Next Article in Journal
Evaluation of Total Female and Male Aedes aegypti Proteomes Reveals Significant Predictive Protein–Protein Interactions, Functional Ontologies, and Differentially Abundant Proteins
Previous Article in Journal
Influence of Temperature and Photoperiod on the Fecundity of Habrobracon hebetor Say (Hymenoptera: Braconidae) and on the Paralysis of Host Larvae, Plodia interpunctella (Hübner) (Lepidoptera: Pyralidae)
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Mitochondrial Genomes of Hestina persimilis and Hestinalis nama (Lepidoptera, Nymphalidae): Genome Description and Phylogenetic Implications

1
School of Environmental Science and Engineering, Taiyuan University of Science and Technology, Taiyuan 030024, China
2
College of Plant Protection, Shanxi Agricultural University, Taiyuan 030031, China
3
Department of Horticulture, Taiyuan University, Taiyuan 030012, China
4
College of Life Sciences, Anhui Normal University, Wuhu 241000, China
*
Authors to whom correspondence should be addressed.
Submission received: 15 July 2021 / Revised: 8 August 2021 / Accepted: 18 August 2021 / Published: 20 August 2021

Abstract

:

Simple Summary

In this study, the mitogenomes of Hestina persimilis and Hestinalis nama were obtained via sanger sequencing. Compared with other mitogenomes of Apaturinae butterflies, conclusions can be made that the mitogenomes of Hestina persimilis and Hestinalis nama are highly conservative. The phylogenetic trees build upon mitogenomic data showing that the relationships among Nymphalidae are similar to previous studies. Hestinalis nama is apart from Hestina, and closely related to Apatura, forming a monophyletic clade.

Abstract

In this study, the complete mitochondrial genomes (mitogenomes) of Hestina persimilis and Hestinalis nama (Nymphalidae: Apaturinae) were acquired. The mitogenomes of H. persimilis and H. nama are 15,252 bp and 15,208 bp in length, respectively. These two mitogenomes have the typical composition, including 37 genes and a control region. The start codons of the protein-coding genes (PCGs) in the two mitogenomes are the typical codon pattern ATN, except CGA in the cox1 gene. Twenty-one tRNA genes show a typical clover leaf structure, however, trnS1(AGN) lacks the dihydrouridine (DHU) stem. The secondary structures of rrnL and rrnS of two species were predicted, and there are several new stem loops near the 5′ of rrnL secondary structure. Based on comparative genomic analysis, four similar conservative structures can be found in the control regions of these two mitogenomes. The phylogenetic analyses were performed on mitogenomes of Nymphalidae. The phylogenetic trees show that the relationships among Nymphalidae are generally identical to previous studies, as follows: Libytheinae\Danainae + ((Calinaginae + Satyrinae) + Danainae\Libytheinae + ((Heliconiinae + Limenitidinae) + (Nymphalinae + (Apaturinae + Biblidinae)))). Hestinalis nama is apart from Hestina, and closely related to Apatura, forming monophyly.

1. Introduction

Hestina persimilis and Hestinalis nama belong to the lepidopteran family Nymphalidae, Apaturinae and mainly distribute in the Palaearctic and Oriental region. Their adults inhabit mountainous broad-leaved forests and present the habit of sipping tree juice and water in wetlands. The larvae were reported as a kind of agriculture pest of the host plants, Ulmaceae. At present, there are only one species (H. nama) of the genus Hestinalis and three species (including Hestina persimilis, Hestina assimilis and Hestina nicevillei) of the genus Hestina distributed in China. Hestinalis nama was originally described as Diadema nama by Doubleday in 1844 [1]. However, in subsequent studies, latter scholars placed it under the genus Hestina [2,3]. The classification of butterflies is mainly based on the characteristics of external genitalia and wing veins. Morphologically, including genitalic structure, Hestina and Hestinalis are easily separable [4]. In recent studies, Hestinalis was treated as a distinct genus [5]. Although most modern literature chooses to separate them, some literature, Wu and Hsu still treats them as one [6]. In this paper, phylogenetic analysis shows that Hestinalis nama is apart from Hestina. Therefore, we also separate them apart.
Mitogenome fragments have been extensively used in phylogenetic analysis for butterflies and moths, particularly for the cox1 gene which was primarily used as a DNA barcoding for animals [7,8,9,10,11]. In the BOLD system [12], lepidopteran insects consist of the largest amount of data being sequenced. However, the phylogenetic relationships at different taxonomic levels are still controversial [13,14,15]. It has been proposed that mitochondrial genomes might provide more genetic information than a single gene fragment [16,17,18]. Therefore, sequencing more mitogenomes might improve our understanding of evolution and phylogeny at different taxonomic levels in Lepidoptera. Furthermore, the mitogenome has been widely used in the areas of population genetic structure, gene drift and phylogenetics, because of its characteristics of maternal inheritance, small genome size (15–20 kb in length) and rapid rate of evolution [19,20]. To date, only one complete mitogenome (H. assimilis) has been sequenced from the genus Hestina; other species and Hestinalis nama were all not sequenced, which is quite limited and will restrict our understanding of evolution in Nymphalidae at the genomic level. In this study, the mitogenomes of H. persimilis and H. nama were obtained, with the aim of: (1) providing a comparative analysis of Apaturinae mitogenomes, including nucleotide composition, codon usage, gene arrangement, prediction of tRNA and rRNA secondary structures and novel features of the control region, and (2) reconstructing the phylogenetic relationships among subfamilies in Nymphalidae based on mitogenomes data.

2. Materials and Methods

2.1. Sampling and DNA Sequencing

Specimens of H. persimilis and H. nama were collected from the Sichuan and Yunnan Provinces of China in 2010. Specimens were being made, followed by morphological identification. One side of the hindfoot for each sample was preserved in absolute ethanol and stored in −20 °C freezer in College of Plant Protection, Shanxi Agricultural University, Taiyuan, China.
The DNA extraction kit and primers [21] (Table S1) were produced by Shanghai Major Biomedical Technology Co., Ltd. (Shanghai, China). The reaction systems of PCR amplifications were 25 μL, including upstream and downstream primers 0.5 μL, respectively, PCR Master Mix 12.5 μL, DNA template 3 μL, and ddH2O 8.5 μL. The amplification reaction conditions were as follows: initial denaturation at 94 °C for 2 min; 35 cycles of denaturation at 94 °C for 1 min, annealing at 53 °C for 45 s, extension at 72 °C for 1 min, and a final extension step at 72 °C for 4 min. PCR products were detected by 1% agarose gel electrophoresis. All the gene fragments were sent to Shanghai Major Biomedical Technology for sequencing.

2.2. Annotation and Analysis of Mitochondrial DNA

The original sequence fragments were assembled with SeqMan (Steve ShearDown, 1998–2001 version reserved by DNASTAR Inc., Madison, WI, USA) to get a complete mitogenome. The secondary structure of tRNA genes were determined by tRNAscan-SE Search Server (http://lowelab.ucsc.edu/tRNAscan-SE/; accessed on 28 June 2021) [22]. Putative tRNA genes, including trnH and trnS1(AGN), which could not be found by tRNAscanSE, were confirmed by comparison with the homologous genes of other Apaturinae species. PCGs and rRNA genes were identified by the MITOS webserver with invertebrate genetic code [23]. The nucleotide composition and codon usage of PCGs were calculated with MEGA-X [24]. Determination of tandem repeat sequences in control regions were performed using the Tandem Repeats Finder online software (http://tandem.bu.edu/trf/trf.html; accessed on 10 May 2020) [25]. The mitogenomes of H. persimilis and H. nama were uploaded to GenBank, with the accession numbers of MT110153 and MT110154.

2.3. Phylogenetic Analysis

Phylogenetic analysis was performed on the dataset of 13 PCGs from 54 complete or nearly complete mitogenomes of Nymphalidae, with two Papilionidae species selected as outgroups (Table S2). All assembled PCGs of 56 mitogenomes were aligned through MEGA-X. The optimal partition tactics and substitution models were selected by PartitionFinder v2 (Tables S3 and S4) [26,27,28]. The maximum likelihood (ML) and Bayesian inference (BI) analyses were conducted through the online CIPRES Science Gateway [29]. The ML analysis was performed with RAxML-HPC2 on XSEDE [30], with GTRGAMMA model applied to all partitions. Bootstrap values were estimated with 1000 replicates. The BI analyses were carried out through two independent Markov chain Monte Carlo (MCMC) chains, which were set for 1,000,000 generations, with sampling per 1000 generations.

3. Results and Discussion

3.1. Mitogenomes Organization

The complete mitogenomes of H. persimilis and H. nama are 15,252 and 15,208 bp. They share the consistent gene organization, order and arrangement with most of other lepidopterans, including 13 PCGs, 22 tRNAs and 2 rRNAs (rrnL and rrnS) (Figure 1 and Figure 2). The mitogenome is circular with two strands. The heavy strand (H-strand) encodes most of the genes (9 PCGs and 14 tRNAs), while the light strand (L-strand) contains the remaining reverse complementary genes (four PCGs, eight tRNAs and two rRNAs), as shown in Table 1 and Table 2. In addition, the nucleotide composition of the two species are both AT-biased, similar to other lepidopterans. The AT contents of the mitogenomes of H. persimilis and H. nama are 80.9 and 79.2%, respectively (Table 3 and Table 4). The obvious AT-biased (Table S5) is generally believed to be related to the evolution of mitochondrial origin [31].

3.2. Protein Coding Genes and Codon Usage

Orthologs from the two Hestina mitogenomes present similar start and stop codons. Most PCGs start with the typical initial codon ATN, but cox1 initiates with CGA. In particular, the putative start codon CGA in the cox1 gene is a common feature of most sequenced lepidopterans, but a few species start with codon ATG, ATT, ATA or TTG. While most PCGs end with the stop codon TAA or TAG, truncated codon T is also detected in cox2 and nad4. It has been proposed that truncated stop codons can be completed by polyadenylation, which was also found in other insectan mitogenomes [32].
Relative synonymous codon usage (RSCU) can directly reflect the preference of codon usage (Table S6). The total number of codons of 13 PCGs are 3703 for H. persimilis and 3709 for H. nama. The RSCU of twelve Apaturinae species show the same codon preference pattern (Figure 3). The mainly used codon families are Leu1 (CUN), Ile, Phe and Met. There are at least 75 codons (CDs) per thousand CDs in each of them (Figure 4), among which Leu1 has the highest utilization rate. Relative synonymous codon usage (RSCU) of Apaturinae show that degeneration codons are biased to use more A/T than G/C. The six most prevalent codons in Apaturinae, including AUU (I), AUA (M), AAU (N), UUU (F), UUA (L) and UAU (Y), are all composed of A and/or T. Conversely, some GC-rich codons are seldom utilized in the Apaturinae species. For example, the codon UCG, CCG are not used in H. persimilis, while CUG, GUC and CCG are absent in Sasakia charonda. This phenomenon is common among lepidopterans [33,34], which indicates that the GC content of genes is closely related to codon preference [35,36].

3.3. Transfer RNAs and Ribosomal RNAs

All 22 tRNAs typical of lepidopteran mitogenomes are found in the mitogenomes. Most tRNA genes are in classical clover-leaf secondary structures except for trnS1(AGN), with its DHU arm forming a simple loop, which is considered as a typical feature in metazoan mitogenomes (Figure 5) [37]. Additionally, the anticodon stem of trnS1(AGN) may be shortened as of base mismatch in some insect mitogenomes [38]. Previous studies showed that not only trnS1(AGN), but also some other tRNAs, such as trnS2(UCN) and trnG, lack a DHU or TΨC arm [39]. Missing a DHU arm and base mismatch are thermodynamically unstable, which indicate that a DHU arm might not really exist. Accordingly, this special structure of trnS1(AGN) still needs further investigation. In addition, it has been shown that some isoforms of tRNAs can be found in control regions or some PCGs on the L-strand of mitogenomes. The isoforms of tRNAs can also be folded into cloverleaf structures. However, it is not clear whether their functions are similar to those of tRNAs [40] or not.
The rrnL gene is found between trnL (CUN) and trnV, while the rrnS gene is located between trnV and the control region. The lengths of rrnL genes of H. persimilis and H. nama mitogenomes are 1334 bp (AT content 84.29%) and 1327 bp (AT content 83.44%). The sizes of rrnS genes of H. persimilis and H. nama are 776 bp (AT content 85.03%) and 774 bp (AT content 84.39%). The secondary structure of rrnL genes include six structural domains except for that domain III is absent in arthropods (Figure 6). The rrnS genes include three structural domains (Figure 7). Both the secondary structures of rrnL and rrnS of the two species are roughly similar to other lepidopterans, such as Amata emma, Apis mellifera [41], Grapholita molesta [42], Manduca sexta [43], etc. The microsatellite sequence (TA)n is not found in rrnL and rrnS, but exists in other insects (e.g., Choristoneura longicellana). There are several new stem loops near the 5′ of rrnL secondary structure, and these loops were not found in other insects.
Dashes, black dots and circles indicate the Watson-Crick base pairings, G-U bonds and U-U, A-A, A-C and A-G bonds, respectively.

3.4. Intergenic and Overlapping Regions

It has been proposed that mitogenomes tend to be highly economized in size by eliminating or reducing intergenic spacers [44]. However, by excluding the control region, 12 intergenic spacers (1 to 91 bp, 150 bp in total) are found in H. persimilis, and 11 intergenic spacers (1 to 69 bp, 118 bp in total) are found in H. nama. It has been reported that Lepidopteran mitogenomes usually have two typical and relatively conservative intergenic spacers. The longer one is located between trnQ and nad2 genes, with the length of 91 and 69 bp in H. persimilis and H. nama. Previous studies found that the nucleotide sequence of the trnQ-nad2 spacer and the nad2 gene have a highly similarity. It has been inferred that the trnQ-nad2 spacer may come from the nad2 gene [45]. The other shorter spacer is located between trnS2(UCN) and nad1 genes, with the length of 22 and 13 bp in H. persimilis and H. nama, respectively, sharing a conserved sequence of ATACTAA.
Comparing with the intergenic spacers, the overlapping regions are more conservative [46]. Fourteen overlapping spacers (1 to 26 bp, 66 bp in total) are found in H. persimilis, and fourteen overlapping spacers (1 to 8 bp, 42 bp in total) are found in H. nama. ATP8 and ATP6 overlap with the ATGATAA motif in the two mitogenomes, which had also been reported in many other lepidopterans [47].

3.5. Putative Control Regions

The control region, also known as the A + T-region or D-loop, is always the largest intergenic spacer in animal mitogenomes and considered as the initial region for replication [48]. The control regions (376 bp in H. persimilis and 390 bp in H. nama) in the two mitogenomes are located between rrnS and trnM. The AT content is also the highest in mitogenomes (91.23% in H. persimilis and 88.72% in H. nama). Previous studies indicated that the control region is the segment with fastest evolutionary rate and can be used as an important molecular marker for animal population genetics.
There are generally four conserved structures in the control region, including a motif of ATAGA located at downstream of rrnS followed by 19 bp Poly-T stretch, a poly-A stretches (9 bp in H. persimilis and 6 bp in H. nama) at the upstream of trnM (Figure 8 and Figure 9), the microsatellite-like repeat regions ((AT)10 in H. persimilis and (AT)6 in H. nama), and the repeated sequences (23 bp in H. persimilis and 25 bp in H. nama). All these characteristics are generally considered to be related to the transcription or replication of mitogenomes [49]. Although the location of initial replication region in complete metamorphosis insects (including lepidopterans) are different, they all located after polyT (about 10–20 bp) (Figure 10) [50]. Accordingly, polyT may be involved in the recognition of the initial replication region [51].

3.6. Phylogenetic Analysis

The phylogenetic analyses are performed on concatenated nucleotide sequences of 13 PCGs (data matrix 10,472 bp) derived from 54 available Nymphalidae mitogenomes, with two Papilionidae species serving as outgroups. These 54 sequences represent 9 subfamilies: Apaturinae; Biblidinae; Calinaginae; Danainae; Heliconiinae; Libytheinae; Limenitidinae; Nymphalinae and Satyrinae. The BI (Figure 11) and ML trees (Figure 12) have roughly the same topology, except for Danainae and Libytheinae, which are in different locations. The overall relationship is generally as follows: Libytheinae\Danainae + ((Calinaginae + Satyrinae) + Danainae\Libytheinae + ((Heliconiinae + Limenitidinae) + (Nymphalinae + (Apaturinae + Biblidinae)))). Earlier, Heliconiinae, Limenitidinae, Biblidinae and Apaturinae were considered as “core nymphalids” [52,53]. Subsequently, there are ten widely recognized subfamilies in Nymphalidae, including Apaturinae, Libytheinae, Danainae, Morphinae, Satyrinae, Calinaginae, Eliconiinae, Limenitidinae, Charaxinae and Nymphalinae [54]. Although mitogenomes of the subfamilies Morphinae, Eliconiinae and Charaxinae are not yet available, the topology is generally identical to those of other studies [55,56,57]. Hestina and Euripus, Sasakia come together to form a branch; Apatura and Hestinalis form a branch; Chitoria, Timelaea and Herona form a branch. Early on, Apatura included Chitoria [58], while in this paper, Chitoria is distant from Apatura. Ohshima et al. (2010) has published an updated work on Apaturinae butterflies, considering that Chitoria should been removed from Apatura [59], their grouping is concordant to this paper. H. persimilis, H. assimilis and the genus Euripus (represented by Euripus nyctelius) form a branch, of which the result is the same as the morphological study. Hestinalis is a clearly distinct lineage within the genus Hestina, not together with H. persimilis and H. assimilis, and closely related to Apatura, that is considered to be a result of mimicry [60], which is also consistent with our paper.

4. Conclusions

The mitogenomes of H. persimilis and H. nama were obtained using sanger sequencing. Comparing them with other mitogenomes of Apaturinae butterflies, the conclusion can be drawn that the mitogenomes are highly conserved, sharing the same gene order, gene location, codon usage, nucleotide composition and AT-biased pattern. The secondary structures of rrnL and rrnS of two species are roughly similar to other lepidopterans. Although the control regions vary greatly in length, their structure has not changed much, which includes four basic conservative regions. The topology of phylogenetic analyses are generally identical to those of other studies. Hestinalis nama is not grouped with Hestina, and is closely related to Apatura, which is consistent with early studies.

Supplementary Materials

The following are available online at https://0-www-mdpi-com.brum.beds.ac.uk/article/10.3390/insects12080754/s1, Figure S1: phylogenetic relationships by BI analysis based on the 13PCGs. Figure S2: phylogenetic relationships by ML analysis based on the 13PCGs. Table S1: sequences of 22 pairs of primers. Table S2: list of species used to construct the phylogenetic tree. Table S3: the starting partitions used to initiate the PartitionFinder analysis. Table S4: evolutionary models from partition strategies start scheme used in the phylogenetic analysis. Table S5: bias of base composition of protein-coding genes in the two mitogenomes. Table S6: codon usage of protein genes of twelve mitogenomes in Apaturinae.

Author Contributions

Conceptualization, Y.W. and T.C.; specimen collection and identification, J.W. (Jiping Wen), J.W. (Juping Wang) and T.C.; methodology, Y.W. and H.F.; software, Y.W., B.H. and H.F.; analysis, J.W. (Juping Wang), Y.W. and H.F.; writing—original draft preparation, Y.W.; writing—review and editing, Y.W. and B.H.; funding acquisition, Y.W. and T.C. All authors have read and agreed to the published version of the manuscript.

Funding

This work was supported by the Science and Technology basic condition platform of Shanxi Province (201605D121020), the Innovational Research Project in Shanxi Academy of Agricultural Sciences, China (YGJPY1908), the open fund of Shanxi Key Laboratory of Integrated Pest Management in Agriculture (YHSW2019001).

Institutional Review Board Statement

Not applicable.

Informed Consent Statement

Not applicable.

Data Availability Statement

All sequences were deposited in the GenBank under accession numbers of MT110153 and MT110154.

Acknowledgments

We sincerely express our gratitude to Tianjuan Su (School of Life Sciences, Jinggangshan University, Jinggangshan, China) for reviewing the manuscript. Thanks to Chunsheng Wu and Qianju Jia (Key Laboratory of Zoological Systematics and Evolution (CAS), Institute of Zoology, Chinese Academy of Sciences, Beijing, China) for the literature inquiry and reviewing the manuscript.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Doubleday, E. List of the Specimens of Lepidopterous Insects in the Collection of the British Museum; Trustees of the British Museum: London, UK, 1844; Volume 1, p. 97. [Google Scholar]
  2. Chou, I. Monograph of Chinese Butterflies; Henan Scientific Technological Publishing House: Zhengzhou, China, 1994; pp. 451–456. [Google Scholar]
  3. Lee, C.L.; Zhu, B.Y. Atlas of Chinese Butterflies; Shanghai Far East Publishers: Shanghai, China, 1992; p. 136. [Google Scholar]
  4. Masui, A.; Bozano, G.C.; Floriani, A. Guide to the Butterflies of the Palearctic Region: Nymphalidae 4: Apaturinae; Omnes Artes: Milan, Italy, 2011; Volume 4, pp. 45–102. [Google Scholar]
  5. Lang, S.Y. The Nymphalidae of China (Lepidoptera, Rhopalocera); Tshikolovets Publications: Pardubice, Czech Republic, 2012; Volume 1, p. 21. [Google Scholar]
  6. Wu, C.S.; Hsu, Y.F. Butterflies of China; The straits Publishing &Distributing Group: Fuzhou, China, 2017; pp. 871–874. [Google Scholar]
  7. Simmons, R.B.; Weller, S.J. Utility evolution of cytochrome b in insects. Mol. Phylogenet. Evol. 2001, 20, 196–210. [Google Scholar] [CrossRef] [Green Version]
  8. Hebert, P.D.N.; Cywinska, A.; Ball, S.L.; de Waard, J.R. Biological identifications through DNA barcodes. Proc. R. Soc. Lond. B Bio. 2003, 270, 313–321. [Google Scholar] [CrossRef] [Green Version]
  9. Regier, J.G.; Mitter, C.; Zwick, A.; Bazinet, A.L.; Cummings, M.P.; Kawahara, A.Y.; Sohn, J.C.; Zwickl, D.J.; Cho, S.; Davis, D.R.; et al. A large–scale, higher–level, molecular phylogenetic study of the insect order Lepidoptera (moths butterflies). PLoS ONE 2013, 8, e58568. [Google Scholar] [CrossRef] [PubMed]
  10. Wu, Y.P.; Zhao, J.L.; Su, T.J.; Luo, A.R.; Zhu, C.D. The complete mitochondrial genome of Choristoneura longicellana (Lepidoptera: Tortricidae) phylogenetic analysis of Lepidoptera. Gene 2016, 591, 161–176. [Google Scholar] [CrossRef] [PubMed]
  11. Dong, W.W.; Dong, S.Y.; Jiang, G.F.; Huang, G.H. Characterization of the complete mitochondrial genome of tea tussock moth, Euproctics pseudoconspersa (Lepidoptera: Lymantriidae) its phylogenetic implications. Gene 2016, 577, 37–46. [Google Scholar] [CrossRef] [PubMed]
  12. Ratnasingham, S.; Hebert, P.D.N. Bold: The Barcode of Life Data System. Mol. Ecol. Notes 2007, 7, 355–364. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  13. Lafontaine, J.D.; Fibiger, M. Revised higher classification of the Noctuoidea (Lepidoptera). Can. Entomol. 2006, 138, 610–635. [Google Scholar] [CrossRef]
  14. Mitter, C.; Davis, D.R.; Cummings, M.P. Phylogeny Evolution of Lepidoptera. Annu. Rev. Entomol. 2017, 62, 265–283. [Google Scholar] [CrossRef]
  15. Yang, M.S.; Song, L.; Shi, Y.X.; Li, J.H.; Zhang, Y.L.; Song, N. The first mitochondrial genome of the family Epicopeiidae higher–level phylogeny of Macroheterocera (Lepidoptera: Ditrysia). Int. J. Biol. Macromol. 2019, 136, 123–132. [Google Scholar] [CrossRef]
  16. Mueller, R.L. Evolutionary rates, divergence dates, the performance of mitochondrial genes in bayesian phylogenetic analysis. Syst. Biol. 2006, 55, 289–300. [Google Scholar] [CrossRef] [Green Version]
  17. Timmermans, M.J.T.N.; Lees, D.C.; Simonsen, T.J. Towards a mitogenomic phylogeny of Lepidoptera. Mol. Phylogenet. Evol. 2014, 79, 169–178. [Google Scholar] [CrossRef]
  18. Liu, N.Y.; Li, N.; Yang, P.Y.; Sun, C.Q.; Fang, J.; Wang, S.Y. The complete mitochondrial genome of Damora sagana phylogenetic analyses of the family Nymphalidae. Genes Genom. 2018, 40, 109–122. [Google Scholar] [CrossRef]
  19. Habib, M.; Lakra, W.S.; Mohindra, V.; Khare, P.; Barman, A.S.; Singh, A.; Lal, K.K.; Punia, P.; Khan, A.A. Evaluation of cytochrome b mtDNA sequences in genetic diversity studies of Channa marulius (Channidae: Perciformes). Mol. Biol. Rep. 2011, 38, 841–846. [Google Scholar] [CrossRef]
  20. Chris, S.; Thomas, R.B.; Francesco, F.; James, B.S.; Andrew, T.B. Incorporating molecular evolution into phylogenetic analysis, a new compilation of conserved polymerase chain reaction primers for animal mitochondrial DNA. Annu. Rev. Ecol. Evol. Syst. 2006, 37, 545–579. [Google Scholar]
  21. Simon, C.; Frati, F.; Bekenbach, A.; Crespi, B.; Liu, H.; Flook, P. Evolution, weighting, phylogenetic utility of mitochondrial genesequences a compilation of conserved polymerase chain reaction primers. Ann. Entomol. Soc. Am. 1994, 87, 651–701. [Google Scholar] [CrossRef]
  22. Lowe, T.M.; Eddy, S.R. tRNAscan–SE: A program for improved detection of transfer RNA genes in genomic sequence. Nucleic Acids Res. 1997, 25, 955–964. [Google Scholar] [CrossRef] [PubMed]
  23. Bernt, M.; Donath, A.; Jühling, F.; Externbrink, F.; Florentz, C.; Fritzsch, G.; Pütz, J.; Middendorf, M.; Stadler, P.F. MITOS: Improved de novo etazoan mitochondrial genome annotation. Mol. Phylogenet. Evol. 2013, 69, 313–319. [Google Scholar] [CrossRef]
  24. 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]
  25. Benson, G. Tandem repeats finder: A program to analyze DNA sequences. Nucleic Acids Res. 1999, 27, 573–580. [Google Scholar] [CrossRef] [Green Version]
  26. Lanfear, R.; Calcott, B.; Ho, S.Y.W.; Guindon, S. Partitionfifinder: Combined selection of partitioning schemes substitution models for phylogenetic analyses. Mol. Biol. Evol. 2012, 29, 1695–1701. [Google Scholar] [CrossRef] [Green Version]
  27. Lanfear, R.; Frandsen, P.B.; Wright, A.M.; Senfeld, T.; Calcott, B. PartitionFinder 2: New methods for selecting partitioned models of evolution for molecular morphological phylogenetic analyses. Mol. Biol. Evol. 2017, 34, 772–773. [Google Scholar] [CrossRef] [Green Version]
  28. Guindon, S.; Dufayard, J.F.; Lefort, V.; Anisimova, M.; Hordijk, W.; Gascuel, O. New algorithms methods to estimate maximum–likelihood phylogenies: Assessing the performance of PhyML 3.0. Syst. Biol. 2010, 59, 307–321. [Google Scholar] [CrossRef] [Green Version]
  29. Miller, M.A.; Pfeiffer, W.T.; Schwartz, T. Creating the CIPRES Science Gateway for Inference of Large Phylogenetic Trees; Gateway Computing Environments Workshop (GCE): New Orleans, LA, USA, 2010; pp. 1–7. [Google Scholar]
  30. Stamatakis, A. RAxML version 8: A tool for phylogenetic analysis post–analysis large phylogenies. Bioinformatics 2014, 30, 1312–1313. [Google Scholar] [CrossRef]
  31. Rodríguez–Trelles, F.; Tarrío, R.; Ayala, F.J. Fluctuating mutation bias the evolution of base composition in Drosophila. J. Mol. Evol. 2000, 50, 1–10. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  32. Clary, D.O.; Wolstenholme, D.R. The mitochondrial DNA molecular of Drosophila yakuba: Nucleotide sequence, gene organization, genetic code. J. Mol. Evol. 1985, 22, 252–271. [Google Scholar] [CrossRef] [PubMed]
  33. Lu, H.F.; Su, T.J.; Luo, A.R.; Zhu, C.D.; Wu, C.S. Characterization of the complete mitochondrion genome of Diurnal moth Amata emma (Butler) (Lepidoptera: Erebidae) its phylogenetic implications. PLoS ONE 2013, 8, e72410. [Google Scholar] [CrossRef] [PubMed]
  34. Yuan, M.L.; Zhang, Q.L.; Guo, Z.L.; Wang, J.; Shen, Y.Y. The complete mitochondrial genome of Corizus tetraspilus (Hemiptera: Rhopalidae) phylogenetic analysis of Pentatomomorpha. PLoS ONE 2015, 10, e0129003. [Google Scholar]
  35. Hershberg, R.; Petrov, D.A. Selection on codon bias. Annu. Rev. Genet. 2008, 42, 287–299. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  36. Plotkin, J.B.; Kudla, G. Synonymous but not the same: The causes consequences of codon bias. Nat. Rev. Genet. 2011, 12, 32–42. [Google Scholar] [CrossRef] [Green Version]
  37. Boore, J.L. Animal mitochondrial genomics. Nucleic Acids Res. 1999, 27, 1767–1780. [Google Scholar] [CrossRef] [Green Version]
  38. Yong, H.S.; Song, S.L.; Lim, P.E.; Eamsobhana, P.; Suana, I.W. Complete Mitochondrial Genome of Three Bactrocera Fruit Flies of Subgenus Bactrocera (Diptera: Tephritidae) and Their Phylogenetic Implications. PLoS ONE 2016, 11, e0148201. [Google Scholar] [CrossRef] [PubMed]
  39. Zhang, K.J.; Zhu, W.C.; Rong, X.; Liu, J.; Ding, X.L.; Hong, X.Y. The complete mitochondrial genome sequence of Sogatella furcifera (Horváth) a comparative mitogenomic analysis of three predominant rice planthoppers. Gene 2014, 533, 100–109. [Google Scholar] [CrossRef]
  40. Nina, V.V.; Sofiya, L.; Derek, W.; Raman, S.; Yury, B.; Dmitrii, Z. Characteristic variability of five complete aphid mitochondrial genomes: Aphis fabae mordvilkoi, Aphis craccivora, Myzus persicae, Therioaphis tenera and Appendiseta robiniae (Hemiptera; Sternorrhyncha; Aphididae). Int. J. Biol. Macromol. 2020, 149, 187–206. [Google Scholar]
  41. Gillespie, J.J.; Johnston, J.S.; Cannone, J.J.; Gutell, R.R. Characteristics of the nuclear (18S, 5.8S, 28S and 5S) and mitochondrial (12S and 16S) rRNA genes of Apis mellifera (Insecta: Hymenoptera): Structure, organization, and retrotransposable elements. Insect Mol. Biol. 2006, 15, 657–686. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  42. Gong, Y.J.; Shi, B.C.; Kang, Z.J.; Zhang, F.; Wei, S.J. The complete mitochondrial genome of the oriental fruit moth Grapholita molesta (Busck) (Lepidoptera: Tortricidae). Mol. Biol. Rep. 2011, 39, 2893–2900. [Google Scholar] [CrossRef] [Green Version]
  43. Moritz, C.; Dowling, T.E.; Brown, W.M. Evolution of animal mitochondrial DNA: Relevance for population biology systematics. Annu. Rev. Ecol. Syst. 1987, 18, 269–292. [Google Scholar] [CrossRef]
  44. Sheffield, N.C.; Song, H.; Cameron, S.L.; Whiting, M.F. A comparative analysis of mitochondrial genomes in coleoptera genome descriptions of six new beetles. Mol. Biol. Evol. 2008, 25, 2499–2509. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  45. Wu, Y.P.; Zhao, J.L.; Su, T.J.; Li, J.; Yu, F.; Chesters, D.; Fan, R.J.; Chen, M.C.; Wu, C.S.; Zhu, C.D. The Complete mitochondrial genome of Leucoptera malifoliella Costa (Lepidoptera: Lyonetiidae). DNA Cell Biol. 2012, 31, 1508–1522. [Google Scholar] [CrossRef] [Green Version]
  46. Cao, Y.Q.; Ma, C.; Chen, J.Y.; Yang, D.R. The complete mitochondrial genomes of two ghost moths, Thitarodes renzhiensis and Thitarodes yunnanensis: The ancestral gene arrangement in Lepidoptera. BMC Genom. 2012, 13, 276. [Google Scholar] [CrossRef] [Green Version]
  47. Yang, M.; Shi, S.; Dai, P.; Song, L.; Liu, X. Complete mitochondrial genome of Palpita hypohomalia (Lepidoptera: Pyraloidea: Crambidae) and its phylogenetic implications. Eur. J. Entomol. 2018, 115, 708–717. [Google Scholar] [CrossRef]
  48. Tayler, M.F.; McKechnie, S.W.; Pierce, N.; Kreitman, M. The lepidopteran mitochondrial control region: Structure evolution. Mol. Biol. Evol. 1993, 10, 1259–1272. [Google Scholar]
  49. Clayton, D.A. Transcription replication of animal mitochondrial DNAs. Int. Rev. Cyt. 1992, 141, 217–232. [Google Scholar]
  50. Saito, S.; Tamura, K.; Aotsuka, T. Replication origin of mitochondrial DNA in insects. Genetics 2005, 171, 1695–1705. [Google Scholar] [CrossRef] [Green Version]
  51. Ye, W.; Dang, J.P.; Xie, L.D.; Huang, Y. Complete mitochondrial genome of Teleogryllus emma (Orthoptera: Gryllidae) with a new gene order in Orthoptera. Zool. Res. 2008, 29, 236–244. [Google Scholar] [CrossRef] [Green Version]
  52. Ehrlich, P.R. The comparative morphology, phylogeny higher classifification of the butterflflies (Lepidoptera: Papilionoidea). Univ. Kans. Sci. Bull. 1958, 39, 305–370. [Google Scholar]
  53. Scott, J.A. The phylogeny of butterflflies (Papilionoidea Hesperoidea). J. Res. Lepid. 1985, 23, 24–281. [Google Scholar]
  54. Kristensen, N.P.; Scoble, M.J.; Karsholt, O. Lepidoptera phylogeny systematic: The state of inventorying moth butterfly diversity. Zootaxa 2007, 1668, 699–747. [Google Scholar] [CrossRef] [Green Version]
  55. Wahlberg, N.; Weingartner, E.; Nylin, S. Towards a better understanding of the higher systematics of Nymphalidae (Lepidoptera: Papilionoidea). Mol. Phylogenet. Evol. 2003, 28, 473–484. [Google Scholar] [CrossRef]
  56. Wahlberg, N.; Wheat, C.W. Genomic outposts serve the phylogenomic pioneers: Designing novel nuclear markers for genomic DNA extractions of Lepidoptera. Syst. Biol. 2008, 57, 231–242. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  57. Wahlberg, N.; Leneveu, J.; Kodandaramaiah, U.; Peña, C.; Nylin, S.; Freitas, A.V.L.; Brower, A.V.Z. Nymphalid butterflies diversify following near demise at the Cretaceous/Tertiary boundary. Proceedings of the Royal Society of London Series B Biological Sciences. Proc. Biol. Sci. 2009, 276, 4295–4302. [Google Scholar] [PubMed] [Green Version]
  58. Freitas, A.V.L.; Brown, K.S. Phylogeny of the Nymphalidae (Lepidoptera). Syst. Biol. 2004, 53, 363–383. [Google Scholar] [CrossRef] [Green Version]
  59. Zhang, M.; Cao, T.W.; Zhang, R.; Guo, Y.P.; Duan, Y.H.; Ma, E.B. Phylogeny of Apaturinae Butterflies (Lepidoptera: Nymphalidae) based on mitochondrial cytochrome oxidase I gene. J. Genet Genom. 2007, 34, 812–823. [Google Scholar] [CrossRef]
  60. Ohshima, I.; Tanikawadodo, Y.; Saigusa, T.; Nishiyama, T.; Kitani, M.; Hasebe, M.; Mohri, H. Phylogeny, biogeography, host–plant association in the subfamily Apaturinae (Insecta: Lepidoptera: Nymphalidae) inferred from eight nuclear seven mitochondrial genes. Mol. Phylogenet. Evol. 2010, 57, 1026–1036. [Google Scholar] [CrossRef] [PubMed]
Figure 1. Linear map of the mitogenome of Hestina persimilis. The J-strand is located on the linear map, and the N-strand is under the linear map.
Figure 1. Linear map of the mitogenome of Hestina persimilis. The J-strand is located on the linear map, and the N-strand is under the linear map.
Insects 12 00754 g001
Figure 2. Linear map of the mitogenome of Hestinalis nama. The J-strand is located on the linear map, and the N-strand is under the linear map.
Figure 2. Linear map of the mitogenome of Hestinalis nama. The J-strand is located on the linear map, and the N-strand is under the linear map.
Insects 12 00754 g002
Figure 3. Relative Synonymous Codon Usage (RSCU) in the mitogenomes of twelve Apaturinae species. Codon families are provided on the X axis.
Figure 3. Relative Synonymous Codon Usage (RSCU) in the mitogenomes of twelve Apaturinae species. Codon families are provided on the X axis.
Insects 12 00754 g003
Figure 4. Codon distributions in the mitogenomes of twelve Apaturinae species. CDspT: codons per thousand codons.
Figure 4. Codon distributions in the mitogenomes of twelve Apaturinae species. CDspT: codons per thousand codons.
Insects 12 00754 g004
Figure 5. Predicted secondary cloverleaf structure for the tRNAs of Hestina persimilis and Hestinalis nama. Dashes (–) and pluses (+) indicate the Watson–Crick base pairings and G-U bonds, respectively.
Figure 5. Predicted secondary cloverleaf structure for the tRNAs of Hestina persimilis and Hestinalis nama. Dashes (–) and pluses (+) indicate the Watson–Crick base pairings and G-U bonds, respectively.
Insects 12 00754 g005
Figure 6. Predicted rrnL secondary structure of Hestina persimilis and Hestinalis nama mitogenomes.
Figure 6. Predicted rrnL secondary structure of Hestina persimilis and Hestinalis nama mitogenomes.
Insects 12 00754 g006
Figure 7. Predicted rrnS secondary structure of Hestina persimilis and Hestinalis nama mitogenomes. Dashes, black dots and circles indicate the Watson–Crick base pairings, G-U bonds and U-U, A-A, A-C, A-G bonds, respectively.
Figure 7. Predicted rrnS secondary structure of Hestina persimilis and Hestinalis nama mitogenomes. Dashes, black dots and circles indicate the Watson–Crick base pairings, G-U bonds and U-U, A-A, A-C, A-G bonds, respectively.
Insects 12 00754 g007
Figure 8. The control region in Hestina persimilis.
Figure 8. The control region in Hestina persimilis.
Insects 12 00754 g008
Figure 9. The control region in Hestinalis nama.
Figure 9. The control region in Hestinalis nama.
Insects 12 00754 g009
Figure 10. Alignment of motif, Poly(T) and poly(A) in control regions of twelve species in Apaturinae.
Figure 10. Alignment of motif, Poly(T) and poly(A) in control regions of twelve species in Apaturinae.
Insects 12 00754 g010
Figure 11. Inferred phylogenetic relationships among Lepidoptera based on the concatenated nucleotide sequences of 13 PCGs using BI. Numbers on branches are Bayesian posterior probabilities. Papilio protenor (KY272622) and Lamproptera curius (KJ141168) are used as outgroups.
Figure 11. Inferred phylogenetic relationships among Lepidoptera based on the concatenated nucleotide sequences of 13 PCGs using BI. Numbers on branches are Bayesian posterior probabilities. Papilio protenor (KY272622) and Lamproptera curius (KJ141168) are used as outgroups.
Insects 12 00754 g011
Figure 12. Inferred phylogenetic relationships among Lepidoptera based on the concatenated nucleotide sequences of 13PCGs using ML. Numbers on branches are bootstrap percentages. Papilio protenor (KY272622) and Lamproptera curius (KJ141168) are used as outgroups.
Figure 12. Inferred phylogenetic relationships among Lepidoptera based on the concatenated nucleotide sequences of 13PCGs using ML. Numbers on branches are bootstrap percentages. Papilio protenor (KY272622) and Lamproptera curius (KJ141168) are used as outgroups.
Insects 12 00754 g012
Table 1. Annotation of Hestina persimilis mitogenome.
Table 1. Annotation of Hestina persimilis mitogenome.
GeneDirectionLocationSizeAnticodonStart CodonStop CodonIntergenic
Nucleotides
trnMF1–6868CAT 32–34
trnIF69–13466GAT 98–100 0
trnQR132–20069TTG 159–161 −3
nad2F292–13051014 ATTTAA91
trnWF1304–137168TCA1335–1337 −2
trnCR1364–142764GCA 1397–1399 −8
trnYR1428–149265GTA 1359–1461 0
cox1F1498–30331536 CGATAA5
trnL2 (UUR)F3029–309567TAA 3059–3061 −5
cox2F3096–3774679 ATGT0
trnKF3772–384271CTT 3802–3804 −3
trnDF3842–390766GTC 3872–3874 −1
atp8F3908–4069162 ATCTAA0
atp6F4063–4737675 ATGTAA−7
cox3F4737–5525789 ATGTAA−1
trnGF5528–559467TCC 5558–5560 2
nad3F5595–5948354 ATTTAG0
trnAF5947–601468TGC 5976–5978 −2
trnRF6014–607764TCG 6040–6042 −1
trnNF6090–615566GTT 6121–6123 12
trnS1 (AGN)F6154–621360GCT 6171–6173 −2
trnEF6216–628065TTC 6245–6247 2
trnFR6279–634264GAA 6310–6312 −2
nad5R6317–80771761 ATTTAA−26
trnHR8075–814167GTG 8109–8111 −3
nad4R8142–94801339 ATGT0
nad4LR9482–9772291 ATATAA1
trnTF9780–984465TGT 9811–9813 7
trnPR9845–990864TGG 9877–9879 0
nad6F9911–10438528 ATATAA2
cobF10,442–11,5931152 ATGTAA3
trnS2 (UCN)F11,596–11,66267TGA 11,625–11,627 2
nad1R11,685–12,626942 ATGTAA22
trnL1 (CUN)R12,628–12,70275TAG 12,671–12,673 1
rrnLR12,703–14,0361334 0
trnVR14,037–14,10064TAC 14,069–14,071 0
rrnSR14,101–14,876776 0
Control region 14,877–15,252376 0
Table 2. Annotation of Hestinalis nama mitogenome.
Table 2. Annotation of Hestinalis nama mitogenome.
GeneDirectionLocationSizeAnticodonStart CodonStop CodonIntergenic
Nucleotides
trnMF1–6868CAT 32–34
trnIF69–13365GAT 99–101 0
trnQR131–19969TTG 158–160 −3
nad2F269–12821013 ATTTAA69
trnWF1281–134868TCA 1312–1314 −2
trnCR1341–140363GCA 1372–1374 −8
trnYR1404–146865GTA 1435–1437 0
cox1F1474–30091536 CGATAA5
trnL2 (UUR)F3005–307167TAA 3035–3037 −5
cox2F3072–3750679 ATGT0
trnKF3748–381871CTT 3778–3780 −3
trnDF3818–388366GTC 3848–3850 −1
atp8F3884–4042159 ATCTAA0
atp6F4036–4713678 ATGTAA−7
cox3F4713–5501789 ATGTAA−1
trnGF5504–556865TCC 5534–5536 2
nad3F5566–5922357 ATATAG−3
trnAF5921–598767TGC 5953–5955 −2
trnRF5987–605266TCG 6014–6016 −1
trnNF6053–611866GTT 6084–6086 0
trnS1 (AGN)F6117–617660GCT 6134–6136 −2
trnEF6180–624364TTC 6108–6210 3
trnFR6244–630865GAA 6276–6278 0
nad5R6308–80441737 ATTTAA−1
trnHR8042–810665GTG 8071–8073 −3
nad4R8107–94451339 ATGT0
nad4LR9447–9731285 ATGTAA1
trnTF9744–980764TGT 9774–9776 12
trnPR9808–987164TGG 9840–9842 0
nad6F9874–10401528 ATATAA2
cobF10,406–11,5541149 ATGTAA4
trnS2 (UCN)F11,561–11,62464TGA 11,589–11,591 6
nad1R11,638–12,579942 ATGTAA13
trnL1 (CUN)R12,581–12,65474TAG 12,623–12,625 1
rrnLR12,655–13,9811327 0
trnVR13,982–14,04463TAC 14,014–14,016 0
rrnSR14,045–14,818774 0
Control region 14,819–15,208390 0
Table 3. Base composition of Hestina persimilis mitogenome.
Table 3. Base composition of Hestina persimilis mitogenome.
Size (bp)A%T%G%C%A + T%G + C%
mtDNA15,25239.741.27.611.580.919.1
PCGs11,22233.845.910.49.979.720.3
tRNA145941.739.711.07.781.418.7
rrnL133444.539.810.55.284.315.7
rrnS77643.641.510.14.985.115
Control region37643.647.62.95.991.28.8
Table 4. Base composition of Hestinalis nama mitogenome.
Table 4. Base composition of Hestinalis nama mitogenome.
Size (bp)A%T%G%C%A + T%G + C%
mtDNA15,20839.939.37.912.979.220.8
PCGs11,19232.844.811.411.077.622.4
tRNA144942.039.310.87.981.318.7
rrnL132740.243.45.311.183.616.4
rrnS77441.243.95.29.785.114.9
Control region39044.943.82.68.788.711.3
Publisher’s Note: MDPI stays neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Share and Cite

MDPI and ACS Style

Wu, Y.; Fang, H.; Wen, J.; Wang, J.; Cao, T.; He, B. Mitochondrial Genomes of Hestina persimilis and Hestinalis nama (Lepidoptera, Nymphalidae): Genome Description and Phylogenetic Implications. Insects 2021, 12, 754. https://0-doi-org.brum.beds.ac.uk/10.3390/insects12080754

AMA Style

Wu Y, Fang H, Wen J, Wang J, Cao T, He B. Mitochondrial Genomes of Hestina persimilis and Hestinalis nama (Lepidoptera, Nymphalidae): Genome Description and Phylogenetic Implications. Insects. 2021; 12(8):754. https://0-doi-org.brum.beds.ac.uk/10.3390/insects12080754

Chicago/Turabian Style

Wu, Yupeng, Hui Fang, Jiping Wen, Juping Wang, Tianwen Cao, and Bo He. 2021. "Mitochondrial Genomes of Hestina persimilis and Hestinalis nama (Lepidoptera, Nymphalidae): Genome Description and Phylogenetic Implications" Insects 12, no. 8: 754. https://0-doi-org.brum.beds.ac.uk/10.3390/insects12080754

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