Next Article in Journal
Enhanced Control of the Fungus Gnat Bradysia odoriphaga (Diptera: Sciaridae) by Co-Application of Clothianidin and Hexaflumuron
Previous Article in Journal
Effects of Helicoverpa armigera Egg Age on Development, Reproduction, and Life Table Parameters of Trichogramma euproctidis
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Characterization of the First Complete Mitochondrial Genome of Cyphonocerinae (Coleoptera: Lampyridae) with Implications for Phylogeny and Evolution of Fireflies

1
The Key Laboratory of Zoological Systematics and Application, School of Life Science, Institute of Life Science and Green Development, Hebei University, Baoding 071002, China
2
College of Agriculture, Yangtze University, Jingzhou 434025, China
*
Authors to whom correspondence should be addressed.
Insects 2021, 12(7), 570; https://doi.org/10.3390/insects12070570
Submission received: 17 May 2021 / Revised: 11 June 2021 / Accepted: 19 June 2021 / Published: 22 June 2021
(This article belongs to the Section Insect Systematics, Phylogeny and Evolution)

Abstract

:

Simple Summary

The classification of Lampyridae has been extensively debated. Although some recent efforts have provided deeper insight into it, few genes have been analyzed for Cyphonocerinae in the molecular phylogenies, which undoubtedly influence elucidating the relationships of fireflies. In this study, we generated the first complete mitochondrial genome for Cyphonocerinae, with Cyphonocerus sanguineus klapperichi as the representative species. The comparative analyses of the mitogenomes were made between C. sanguineus klapperichi and that of well-characterized species. The results showed that the mitogenome of Cyphonocerinae was conservative in the organization and characters, compared with all other fireflies. Like most other insects, the cox1 gene was most converse, and the third codon positions of the protein-coding genes were more rate-heterogeneous than the first and second ones in the fireflies. The phylogenetic analyses suggested that Cyphonocerinae as an independent lineage was more closely related to Drilaster (Ototretinae). Nevertheless, more sampler species are needed in the reconstruction of fireflies’ phylogeny to verify this result.

Abstract

Complete mitochondrial genomes are valuable resources for phylogenetics in insects. The Cyphonoceridae represents an important lineage of fireflies. However, no complete mitogenome is available until now. Here, the first complete mitochondrial genome from this subfamily was reported, with Cyphonocerus sanguineus klapperichi as a representative. The mitogenome of C. sanguineus klapperichi was conserved in the structure and comparable to that of others in size and A+T content. Nucleotide composition was A+T-biased, and all genes exhibited a positive AT-skew and negative GC-skew. Two types of tandem repeat sequence units were present in the control region (136 bp × 2; 171 bp × 2 + 9 bp). For reconstruction of Lampyridae’s phylogeny, three different datasets were analyzed by both maximum likelihood (ML) and Bayesian inference (BI) methods. As a result, the same topology was produced by both ML analysis of 13 protein-coding genes and 2rRNA and BI analysis of 37 genes. The results indicated that Lampyridae, Lampyrinae, Luciolinae (excluding Emeia) were monophyletic, but Ototretinae was paraphyletic, of which Stenocladius was recovered as the sister taxon to all others, while Drilaster was more closely related to Cyphonocerinae; Phturinae + Emeia were included in a monophyletic clade, which comprised sister groups with Lampyridae. Vesta was deeply rooted in the Luciolinae.

1. Introduction

Lampyridae Rafinesque, 1815 is a cosmopolitan family consisting of about 100 genera and 2200 species [1,2,3]. It is an amazing bioluminescent beetle group, with all species found to be luminous, at least at the larval stage [2]. However, some lineages exhibit no or weak bioluminescence in adulthood, leading to considerable confusion in their taxonomic positions. One of such non-luminescent groups is the monotypic subfamily Cyphonocerinae Crowson, 1972, which represents an important lineage with the type of genus Cyphonocerus Kiesenwetter, 1879 [3]. The adults of Cyphonocerus are easily recognized by the bipectinate antennae [4]. It is a small genus and has 17 species (subspecies) hitherto known from Japan, China, Nepal, and N. India [5]. However, there has been inconsistency in its family-group assignment since its establishment.
The genus Cyphonocerus was originally established under the family Drilidae Blanchard, 1845 [6,7,8,9,10] (now as a tribe of Elateridae Leach, 1825 [11]), and transferred to Lampyridae by Nakane [12]. It was once placed in the subfamily Amydetinae Olivier in Wytsman, 1907 [13] and later as the type genus to establish the subfamily Cyphonocerinae by Crowson [14], which was recognized as a subjective synonym of Psilocladinae McDermott, 1964 by Jeng et al. [4], and synonymized with Lampyrinae by Lawrence et al. [15], but most recently revalidated by Martin et al. [3] with Cyphonocerus left as the sole member.
Recent phylogenetic efforts have provided deeper insight into the classification of fireflies by expanding morphological [16,17], or molecular [2,3,18,19], or both datasets [19]. All of these studies highlighted the need to update the higher-level classification of Lampyridae within a phylogenetic framework. However, in the previous molecular phylogenetic studies, none [3,19] or limited genes [20,21] have been analyzed for Cyphonocerinae, or not all sampler species were provided with the same number of gene markers in reconstructing phylogenetic trees [2], this undoubtedly influenced elucidating the relationships of fireflies. Until now, only 16S, cox1, nad5, 18S genes are available for Cyphonocerus in the public database [21,22,23], so more thorough gene analyses are needed.
Complete mitochondrial genomes have been widely used in investigating molecular evolution and phylogenetic relationships among different lineages of insects due to their highly conserved structure in evolution, rare recombination, and rapid evolutionary rate [24,25,26]. In mitochondrial genomes, compositional bias and substitution rate variation have been extensively investigated in comparative analyses since that they provide critical information for molecular evolution [27,28,29,30,31]. Furthermore, phylogenomic analysis with higher numbers of genes, up to all of the 37 mitochondrial genes, has been tested to get more highly supported nodal confidence, compared with a single or a few locus phylogenetics [28,32,33].
In this study, we generated and analyzed the first complete mitochondrial genome for Cyphonocerus sanguineus klapperichi Pic, 1955. This enabled us to provide the comparative analysis of the genomic structure, base composition, substitution, and evolutionary rates between Cyphonocerinae and that of well-characterized complete mitogenomes of other fireflies, as well as a comprehensive molecular phylogenetic analysis of Lampyridae based on complete mitogenomes. The complete mitogenome reported here will contribute to fireflies’ higher phylogeny reconstruction based on mitogenome sequences and promote comparative mitogenome studies, which should help understand the mitogenome evolution across different lineages of Lampyridae.

2. Materials and Methods

2.1. Taxon Sampling

The material of C. sanguineus klapperichi was collected from China, Fujian Province, Wuyishan, Tongmu, San’gang, 117°41′16′′ E, 27°45′10′′ N, on 23 May 2018. Specimens were preserved in 100% ethanol at ‒20 °C before molecular experiments. The analyzed specimen was identified using the identification key provided by Jeng et al. [5].

2.2. DNA Extraction, Mitochondrial Genome Sequencing, and Assembly

Total genomic DNAs were extracted using a DNeasy Blood & Tissue kit (QIAGEN, Beijing, China), according to the manufacturer’s instructions. DNA was stored at −20 °C for long-term storage and further molecular analyses, which were deposited in the Museum of Hebei University (MHBU, accession No. 2CAN0196).
The whole mitochondrial genome sequence was sequenced using an Illumina Novaseq 6000 platform with 150 bp paired-end reads at BerryGenomics, China. The sequence reads were first filtered by the programs following Zhou et al. [34], and then the remaining high-quality reads were assembled using IDBA-UD [35], under similarity threshold 98%, and k values minimum 40 and maximum 160 bp. The gene cox1 was amplified by polymerase chain reaction (PCR) using universal primers as ‘reference sequences’ to target mitochondrial scaffolds by IDBA-UD [35] to acquire the best-fit, which is under at least 98% similarity. Geneious 2019.2 [36] software was used to manually map the clean readings to the obtained mitochondrial scaffolds to check the accuracy of the assembly.

2.3. Genome Annotation and Analyses

Gene annotation was done by Geneious 2019.2 [36] software and the MITOS web server (http://mitos.bioinf.uni-leipzig.de/index.py, accessed on 20 March 2021) [37]. The positions and secondary structures of 22 tRNAs were estimated by a combination of the results predicted by an ARWEN and tRNAscan-SE Search Server v.1.21 [38,39]. The mitogenomic circular map was produced using a CGView Server (http://stothard.afns.ualberta.ca/cgview_server, accessed on 20 March 2021) [40]. The skewness was determined with base composition of nucleotide sequences by using the formula: AT skew = [A − T]/[A + T], GC skew = [G − C]/[G + C] [41]. The Tandem Repeat Finder program was used to predicted tandem repeats in A + T-rich region [42]. The relative synonymous codon usage (RSCU) was analyzed by MEGA 7.0 [43]. DnaSP v5.10.01 [44] was used to calculate the nucleotide diversity (Pi) and sliding window analysis (a sliding window of 200 bp and a step size of 20 bp) based on 13 aligned protein-coding genes (PCGs) and non-synonymous (Ka)/synonymous (Ks) substitution rates among the 13 PCG. The base composition and component skew were analyzed using PhyloSuite v1.2.2 [45]. The genetic distances were computed using MEGA 7.0 with the Kimura-2-parameter model. SymTest v2.0.47 [46] with Bowker’s matching pair symmetry test was used to analyze the differences of heterogeneous sequences in the datasets, and the heat maps were generated according to the inferred p-values.

2.4. Phylogenetic Analysis

We followed the classification of Lampyridae by Martin et al. [3]. In addition to the newly sequenced genome here for Cyphonocerinae, another 32 species representing four subfamilies of Lampyridae were selected as the ingroups, which are the previously published complete or almost complete mitochondrial genomes downloaded from GenBank (Table 1). Two species of Rhagophthalmidae (Rhagophthalmus ohbai Wittmer, 1994) and Phengodidae (Phrixothrix hirtus Olivier, 1909) were chosen as the outgroups [2].
Data standardization and information extraction were performed by PhyloSuite v 1.2.2 [45]. The 13 PCGs were aligned using the MAFFT algorithm implemented in TranslatorX [47] with the L-INS-i strategy. The 2 rRNAs and 22 tRNAs were aligned with MAFFT version 7 online services using the G-INS-i strategy. Gblocks v 0.91b [48] was used to remove the gaps and ambiguously aligned sites. The aligned data were concatenated with Sequence Matrix v.1.7.8 [49] and PhyloSuite v 1.2.2. The alignment of the individual gene was concatenated into three datasets: (i) the PCGrRNA matrix, including 13 PCGs and 2 rRNA genes (12,875 bp), (ii) the PCG12RNA matrix, including the first and second codon positions of the PCGs and 2 rRNA genes (9236 bp), and (iii) the PCGRNA matrix, including 13 PCGs, 2 rRNA genes and 22 tRNA genes (14,321 bp).
All datasets were analyzed using maximum likelihood (ML) on the IQ-TREE web server (http://iqtree.cibiv.univie.ac.at/, accessed on 6 April 2021) [50] with the GTR+G+I substitution model (Supplementary Tables S1–S3). Bayesian inference (BI) was also used for the phylogenetic analyses either by PhyloBayes MPI v.1.7a [51] (for the PCG12RNA and PCGrRNA matrixes) with the site-heterogeneous mixture CAT + GTR model (Supplementary Table S4), or by MrBayes 3.2.6 [52] (for the PCGRNA matrix) with two independent Markov Chain Monte Carlo chains run of 2 × 106 generations, of which the tree was sampled every 1000 generations, and the initial 25% of sampled data were discarded as burn. ITOL (http://itol.embl.de/, accessed on 10 April 2021) [53] was used to annotate and beautify the phylogenetic tree.

3. Results

3.1. Genomic Structure and Base Compositions

As in other fireflies, the mitogenome of C. sanguineus klapperichi (Figure 1) is a typical double-strand circular molecule and contains 13 protein-coding genes (PCGs), 22 transfer RNA genes (tRNAs), 2 ribosomal RNA(rRNAs) genes, and a control region (CR) or AT-rich region, in which 14 genes (8 tRNAs, 4 PCGs, and 2 rRNAs) are transcribed from the minority strand (N-strand), while others (14 tRNAs and 9 PCGs) from the majority strand (J-strand). The annotated sequence was registered in GenBank with accession number MW365445. This is the first complete mitogenome record for Cyphonocerinae.
Seven gene overlaps are present in C. sanguineus klapperichi mitogenome (Supplementary Table S5), ranging from 1 to 4 bp, with the longest overlap (4 bp) occurring between atp6 and atp8, and also nad4 and nad4l, respectively. The overlap between atp6 and atp8 is also found in mitogenomes of other arthropods [66,67,68]. Moreover, there are 13 intergenic spacer regions between genes (Supplementary Table S5), of which the total length is 211 bp, with the longest intergenic spacer (72 bp) exists between trnC and trnW. This result shows that the number and length of gene spacers are significantly higher than those of gene overlaps.
Known complete mitogenomes of fireflies range from 15,950 bp (Asymmetricata circumdata) to 18,054 bp (Pyrocoelia thibetana) (Supplementary Table S6). The mitogenome of C. sanguineus klapperichi is 16,443 bp in length (Supplementary Table S5) and slightly shorter than most others, of which the average length is 16,855 bp.
For fireflies’ mitogenomes, the sizes of the control region vary greatly among different species, whereas the PCGs, tRNAs, and rRNAs show little variation in length (Figure 2A, Supplementary Table S7). This suggests that the mitogenome size of different fireflies is largely determined by the size of control regions, like other insects [68].
The base composition of C. sanguineus klapperichi is A (42.5%), T (34.2%), C (13.9%), and G (9.4%), respectively (Table 2). It contains a slightly lower A+T content (76.7%) and a higher G + C content (23.3%), compared to that of other firefly species, which have an average value of A + T content being 78.0%, varying from 75.7% to 80.7% (Supplementary Table S6). This lower A + T bias in C. sanguineus klapperichi is reflected in all components of its genome, except tRNAs are near to the average value (Figure 2B).
The nucleotide skew analysis shows that the full mitogenome of C. sanguineus klapperichi exhibits a positive AT-skew (0.11) and a negative GC-skew (−0.19) (Table 2). A similar pattern is found in all other firefly mitogenomes, with the AT-skew ranging from 0.06 (Photinus pyralis) to 0.14 (Luciola cruciata) and a GC-skew varying from −0.08 (Diaphanes nubilus) to −0.24 (Lamprigera yunnana) (Supplementary Table S6). These results indicate that Ototretinae (Drilaster sp. and Stenocladius sp.) has the strongest A skew and weakest C skew, while Cyphonocerinae (C. sanguineus klapperichi) has an average value of AT-skew and GC-skew in comparison with other known firefly mitogenomes (Figure 2C,D). This base composition bias has been suggested to be associated with replication and transcription of the mitochondrial genome [67].

3.2. Protein-Coding Genes

The overall size (excluding stop codons) of 13 PCGs of C. sanguineus klapperichi is 11,008 bp in length, accounting for 66.95% of the total genome (Table 2). Like the full mitogenome, the whole PCGs show a slightly lower A+T content (75.1%), of which the third codon position (80.2%) is higher than those of the first and second codon positions (72.3% and 72.9%, respectively). The AT-skew (0.09) is positive, while GC-skew (−0.17) is negative for the PCGs, reflecting a bias towards nucleotides A and C than their counterparts.
All PCGs of C. sanguineus klapperichi are initiated with the standard ATN codons and terminated with TAA/TAG or a truncated termination codon T (Supplementary Table S5). These incomplete stop codons are thought to be ubiquitous in metazoan [26] and have been supposed to be completed through posttranscriptional polyadenylation [69].
The codon usage analysis of C. sanguineus klapperichi shows that the most frequently used codons are UUA-Leu (352), AUU-Ile (346), UUU-Phe (326), and AUA-Met (235) (Figure S1A, Supplementary Table S8). The UUA-Leu also has the highest RSCU value (3.9), further indicating that UUA is the most preferred codon. The RSCU values of the PCGs reveal that there is a higher frequency in the usage of AT than that of GC in the third codon positions (Figure S1B, Supplementary Table S8).
Sliding window analysis was implemented to study the nucleotide diversity of 13 PCGs among fireflies exhibited in Figure 3A. Nucleotide diversity values range from 0.187 (cox1) to 0.337 (atp8). Among the genes, atp8 (Pi = 0.337) has the highest variability, followed by nad6 (Pi = 0.303), nad2 (Pi = 0.282), and nad3 (Pi = 0.257) (Supplementary Table S9). In contrast, cox1 (Pi = 0.187) and nad1 (Pi = 0.198) have relatively low values and are the most conserved of the 13 PCGs. This result indicates that the nucleotide diversity is highly variable among the 13 PCGs.
Pairwise genetic distances among the mitogenomes of fireflies (Figure 3B, Supplementary Table S9) show that atp8 (0.439), nad6 (0.423), and nad2 (0.360) evolve comparatively faster, while cox1 (0.216) and nad1 (0.234) evolve comparatively slowly.
The ratio (ω) of non-synonymous (Ka) to synonymous (Ks) substitution rates, which is a diagnostic statistical method to detect molecular adaption [70,71], is used to estimate the evolutionary rate among insects. In Lampyridae, the genes atp8 (0.777), nad6 (0.641), and nad2 (0.501) have comparatively high Ka/Ks ratios, while cox1 (0.165), cox2 (0.266), and cox3 (0.282) have relatively low values (Figure 3B, Supplementary Table S9). The average Ka/Ks (ω) of 13 PCGs of the fireflies are all less than 1, indicating that these genes are under purifying selection [72].
Heterogeneity of nucleotide divergence was examined under pairwise comparisons in a multiple sequence alignment (Figure 4). The datasets PCGrRNA and PCGRNA exhibit higher heterogeneous sequence divergence than PCG12rRNA, indicating that the third codon positions are more rate-heterogeneous than the first and second ones. The higher compositional heterogeneity may result in systematic errors in phylogenetic analyses [27].

3.3. Transfer and Ribosomal RNA Genes

The complete set of 22 typical tRNAs were all found in the mitochondrial genome of C. sanguineus klapperichi, and their secondary structures are shown in Supplementary Figure S2. The total length of tRNAs is 1432 bp, ranging from 62 bp (trnF, trnG, trnH, and trnL1) to 71 bp (trnK) in size (Supplementary Table S5). The AT-skew (0.07) is positive, and GC-skew (−0.15) is negative (Table 2), indicating a preference for using A base over T and G over C.
Most tRNAs exhibit the typical clover-leaf structures, except trnS1 missing the dihydrouridine (DHU) arm (Supplementary Figure S2). Lacking the DHU arm in trnS1 is a common feature for most metazoan mitogenomes [26]. These aberrant tRNAs are supposed to sustain their function via a posttranscriptional RNA editing mechanism [73,74]. In the tRNAs, except the classic base pairs (A-U and C-G), 22 non-canonical base pairings (G-U and A-C), and 8 other mismatched base pairs (U-U, A-A, A-G) were found in the arms (Supplementary Figure S2).
In the mitogenome of C. sanguineus klapperichi, the rrnL (1268 bp) and rrnS (766 bp) are located in the conserved positions between trnL and trnV, and trnV and the control region, respectively (Figure 1). There is a little variation in the sizes of both rRNAs among firefly species (Figure 2A, Supplementary Table S7). The A+T content of rrnL and rrnS are 81.5% and 78.7%, respectively (Supplementary Table S7), which are slightly lower than most other firefly species (Figure 2B). The overall rRNA shows a positive AT-skew (0.17) and a negative GC-skew (−0.34), which shows a slight bias toward using A and an obvious bias toward G (Table 2).

3.4. Control Region

The control region of C. sanguineus klapperichi was identified by the position between rrnS and trnI, spanning 1776 bp in length (Figure 1, Table 2). This is comparable to that of most other fireflies, which have an average length of 1810 bp (Supplementary Table S7), ranging from 1400 bp (Photinus pyralis) to 2341 bp (Diaphanes sp.). Generally, the length of the control region varies more than other components in the fireflies (Figure 2A). The control region is supposed to be involved in the initiation of replication and transcription of mitogenomes [75].
The control region of mitogenome in C. sanguineus klapperichi has a slightly lower A+T content (81.2%) compared to most other firefly species (Figure 2B), of which the average value is 86.6% (Supplementary Table S7). The AT-skew (0.12) is positive, while the GC-skew (−0.23) is negative (Table 2), showing a bias towards using A and G.
Within the control region, two tandem repeat sequence units are present in the mitogenome of C. sanguineus klapperichi; their positions and length are shown in Figure S3. They are a 136 bp-sequence tandemly repeated twice and a 171 bp-sequence tandemly repeated twice with a partial third repeat (9 bp), respectively. In Lampyridae, the tandem repeat sequences within the control region are quite diverse. They have been found in most firefly species which are all located between rrnS and trnI and vary widely in size and number of repeat units but are absent in Luciola substriata (Supplementary Figure S3). The species has two different types of repeat units at least, and every unit repeated at least twice, except Ellychnia corrusca, Bicellonycha lividipennis, and Asymmetricata circumdata, which have only one type of repeat unit. The most diverse tandem repeat sequence units happen in Pyrocoelia praetexta, which includes five types of repeat units. The longest repeat unit was found in Diaphanes mendax, which contained two 303 bp repeats. However, in Diaphanes pectinealis, the tandem repeat region is the shortest, with a 9 bp repeat unit tandemly repeated six times plus a 6 bp partial sequence.

3.5. Phylogenetic Analysis

Analyses of the three datasets resulted in nearly identical and fully resolved topologies with high nodal support values under ML and BI methods. What is noted, the BI reconstruction of 37 genes dataset and ML reconstruction of the 13 PCGs and 2rRNA dataset produced the same topology shown in Figure 5.
In all topologies (Figure 5, Supplementary Figures S4–S7), the monophyly of Lampyridae was well supported based on mitogenomes of different genes datasets (PP = 1/0.98/0.92, BS = 100/100/96). Furthermore, the monophyly of Lampyrinae and Luciolinae Lacordaire, 1857 (with Emeia Fu, Ballantyne et Lambkin, 2012 excluded), including 11 and 16 representative species respectively, were highly supported (PP = 1, BS = 100). The clade composing Photurinae Lacordaire, 1857 (only one species included) and Emeia was suggested to be a monophyly (PP = 1, BS = 100), which was a recovered sister to Lampyrinae with high support value (PP = 1, BS = 100). However, the monophyly of Ototretinae McDermott, 1964 was not recovered, with the sampler genera Drilaster Kiesenwetter, 1879 and Stenocladius Fairmaire in Deyrolle and Fairmaire, 1878 splitting in different clades. The monophyly of the monotypic Cyphonocerinae with a single species here could not be tested.
Except for the ML reconstruction of PCGRNA dataset (Supplementary Figure S4), Stenocladius was recovered as the sister taxon to all other fireflies, albeit with comparatively high support values (PP = 0.854/0.74/0.59, BS = 90/94).
In the topologies produced by ML analyses of all datasets (Figure 5, Supplementary Figures S4 and S5) and BI analysis of PCGRNA dataset (Figure 5), C. sanguineus klapperichi (Cyphonocerinae) was always grouped with Drilaster (Ototretinae), with a high support value (BS = 98/93/83; PP = 1); then except ML analysis of PCGRNA dataset, which together with a sister to Lampyrinae + (Photurinae + Emeia) (Figure 5, Supplementary Figure S5), was highly supported (BS = 76/81; PP = 0.924), while in the latter (Supplementary Figure S4), they were a recovered sister to Luciolinae but with very low support (BS = 35). However, based on the other two datasets of BI analyses (Supplementary Figures S6 and S7), C. sanguineus klapperichi was solely suggested as a sister to Lampyrinae + (Photurinae + Emeia), also with high support values (PP = 0.99/0.89), while Drilaster was in uncertain relationships with or sister to these taxa.
Based on the BI reconstruction of PCGrRNA and PCG12rRNA datasets (Supplementary Figures S6 and S7) and ML reconstruction of PCG12rRNA dataset (Supplementary Figure S5), Lamprigera Motschulsky, 1853 was a recovered sister to Luciolinae, with comparatively high support values (PP = 0.99/0.99; BS = 63). However, it was a recovered sister to the remaining fireflies except Stenocladius under ML analysis of PCGrRNA dataset and BI analysis of PCGRNA dataset (Figure 5), also with high support (BS = 90; PP = 0.854). Moreover, it was possibly a sister to Stenocladius based on ML analysis of PCGRNA dataset (Supplementary Figure S4), but with a low support value (BS = 56).
In all topologies (Figure 5, Supplementary Figures S4–S7), Vesta Laporte, 1833 was deeply rooted in Luciolinae and a sister to Pristolycus Gorham, 1883, which was greatly supported (PP = 1, BS = 100). Surprisingly, Emeia was always grouped with Bicellonycha Motschulsky, 1853 (Photurinae), which was highly supported (PP = 1, BS = 100) in all analyses.

4. Discussion

4.1. Features of Mitochondrial Genomes in Lampyridae

In all known mitochondrial genomes of Lampyridae, both the size and A+T content vary greatly for the control region, but a less variation for PCGs, tRNAs, and rRNAs, respectively, indicating that the control region is an important component that heavily affects the size and total A+T content of fireflies’ mitogenome.
Fireflies exhibit the typical A+T-biased composition of insect mitogenomes [26,76,77], in either the full genome or each component, all over 75.1%. The biological reasons for such A+T-biased compositional heterogeneity have been extensively investigated [78,79,80], and one of the hypotheses is the energy efficiency trade-offs which has been experimentally rested [79]. The hypothesis suggests that the resources for nucleotide production are limited, and synthesis of G+C consumes more energy and nitrogen than A+T, so A and T are preferred nucleotides [79]. Although the mitogenome of fireflies is A+T-biased, Cyphonocerinae have a slightly lower value of A+T content in comparison with most others (Figure 2A). What is more interesting, the size of the full genome of Cyphonocerinae is about 400 bp shorter than those of the other fireflies (Supplementary Table S6). Therefore, it is presumed that the higher G+C content in this group seems to be compensated by the shortened mitogenome, which is likely to be shaped by selection for efficient usage [30]. Additionally, this is suggested to be a molecular strategy to ensure a reliable protein synthesis under high temperatures [76].
In Lampyridae, the AT-skew is all positive while GC-skew is negative, indicating the base composition bias towards A and C than their counterparts. Although the cases for such skewed strand composition are multifactorial, most of the hypotheses suggest that the strand asymmetry is the result of mutations and selection pressures [81], and the value of GC-skew of insect mitogenomes seem to be associated with replication orientation [79].
The analyses of nucleotide diversity, pairwise genetic distances, and Ka/Ks (ω) all showed that in the Lampyridae, the genes atp8, nad6, and nad2 to be more variable and evolve faster, while cox1 is more conserved and evolves comparatively slowly. Nucleotide diversity analyses are useful for designing species-specific markers, especially in taxa where morphological identification is difficult and ambiguous [82,83]. The cox1 gene is often used as a universal barcode for species identification for the insects [84,85,86], but for the fireflies, its low variability indicates that it is more suitable for exploring the phylogenetic relationships among the higher grades. In contrast, those genes exhibiting an optimal combination of fast evolution and sufficiently large size, such as nad6, should be evaluated as potential DNA markers for species and/or population identification.

4.2. Phylogenetic Implications of Mitochondrial Genomes in Lampyridae

As a morphologically and biologically diverse group [16,17,18,87], the taxonomy and classification of fireflies have been extensively debated [2,3,7,14,15,20,88,89,90,91], particularly the non-luminescent groups are more controversial, such as Ototretinae and Cyphonocerinae.
The subfamily Ototretinae was established by McDermott [92] and redefined in a broad sense by Crowson [14], which was followed by Lawrence and Newton [91] and Brancucci and Geiser [93]. This group consists of several genera, including Drilaster and Stenocladius. However, Branham and Wenzel [16] excluded the latter taxa from Lampyridae on the basis of a morphological phylogeny, which was supported by Lawrence et al. [15], who placed them in Elateriformia incertae sedis, but against the molecular phylogenetic analyses by Bocakova et al. [22] and Sagegami-Oba et al. [94], also not adopted by Geisthardt and Satô [95]. Recently, this subfamily was revised by Janisova and Bocakova [96], based on the comparative morphology of adults, and transferred the above genera to Lampyridae again. In the most recent molecular phylogenetics, the members of Ototretinae were deeply rooted in Lampyridae with high support value [2,3,20] but often suggested to be a paraphyletic group [2,3], which is congruent with the present study.
Furthermore, our analyses indicated that Stenocladius was most probably the basal lineage of Lampyridae, which correlated with that of Li et al. [97]. In addition, Cyphonocerus (Cyphonocerinae) was recovered more closely related to Drilaster, which is in agreement with that of Martin et al. [20]. The clade of Drilaster and Cyphonocerus seemed more closely related to Lampyrinae but Luciolinae, which is in agreement with Suzuki [21], but against those results of Martin et al. [20] or Chen et al. [2].
Unfortunately, the monophyly of Cyphonocerinae could not be tested in this study due to a lack of more material of Cyphonocerus. Furthermore, the phylogenetic relationships among Psilocladus Blanchard, 1846 and Pollaclasis Newman, 1838 were not evaluated because of a deficiency of the complete mitogenome data. The three taxa were included in the subfamily Psilocladinae by Jeng [4,5], which was redefined by Martin et al. [3] as a monotypic subfamily (including only Psilocladus) and Pollaclasis in Lampyridae incertae sedis, meanwhile, Cyphonocerus left as the sole member of Cyphonocerinae. The efforts need to be made to clarify their relationships in the future when the material is available for all these genera.
Previous works have recovered Lamprigera in various positions within Lampyridae [3,20,61,97]. In this study, Lamprigera was either a recovered sister to Luciolinae, which is congruent with Martin et al. [3] and Chen et al. [2], or a sister to the remaining fireflies except for Stenocladius, similar to that of Li et al. [97], but never be a member of Lampyrinae [98]. Given this incongruence, the exact position of Lamprigera remains uncertain, as what has been done by Martin et al. [3,20], placing it as Lampyridae incertae sedis. To rigorously test the classification of Lamprigera relative to other subfamilies, an expanded taxon sampling including deeper species coverage of this genus will be needed.
Additionally, here we recovered Vesta in the Luciolinae for the first time and sister to Pristolycus with strong support and congruence among all of our analyses. It was once placed in Amydetinae [92] or Lampyrinae [90] and was noted to be a paraphyletic group with some species from Photurinae and Lampyrinae by Jeng [98]. Evidence from individual or multi-molecular markers supported the position of Vesta near Photurinae in the phylogenetics [2,3,20,97] and was placed in the Lampyridae incertae sedis. However, this placement was based on a single Vesta species; as a specious genus, more taxa should be included in the future study to verify this result.
It is surprising that the monotypic Emeia and Bicellonycha (Photurinae) comprised a monophyletic clade in all of our analyses. Regardless of this result, their morphological characteristics, biology, and luminous behaviors are substantially different [99]. In addition, its placement in Luciolinae has been well supported by both morphological [99] and molecular phylogenies [2,3]. Since only one species of Photurinae was included in our analyses, an expanded sampling taxon will be needed to test this placement, so we do not make any taxonomic change here.
Above all, the molecular phylogenies, including this study, were analyzed on the basis of a minority part of species or limited molecular data in comparison to an estimated 2200 species of fireflies worldwide. Therefore, many more species need to be included in future analysis to establish a solid and dependable classification of Lampyridae. Particularly, the complete mitochondrial genomes should be encouraged to accumulate more for Lampyridae, in view of its high value in investigating phylogenetic relationships of the insects.

5. Conclusions

In the present study, we generated and analyzed the first complete mitochondrial genome for Cyphonocerinae, with C. sanguineus klapperichi as a representative. Compared with that of all well-characterized mitochondrial genomes of fireflies, the mitogenome of Cyphonocerinae is highly conserved in structure and size. It is a highly A+T-biased composition, with a positive AT-skew and negative GC-skew. It has conserved codon usage of protein-coding genes and secondary structures of tRNAs, as well as a unique type of tandem repeat sequence units present in the control region. This provides the basic information to perform comparative analyses and further discussion of the mitogenomes’ evolution of Lampyridae.
Furthermore, the nucleotide diversity, genetic distance substitution rates, and heterogeneity of nucleotide divergence were analyzed and examined. The result indicates that in Lampyridae, the genes atp8, nad6, and nad2 are more variable and evolve faster, while cox1 is more conserved and evolves comparatively slowly. Furthermore, the third codon positions are more rate-heterogeneous than the first and second ones.
Moreover, the phylogenetic trees of Lampyridae were reconstructed based on three different datasets by both maximum likelihood (ML) and Bayesian inference (BI) methods. The result suggests that Lampyridae, Lampyrinae, Luciolinae (excluding Emeia) are monophyletic. Ototretinae is paraphyletic, of which Stenocladius is at the basal lineage and sister to all others, while Drilaster is more closely related to Cyphonocerinae. Lampyridae + (Photurniae + Emeia) comprises sister groups, and the latter two are a monophyletic clade. Here Vesta is recovered in Luciolinae for the first time. Nevertheless, large-scale analyses with denser taxon sampling are needed to confirm the present results.

Supplementary Materials

The following are available online at https://www.mdpi.com/article/10.3390/insects12070570/s1, Table S1: The best partitioning schemes and models for the Maximum likelihood (ML) method in the analysis of the PCG12rRNA matrix; Table S2: The best partitioning schemes and models for the Maximum likelihood (ML) method in the analysis of the PCGrRNA matrix; Table S3: The best partitioning schemes and models for the Maximum likelihood (ML) method in the analysis of the PCGRNA matrix; Table S4: The best partitioning schemes and models for the Bayesian inference (BI) method in the analysis of the PCGRNA matrix; Table S5: Mitogenomic organization of C. sanguineus klapperichi. Table S6: Comparison of length, nucleotide composition, and skewness of the representative fireflies’ mitochondrial genomes; Table S7: Comparison of length and A+T% in different components of mitogenomes of the representative fireflies; Table S8: The count and relative synonymous codon usage (RSCU) of C. sanguineus klapperichi; Table S9: Nucleotide diversity, synonymous and non-synonymous substitution and genetic distance of representative fireflies. Figure S1: The codon usage numbers (A) and the RSCU (B) of amino acids used for the construction of 13 protein-coding genes of C. sanguineus klapperichi. Codon families are labeled below the X-axis. Figure S2: Secondary structures of the tRNAs in the mitochondrial genome of C. sanguineus klapperichi. Non-canonical base pairings are indicated in green and mismatched base pairs in red. Figure S3: Repeat units of the control region of the mitochondrial genome from C. sanguineus klapperichi and the representative species from other firefly subfamilies. Different colors of square bars indicate different tandem repeats. rrnS and tRNAs are indicated by the gray and purple square bars, respectively. Figure S4: Phylogenetic tree of Lampyridae produced from ML analysis of PCGRNA dataset; Figure S5: Phylogenetic tree of Lampyridae produced from ML analysis of PCG12rRNA dataset; Figure S6: Phylogenetic tree of Lampyridae produced from BI analysis of PCGrRNA dataset; Figure S7: Phylogenetic tree of Lampyridae produced from BI analysis of PCG12rRNA dataset.

Author Contributions

Conceptualization, Y.Y. and H.L.; Methodology, X.G., L.Y., Y.K., T.L., and Y.Y.; Investigation, Y.Y. and X.G.; Validation, X.G. and Y.Y.; Writing―Original Draft Preparation, X.G., L.Y., and Y.Y.; Writing―Review and Editing, X.G., H.L., and Y.Y; Funding acquisition, Y.Y. and H.L. All authors have read and agreed to the published version of the manuscript.

Funding

This research was funded by the National Natural Science Foundation of China (Nos 31772507, 41401064), the Natural Science Foundation of Hebei Province (No. C201720112), and the Science and Technology Project of Hebei Education Department (No.BJ2017030) to Y.Y., also the Biodiversity Survey and Assessment Project of the Ministry of Ecology and Environment, China (No. 2019HJ2096001006) and the Natural Science Foundation of Hebei Province (No. C2019201192) to H.L.

Institutional Review Board Statement

Not applicable.

Informed Consent Statement

Not applicable.

Data Availability Statement

The sequence generated in this study is deposited in GenBank with accession number (MW365445).

Acknowledgments

We are grateful to Dong Zhang and Liping Yan (Beijing Forestry University, Beijing) for their valuable help in the examination of heterogeneity of nucleotide divergence of the mitochondrial genome. In addition, thanks are given to Ming Bai and Yuanyuan Lu (Institute of Zoology, Chinese Academy of Sciences, Beijing) for providing us with the material.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Branhm, M.A. Lampyridae Latreille, 1817. In Coleoptera, Beetles: Morphology and Systematics (Elateroidea, Bostrichiformia, Cucujiformia partim); Leschen, R.A.B., Beutel, R.G., Lawrence, J.F., Eds.; Walter de Gruyter: Berlin, Germany, 2010; Volume 2, pp. 141–149. [Google Scholar]
  2. Chen, X.; Dong, Z.; Liu, G.; He, J.; Zhao, R.; Wang, W.; Peng, Y.; Li, X. Phylogenetic analysis provides insights into the evolution of Asian fireflies and adult bioluminescence. Mol. Phylogenet. Evol. 2019, 140, 1–14. [Google Scholar] [CrossRef]
  3. Martin, G.J.; Stanger-Hall, K.F.; Branham, M.A.; Silveira, L.F.L.; Lower, S.E.; Hall, D.W.; Li, X.Y.; Lemmon, A.R.; Lemmon, E.M.; Bybee, S.M. Higher-level phylogeny and reclassification of Lampyridae (Coleoptera: Elateroidea). Insect Syst. Divers. 2019, 3, 1–15. [Google Scholar] [CrossRef]
  4. Jeng, M.L.; Yang, P.S.; Satô, M. The genus Cyphonocerus (Coleoptera, Lampyridae) from Taiwan and Japan, withnotes on the subfamily Cyphonocerinae. Elytra 1998, 26, 379–398. [Google Scholar]
  5. Jeng, M.L.; Yang, P.S.; Satô, M. Synopsis of Cyphonocerus (Coleoptera: Lampyridae) with the description of four new species and a key to the genus. Zool. Stud. 2006, 45, 157–167. [Google Scholar]
  6. Lewis, G. On the Dascillidae and Malacoderm Coleoptera of Japan. Ann. Mag. Nat. Hist. 1895, 16, 98–122. [Google Scholar] [CrossRef] [Green Version]
  7. Olivier, E. Lampyridae. In Coleopterorum Catalogus; Pars. 9; Schenkling, S., Ed.; W Junk: Berlin, Germany, 1910; pp. 1–68. [Google Scholar]
  8. Pic, M. Coléoptèresexotiques nouveaux oupeuconnus. Échange 1911, 27, 142–144. [Google Scholar]
  9. Pic, M. Coléoptèresnouvaeux de Chine. Bull. Soc. Entomol. Mulhouse 1955, 1955, 25–26. [Google Scholar]
  10. Wittmer, W. Catalogue des Drilidae, E. Oliv. (Coleoptera, Malacodermata). Rev. Soc. Entomol. Argentina 1944, 12, 203–221. [Google Scholar]
  11. Kundrata, R.; Bocak, L. The phylogeny and limits of Elateridae (Insecta, Coleoptera): Is there a common tendency of click beetles to soft-bodiedness and neoteny? Zoologica Scripta. 2011, 40, 364–378. [Google Scholar] [CrossRef]
  12. Nakane, T. On the genus Cyphonocerus Kiesenwetter in Japan and Formosa (Insecta, Coleoptera, Lampyridae). Bull. Nat. Sci. Mus. Tokyo 1967, 10, 7–9. [Google Scholar]
  13. Nakane, T. The classification of Lampyridae. Nat. Insects. 1968, 3, 3–6. [Google Scholar]
  14. Crowson, R.A. A review of the classification of Cantharoidea (Coleoptera), with the definition of two newfamilies, Cneoglossidae and Omethidae. Rev. Univ. Madrid. 1972, 21, 35–77. [Google Scholar] [CrossRef]
  15. Lawrence, J.F.; Hastings, A.M.; Dallwitz, M.J.; Paine, T.A.; Zurcher, E.J. 2000 (onwards). Elateriformia (Coleoptera): Descriptions, illustrations, identification, and information retrieval for families and sub-families. Version: 9th October 2005. Available online: http://delta-intkey.com./delta/elateria/www/lamplamp.htm (accessed on 20 March 2021).
  16. Branham, M.A.; Wenzel, J.W. The evolution of bioluminescence in cantharoids (Coleoptera: Elateroidea). Fla. Entomol. 2001, 84, 565–586. [Google Scholar] [CrossRef]
  17. Branham, M.A.; Wenzel, J.W. The origin of photic behavior and the evolution of sexual communication in fireflies (Coleoptera: Lampyridae). Cladistics 2003, 19, 1–22. [Google Scholar] [CrossRef]
  18. Stanger-Hall, K.F.; Lloyd, J.E.; Hillis, D.M. Phylogeny of North American fireflies (Coleoptera: Lampyridae): Implications for the evolution of light signals. Mol. Phylogenet. Evol. 2007, 45, 33–49. [Google Scholar] [CrossRef]
  19. Sriboonlert, A.; Wonnapinij, P. Comparative mitochondrial genome analysis of the firefly, Inflata indica (Coleoptera: Lampyridae) and the first evidence of heteroplasmy in fireflies. Int. J. Biol. Macromol. 2019, 121, 671–676. [Google Scholar] [CrossRef]
  20. Martin, G.J.; Branham, M.A.; Whiting, M.F.; Bybee, S.M. Total evidence phylogeny and the evolution of adult bioluminescence in fireflies (Coleoptera: Lampyridae). Mol. Phylogenet. Evol. 2017, 107, 564–575. [Google Scholar] [CrossRef] [Green Version]
  21. Suzuki, H. Molecular phylogenetic studies of Japanese fireflies and their mating systems (Coleoptera: Cantharoidea). Tokyo Metro. Univ. Bull. Nat. Hist. 1997, 3, 1–53. [Google Scholar]
  22. Bocakova, M.; Bocak, L.; Hunt, T.; Teraväinen, M.; Vogler, A.P. Molecular phylogenetics of Elateriformia (Coleoptera): Evolution of bioluminescence and neoteny. Cladistics 2007, 23, 477–496. [Google Scholar] [CrossRef]
  23. Oba, Y.; Branham, M.A. The terrestrial bioluminescent animals of Japan. Zool Sci. 2011, 28, 771–789. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  24. Boore, J.L. Animal mitochondrial genomes. Nucleic Acids Res. 1999, 27, 1767–1780. [Google Scholar] [CrossRef] [Green Version]
  25. Boore, J.L.; Collins, T.M.; Stanton, D.; Daehler, L.L.; Brown, W.M. Deducing the pattern of arthropod phylogeny from mitochondrial DNA rearrangements. Nature 1995, 376, 163–165. [Google Scholar] [CrossRef]
  26. Cameron, S.L. Insect mitochondrial genomics: Implications for evolution and phylogeny. Annu. Rev. Entomol. 2014, 59, 95–117. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  27. Song, F.; Li, H.; Jiang, P.; Zhou, X.; Liu, J.; Sun, C.; Vogler, A.P.; Cai, W. Capturing the phylogeny of holometabola with mitochondrial genome data and Bayesian siteheterogeneous mixture models. Genome Biol. Evol. 2016, 8, 1411–1426. [Google Scholar] [CrossRef] [PubMed]
  28. 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 Biol. Sci. 2017, 284. [Google Scholar] [CrossRef] [PubMed]
  29. Castro, L.R.; Austin, A.D.; Dowton, M. Contrasting rates of mitochondrial molecular evolution in parasitic Diptera and Hymenoptera. Mol. Biol. Evol. 2002, 19, 1100–1113. [Google Scholar] [CrossRef]
  30. Oliveira, D.C.S.G.; Raychoudhury, R.; Lavrov, D.V.; Werren, J.H. Rapidly evolving mitochondrial genome and directional selection in mitochondrial genes in the parasitic wasp Nasonia (Hymenoptera: Pteromalidae). Mol. Biol. Evol. 2008, 25, 2167–2180. [Google Scholar] [CrossRef] [Green Version]
  31. Li, X.; Yan, L.; Pape, T.; Gao, Y.; Zhang, D. Evolutionary insights into bot flies (Insecta: Diptera: Oestridae) from comparative analysis of the mitochondrial genome. Int. J. Biol. Macromol. 2020, 149, 371–380. [Google Scholar] [CrossRef]
  32. Cameron, S.L.; Lambkin, C.L.; Barker, S.C.; Whiting, M.F. A mitochondrial genome phylogeny of Diptera: Whole genome sequence data accurately resolve relationships over broad timescales with high precision. Syst. Entomol. 2007, 32, 40–59. [Google Scholar] [CrossRef]
  33. Du, Z.; Hasegawa, H.; Cooley, J.R.; Simon, C.; Yoshimura, J.; Cai, W.; Sota, T.; Li, H. Mitochondrial genomics reveals shared phylogeographic patterns and demographic history among three periodical cicada species groups. Mol. Biol. Evol. 2019, 36, 1187–1200. [Google Scholar] [CrossRef]
  34. Zhou, X.; Li, Y.; Liu, S.; Yang, Q.; Su, X.; Zhou, L.; Tang, M.; Fu, R.; Li, J.; Huang, Q. Ultra-deep sequencing enables high-fidelity recovery of biodiversity for bulk arthropod samples without PCR amplification. GigaScience 2013, 2, 1–12. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  35. Peng, Y.; Leung, H.C.M.; Yiu, S.M.; Chin, F.Y.L. IDBA-UD: A de novo assembler for single-cell and metagenomic sequencing data with highly uneven depth. Bioinformatics 2012, 28, 1420–1428. [Google Scholar] [CrossRef] [Green Version]
  36. Kearse, M.; Moir, R.; Wilson, A.; Stones-Havas, S.; Cheung, M.; Sturrock, S.; Buxton, S.; Cooper, A.; Markowitz, S.; Duran, C.; et al. Geneious Basic: An integrated and extendable desktop software platform for the organization and analysis of sequence data. Bioinformatics 2012, 28, 1647–1649. [Google Scholar] [CrossRef] [PubMed]
  37. Bernt, M.; Donath, A.; Jühling, F.; Gärtner, F.; Florentz, C.; Fritzsch, G.; Pütz, J.; Middendorf, M.; Stadler, P.F. MITOS: Improved de novo metazoan mitochondrial genome annotation. Mol. Phylogenetics Evol. 2013, 69, 313–319. [Google Scholar] [CrossRef]
  38. Laslett, D.; Canbäck, B. ARWEN: A program to detect tRNA genes in metazoan mitochondrial nucleotide sequences. Bioinformatics 2008, 24, 172–175. [Google Scholar] [CrossRef] [Green Version]
  39. 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]
  40. Grant, J.R.; Stothard, P. The CGView server: A comparative genomics tool for circular genomes. Nucleic Acids Res. 2008, 36, 181–184. [Google Scholar] [CrossRef]
  41. Perna, N.T.; Kocher, T.D. Patterns of nucleotide composition at fourfold degenerate sites of animal mitochondrial genomes. J. Mol. Evol. 1995, 41, 353–358. [Google Scholar] [CrossRef]
  42. Benson, G. Tandem repeats finder: A program to analyze DNA sequences. Nucleic Acids Res. 1999, 37, 573–580. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  43. Kumar, S.; Stecher, G.; Tamura, K. MEGA7: Molecular evolutionary genetics analysis version 7.0 for bigger datasets. Mol. Biol. Evol. 2016, 33, 1870–1874. [Google Scholar] [CrossRef] [Green Version]
  44. Librado, P.; Rozas, J. DnaSPv5: A software for comprehensive analysis of DNA polymorphism data. Bioinformatics 2009, 25, 1451–1452. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  45. Zhang, D.; Gao, F.; Jakovli’c, I.; Zou, H.; Zhang, J.; Li, W.X.; Wang, G.T. PhyloSuite: An integrated and scalable desktop platform for streamlined molecular sequence data management and evolutionary phylogenetics studies. Mol. Ecol. Resour. 2020, 20, 348–355. [Google Scholar] [CrossRef]
  46. Bowker, A.H. A test for symmetry in contingency tables. J. Am. Stat. Assoc. 1948, 43, 572–574. [Google Scholar] [CrossRef]
  47. Abascal, F.; Zardoya, R.; Telford, J.M. TranslatorX: Multiple alignment of nucleotide sequences guided by amino acid translations. Nucleic Acids Res. 2010, 38, 7–13. [Google Scholar] [CrossRef] [Green Version]
  48. Talavera, G.; Castresana, J. Improvement of phylogenies after removing divergent and ambiguously aligned blocks from protein sequence alignments. Syst. Biol. 2007, 56, 564–577. [Google Scholar] [CrossRef] [Green Version]
  49. 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 2011, 27, 171–180. [Google Scholar] [CrossRef]
  50. Nguyen, L.-T.; Schmidt, H.A.; von Haeseler, A.; Minh, B.Q. IQ-TREE: A fast and effective stochastic algorithm for estimating maximum-likelihood phylogenies. Mol. Biol. Evol. 2014, 32, 268–274. [Google Scholar] [CrossRef] [PubMed]
  51. Lartillot, N.; Rodrigue, N.; Stubbs, D.; Richer, J. PhyloBayes MPI: Phylogenetic reconstruction with infinite mixtures of profiles in a parallel environment. Syst. Biol. 2013, 62, 611–615. [Google Scholar] [CrossRef] [Green Version]
  52. Ronquist, F.; Huelsenbeck, J.P. MrBayes 3: Bayesian phylogenetic inference under mixed models. Bioinformatics 2003, 19, 1572–1574. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  53. Letunic, I.; Bork, P. Interactive tree of life (iTOL) v4: Recent updates and new developments. Nucleic Acids Res. 2019, 47, W256–W259. [Google Scholar] [CrossRef] [Green Version]
  54. Amaral, D.T.; Mitani, Y.; Ohmiya, Y.; Viviani, V.R. Organization and comparative analysis of the mitochondrial genomes of bioluminescent Elateroidea (Coleoptera: Polyphaga). Gene 2016, 586, 254–262. [Google Scholar] [CrossRef] [PubMed]
  55. Li, X.-Y.; Ogoh, K.; Ohba, N.; Liang, X.-C.; Ohmiya, Y. Mitochondrial genomes of two luminous beetles, Rhagophthalmus lufengensis and R. ohbai (Arthropoda, Insecta, Coleoptera). Gene 2007, 392, 196–205. [Google Scholar] [CrossRef]
  56. Linard, B.; Crampton-Platt, A.; Morinière, J.; Timmermans, M.; Arribas, P.; Miller, K.; Lipecki, J.; Favreau, E.; Hunter, A. The contribution of mitochondrial metagenomics to large-scale data mining and phylogenetic analysis of Coleoptera. Mol. Phylogenet. Evol. 2018, 128, 1–11. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  57. Fallon, T.R.; Lower, S.E.; Chang, C.H.; Bessho-Uehara, M.; Weng, J.K. Firefly genomes illuminate parallel origins of bioluminescence in beetles. eLife 2018, 7. [Google Scholar] [CrossRef]
  58. Bae, J.S.; Kim, I.; Sohn, H.D.; Jin, B.R. The mitochondrial genome of the firefly, Pyrocoelia rufa: Complete DNA sequence, genome organization, and phylogenetic analysis with other insects. Mol. Phylogenet. Evol. 2004, 32, 978–985. [Google Scholar] [CrossRef] [PubMed]
  59. Hu, J.; Fu, X. The complete mitochondrial genome of the firefly, Abscondita anceyi (Olivier) (Coleoptera: Lampyridae). Mitochondrial DNA B 2018, 3, 442–443. [Google Scholar] [CrossRef]
  60. Jiao, H.; Ding, M.; Zhao, H. Sequence and organization of complete mitochondrial genome of the firefly, Aquatica leii (Coleoptera: Lampyridae). Mitochondrial DNA 2015, 26, 775–776. [Google Scholar] [CrossRef]
  61. Wang, K.; Hong, W.; Jiao, H.; Zhao, H. Transcriptome sequencing and phylogenetic analysis of four species of luminescent beetles. Sci. Rep. 2017, 7, 1814. [Google Scholar] [CrossRef]
  62. Maeda, J.; Kato, D.I.; Arima, K.; Ito, Y.; Toyoda, A.; Noguchi, H. The complete mitochondrial genome sequence and phylogenetic analysis of Luciola lateralis, one of the most famous firefly in Japan (Coleoptera: Lampyridae). Mitochondrial DNA B 2017, 2, 546–547. [Google Scholar] [CrossRef] [Green Version]
  63. Hu, J.; Fu, X. The complete mitochondrial genome of the firefly, Luciola curtithorax (Coleoptera: Lampyridae). Mitochondrial DNA B 2018, 3, 378–379. [Google Scholar] [CrossRef] [Green Version]
  64. Fan, Y.; Fu, X. The complete mitochondrial genome of the firefly, Pteroptyx maipo (Coleoptera: Lampyridae). Mitochondrial DNA B 2017, 2, 795–796. [Google Scholar] [CrossRef] [Green Version]
  65. Mu, F.J.; Liang, A.; Zhao, H.B.; Kai, W. Characterization of the complete mitochondrial genome of the firefly, Luciola substriata (Coleoptera: Lampyridae). Mitochondrial DNA A 2016, 27, 3360–3362. [Google Scholar] [CrossRef]
  66. Zhang, B.; Ma, C.; Edwards, O.; Fuller, S.; Kang, L. The mitochondrial genome of the Russian wheat aphid Diuraphis noxia: Large repetitive sequences between trnE and trnF in aphids. Gene 2014, 533, 253–260. [Google Scholar] [CrossRef] [PubMed]
  67. Lavrov, D.V.; Boore, J.L.; Brown, W.M. The complete mitochondrial DNA sequence of the horseshoe crab Limulus polyphemus. Mol. Biol. Evol. 2000, 17, 813–824. [Google Scholar] [CrossRef] [Green Version]
  68. Zhang, H.; Liu, Q.; Lu, C.; Deng, J.; Huang, X. The first complete mitochondrial genome of Lachninae species and comparative genomics provide new insights into the evolution of gene rearrangement and the repeat region. Insects 2021, 12, 55. [Google Scholar] [CrossRef] [PubMed]
  69. Ojala, D.; Montoya, J.; Attardi, G. tRNA punctuation model of RNA processing in human mitochondria. Nature 1981, 290, 470–474. [Google Scholar] [CrossRef]
  70. Hurst, L.D. The Ka/Ks ratio: Diagnosing the form of sequence evolution. Trends Genet. 2002, 18, 486–487. [Google Scholar] [CrossRef]
  71. Yang, Z.; Bielawski, J.R. Statisticalmethods for detecting molecular adaptation. Trends Ecol. Evol. 2000, 15, 496–503. [Google Scholar] [CrossRef]
  72. Mori, S.; Matsunami, M.J.G. Signature of positive selection in mitochondrial DNA in Cetartiodactyla. Genes Genet. Syst. 2018, 93, 65–73. [Google Scholar] [CrossRef] [Green Version]
  73. Lavrov, D.V.; Brown, W.M.; Boore, J.L. A novel type of RNA editing occurs in the mitochondrial tRNAs of the centipede Lithobius forficatus. Proc. Natl. Acad. Sci. USA 2000, 97, 13738–13742. [Google Scholar] [CrossRef] [Green Version]
  74. Masta, S.E.; Boore, J.L. The complete mitochondrial genome sequence of the spider Habronattus oregonensis reveals rearranged and extremely truncated tRNAs. Mol. Biol. Evol. 2004, 21, 893–902. [Google Scholar] [CrossRef] [Green Version]
  75. 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]
  76. Yan, L.; Pape, T.; Elgar, M.A.; Gao, Y.; Zhang, D. Evolutionary history of stomach bot flies in the light of mitogenomics. Syst. Entomol. 2019, 44, 797–809. [Google Scholar] [CrossRef]
  77. Song, F.; Li, H.; Liu, G.; Wang, W.; James, P.; Colwell, D.D.; Tran, A.; Gong, S.; Cai, W.; Shao, R. Mitochondrial genome fragmentation unites the parasitic lice of Eutherian mammals. Syst. Biol. 2019, 68, 430–440. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  78. Foerstner, K.U.; von Mering, C.; Hooper, S.D.; Bork, P. Environments shape the nucleotide composition of genomes. EMBO Rep. 2005, 6, 1208–1213. [Google Scholar] [CrossRef]
  79. Chen, W.-H.; Lu, G.; Bork, P.; Hu, S.; Lercher, M.J. Energy efficiency trade-offs drive nucleotide usage in transcribed regions. Nat. Commun. 2016, 7, 11334. [Google Scholar] [CrossRef] [Green Version]
  80. Rocha, E.P.C.; Danchin, A. Base composition bias might result from competition for metabolic resources. Trends Genet. 2002, 18, 291–294. [Google Scholar] [CrossRef]
  81. Kono, N.; Tomita, M.; Arakawa, K. Accelerated laboratory evolution reveals the influence of replication on the GC skew in Escherichia coli. Genome Biol. Evol. 2018, 10, 3110–3117. [Google Scholar] [CrossRef] [Green Version]
  82. Jia, W.Z.; Yan, H.B.; Guo, A.J.; Zhu, X.Q.; Wang, Y.C.; Shi, W.G.; Chen, H.T.; Zhan, F.; Zhang, S.H. Complete mitochondrial genomes of Taenia multiceps, T. hydatigena and T. pisiformis: Additional molecular markers for a tapeworm genus of human and animal health significance. BMC Genomics 2010, 11, 447. [Google Scholar] [CrossRef] [Green Version]
  83. Ye, F.; Easy, R.H.; King, S.D.; Cone, D.K.; You, P. Comparative analyses within Gyrodactylus (Platyhelminthes: Monogenea) mitochondrial genomes and conserved polymerase chain reaction primers for gyrodactylid mitochondrial DNA. J. Fish. Dis. 2017, 40, 541–555. [Google Scholar] [CrossRef]
  84. Brabec, J.; Kostadinova, A.; Scholz, T.; Littlewood, D.T.J. Complete mitochondrial genomes and nuclear ribosomal RNA operons of two species of Diplostomum (Platyhelminthes: Trematoda): A molecular resource for taxonomy and molecular epidemiology of important fish pathogens. Parasit. Vectors. 2015, 8, 336. [Google Scholar] [CrossRef] [Green Version]
  85. Demari-Silva, B.; Foster, P.G.; de Oliveira, T.M.P.; Bergo, E.S.; Sanabani, S.S.; Pessôa, R.; Sallum, M.A.M. Mitochondrial genomes and comparative analyses of Culex camposi, Culex coronator, Culex usquatus and Culex usquatissimus (Diptera:Culicidae), members of the coronator group. BMC Genomics 2015, 16, 831. [Google Scholar] [CrossRef] [Green Version]
  86. Du, Y.; Dai, W.; Dietrich, C.H. Mitochondrial genomic variation and phylogenetic relationships of three groups in the genus Scaphoideus (Hemiptera: Cicadellidae: Deltocephalinae). Sci. Rep. 2017, 7, 16908. [Google Scholar] [CrossRef] [Green Version]
  87. Nunes, V.C.S.; Souto, P.M.; Minelli, A.; Stanger-Hall, K.F.; Silveira, L.F.L. Antennomere numbers in fireflies (Coleoptera: Lampyridae): Unique patterns and tentative explanations. Zool. Anz. 2020, 286, 1–10. [Google Scholar] [CrossRef]
  88. Green, J.W. Revision of the Species of Microphotus, with an emendation of the Lampyrini (Lampyridae). Coleopt. Bull. 1959, 13, 80–96. [Google Scholar]
  89. McDermott, F.A. Coleopterorum Catalogus Supplementa, Pars 9: Lampyridae, 2nd ed.; Steel, W.O., Ed.; W Junk: The Hague, The Netherlands, 1966; pp. 1–149. [Google Scholar]
  90. Nakane, T. Lampyrid insects of the world. In The Reconstruction of Firefly Environments. Special No. 1; The Association of Natural Restoration of Japan, Ed.; Saiteku: Tokyo, Japan, 1991; pp. 3–11. [Google Scholar]
  91. Lawrence, J.F.; Newton, A.F. Families and subfamilies of Coleoptera (with selected genera, notes, references and data on family-group names). In Biology, Phylogeny, and the Classification of Coleoptera: Papers Celebrating the 80th Birthday of Roy A. Crowson; Pakaluk, J., Ślipiński, S.A., Eds.; Muzeum I InstytutZoologii PAN: Warsaw, Poland, 1995; pp. 779–1092. [Google Scholar]
  92. McDermott, F.A. The taxonomy of the Lampyridae (Coleoptera). Trans. Am. Entomol. Soc. 1964, 90, 1–72. [Google Scholar] [CrossRef]
  93. Brancucci, M.; Geiser, M. A revision of the genus Lamellipalpus Maulik, 1921 (Coleoptera, Lampoyridae). Zootaxa 2009, 2080, 1–20. [Google Scholar] [CrossRef] [Green Version]
  94. Sagegami-Oba, R.; Takahashi, N.; Oba, Y. The evolutionary process of bioluminescence and aposematism in cantharoid beetles (Coleoptera: Elateroidea) inferred by the analysis of 18S ribosomal DNA. Gene 2007, 400, 104–113. [Google Scholar] [CrossRef]
  95. Geisthardt, M.; Satô, M. Lampyridae. In Catalogue of Palaearctic Coleoptera; Löbl, I., Smetana, A., Eds.; Apollo Books: Stenstrup, Denmark, 2007; Volume 4, pp. 225–233. [Google Scholar]
  96. Janisova, K.; Bocakova, M. Revision of the subfamily Ototretinae (Coleoptera: Lampyridae). Zool. Anz. 2013, 252, 1–19. [Google Scholar] [CrossRef]
  97. Li, X.Y.; Yang, S.; Xie, M.; Liang, X.C. Phylogeny of fireflies (Coleoptera: Lampyridae) inferred from mitochondrial 16S ribosomal DNA, with references to morphological and ethological traits. Prog. Nat. Sci. 2006, 16, 817–826. [Google Scholar] [CrossRef]
  98. Jeng, M.L. Comprehensive phylogenetics, systematics, and evolution of neoteny of Lampyridae (Insecta: Coleoptera). Ph.D. Thesis, University of Kansas, Lawrence, KS, USA, 2008; 388p. [Google Scholar]
  99. Fu, X.H.; Ballantyne, L.A.; Lambkin, C.L. Emeia gen. nov., a new genus of Luciolinae fireflies from China (Coleoptera: Lampyridae) with an unusual trilobite-like larva, and a redescription of the genus Curtos Motschulsky. Zootaxa 2012, 3403, 1–53. [Google Scholar] [CrossRef]
Figure 1. Circle map of the complete mitogenome of C. sanguineus klapperichi. Different colors indicate different types of genes and regions. Genes shown at the outer circle are located on the J-strand, and those at the inner circle are located at the N-strand.
Figure 1. Circle map of the complete mitogenome of C. sanguineus klapperichi. Different colors indicate different types of genes and regions. Genes shown at the outer circle are located on the J-strand, and those at the inner circle are located at the N-strand.
Insects 12 00570 g001
Figure 2. The size (A) and AT content (B) of PCGS, rrnL, rrnS, CR, and tRNA of 33 firefly species with the red dots representing C. sanguineus klapperichi (lower edge of the gray rectangle, 25 percentile; central black bar within the rectangle, median; upper edge of the rectangle, 75 percentile); Nucleotide composition of 33 complete firefly species, with the black dots representing the C. sanguineus klapperichi: (C) the A+T content and AT skew; (D) the G+C content and GC-skew.
Figure 2. The size (A) and AT content (B) of PCGS, rrnL, rrnS, CR, and tRNA of 33 firefly species with the red dots representing C. sanguineus klapperichi (lower edge of the gray rectangle, 25 percentile; central black bar within the rectangle, median; upper edge of the rectangle, 75 percentile); Nucleotide composition of 33 complete firefly species, with the black dots representing the C. sanguineus klapperichi: (C) the A+T content and AT skew; (D) the G+C content and GC-skew.
Insects 12 00570 g002
Figure 3. (A) The nucleotide diversity (Pi) of 13 protein-coding genes of mitogenome among 33 species of Lampyridae in a sliding window analysis (a sliding window of 200 bp with the step size of 20 bp); the Pi value of each gene is shown under the gene name. (B) Genetic distances and the ratio of non-synonymous (Ka) to synonymous (Ks) substitution rates of 13 protein-coding genes among 33 species of Lampyridae. The average value for each PCGs is shown under the gene name.
Figure 3. (A) The nucleotide diversity (Pi) of 13 protein-coding genes of mitogenome among 33 species of Lampyridae in a sliding window analysis (a sliding window of 200 bp with the step size of 20 bp); the Pi value of each gene is shown under the gene name. (B) Genetic distances and the ratio of non-synonymous (Ka) to synonymous (Ks) substitution rates of 13 protein-coding genes among 33 species of Lampyridae. The average value for each PCGs is shown under the gene name.
Insects 12 00570 g003
Figure 4. Heterogeneous sequence divergence of mitochondrial genomes of Lampyridae resulting from pairwise comparison of three aligned datasets: (A) PCG12rRNA; (B) PCGrRNA; (C) PCGRNA. The dark colors indicate the higher randomized accordance, while the lighter colors indicate the opposite. All taxa names (indicated by the abbreviations consist of the first two letters of the genera and species names, respectively) are listed on top and to the right of the heat map. While cells specify p-values > 0.05, indicating that corresponding pairs of nucleotide sequences do not violate the assumption of global stationary, reversibility, and homogeneity conditions.
Figure 4. Heterogeneous sequence divergence of mitochondrial genomes of Lampyridae resulting from pairwise comparison of three aligned datasets: (A) PCG12rRNA; (B) PCGrRNA; (C) PCGRNA. The dark colors indicate the higher randomized accordance, while the lighter colors indicate the opposite. All taxa names (indicated by the abbreviations consist of the first two letters of the genera and species names, respectively) are listed on top and to the right of the heat map. While cells specify p-values > 0.05, indicating that corresponding pairs of nucleotide sequences do not violate the assumption of global stationary, reversibility, and homogeneity conditions.
Insects 12 00570 g004
Figure 5. Phylogenetic tree of Lampyridae inferred from the BI analysis of the PCGRNA dataset or ML analysis of PCGrRNA dataset. Numbers under the branches are bootstrap values (right) or posterior probabilities (left).
Figure 5. Phylogenetic tree of Lampyridae inferred from the BI analysis of the PCGRNA dataset or ML analysis of PCGrRNA dataset. Numbers under the branches are bootstrap values (right) or posterior probabilities (left).
Insects 12 00570 g005
Table 1. Information for the representative species’ mitogenomes used for phylogenetic analysis.
Table 1. Information for the representative species’ mitogenomes used for phylogenetic analysis.
FamilySubfamilySpeciesAccession NumberReference
Phengodidae Phrixothrix hirtusKM923891.1[54]
RhagophthalmidaeRhagophthalmus ohbaiNC_010964.1[55]
LampyridaeLampyrinaeDiaphanes nubilusMK292094.1[2]
Diaphanes sp. MK292095.1[2]
Diaphanes citrinusMK292103.1[2]
Diaphanes mendaxMK292116.1[2]
Diaphanes pectinealisMK292118.1[2]
Ellychnia corruscaMG242622.1Unpublished
Lampyris noctilucaKX087302.1[56]
Photinus pyralisKY778696.1[57]
Pyrocoelia rufaAF452048.1[58]
Pyrocoelia praetextaMK292115.1[2]
Pyrocoelia thibetanaMK292117.1[2]
LuciolinaeAbscondita anceyiMH020192.1[59]
Abscondita terminalisMK292092.1[2]
Aquatica leiiKF667531.1[60]
Aquatica fictaKX758085.1[61]
Aquatica wuhanaKX758086.1[61]
Aquatica lateralisNC_035755.1[62]
Asymmetricata circumdataMK292113.1[2]
Curtos bilineatusMK292114.1[2]
Emeia pseudosauteriMK292112.1[2]
Luciola curtithoraxMG770613.1[63]
Luciola cruciataNC_022472.1[19]
Pristolycus sp.MK292099.1[2]
Pteroptyx maipoMF686051.1[64]
Pygoluciola qingyuMK292093.1[2]
Pygoluciola sp. MK292102.1[2]
Luciola substriataKP313820.1[65]
Incertae sedisVesta saturnalisMK292111.1[2]
Lamprigera yunnanaMK292091.1[2]
PhoturinaeBicellonycha lividipennisKJ922151.1[53]
OtotretinaeDrilaster sp. MK292100.1[2]
Stenocladius sp. MK292101.1[2]
CyphonocerinaeCyphonocerus sanguineus klapperichiMW365445In this study
Note: “unpublished“ means the sequence with an accession number of the species could be download from the NCBI, but the publication could not be found.
Table 2. Nucleotide composition and skewness of mitogenomes of C. sanguineus klapperichi.
Table 2. Nucleotide composition and skewness of mitogenomes of C. sanguineus klapperichi.
RegionsSize
(bp)
A
(%)
C
(%)
G
(%)
T
(%)
AT
(%)
GC
(%)
AT
Skew
GC
Skew
Full genome16,44342.513.99.434.276.723.30.11−0.19
PCGs11,00841.114.610.33475.124.90.09−0.17
1st codon position366933.212.215.539.172.327.7−0.080.12
2nd codon position366925.214.912.347.772.927.2−0.31−0.10
3rd codon position366937.810.69.342.480.219.9−0.06−0.07
tRNAs143241.612.99.635.977.522.50.07−0.15
rRNAs20344713.16.433.580.519.50.17−0.34
CR177645.611.67.235.681.218.80.12−0.23
Publisher’s Note: MDPI stays neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Share and Cite

MDPI and ACS Style

Ge, X.; Yuan, L.; Kang, Y.; Liu, T.; Liu, H.; Yang, Y. Characterization of the First Complete Mitochondrial Genome of Cyphonocerinae (Coleoptera: Lampyridae) with Implications for Phylogeny and Evolution of Fireflies. Insects 2021, 12, 570. https://doi.org/10.3390/insects12070570

AMA Style

Ge X, Yuan L, Kang Y, Liu T, Liu H, Yang Y. Characterization of the First Complete Mitochondrial Genome of Cyphonocerinae (Coleoptera: Lampyridae) with Implications for Phylogeny and Evolution of Fireflies. Insects. 2021; 12(7):570. https://doi.org/10.3390/insects12070570

Chicago/Turabian Style

Ge, Xueying, Lilan Yuan, Ya Kang, Tong Liu, Haoyu Liu, and Yuxia Yang. 2021. "Characterization of the First Complete Mitochondrial Genome of Cyphonocerinae (Coleoptera: Lampyridae) with Implications for Phylogeny and Evolution of Fireflies" Insects 12, no. 7: 570. https://doi.org/10.3390/insects12070570

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