Dmitriev, 2006), they are difficult to identify. In recent years, the development of molecular technology has been applied to the classification of insects (Singh et al., 2017; Lu et al., 2018; Matsuia et al., 2022) and to support the results and correct the attribution of traditional classification. Our findings here on the complete mitochondrial genome supports the classification of two new species.

In addition, we also sequenced and analyzed the mitochondria genomes of two Mitjaevia genera, this result enriches the mitochondrial database information of Cicadellidae family and is consistent with the results of previous articles published by our research group (Chen, Li & Song, 2021). In the Typhlocybinae, each genus is divided into a separate branch, this result is consistent with previous results (Lin, Huang & Zhang, 2021), and all Mitjaevia are clustered in one branch. The phylogenetic relationship between the two new species is closer to that of Mitjaevia. This may be due to the limited mitochondrial data currently sequenced in Erythroneurini, which requires more and more extensive mitochondrial data to support and elucidate the phylogenetic relationship of the new species. The mitochondrial data can not only confirm the correctness of traditional classification, but also establish a large database, and provide simpler, faster, and more efficient results for subsequent species classification.

Conclusions

Two new leafhopper species discovered in Chongqing, A. (A.) rongchangensis sp. nov. and T. (T.) jiulongensis sp. nov. are described and illustrated. The mitochondrial genomes of these species together with M. bifurcata and M. diana are assembled and annotated in this work. The study shows that their mitogenomes are conserved in structure, with length of 15,596 bp, 15,676 bp, 16,813 bp and 16,589 bp, including 13 protein-coding genes, 22 tRNA genes, and two rRNA genes. The PCGs begin with ATA/ATG/ATT/TTG, and cease with TAA/TAG/T. All tRNAs are folded into a typical clover-leaf secondary structure, except a few tRNAs with a reduced arm, offering a simple loop or constituted unpaired bases. In this work, the phylogenetic analysis showed a well-supported, Arboridia has a closer relationship with Mitjaevia. M. bifurcata, M. protuberanta and M. diana are gathered into one clade, while M. bifurcata and M. protuberanta are sister groups of each other. In addition, Zyginellini can consider as a junior synonym of Typhlocybini. Based on the similarity in appearance of tribe Erythroneurini, the complete mitochondrial genome can provide faster and more convincing evidence for traditional classification.

Supplemental Information

Mitiaevia bifurcate

Chromas

DOI: 10.7717/peerj.16853/supp-1

Mitjaevia diana

Chromas

DOI: 10.7717/peerj.16853/supp-2

Inferred secondary structures of 22 tRNA from four species

Watson–Crick base pairings are illustrated by lines (-), whereas GU base pairings are illustrated by ⋆.

DOI: 10.7717/peerj.16853/supp-3
  Visitors   Views   Downloads