Next Article in Journal
Identification of LncRNAs and Functional Analysis of ceRNA Related to Fatty Acid Synthesis during Flax Seed Development
Next Article in Special Issue
Structural Characteristics and Phylogenetic Analysis of the Mitochondrial Genomes of Four Krisna Species (Hemiptera: Cicadellidae: Iassinae)
Previous Article in Journal
Hepatic Transcriptomics Reveals Reduced Lipogenesis in High-Salt Diet Mice
Previous Article in Special Issue
Full-Length Transcriptome Profiling of Coridius chinensis Mitochondrial Genome Reveals the Transcription of Genes with Ancestral Arrangement in Insects
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

The First Complete Mitochondrial Genome of Genus Isocapnia (Plecoptera: Capniidae) and Phylogenetic Assignment of Superfamily Nemouroidea

1
College of Plant Protection & Institute of Applied Entomology, Yangzhou University, Yangzhou 225009, China
2
Joint International Research Laboratory of Agriculture and Agri-Product Safety, The Ministry of Education, Yangzhou University, Yangzhou 225009, China
*
Author to whom correspondence should be addressed.
Genes 2023, 14(5), 965; https://doi.org/10.3390/genes14050965
Submission received: 1 March 2023 / Revised: 18 April 2023 / Accepted: 21 April 2023 / Published: 24 April 2023

Abstract

:
Capniidae are a family of stoneflies, also known as snow flies, who emerge in winter. The phylogeny of Capniidae is widely accepted to be based on morphological analysis. Until now, only five Capniidae mitochondrial genomes have been sequenced so far. In addition, sampling is required to determine an accurate phylogenetic association because the generic classification of this family is still controversial and needs to be investigated further. In this study, the first mitogenome of genus Isocapnia was sequenced with a length of 16,200 bp and contained 37 genes, including a control region, two rRNAs, 22 tRNAs, and 13 PCGs, respectively. Twelve PCGs originated with the common start codon ATN (ATG, ATA, or ATT), while nad5 used GTG. Eleven PCGs had TAN (TAA or TAG) as their last codon; however, cox1 and nad5 had T as their final codon due to a shortened termination codon. All tRNA genes demonstrated the cloverleaf structure, which is distinctive for metazoans excluding the tRNASer1 (AGN) that missed the dihydrouridine arm. A Phylogenetic analysis of the superfamily Nemouroidea was constructed using thirteen PCGs from 32 formerly sequenced Plecoptera species. The Bayesian inference and maximum likelihood phylogeny tree structures derived similar results across the thirteen PCGs. Our findings strongly supported Leuctridae + ((Capniidae + Taeniopterygidae) + (Nemouridae + Notonemouridae)). Ultimately, the best well-supported generic phylogenetic relationship within Capniidae is as follows; (Isocapnia + (Capnia + Zwicknia) + (Apteroperla + Mesocapnia)). These findings will enable us to better understand the evolutionary relationships within the superfamily Nemouroidea and the generic classification and mitogenome structure of the family Capniidae.

1. Introduction

The mitogenome is a fully functional and relatively independent organelle encompassing general information ranging from molecular studies to genetic variants [1]. Mitogenome research is frequently used in taxonomy, phylogenetics, and evolutionary studies in many animals. It has been extensively studied to resolve insect phylogeny [2,3,4]. Metazoans often have circular, double-standard molecules in their mitogenomes. It comprises two ribosomal RNA genes (rRNAs), 22 transfer RNA genes (tRNAs), 13 protein-coding genes (PCGs), and a non-coding region (CR) [5,6,7].
Order Plecoptera is the earliest, most diverse group of hemimetabolous insects, comprising more than 4000 described species under 17 families [8]. Their larvae live in cold lakes, pools, clean rocky creeks, rivers, and grassy places. The freshwater habitats and sandy regions are ideal for their larval development because the larval stages are highly susceptible to water pollution. The presence of stoneflies can be considered as a good indicator of water quality [9,10,11]. The genus Isocapnia [12] belongs to the Capniidae family and is predominantly distributed in North America and Asia [13,14]. The males of the genus Isocapnia often have an elongated epiproct, an apex with a distinctive form and a dark ventral vesicle extending from the sternum 9. Moreover, the female Isocapnia typically exhibits diagnostic sclerotization patterns on its subgenital plate [15]. To date, only 21 species of the genus Isocapnia have been described based on a morphological basis [8], but the molecular studies on this genus remain neglected; however, the current research provides initial molecular insights into the mitogenome features and phylogenetic placement. Currently, the most conventional classification of stoneflies was suggested by Zwick [16] using morphological research. In this study, Capniidae and four other families (Taeniopterygidae, Nemouridae, Notonemouridae, and Leuctridae) were grouped into Nemouroidea. The research demonstrated that Leuctridae and Capniidae were sister groups, which were named Leuctroidea, but the molecular studies did not confirm the monophyletic group Leuctroidea. Ding et al. [17] found that the Taeniopterygidae and Capniidae are sister groups, and Leuctridae was related to the rest of the Nemouroidea. In addition, Davis [18] also placed Taeniopterygidae and Capniidae as sister groups based on transcriptomic analysis, and several studies agreed with the research findings of Davis [19,20,21,22,23,24]. The initiation of next-generation sequencing techniques has led to a tremendous increase in the understanding of mitogenome studies; nonetheless, the mitogenome of Capniidae is still less documented. These limited data hinder accurate phylogenetic analysis [25,26]. More than 80 stoneflies have mitogenomes that are complete or almost complete, and these are now accessible in GenBank; however, only five mitogenomes of the Capniidae family were available [27,28]. While numerous phylogeny studies have been undertaken to identify the phylogenetic affinities amongst stoneflies, many of these affinities are still unclear. Nevertheless, inadequate taxon sampling renders these studies unsuitable for resolving exact phylogenetic associations within the Nemouroidea. The current study analyzed the first mitogenome of Isocapnia anguis, which will help to identify the genus classification as well as provide significant support for the evolutionary analysis of Nemouroidea. In addition, the composition of nucleotides, tRNAs secondary structure, and codon usage were also analyzed. Finally, a phylogenetic study of the superfamily Nemouroidea was undertaken using thirteen PCGs from accessible genomes. Hence, the current study aims to better understand the evolution of the Nemouroidea superfamily and the generic classification of the Capniidae family.

2. Materials and Methods

2.1. Sampling Collection and DNA Extraction

The specimens of Isocapnia anguis (Figure S1) [14] were collected from the China Sichuan Province (32°9.27′ N 104°1.44′ E, 2441 m) in 2016 and preserved in 75% ethanol. The terminalia was examined by KEYENCE VHX−5000 digital microscope and the final images were prepared using Adobe Photoshop CS6. The specimens were placed in the Insect collection of Yangzhou University (ICYZU), Jiangsu Province, China. This study was carried out without causing harm to endangered or threatened species and all scientific investigations were permitted. All specimens were kept in 100% ethanol at −20 °C. The genomic DNA was pulled out from the thoracic muscles and legs of adults using the column mtDNAoutKit (Axygen Biotechnology, Hangzhou, China), as endorsed by the supplier, and stored at −20 °C until used for PCR.

2.2. PCR Amplification and Sequencing

Mitogenome was amplified using LA-PCR and clear PCR implication under the following conditions: achieve initial denaturation at 93 °C for 02 min, then perform 40 cycles at 92 °C for 10 s; soften at 54 °C for 30 s; and extend at 68 °C for 8 min. The elongation rate was increased by 20 s for each cycle in the last 20 cycles, resulting in a final extension of 10 min at 68 °C. High throughput sequencing and universal primers were used to amplify mitochondrial genes in long overlapping fragments. The Axygen DNA gel Extraction Kit (Axygen Biotechnology, Hangzhou, China) was utilized to purify PCR results [19]. When the DNA had been extracted and refined, it was subjected to quality control, which involved testing via qubit 3.0 as well as electrophoresis on a 1% agarose gel. The NEBNext Ultra DNA Library Prep Kit for sequence analysis was used to generate a 500 bp matched library from high-quality DNA samples, and sequencing was performed on the Illumina NovaSeq 6000 platform (BIOZERON Co., Ltd., Shanghai, China). The Contigs of the mitogenome were generated by de novo assembly using GetOrganelle v1.6.5 as a reference for the mitochondrial genomes of closely related species. The potential mitochondrial readings were isolated from Illumina reads using BLAST against the mitogenomes species and GetOrganelle outcome. The mitochondrial Illumina states were achieved to complete mitogenome de novo assembly utilizing the SPAdes-3.13.1 program. The GetOrganelle contig was enhanced by the scaffolds from the SPAdes-3.13.1 program. To complete the process, the accumulated sequences were rearranged and orientated to match the reference mitochondrial genome, providing a complete genetic sequence of mitochondria (BIOZERON Co., Ltd., Shanghai, China).

2.3. Mitogenome Assembly and Annotation

The mitogenome assembly was performed using CodonCode Aligner (accessed on 5 February 2023). The mitogenomes from other species of order Plecoptera were used to recognize rRNAs, tRNAs, and PCGs genes, and ORFs were enclosed through ORF finder (accessed on 5 February 2022). The CGview tool server was used to create a circular map of the mitogenome [29]. We used the online MITOS program [30] to characterize the genes, with the default settings being used to predict PCGs, tRNA, and RNA (rRNA) genes. The structure of tRNA was predicted by the ARWEN algorithm using its default settings [31]. The codon usage, RSCU (relative synonymous codon usage) and nucleotide compositions were examined via MEGA V. 7.0 [32]. The AT-skew = (A − T)/(A + T) and the GC-skew = (G − C)/(G + C) formulas [33] were used to evaluate the skew of AT and GC data, respectively. The DNAMAN program was used to predict stem-loop structures in the CR, and tandem repeats of CR were assessed using the Tandem repeats finder online [34]. Mitogenome sequences of I. anguis were deposited in the GenBank as OQ735414 (Table 1).

2.4. Phylogenetic Analysis

We used thirty-one published Plecoptera and one newly sequenced mitogenome for the phylogenetic assessment, and two species of the family Perlodidae were treated as an outgroup (Table 1). The thirteen PCGs in thirty-two mitogenomes were synthesized using SequenceMatrix v. 1.7.8 [46] and aligned with the MAFFT algorithm [47], excluding stop codon. Before creating phylogenetic trees, the DAMBE v. 5.2 was used to check for nucleotide saturation. The MEGA V. 7.0 [32] was used to choose the GTR+G+I model as the optimal nucleotide substitution analytical model. The MrBayes (version 3.1.2) was used to conduct a Bayesian inference (BI) study under the following conditions: 10 million generations with 100-generation intervals, 4 chains, and a burn-in of 25% of the tree [23,48,49]. The maximum likelihood (ML) was conducted using IQTree v. 1.6.12 [50] with 1000 ultrafast bootstrap estimates, and the tree was visualized with Fig Tree v. 1.4.2.

3. Results and Discussion

3.1. Characteristics of the Mitogenome

The length of the whole mitogenome of I. anguis is 16,200 base pairs (bp) (Figure 1), which is virtually identical to the genome of Capnia zijinshana (16,310 bp) and comparable to prior researches attempting to study the entire mitogenomes of stoneflies (Table 1). Among the Capniidae complete mitogenomes, the length variation was minimal in PCGs, tRNAs, and rRNAs, but differences in the control region are high. The mitogenome is a circular DNA; its components include 13 PCGs, 22 tRNAs, 2 rRNA genes, and a control region, also recognized as a non-coding region (Figure 1, Table 2). Twenty-three genes (9 PCGs and 14 tRNAs) were detected on the J−strand (major strand), the remaining genes (4 PCGs, 8 tRNAs, and 2 rRNAs) being oriented on the N−strand (minor strand) (Table 2, Figure 1). The gene order of I. anguis is identical to other sequenced stoneflies and to Drosophila yakuba, which is reflected in the hypothesized ancestral mitogenome [51,52]. In I. anguis, 53 overlapping nucleotides were positioned in 15 pairs of adjacent genes, ranging from 1 to 8 bp; the most extended overlap (8 bp) was found among trnTrp (W) and trnCys (C). Only 41 IGN (intergenic nucleotides) were observed in eight various locations, and their lengths ranged from 1 to 16 base pairs (bp), with the longest IGN (16 bp) located amongst trnSer2 (UCN) and nad1.

3.2. Base Composition

In I. anguis, the overall nucleotide composition is 31.28% T, 20.64% C, 34.91% A, and 13.17% G, correspondingly (Table 2). The A + T content of the entire mitogenome, PCGs, PCG−J, PCG−N, rRNAs, tRNAs, and the non-coding region is 66.19%, 64.78%, 62.95, 66.52%, 71.20%, 69.02% and 66.62%, respectively (Table 2). The A + T content is 86.15% for trnGlu (E), and the lowest was 56.25% for trnArg (R) (Table S1). The A + T content of mitogenome is identical with Apteroperla tikumana (66.51%), while a slightly increase was seen in C. zijinshana (68.5%) compared to I. anguis species. The bias in the nucleotide makeup was assessed by AT- and GC-skew, as previously described [33]. Typically, the nucleotide contents of metazoan mitogenomes have a distinct strand bias [53,54].
In this study, the I. anguis whole mitogenome exhibited a positive AT-skew (0.05) and negative GC-skew (−0.22), while the PCGs and PCG−J (PCGs encoded by the majority strand) both showed a negative AT- and GC-skew and PCG−N (PCGs encoded by the minority strand) showed positive GC-skew and negative AT-skew (Table 2). For the J−strand, the mitogenomes for most insects showed a positive AT-skew and negative GC-skew [55], while in this study, PCG and PCG−J both showed negative AT- and GC- skew. The A + T bias in I. anguis was linked to asymmetric mutation and selection pressure during transcription and replication, which was demonstrated to be functional in mitogenome studies [33,34,56].

3.3. Protein Coding Genes

The full size of the thirteen PCGs of I. anguis was 11,219 bp, and its A + T content stood at 64.78% (Table 2). In the I. anguis mitogenome, twelve PCGs initiated with the conventional start codon ATN (ATG, ATA or ATT), whereas nad5 gene began with codon GTG. Eleven PCGs had TAN (TAA or TAG) as their last codon; however, cox1 and nad5 both had shortened termination codons and concluded with T (Table S1). The use of incomplete stop codon T is mutual in most of the stoneflies [20,23,27,28,40,57] and other animals’ mitogenome [5], forming a comprehensive TAA terminal signal via post-transcriptional polyadenylation [58,59].
In the entire mitogenome, the A + T content ranged from 65.19% to 68.89% in the Capniidae species (Figure 2, Table S2). Comparison among the PCGs (Figure 2, Table S3) of all Capniidae showed that the A + T content of cox3 was the lowest (62.10%), followed by cox1 (62.83%) and cytb (63.63%), whereas nd4l, nd6, and atp8 were the highest (73.85%, 70.69%, and 69.285%), respectively (Figure 2, Table S3). Furthermore, PCGs have a substantially greater A + T concentration in the third codon position (ranging from 65.27% to 73.27%) than in the first and second (ranging from 65.08% to 69.57%) (Figure S1). We calculated the RSCU values of the I. anguis (Figure 3). The GCT (Ala), TGT (Cys), CAA (Gln), ATT (Ile), TTA (Leu), CCT (Pro), and TCT (Ser) were comparatively high, whereas UUG (Leu1) and CCG (Pro) were used the least (Figure 3). This differs from other species of stoneflies [19,28], so we believe that perhaps this was unique to the genus Isocapnia. Nevertheless, further research and data are necessary to provide confirmation.

3.4. Transfer RNA Genes

The I. anguis mitogenome had 22 tRNA genes scattered all over the PCGs and rRNAs. The 22 tRNAs were 1475 bp with a length ranging from 64 to 71 bp and had an A + T content of 69.02% (Table 2). Most tRNAs had a distinctive cloverleaf structure (Figure 4), with the exception of trnSer (AGN), which lacked the dihydrouridine (DHU) arm and formed a small loop, as is usually the case in mammals and in numerous metazoans [60]. The anticodon of all tRNAs was identical to that of stoneflies. A total of 41 miss-matched base pairs were detected in the tRNAs, which are 2 bp U−U, 5 bp A−U, 2 bp U−C, 1 bp A−C, 1 bp A−G, and 30 bp weak G−U pairs (Figure 4), which produced weak bonds in tRNAs and non-canonical pairs in the secondary structure of tRNAs [61]. The well-maintained and variable position in tRNA structures were detected in the TΨC loop and anticodon arm, respectively (Figure 4).
There were 2 rRNAs in the I. anguis mitogenome with a total size of 2125 bp and an A + T content of 71.20% (Table 2). The boundaries of the small and large subunit ribosomal RNAs (srRNA and lrRNA) were found using previously published plecopteran sequence alignment. The large rRNA subunit gene (lrRNA) is 1333 bp long, has an A + T content of 72.47%, and was found between trnLeu1 (CUN) and trnVal (V). The small rRNA gene (srRNA) was 792 bp long and had a 69.07% A + T content, which is located between trnVal (V) and the non-coding region (Table S1). Both lrRNA and srRNA size and characteristics are consistent with those reported in well-documented Plecoptera species [19,20,21,22,23,24,25,26,27,28].

3.5. The Control Region

The control region is the most variable part of the mitogenome due to the insertion and deletion measures, deviation in domain length fluctuation, and the number of tandem repeats [62,63], which contain vital components for the initiation of transcription and replication [64]. The CR of the I. anguis mitogenome was 1392 bp with an A + T content of 66.6%, which was located amongst rrnS and trnlle (Table S1). The length of the CR in Capniidae mitogenomes was more prolonged compared to other stoneflies with a lower A + T content. The CR of C. zijinshana was substantially longer among Capniidae mitogenomes 1513 bp in length, with a lower A + T content of 62.0% [28].
The CR of the I. anguis can be divided into three parts; (1) a leading sequence 819 bp adjacent to srRNA with high A + T content; (2) one tandem repeat (TR) sequence blocks consisting of repeats units; (3) the rest of the control region (457 bp) (Figure 5). The tandem repeats are located between 15,626–15,742 bp and four stem-loop structures (14,948–14,974 bp, 15,152–15,177 bp, 15,467–15,497 bp, 15,496–15517 bp) (Figure 5). The SL structure was a single root with (TA)n structures on the left side and CAT or C (T)nA structures on the right side (Figure 5). The study suggested that these SL structures and CR tandem repeats may have stimulatory effects on mitogenome replication and transcription [65,66].

3.6. Phylogenetic Analysis

The phylogenetic analysis of the whole Nemouroidea was constructed using thirteen PCGs from thirty-two Plecoptera species. These mitogenomes included seventeen species from the family Nemouridae, one species of the Notonemouridae, six species from the Capniidae, three species each from Taeniopterygidae and Leuctridae, while two species of Perlodidae (Isoperla eximia and Pseudomegarcys japonica) were included as outgroups (Table 1). The phylogeny tree patterns were analogous to dendrograms derived from BI and ML analysis. The topological tree structures were similar for both ML and BI analysis. The monophyly of every group was highly supported (Post probability values (PP) = 1.00 and bootstrap values (BP) > 99). This study represented the phylogeny of the superfamily Nemouroidea, which includes five families, such as Nemouridae, Notonemouridae, Capniidae, Taeniopterygidae, and Leuctridae. Our analysis substantially supported Leuctridae + ((Capniidae + Taeniopterygidae) + (Nemouridae +Notonemouridae)). Moreover, the Leuctridae family was the initial branch to be developed within the Nemouroidea superfamily. It was determined in our study that Nemouridae is the sister group of Notonemouridae (BP = 100, PP = 1.00) (Figure 6). These results correlate with the morphological hypothesis [16] and mitogenome studies of [20,21,22,23,24], but few molecular analyses disagree with this placement [67,68]. According to the studies of Zwick [16], the Capniidae family was initially positioned as a sister to the Leuctridae. After this, the Capniidae was clustered with the Nemouridae and Notonemouridae. The current results supported Capniidae as a sister taxon to Taeniopterygidae (BP = 98, PP = 1.00) (Figure 6). This hypothesis varies from the morphological hypothesis that Leuctridae and Capniidae are sister taxon [16], but until now, no molecular study has supported this placement. In previous mitogenome studies [17,19,20,21,22,23,24], the Capniidae and Taeniopterygidae were considered as sister groups, and Leuctridae to the remaining Nemouroidea. However, the results of this study demonstrated more stability (BP =99, PP = 1.00) (Figure 6) than prior studies [17,24].
In the generic level of the Capniidae family, the genus Isocapnia is a sister to Capnia (BP = 98, PP = 1.00) (Figure 6). The monophyly of each genus is often strongly supported (BP > 97 and PP > 0.99), except for Zwicknia and Capnia genera. Zwcknia and Capnia were more closely related, yet their bootstrap probability values are lower than those of other genera. Zwicknia is a newly erected genus from the Capnia bifrons species group [69], and each genus has just one species sequenced so far. As data for only a few taxa of the Capniidae family are currently available, additional molecular testing from a wider variety of genera and an increased number of species are required to better resolve the phylogeny of this group. Finally, in this study, the best well-supported generic phylogenetic relationship within Capniidae is as follows: (Isocapnia + (Capnia + Zwicknia) + (Apteroperla + Mesocapnia)). However, to form a better understanding, more molecular investigations are needed to resolve the phylogeny of the Capniidae family.

4. Conclusions

The contemporary phylogeny of Capniidae is based on traditional classification and is universally recognized. However, molecular evidence supports different hypotheses. Since only five mitogenomes are currently available, more research data on other genera and species are required to determine the precise phylogenetic affiliation of this family. In this study, the first mitogenome of the genus Isocapnia was sequenced. The Isocapnia mitogenome was found to be highly conserved in terms of size, gene order, and nucleotide sequence when compared to other members of the Capniidae family. Both the ML and BI approaches were used to infer phylogenetic trees, and these trees created identical topologies across all 13 PCGs. The best well-supported generic phylogenetic relationship within Capniidae is as follows: (Isocapnia + (Capnia + Zwicknia) + (Apteroperla + Mesocapnia)). The relationship of the superfamily Nemouroidea recovered as strongly supported Leuctridae + ((Capniidae + Taeniopterygidae) + (Nemouridae + Notonemouridae)). These findings will help us to understand the general classification and mitogenome structure of the Capniidae family and their relationship within Nemouroidea.

Supplementary Materials

The following supporting information can be downloaded at: https://www.mdpi.com/article/10.3390/genes14050965/s1, Table S1: Organization of the I. anguis mitochondrial genome; Table S2: A + T content of different components or positions among mitogenome of the family Capniidae; Table S3. A + T content of different PCGs among mitogenome of the family Capniidae; Figure S1. (A) I. anguis adult habitus dorsal view, (B) head and thorax dorsal view, (C) terminalia dorsal view.

Author Contributions

Conceptualization, A.R. and Y.-Z.D.; methodology, A.R.; software, A.R.; validation, Y.-Z.D. and Q.-B.H.; formal analysis, A.R.; resources, Y.-Z.D.; data curation, Y.-Z.D.; writing—original draft preparation, A.R.; writing—review and editing, Y.-Z.D. and Q.-B.H.; visualization, A.R.; supervision, Y.-Z.D.; project administration, Y.-Z.D.; funding acquisition, Y.-Z.D. All authors have read and agreed to the published version of the manuscript.

Funding

This research was supported by the National Natural Science Foundation of China (No. 32170459; 31572295).

Institutional Review Board Statement

Not applicable.

Informed Consent Statement

Not applicable.

Data Availability Statement

Data are available on reasonable request.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Simon, C.; Buckley, T.R.; Frati, F.; Stewart, J.B.; Beckenbach, A.T. Incorporating molecular evolution into phylogenetic analysis, and a new compilation of conserved polymerase chain reaction primers for animal mitochondrial DNA. Annu. Rev. Ecol. Evol. Syst. 2006, 37, 545–579. [Google Scholar] [CrossRef]
  2. Cameron, S.L. Insect mitochondrial genomics: Implications for evolution and phylogeny. Annu. Rev. Entomol. 2014, 59, 95–117. [Google Scholar] [CrossRef]
  3. Yuan, M.L.; Zhang, Q.L.; Zhang, L.; Guo, Z.L.; Liu, Y.J.; Shen, Y.Y.; Shao, R. High-level phylogeny of the Coleoptera inferred with mitochondrial genome sequences. Mol. Phylogenet. Evol. 2016, 104, 99–111. [Google Scholar] [CrossRef] [PubMed]
  4. Li, H.; Leavengood, J.M.; Chapman, E.G.; Burkhardt, D.; Song, F.; Jiang, P.; Liu, J.; Zhou, X.; Cai, W. Mitochondrial phylogenomics of Hemiptera reveals adaptive innovations driving the diversification of true bugs. Proc. R. Soc. B 2017, 284, 20171223. [Google Scholar] [CrossRef]
  5. Wolstenholme, D.R. Animal mitochondrial DNA: Structure and evolution. Int. Rev. Cytol. 1992, 141, 173–216. [Google Scholar] [PubMed]
  6. Simon, C.; Frati, F.; Beckenbach, A.; Crespi, B.; Liu, H.; Flook, P. Evolution, weighting, and phylogenetic utility of mitochondrial gene sequences and a compilation of conserved polymerase chain reaction primers. Ann. Entomol. Soc. Am. 1994, 87, 651–701. [Google Scholar] [CrossRef]
  7. Gissi, C.; Iannelli, F.; Pesole, G. Evolution of the mitochondrial genome of metazoa as exemplified by comparison of congeneric species. Heredity 2008, 101, 301–320. [Google Scholar] [CrossRef]
  8. DeWalt, R.E.; Maehr, M.D.; Hopkins, H.; Neu-Becker, U.; Stueber, G. Plecoptera Species File Online. Version 5.0/5.0. Available online: http://Plecoptera.SpeciesFile.org (accessed on 13 February 2023).
  9. Stewart, K.W.; Stark, B.P. Nymphs of North American Stonefly Genera (Plecoptera), 2nd ed.; The Caddis Press: Columbus, OH, USA, 2002. [Google Scholar]
  10. Stewart, K.W.; Stark, B.P. Plecoptera. In An Introduction to the Aquatic Insects of North America, 4th ed.; Merritt, R.W., Cummins, K.W., Berg, M.B., Eds.; Kendall/Hunt Publishing Company: Dubuque, IA, USA, 2008; Chapter 14; pp. 311–384. [Google Scholar]
  11. Fochetti, R.; Figueroa, J.M.T. Global diversity of stoneflies (Plecoptera, Insecta) in freshwater. Hydrobiologia 2008, 595, 365–377. [Google Scholar] [CrossRef]
  12. Banks, N. New native neuropteroid insects. Psyche J. Entomol. 1938, 45, 72–79. [Google Scholar] [CrossRef]
  13. Zhiltzova, L.A. Plecoptera, Gruppe Euholognatha. Fauna Russ. Neighboring Ctries New Ser. 2003, 145, 1–538. [Google Scholar]
  14. Chen, Z.T.; Du, Y.Z. First record of Isocapnia (Plecoptera: Capniidae) from China with description of a new species. Ann. Soc. Entomol. Fr. 2017, 53, 42–46. [Google Scholar] [CrossRef]
  15. Judson, S.W.; Nelson, C.R. A guide to Mongolian stoneflies (Insecta: Plecoptera). Zootaxa 2012, 3541, 1–118. [Google Scholar] [CrossRef]
  16. Zwick, P. Phylogenetic System and Zoogeography of the Plecoptera. Annu. Rev. Entomol. 2000, 45, 709–746. [Google Scholar] [CrossRef] [PubMed]
  17. Ding, S.M.; Li, W.H.; Wang, Y.; Cameron, S.L.; Murányi, D.; Yang, D. The phylogeny and evolutionary timescale of stoneflies (Insecta: Plecoptera) inferred from mitochondrial genomes. Mol. Phylogenet. Evol. 2019, 135, 123–135. [Google Scholar] [CrossRef] [PubMed]
  18. Davis, N.G. Application of Next-Generation Transcriptomic Tools for Non-Model Organisms: Gene Discovery and Marker Development within Plecoptera (Insecta). Master’s Thesis, Brigham Young University, Provo, UT, USA, 2013; pp. 1–39. [Google Scholar]
  19. Chen, Z.T.; Du, Y.Z. The first two mitochondrial genomes from Taeniopterygidae (Insecta: Plecoptera): Structural features and phylogenetic implications. Int. J. Biol. Macromol. 2018, 111, 70–76. [Google Scholar] [CrossRef]
  20. Cao, J.J.; Wang, Y.; Li, W.H. Comparative mitogenomic analysis of species in the subfamily Amphinemurinae (Plecoptera: Nemouridae) reveal conserved mitochondrial genome organization. Int. J. Biol. Macromol. 2019, 138, 292–301. [Google Scholar] [CrossRef]
  21. Cao, J.J.; Wang, Y.; Guo, X.; Wang, G.Q.; Li, W.H.; Murányi, D. Two complete mitochondrial genomes from Leuctridae (Plecoptera: Nemouroidea): Implications for the phylogenetic relationships among stoneflies. J. Insect Sci. 2021, 21, 16. [Google Scholar] [CrossRef]
  22. Shen, Y.; Du, Y.Z. The mitochondrial genome of Leuctra sp. (Plecoptera: Leuctridae) and its performance in phylogenetic analyses. Zootaxa 2019, 4671, 571–580. [Google Scholar] [CrossRef]
  23. Zhao, M.Y.; Huo, Q.B.; Du, Y.Z. Molecular phylogeny inferred from the mitochondrial genomes of Plecoptera with Oyamia nigribasis (Plecoptera: Perlidae). Sci. Rep. 2020, 10, 20955. [Google Scholar] [CrossRef]
  24. Guo, X.; Guo, C.; Dong, X.; Zhang, H.; Murányi, D.; Li, W.; Wang, Y. Mitochondrial Genome of Strophopteryx fasciata (Plecoptera: Taeniopterygidae), with a Phylogenetic Analysis of Nemouroidea. Genes 2022, 13, 1116. [Google Scholar] [CrossRef]
  25. Wu, H.Y.; Ji, X.Y.; Yu, W.W.; Du, Y.Z. Complete mitochondrial genome of the stonefly Cryptoperla stilifera Sivec (Plecoptera: Peltoperlidae) and the phylogeny of Polyneopteran insects. Gene 2014, 537, 177–183. [Google Scholar] [CrossRef] [PubMed]
  26. Mo, R.R.; Wang, Y.; Cao, J.J.; Wang, G.Q.; Li, W.H.; Murányi, D. Two complete mitochondrial genomes of the subfamily Chloroperlinae (Plecoptera: Chloroperlidae) and their phylogenetic implications. Arthropod Syst. Phylogeny 2022, 80, 155–168. [Google Scholar] [CrossRef]
  27. Wang, Y.; Cao, J.; Li, W.; Chen, X. The mitochondrial genome of Mesocapnia daxingana (Plecoptera: Capniidae). Conserv. Genet. Resour. 2017, 9, 639–642. [Google Scholar] [CrossRef]
  28. Chen, Z.T.; Du, Y.Z. Complete mitochondrial genome of Capnia zijinshana (Plecoptera: Capniidae) and phylogenetic analysis among stoneflies. J. Asia-Pac. Entomol. 2017, 20, 305–312. [Google Scholar] [CrossRef]
  29. Grant, J.R.; Stothard, P. The CGView server: A comparative genomics tool for circular genomes. Nucl. Acids Res. 2008, 36, 181–184. [Google Scholar] [CrossRef]
  30. Bernt, M.; Donath, A.; Jühling, F.; Externbrinka, F.; Florentz, C.; Fritzsch, G.; Pützh, J.; Middendorf, M.; Stadler, P.F. MITOS: Improved de novo metazoan mitochondrial genome annotation. Mol. Phylogenet. Evol. 2013, 69, 313–319. [Google Scholar] [CrossRef]
  31. Laslett, D.; Canbäck, B. ARWEN, a program to detect tRNA genes in metazoan mitochondrial nucleotide sequences. Bioinformatics 2008, 24, 172–175. [Google Scholar] [CrossRef]
  32. Kumar, S.; Stecher, G.; Tamura, K. MEGA 7: Molecular evolutionary genetics analysis version 7.0 for bigger datasets. Mol. Biol. Evol. 2016, 33, 1870–1874. [Google Scholar] [CrossRef]
  33. Perna, N.T.; Kocher, T.D. Patterns of nucleotide composition at fourfold degenerate sites of animal mitogenome. J. Mol. Evol. 1995, 41, 353–358. [Google Scholar] [CrossRef]
  34. Benson, G. Tandem repeats finder: A program to analyze DNA sequences. Nucleic Acids Res. 1999, 27, 573–580. [Google Scholar] [CrossRef]
  35. National Center for Biotechnology Information (NCBI). Available online: https://www.ncbi.nlm.nih.gov (accessed on 5 February 2023).
  36. Wei, X.; Cao, J.; Kong, F.; Wang, Y. The mitochondrial genome analysis of Sphaeronemoura elephas (Plecoptera: Nemouridae) from Jiangxi Province of southeastern China. Mitochondrial DNA Part B Resour. 2020, 5, 1107–1108. [Google Scholar] [CrossRef] [PubMed]
  37. Cao, J.J.; Wang, Y.; Huang, Y.R.; Li, W.H. Mitochondrial genomes of the stoneflies Mesonemoura metafiligera and Mesonemoura tritaenia (Plecoptera, Nemouridae), with a phylogenetic analysis of Nemouroidea. ZooKeys 2019, 835, 43–63. [Google Scholar] [CrossRef] [PubMed]
  38. Wang, J.; Tian, T.; Zhou, M.; Li, W.; Cao, J. The complete mitochondrial genome of a stonefly species, Protonemura datongensis (Plecoptera: Nemouridae). Mitochondrial DNA Part B 2020, 5, 2006–2007. [Google Scholar] [CrossRef]
  39. Chen, J.; Cao, J.; Chen, M.; Chen, S.; Li, W.; Wang, Y. The complete mitochondrial genome of Amphinemura claviloba (Wu, 1973)(Plecoptera: Nemouridae). Zootaxa 2020, 4732, 4732. [Google Scholar] [CrossRef] [PubMed]
  40. Wang, Y.; Cao, J.; Murányi, D.; Chen, X.; Yan, F. The complete mitochondrial genome of Amphinemura bulla Shimizu, 1997 (Plecoptera: Nemouridae) from Japan. Mitochondrial DNA Part B 2021, 6, 846–847. [Google Scholar] [CrossRef]
  41. Cao, J.; Wang, Y.; Ma, G.; Li, W. Complete mitochondrial genome of a stonefly, Nemoura papilla (Plecoptera: Nemouridae). Mitochondrial DNA Part B 2019, 4, 806–807. [Google Scholar] [CrossRef]
  42. Chen, Z.T.; Du, Y.Z. First Mitochondrial Genome from Nemouridae (Plecoptera) Reveals novel features of the elongated control region and phylogenetic implications. Int. J. Mol. Sci. 2017, 18, 996. [Google Scholar] [CrossRef]
  43. Chen, S.; Cao, J.; Li, W.; Wang, Y. Characterization of the complete mitochondrial genome of Nemoura meniscata Li & yang (Plecoptera: Nemouridae) from China. Mitochondrial DNA Part B Resour. 2020, 5, 1052–1053. [Google Scholar]
  44. Elbrecht, V.; Leese, F. The mitochondrial genome of the Arizona Snowfly Mesocapnia arizonensis (Plecoptera, Capniidae). Mitochondrial DNA Part A 2016, 27, 3365–3366. [Google Scholar] [CrossRef]
  45. Wang, Y.; Cao, J.; Murányi, D.; Li, W. Comparison of two complete mitochondrial genomes from Perlodidae (Plecoptera: Perloidea) and the family-level phylogenetic implications of Perloidea. Gene 2018, 675, 254–264. [Google Scholar] [CrossRef]
  46. Vaidya, G.; Lohman, D.J.; Meier, R. SequenceMatrix: Concatenation software for the fast assembly of multi-gene datasets with character set and codon information. Cladistics 2010, 27, 171–180. [Google Scholar] [CrossRef] [PubMed]
  47. Katoh, K.; Standley, D.M. MAFFT multiple sequence alignment software version 7: Improvements in performance and usability. Mol. Biol. Evol. 2013, 30, 772–780. [Google Scholar] [CrossRef] [PubMed]
  48. Ronquist, F.; Huelsenbeck, J.P. MrBayes 3: Bayesian phylogenetic inference under mixed models. Bioinformatics 2003, 19, 1572–1574. [Google Scholar] [CrossRef] [PubMed]
  49. Xiang, Y.; Zhao, M.; Huo, Q.; Du, Y. Mitochondrial Genomes of the Genus Claassenia (Plecoptera: Perlidae) and Phylogenetic Assignment to Subfamily Perlinae. Genes 2021, 12, 1986. [Google Scholar] [CrossRef]
  50. Nguyen, L.T.; Schmidt, H.A.; Haeseler, A.V.; 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]
  51. Stewart, J.B.; Beckenbach, A.T. Insect mitochondrial genomics: The complete mitogenome sequence of the meadow spittlebug Philaenus spumarius (Hemiptera: Auchenorrhyncha: Cercopoidae). Genome 2005, 48, 46–54. [Google Scholar] [CrossRef]
  52. Clary, D.O.; Wolstenholme, D.R. The ribosomal RNA genes of Drosophila mitochondrial DNA. Nucleic Acids Res. 1985, 13, 4029–4045. [Google Scholar] [CrossRef]
  53. Hassanin, A.; Léger, N.; Deutsch, J. Evidence for multiple reversals of asymmetric mutational constraints during the evolution of the mitochondrial genome of Metazoa, and consequences for phylogenetic inferences. Syst. Biol. 2005, 54, 277–298. [Google Scholar] [CrossRef]
  54. Hassanin, A. Phylogeny of Arthropoda inferred from mitochondrial sequences: Strategies for limiting the misleading effects of multiple changes in pattern and rates of substitution. Mol. Phylogenet. Evol. 2006, 38, 100–116. [Google Scholar] [CrossRef]
  55. Wei, S.J.; Shi, M.; Chen, X.X.; Sharkey, M.J.; van Achterberg, C.; Ye, G.Y.; He, J.H. New views on strand asymmetry in insect mitochondrial genomes. PLoS ONE 2010, 5, e12708. [Google Scholar] [CrossRef]
  56. Lindahl, T. Instability and decay of the primary structure of DNA. Nature 1993, 362, 709–715. [Google Scholar] [CrossRef] [PubMed]
  57. Wang, K.; Ding, S.; Yang, D. The complete mitochondrial genome of a stonefly species, Kamimuria chungnanshana Wu, 1948 (Plecoptera: Perlidae). Mitochondrial DNA Part A 2016, 27, 3810–3811. [Google Scholar] [CrossRef] [PubMed]
  58. Boore, J.L. Complete mitochondrial genome sequence of Urechis caupo, a representative of the phylum Echiura. BMC Genome 2004, 5, e67. [Google Scholar] [CrossRef] [PubMed]
  59. Cao, J.; Guo, X.; Guo, C.; Wang, X.; Wang, Y.; Yan, F. Complete Mitochondrial Genome of Malenka flexura (Plecoptera: Nemouridae) and Phylogenetic Analysis. Genes 2022, 13, 911. [Google Scholar] [CrossRef]
  60. Garey, J.R.; Wolstenholme, D.R. Platyhelminth mitochondrial DNA: Evidence for early evolutionary origin of a tRNAserAGN that contains a dihydrouridine arm replacement loop, and of serine-specifying AGA and AGG codons. J. Mol. Evol. 1989, 28, 374–387. [Google Scholar] [CrossRef]
  61. Gutell, R.R.; Lee, J.C.; Cannone, J.J. The accuracy of ribosomal RNA comparative structure models. Curr. Opin. Struct. Biol. 2002, 12, 301–310. [Google Scholar] [CrossRef]
  62. Zhang, D.X.; Szymura, J.M.; Hewitt, G.M. Evolution and structure conservation of the control region of insect mitochondrial DNA. J. Mol. Evol. 1995, 40, 382–391. [Google Scholar] [CrossRef]
  63. Taanman, J.W. The mitochondrial genome: Structure, transcription, translation and replication. BBA Bioenerg. 1999, 1410, 103–123. [Google Scholar] [CrossRef]
  64. Zhang, D.X.; Hewitt, G.M. Insect mitochondrial control region: A review of its structure, evolution and usefulness in evolutionary studies. Biochem. Syst. Ecol. 1997, 25, 99–120. [Google Scholar] [CrossRef]
  65. Yang, F.; Du, Y.Z.; Wang, L.P.; Cao, J.M.; Yu, W.W. The complete mitochondrial genome of the leafminer Liriomyza sativae (Diptera: Agromyzidae): Great difference in the A + T-rich region compared to Liriomyza trifolii. Gene 2011, 485, 7–15. [Google Scholar] [CrossRef]
  66. Crozier, R.H.; Crozier, Y.C. The mitochondrial genome of the honeybee Apis mellifera: Complete sequence and genome organization. Genetics 1993, 133, 97–117. [Google Scholar] [CrossRef] [PubMed]
  67. Thomas, M.A.; Walsh, K.A.; Wolf, M.R.; McPheron, B.A.; Marden, J.H. Molecular phylogenetic analysis of evolutionary trends in stonefly wing structure and locomotor behavior. Proc. Natl. Acad. Sci. USA 2000, 97, 13178–13183. [Google Scholar] [CrossRef] [PubMed]
  68. Terry, M.D. Phylogeny of the Polyneopterous Insects with Emphasis on Plecoptera: Molecular and Morpological Evidence. Ph.D. Thesis, Brigham Young University, Provo, UT, USA, 2003; pp. 1–118. [Google Scholar]
  69. Muranyi, D.; Gamboa, M.; Orci, K.M. Zwicknia Gen. n., a New Genus for the Capnia Bifrons Species Group, with Descriptions of Three New Species Based on Morphology, Drumming Signals and Molecular Genetics, and a Synopsis of the West Palaearctic and Nearctic Genera of Capniidae (Plecoptera). Zootaxa 2014, 3812, 1–82. [Google Scholar] [CrossRef] [PubMed]
Figure 1. Map of the mitogenome of I. anguis. Genes outside the map are recorded clockwise, while genes inside the map are counterclockwise. The interior spheres showed GC content (Guanine and Cytosine nucleotides) and the GC-skew, plotted as the deviation from of overall mean value of the whole sequence.
Figure 1. Map of the mitogenome of I. anguis. Genes outside the map are recorded clockwise, while genes inside the map are counterclockwise. The interior spheres showed GC content (Guanine and Cytosine nucleotides) and the GC-skew, plotted as the deviation from of overall mean value of the whole sequence.
Genes 14 00965 g001
Figure 2. Comparison of nucleotide components of mitogenome across the Capniidae family: (a) A + T% of different positions; (b) A + T% of PCGs.
Figure 2. Comparison of nucleotide components of mitogenome across the Capniidae family: (a) A + T% of different positions; (b) A + T% of PCGs.
Genes 14 00965 g002
Figure 3. The relative synonymous codon usage (RSCU) in the I. anguis mitogenome.
Figure 3. The relative synonymous codon usage (RSCU) in the I. anguis mitogenome.
Genes 14 00965 g003
Figure 4. Secondary structures of 22 tRNAs of I. anguis. All tRNAs are labeled with the abbreviations of their consistent amino acids. Dashes (-) specify Watson–Crick base combination and dots (•) specify G-U base pairing.
Figure 4. Secondary structures of 22 tRNAs of I. anguis. All tRNAs are labeled with the abbreviations of their consistent amino acids. Dashes (-) specify Watson–Crick base combination and dots (•) specify G-U base pairing.
Genes 14 00965 g004
Figure 5. Control region of I. anguis mitogenome. (a) Structure elements of CR. (b) Stem-loop structures in the CR of I. anguis. The two-sided nucleotide motifs of stem-loop (TA)n, CAT, C(T)nA are specified by squares.
Figure 5. Control region of I. anguis mitogenome. (a) Structure elements of CR. (b) Stem-loop structures in the CR of I. anguis. The two-sided nucleotide motifs of stem-loop (TA)n, CAT, C(T)nA are specified by squares.
Genes 14 00965 g005
Figure 6. Phylogenetic tree of superfamily Nemouroidea based on the mitogenomes of 32 stoneflies using BI (Bayesian inference) and ML (Maximum Likelihood). Numbers at nodes represent posterior probabilities (left) and bootstrap values (right). Families are written to the right side of each species. Pseudomegarcys japonica, Isoperla eximia served as outgroup species.
Figure 6. Phylogenetic tree of superfamily Nemouroidea based on the mitogenomes of 32 stoneflies using BI (Bayesian inference) and ML (Maximum Likelihood). Numbers at nodes represent posterior probabilities (left) and bootstrap values (right). Families are written to the right side of each species. Pseudomegarcys japonica, Isoperla eximia served as outgroup species.
Genes 14 00965 g006
Table 1. Common information of Nemouroidea species analyzed in this study.
Table 1. Common information of Nemouroidea species analyzed in this study.
FamilySpeciesNumber (bp)GenBank NumberReferences
NemouridaeSphaeronemoura grandicauda15,661MH085454[20]
Sphaeronemoura acutispina15,016MH085455 *[35]
Sphaeronemoura hainana15,260MK111420 *[17]
Sphaeronemoura elephas15,846MN944385[36]
Mesonemoura tritaenia15,778MH085451[37]
Mesonemoura metafiligera15,739MH085450[37]
Protonemura kohnoae15,707MH085452[20]
Protonemura datongensis15,756MT276842[38]
Protonemura orbiculata15,758MH085453[20]
Amphinemura claviloba15,707MN720741[39]
Amphinemura bulla15,827MW339348[40]
Amphinemura longispina15,709MH085446[20]
Indonemoura nohirae15,738MH085449[20]
Indonemoura jacobsoni15,642MH085448[20]
Nemoura papilla15,774MK290826[41]
Nemoura nankinensis16,602KY940360[42]
Nemoura meniscata15,895MN944386[43]
NotonemouridaeNeonemura barrosi14,852MK111418 *[17]
CapniidaeMesocapnia daxingana15,524KY568983 *[27]
Mesocapnia arizonensis14,921KP642637 *[44]
Capnia zijinshana16,310KX094942[28]
Apteroperla tikumana15,564NC_027698[35]
Zwicknia bifrons15,380MT872688*[35]
Isocapnia anguis16,200OQ735414This study
TaeniopterygidaeDoddsia occidentalis16,020MG589787[19]
Strophopteryx fasciata15,527ON500674[24]
Taeniopteryx ugola15,353MG589786[19]
LeuctridaeParaleuctra cercia15,625MK492251[21]
Rhopalopsole bulbifera15,599MK111419 *[17]
Perlomyia isobeae15,795MK492252[21]
Perlodidae(Outgroup)Isoperla eximia16,034MG910457[45]
Pseudomegarcys japonica16,067MG910458[45]
* Incomplete mitogenome sequence.
Table 2. The nucleotide components of the I. anguis mitogenome.
Table 2. The nucleotide components of the I. anguis mitogenome.
FeatureProportion of NucleotidesNo. of
Nucleotides
T%C%A%G%AT%GC%AT SkewGC Skew
Genome31.2820.6434.9113.1766.1933.810.05−0.2216,200
Protein coding genes37.6318.6227.1516.6064.7835.22−0.16−0.0611,219
First position36.5818.6127.3817.4163.9636.02−0.14−0.033739
Second position37.9418.0027.1416.9065.0834.90−0.170.033739
Third position38.3719.2526.9015.4665.2734.71−0.18−0.113739
PCG−J34.3422.5528.6114.5062.9537.05−0.09−0.226904
First codon position28.6722.4234.0114.9062.6837.320.09−0.202302
Second codon position35.1622.5624.1218.1759.2840.72−0.19−0.112301
Third codon position39.2022.6927.6810.4366.8833.12−0.17−0.372301
PCG−N40.8814.8225.6418.6766.5233.48−0.230.125609
First codon position40.9112.6226.6819.7967.5932.41−0.210.221870
Second codon position40.7717.7120.6520.8761.4238.58−0.330.081869
Third codon position40.9614.1229.5715.3570.5329.47−0.160.041870
tRNA genes34.2413.6934.7817.2969.0230.980.010.121475
rRNA genes37.5110.4533.6918.3571.2028.80−0.050.272125
lrRNA37.669.0834.8118.4572.4727.53−0.040.341333
srRNA37.2512.7531.8218.1869.0730.93−0.080.18792
CR30.0820.6736.5412.7166.6233.380.10−0.241393
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

Rehman, A.; Huo, Q.-B.; Du, Y.-Z. The First Complete Mitochondrial Genome of Genus Isocapnia (Plecoptera: Capniidae) and Phylogenetic Assignment of Superfamily Nemouroidea. Genes 2023, 14, 965. https://doi.org/10.3390/genes14050965

AMA Style

Rehman A, Huo Q-B, Du Y-Z. The First Complete Mitochondrial Genome of Genus Isocapnia (Plecoptera: Capniidae) and Phylogenetic Assignment of Superfamily Nemouroidea. Genes. 2023; 14(5):965. https://doi.org/10.3390/genes14050965

Chicago/Turabian Style

Rehman, Abdur, Qing-Bo Huo, and Yu-Zhou Du. 2023. "The First Complete Mitochondrial Genome of Genus Isocapnia (Plecoptera: Capniidae) and Phylogenetic Assignment of Superfamily Nemouroidea" Genes 14, no. 5: 965. https://doi.org/10.3390/genes14050965

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