The Complete Mitochondrial Genome of Sphaeniscus atilius (Walker, 1849) (Diptera: Tephritidae) and Implication for the Phylogeny of Tephritidae
The Complete Mitochondrial Genome of Sphaeniscus atilius (Walker, 1849) (Diptera: Tephritidae) and Implication for the Phylogeny of Tephritidae
Shibao Guo1, Junhua Chen1, Nan Song2, Fangmei Zhang1*
1Xinyang Agriculture and Forestry University, Xinyang 464000, China
2Henan Agriculture University, Zhengzhou 450002, China
ABSTRACT
The complete mitochondrial genome of Sphaeniscus atilius was characterized and annotated in this study. The mitogenome was 16,854 bp in length and encoded 37 typical mitochondrial genes, including 13 protein-coding genes, 22 tRNA genes, 2 ribosomal RNA genes, and 1 control regions. The total length of the 13 PCGs was 11,140 bp, and the AT content was 79.8%. There were five types of start codons, ATT (nad2, nad3, nad5, and nad6), ATG (cox2, cox3, atp6, nad4, nad4l, and cob), CGA (cox1), as well as ATC (atp8) and ATA (nad1). Most of the PCGs had typical TAA stop codons, except nad5 which terminated with incomplete forms T-. Ile, Phe, Leu and Asn were the most frequently used amino acids in mitochondrial PCGs. Most tRNA genes could be folded into the typical cloverleaf structure, except trnS1 and trnT which lacked the dihydrouridine (DHU) and TΨC arms, respectively. Phylogenetic analyses based on 13 protein-coding genes among the available sequenced species of family Tephritidae by maximum likelihood and bayesian inference methods suggested the genus relationship of Tephritidae: ((Bactrocera, Dacus, Zeugodacus), Felderimyia, Anastrepha), (Acrotaeniostola, (Neoceratitis, Ceratitis), Euleia, Rivellia), (Procecidochares, (Tephritis, Sphaenisscus))))). Our results presented the first mitogenome from Sphaeniscus and provide insights into the species identification, taxonomy and phylogeny of S. atilius.
Article Information
Received 12 September 2023
Revised 09 December 2023
Accepted 26 December 2023
Available online 04 April 2024
(early access)
Published 08 November 2024
Authors’ Contribution
Data curation: SG, NS. Formal analysis: FZ, NS. Investigation: JC. Methodology: FZ, NS. Funding acquisition: SG. Supervision: JC. Writing-original draft: SG, FZ. Writing-review and editing: SG, FZ, NS.
Key words
Tephritidae, Sphaeniscus atilius, Mitochondrial genome, Phylogeny
DOI: https://dx.doi.org/10.17582/journal.pjz/20230912053433
* Corresponding author: zhangfm@xyafu.edu.cn
0030-9923/2024/0006-2929 $ 9.00/00
Copyright 2024 by the authors. Licensee Zoological Society of Pakistan.
This article is an open access article distributed under the terms and conditions of the Creative Commons Attribution (CC BY) license (https://creativecommons.org/licenses/by/4.0/).
INTRODUCTION
Tephritidae, one of the largest families of Diptera, consists of more than 500 genera and almost 5,000 named species and predominantly distributes throughout the temperate and tropical areas of the world (Pape et al., 2009; Aluja and Norrbom, 2000; Mazzon et al., 2021). This family is also referred to as true fruit flies, with hundreds of fruit-eating species accounting for about 40% of the species. It has been reported to attack a great variety of fruit plants, bamboo culms, vegetables, flowers, and seeds (Korneyev, 1999; Dohm et al., 2014). In practice, some of the fruit-eating species in the Anastrepha, Bactrocera, Ceratitis, Dacus, and Rhagoletis genera have been considered serious agricultural pests due to their significant economic impact on the production of fruit crops and stored fruit (White and Elson-Harris, 1992; Aluja and Mangan, 2008). Melon fly, Zeugodacus cucurbitae (Siderhurst and Jang, 2010), medfly, Ceratitis fasciventris (Drosopoulou et al., 2017), together with Bactrocera latifrons (Yong et al., 2016), are well-known examples.
The insect mitochondrial genome has been regarded as a useful molecular marker in studies of phylogenetic and evolutionary analysis, genetic diversity, and species delimitation at the genus or species level, due to its small size, high copy numbers, maternal inheritance, unambiguous orthologous genes, conserved gene composition, and high evolutionary rate (Cameron, 2014; Song et al., 2016; Wilson and Xu, 2012). Generally, the typical insect mitochondrial genome is a highly conserved circular molecule ranging in size from approximately 14 to 40 kbp, encoding a fixed set of 37 genes, including 13 protein-coding genes (PCGs), 22 transfer RNA (tRNA) genes, 2 ribosomal RNA (rRNA) genes, and a control region (CR) or the A+T-rich region (Du et al., 2021; Wolstenholme, 1992), with only a few exceptions. For example, long gene intergenic spacers, gene rearrangements, and gene loss have also been reported in different orders of insects (Du et al., 2017; Gong et al., 2018; Yan et al., 2022).
Partial mitochondrial gene sequences have become a preferred approach for inferring phylogenetic and molecular systematic studies in several insect groups, such as Arma custos and Picromerus lewisi (Hemiptera) (Mu et al., 2022), Episymploce splendens (Blattodea) (Yan et al., 2022), and Haematopinus tuberculatus (Psocodea) (Fu et al., 2022), Coomaniella copipes, Coomaniella dentata, and Dicerca corrugate (Coleoptera) (Huang et al., 2022), including the Tephritidae family (Ceratitis fasciventris (Drosopoulou et al., 2017), Bactrocera carambolae (Drosopoulou et al., 2019), Bactrocera biguttula (Teixeira et al., 2019), Zeugodacus cucurbitae (Zhou et al., 2020), and Lepidotrigona flavibasis (Wang et al., 2021). The aforementioned studies of mitochondrial gene sequences explored the origin and evolution of insects, explained the species and evolution of the system, and revealed the geographical distribution of intraspecific polymorphism.
Sphaeniscus atilius (Diptera: Tephritidae: Sphaeniscus), which is distributed from India to Russia, Korea, Japan, Australasian and Oceanian regions, can be morphologically distinguished from any other tephritid species based on clear diagnostic morphological features, including an almost entirely dark brown body, the number of orbital frontal setae, and an almost perpendicular base of the discal band of the wing (Han et al., 2010). To date, the complete mitochondrial genomes (mitogenomes) of 46 species, belonging to 14 genera of Tephritidae, are available in GenBank (https://www.ncbi.nlm.nih.gov/nucleotide/). All of these species belong to 5 subfamilies, Sciomyzidae, Sepsidae, Lauxaniidae, Celyphidae, Platystomatidae, and Tephritidae, respectively. However, there are no reports on the molecular phylogeny studies of the mitochondrial genome information in S. atilius, which limits our comprehensive understanding of the evolutionary and phylogenetic relationships of S. atilius.
In the current study, we sequenced, annotated, and described the complete mitogenome of S. atilius using next-generation sequencing, which is the first complete mitogenome sequence reported in the genus Sphaeniscus. We predicted and analyzed the gene organization, base composition, PCGs, codon usage, and the structure of the tRNAs and rRNAs of its mitochondrial genome. Additionally, we carried out phylogenetic analyses based on maximum likelihood (ML) and Bayesian Inference (BI) methods to assess the phylogenetic position of S. atilius. These results will be greatly helpful for clarifying the phylogenetic status and relationships between different species of Tephritidae.
MATERIALS AND METHODS
Taxon sampling and DNA extraction
Specimens of S. atilius were collected in Mount Jigong, Xinyang, Henan Province, China (31°48′43″N, 114°05′43″E), in June 2020. Specimens were preserved in 100% ethanol, and stored at -20°C. After morphological identification, total genomic DNA was extracted from muscle tissue of pre-thoraxes using Tissue DNA kit (TIANGEN Biotech, Beijing, China) according to the manufacture’s protocol. The impurities and concentration were detected by agarose (1%) electrophoresis and Nanodrop spectrophotometer (ThermoFisher Scientific, Waltham, MA), respectively.
Sequencing and assembling of mitochondrial genome
Genome sequencing was performed on an Illumina HiSeq 2500 platform (150 bp paired-end reads). The TruSeq library was prepared with an insert size of 400 bp. Quantity of sequencing data for each sample was at least 20 Gb. Sequencing was performed at Beijing Novogene Bioinformatics Technology Co., Ltd, China. A total of 2 Gb raw paired reads were generated. Data filtering was conducted using NGS QC-Toolkit v2.5 (Patel and Jain, 2012). After removing the connector and the unmatched, short, and poor-quality reads, the high-quality reads (Q20 > 90% and Q30 > 80%) were used for genome assembly.
De novo assembly was performed with IDBA-UD v. 1.1.1 (Peng et al., 2012). The parameter settings were as follows: 200 minimum sizes of contig, 41 as the minimum k-mer size, 10 as an iteration size, and 91 as the maximum k-mer size. The pre sequenced mitochondrial cox1 gene was used to bait the mitogenome from the assembled contigs.
Mitochondrial genome annotation and analysis
Preliminary annotation of mitogenome was conducted using MITOS (http://mitos.bioinf.uni-leipzig.de/index.py) (Bernt et al., 2013). The gene boundaries were refined by blasting against closely related species. The secondary structures of tRNA genes were predicted in MITOS. The structure map of the mitogenome was drawn using OGDRAW v1.3.1 (Greiner et al., 2019). The annotated mitogenome sequence of S. atilius was deposited at GenBank (accession number OQ909100).
The mitogenome nucleotide composition and relative synonymous codon usage (RSCU) were computed using MEGA 7.0 (Kumar et al., 2016). AT and GC skews were calculated following the formula: AT Skew = (A − T)/(A + T), GC Skew = (G − C)/(G + C) (Perna and Kocher, 1995). The nucleotide diversity (Pi) and nonsynonymous (Ka)/synonymous (Ks) mutation rate ratios were calculated by DnaSP v5.10.01 (Librado and Rozas, 2009).
Phylogenetic analysis
The new mitogenome sequence of S. atilius was merged with the existing dipteran mitogenome sequences. The amino acid sequences of 13 protein coding genes of 38 dipteran species were aligned individually using MAFFT with default parameters (Katoh and Standley, 2013). We chose Pachycerina (Pachycerina decemlineata) and Sciomyza (Sciomyza simplex) as outgroups. The alignments were trimmed by trimAl (Capella-Gutierrez et al., 2009). The resulting alignments were concatenated using SequenceMatrix 1.8 (Vaidya et al., 2011).
The phylogenetic analysis was conducted using ML and BI methods. ML tree was reconstructed using IQ-TREE v2.2.3 (Minh et al., 2020). The best-fitting model for the alignment was estimated by ModelFinder (Kalyaanamoorthy et al., 2017). Branch support was assessed by 10,000 ultrafast bootstrap replicates (Hoang et al., 2018). MrBayes 3.1 was used to reconstruct the BI tree, and four independent Markov chain runs were performed for 1,000,000 metropolis–coupled (MCMC) generations, sampling a tree every 100 generations.
RESULTS
Genome organization
The length of the S. atilius mitogenome was 16,854 bp in length, which consisted of the typical 37 genes including PCGs, 22 tRNA genes, and two rRNA genes (Fig. 1, Table I). In addition, a major non-coding region known as the control region (CR) or A+T-rich region was found between rrnS and trnI. The heavy chain (H-strand) encoded 23 genes (nine PCGs and 14 tRNAs). The remaining 4 PCGs, 8 tRNAs and 2 rRNAs were transcribed in the light chain (L-strand).
Table I. Summary of the mitogenome of S. atilius.
Gene |
Strand |
Location |
Size (bp) |
Anti codon |
Start codon |
Stop codon |
Intergenic nucleotides |
trnI |
H |
1-67 |
67 |
GAT |
7 |
||
trnQ |
L |
67-135 |
69 |
TTG |
-1 |
||
trnM |
H |
157-228 |
72 |
CAT |
21 |
||
nad2 |
H |
243-1250 |
1008 |
ATT |
TAA |
14 |
|
trnW |
H |
1242-1309 |
68 |
TCA |
-8 |
||
trnC |
L |
1311-1376 |
66 |
GCA |
1 |
||
trnY |
L |
1377-1443 |
67 |
GTA |
47 |
||
cox1 |
H |
1491-3026 |
1536 |
CGA |
TAA |
-5 |
|
trnL2 |
H |
3021-3086 |
66 |
TAA |
10 |
||
cox2 |
H |
3110-3793 |
684 |
ATG |
TAA |
23 |
|
trnK |
H |
3810-3880 |
71 |
CTT |
16 |
||
trnD |
H |
3881-3954 |
74 |
GTC |
0 |
||
atp8 |
H |
3947-4111 |
165 |
ATC |
TAA |
-7 |
|
atp6 |
H |
4240-4917 |
678 |
ATG |
TAA |
28 |
|
cox3 |
H |
4940-5725 |
786 |
ATG |
TAA |
22 |
|
trnG |
H |
5726-5791 |
66 |
TCC |
0 |
||
nad3 |
H |
5864-6217 |
354 |
ATT |
TAA |
72 |
|
trnA |
H |
6245-6313 |
69 |
TGC |
27 |
||
trnR |
H |
6342-6408 |
68 |
TCG |
28 |
||
trnN |
H |
6409-6471 |
65 |
GTT |
0 |
||
trnS1 |
H |
6474-6540 |
67 |
GCT |
2 |
||
trnE |
H |
6532-6997 |
66 |
TTC |
-8 |
||
trnF |
L |
7058-7125 |
68 |
GAA |
60 |
||
nad5 |
L |
7141-8854 |
1714 |
ATT |
T(AA) |
15 |
|
trnH |
L |
8855-8920 |
66 |
GTG |
0 |
||
nad4 |
L |
8919-10259 |
1341 |
ATG |
TAA |
-1 |
|
nad4L |
L |
10262-10552 |
291 |
ATG |
TAA |
2 |
|
trnT |
H |
10553-10616 |
64 |
TGT |
0 |
||
trnP |
L |
10637-10702 |
66 |
TGG |
20 |
||
nad6 |
H |
10700-11206 |
507 |
ATT |
TAA |
-2 |
|
cob |
H |
11265-11401 |
1137 |
ATG |
TAA |
58 |
|
trnS2 |
H |
11419-11485 |
67 |
TGA |
17 |
||
nad1 |
L |
11496-12423 |
939 |
ATA |
TAA |
10 |
|
trnL1 |
L |
12424-12505 |
66 |
TAG |
16 |
||
16sRNA |
L |
12547-13855 |
1309 |
41 |
|||
trnV |
L |
13854-13925 |
72 |
TAC |
-1 |
||
12sRNA |
L |
13913-14711 |
799 |
-12 |
|||
CR |
14712-16854 |
2143 |
0 |
Note: Strand of the genes is presented as L for majority and H for minority strand. In the column for intergenic length, a positive sign indicates the interval in base pairs between genes, while the negative sign indicates overlapping base pairs between genes.
Table II. Composition and skewness of the S. atilius mitogenome.
S. atilius |
Size (bp) |
A (%) |
T (%) |
G (%) |
C (%) |
A+T (%) |
G+C (%) |
AT-skew |
GC-skew |
Genome |
16,854 |
41.80 |
39.92 |
7.67 |
10.61 |
81.72 |
18.28 |
0.02 |
-0.16 |
Protein-coding genes |
11,140 |
34.15 |
45.67 |
10.30 |
9.88 |
79.82 |
20.18 |
-0.14 |
0.02 |
tRNA genes |
1,490 |
41.41 |
39.06 |
10.94 |
8.59 |
79.47 |
19.53 |
0.03 |
0.12 |
rRNA genes |
2108 |
39.28 |
42.88 |
11.48 |
6.36 |
82.16 |
17.84 |
-0.04 |
0.28 |
Detected CR |
1674 |
45.65 |
42.60 |
1.61 |
7.46 |
88.25 |
9.07 |
0.03 |
0.64 |
CR, control region
There were five intergenic overlapping regions totaling 29 bp, with varying lengths of 1-8 bp, which ware mainly present in the tRNA genes. The two longest overlapping regions, both with a length of 8 bp, occured between nad2 and trnW and between trnS1 and trnE. Sixteen intergenic spacer regions were identified, totaling 555 bp in length, with the longest spacer sequence (72 bp) located between trnG and nad3, followed by a 60 bp spacer between trnE and trnF. There were also three regions without gene overlaps or intergenic spacers.
The mitogenome nucleotide composition was 41.80% for A, 39.92% for T, 7.67% for G, and 10.61% for C, respectively. The mitogenome was significantly biased to A+T (81.72%). The whole mitogenome of S. atilius exhibited a positive AT skew (0.02) and a negative GC skew (-0.16) (Table II).
Protein-coding genes
The total length of the 13 PGGs of in the S. atilius mitogenome was 11,140 bp, accounting for 66.10% of the whole mitogenome sequence, and encoding a total of 3,612 codons. Among them, nad5 (1,714 bp) was found to be the longest sequence, and nad4L (291 bp) was the shortest (Table I). Nine PCGs (nad2, cox1, cox2, atp8, atp6, cox3, nad3, nad6, and cob) were coded on the H-strand, while the remaining four PCGs (nad5, nad4, nad4L, and nad1) were located on the L-strand. The content of AT and GC was 79.82% and 20.18% in the 13 PCGs, exhibiting a highly AT bias (Table II). The AT skew was negative (-0.14), while the GC skew was positive (0.02) in the PCGs. Additionally, six PCGs (cox2, cox3, atp6, nad4, nad4L, and cob) initiated with an ATG start codons, four PCGs (nad2, nad3, nad5 and nad6) used ATT as the start codon, nad1 used ATA, atp8 used ATC, and cox1 used CGA. The termination codons of 12 PCGs were TAA. Only the nad5 used an incomplete stop codon T.
The amino acid usage of the 13 PCGs and the relative synonymous codon usage (RSCU) frequency are shown in Figure 2 and Table III. The most frequently used amino acids in mitochondrial PCGs were Ile, Phe, Leu, and Asn, accounting for 16.68%, 12.51%, 10.86%, and 8.13% of the total amino acids, respectively (Fig. 2A). The most frequently used codons were UUU (436), AUU (436), UUA (342), and AAU (277). The relatively scarce amino acids were Met (0.62%), Trp (0.67%), Arg1 (0.93%), and Asp (1.23%). The most frequent synonymous codons were UUA, AGA, and GUU, with UUA having the highest frequency of relative synonymous codons (RSCU = 4.86) (Fig. 2B).
.
The nucleotide diversity (Pi) of the PCGs among 46 species was calculated, ranging from 0.14 to 0.22 (Fig. 3). Among them, nad6 (Pi=0.22) showed the most diverse nucleotide variability among all PCGs, followed by nad2 (Pi=0.21), nad3 (Pi=0.19), and atp8 (Pi=0.18). The nad5 (Pi=0.14), nad1 (Pi=0.15), and cox3 (0.15) genes exhibited relatively low values of nucleotide variability. The ratio of Ka/Ks was calculated for each gene of the 13 PCGs (Fig. 3). The value of the cox3 gene (Ka/Ks=1.26) was higher than others. Meanwhile, the ratio of Ka/Ks of other 12 PCGs were all significantly less than 1, with the value of cox1 gene being the lowest (Ka/Ks=0.06).
Table III. Codon number and RSCU in S. atilius mitochondrial PCGs.
Codon |
Count |
RSCU |
Codon |
Count |
RSCU |
Codon |
Count |
RSCU |
Codon |
Count |
RSCU |
UUU(F) |
431 |
1.77 |
UCU(S) |
82 |
1.65 |
UAU(Y) |
183 |
1.8 |
UGU(C) |
36 |
1.18 |
UUC(F) |
55 |
0.23 |
UCC(S) |
27 |
0.54 |
UAC(Y) |
20 |
0.20 |
UGC(C) |
25 |
0.82 |
UUA(L) |
342 |
4.86 |
UCA(S) |
109 |
2.19 |
UAA(*) |
60 |
1.05 |
UGA(*) |
93 |
1.62 |
UUG(L) |
42 |
0.60 |
UCG(S) |
13 |
0.26 |
UAG(*) |
19 |
0.33 |
UGG(W) |
26 |
1.00 |
CUU(L) |
23 |
0.33 |
CCU(P) |
32 |
1.58 |
CAU(H) |
56 |
1.93 |
CGU(R) |
8 |
0.35 |
CUC(L) |
0 |
0 |
CCC(P) |
8 |
0.40 |
CAC(H) |
2 |
0.07 |
CGC(R) |
0 |
0 |
CUA(L) |
15 |
0.21 |
CCA(P) |
40 |
1.98 |
CAA(Q) |
47 |
1.96 |
CGA(R) |
27 |
1.17 |
CUG(L) |
0 |
0 |
CCG(P) |
1 |
0.05 |
CAG(Q) |
1 |
0.04 |
CGG(R) |
1 |
0.04 |
AUU(I) |
436 |
2.01 |
ACU(T) |
68 |
1.59 |
AAU(N) |
277 |
1.75 |
AGU(S) |
44 |
0.88 |
AUC(I) |
33 |
0.15 |
ACC(T) |
22 |
0.51 |
AAC(N) |
39 |
0.25 |
AGC(S) |
24 |
0.48 |
AUA(I) |
183 |
0.84 |
ACA(T) |
67 |
1.57 |
AAA(K) |
123 |
1.73 |
AGA(R) |
59 |
2.57 |
AUG(M) |
24 |
1.00 |
ACG(T) |
14 |
0.33 |
AAG(K) |
19 |
0.27 |
AGG(R) |
43 |
1.87 |
GUU(V) |
52 |
2.31 |
GCU(A) |
48 |
2.13 |
GAU(D) |
43 |
1.79 |
GGU(G) |
32 |
0.98 |
GUC(V) |
2 |
0.09 |
GCC(A) |
4 |
0.18 |
GAC(D) |
5 |
0.21 |
GGC(G) |
1 |
0.03 |
GUA(V) |
33 |
1.47 |
GCA(A) |
36 |
1.60 |
GAA(E) |
54 |
1.96 |
GGA(G) |
91 |
2.78 |
GUG(V) |
3 |
0.13 |
GCG(A) |
2 |
0.09 |
GAG(E) |
1 |
0.04 |
GGG(G) |
7 |
0.21 |
tRNAs and rRNAs
The 22 tRNA genes of the mitogenome of S. atilius had a total of 1,490 bp in length, 9.51% of the entire mitogenome, ranging from 65 bp (trnN) to 74 bp (trnD) (Table I). Among these, 14 genes (trn1, trnM, trnW, trnL2, trnK, trnD, trnG, trnA, trnR, trnN, trnS1, trnE, trnT, and trnS2) were located on the H-strand and the remaining eight genes (trnQ, trnC, trnY, trnF, trnH, trnP, trnL1, and trnV) were located on the L-strand. Through the analysis of the secondary structure of the tRNAs (Fig. 4), most tRNA genes could be folded into the typical cloverleaf secondary structure, while trnS1 and trnT lacked the dihydrouridine (DHU) and TΨC arms, respectively. In the secondary structures of tRNAs of S. atilius (Fig. 3), three or four base pairs in the DHU arms, and four or five base pairs in the TΨC arms. Except the classic base pairs (A-U and C-G), fourteen wobble base pairs (G-U) were detected in nine genes (trnA, trnC, trnF, trnG, trnH, trnP, trnQ, trnT, and trnV), which occured in the amino acid-accepting arms, anticodon arms, TψC arms or DHU arms. Of them, trnH had the highest rate (three pairs each). Besides, five pairs of U-U base mismatches in trnA, trnG, trnL1, trnR, trnV, and trnW and one mismatches base in trnN were found in the TψC arms. The AT and GC content were 79.47% and 19.53% in the 22 tRNA genes, respectively, with a positive AT skew (0.03) and GC skew (0.12) (Table II).
There were two rRNAs in the mitogenome of S. atilius: a 1,309 bp 16S rRNA (rrnL) and a 799 bp 12S rRNA (rrnS) (Table I). The 16S rRNA gene was located between trnL1 and trnV, while the 12S rRNA gene was located between trnV and the control region. Both 16S rRNA and 12S rRNA were embedded in the L-strand. The AT content was 82.16%, with a negative AT skew (-0.04) and positive GC skew (0.28) (Table II).
Control region
The mitogenome of S. atilius contained one large putative control region (CR), and the length of the CR was 2143 bp. It was located between the rrnS and trnI, trnE and trnF, nad6 and cob, and rrnS and trnI, respectively. The AT content of non-coding regions (88.25%) was obviously higher than other regions. Additionally, the positive AT skew (0.03) and GC skew (0.64) were detected in the control region (Table II).
Phylogenetic analysis
The same phylogenetic relationships of S. atilius and other Tephritidae species were produced by ML and BI methods. The result revealed that the thirteen genera of Tephritidae species followed the following monophyletic relationships: ((Bactrocera, Dacus, Zeugodacus), Felderimyia, Anastrepha), (Acrotaeniostola, (Neoceratitis, Ceratitis), Euleia, Rivellia), (Procecidochares, (Tephritis, Sphaenisscus))))). Of them, Bactrocera (12 exemplars) formed a separate clade at the top of phylogenetic tree, and formed the sister group of a clade including Dacus (4 exemplars), Zeugodacus (8 exemplars), Felderimyia and Anastrepha. Ceratitis (4 exemplars) formed a monophyletic group and was sister to a clade comprising Acrotaeniostola and Neoceratit. Euleia and Rivellia formed the separate clades, respectively. Sphaenisscus and Tephritis clustered together showing a high statistical support value (PP=1, BS=100), and formed a sister group to Procecidochares (1 exemplar). S. atilius and T. femoralis were closely related and formed a sister group to Procecidochares utilis.
DISCUSSION
The complete mitogenome of S. atilius was a circular, double-stranded DNA molecule with a total length of 16, 854 bp, which was similar to that of other Tephritoidea insects analyzed, ranging from 15,117 bp (Tephritis femoralis) to 16,739 bp (Anastrepha fraterculus) (Table IV). It has the typical organization and composition of an insect mitochondrion, including 13 PCGs, 22 tRNA genes, and 2 rRNA genes, which was consistent with the existing mitogenomes of Tephritoidea, such as Bactrocera carambolae (Drosopoulou et al., 2019), Zeugodacus cucurbitae (Zhou et al., 2020), Dacus haikouensis (Wang et al., 2022). The nucleotide composition of all regions showed a strong AT bias (Nguyen et al., 2020), as seen in other insects (Yan et al., 2022; Mu et al., 2022; Lv et al., 2021). The AT-skew of the entire mitogenome was positive (0.02), while the GC skew is negative (-0.16), indicating that the content of bases C was higher than that of G, and A was higher than T in the whole genome.
The mitogenome commonly exhibited compact arrangement, such as small intergenic spacers or overlapping genes (Ojala et al., 1981). In the present study, intergenic overlapping regions ranging from 1 to 8 bp, with a total length of 29 bp. Overlapping regions of similar size were common among Tephritoidea insects (Drosopoulou et al., 2017, 2019; Yong et al., 2016; Teixeira et al., 2019), while their positions varied across species. For example, the longest overlaps were found between atp8 and atp6 in B. carambolae (Drosopoulou et al., 2019), and between nad2 and trnW, and between trnS1 and trnE in S. atilius. Most of the gene overlaps occurred in tRNA genes due to the lower evolutionary constraints of these genes (Yuan et al., 2021).
The non-coding intergenic spacers, which were composed of less than 10 non-coding nucleotides in the mitochondria of most animals, which contributed to species identification and the evolution of insect mitochondrial genomes (Yan et al., 2022). Sixteen intergenic spacer regions were examined, with a total length of 555 bp, and the longest spacer sequence (72 bp) was located between trnG and nad3. Larger intergenic spacers in mitogenomes had been reported in Tephritoidea insects, such as 94 bp in Bactrocera latifrons, 82 bp in Bactrocera melastomatos and 79 bp in Bactrocera umbrosa (Yong et al., 2016). Generally, the duplication/random loss model and slipped-strand mispairing can be used to explain the origin of mitogenome intergenic spacers (Du et al., 2017; Cheng et al., 2016). Whether these long spacer regions being functional was controversial (Yan et al., 2022).
The 13 PCGs of the S. atilius mitogenome were found to be 11,140 bp in length and used a variety of start codons, including ATG for cox2, cox3, atp6, nad4, nad4L, and cob; ATT for nad2, nad3, nad5, and nad6; ATA for nad1; CGA for cox1; and ATC for atp8, as reported for Bactrocera biguttula (Teixeira et al., 2019) and Bactrocera carambolae PCGs (Drosopoulou et al., 2019). Alternative start codons had also been found in other insects, such as TTG in Arma custos and Picromerus lewisi (Mu et al., 2022) and GTG in Nisia fuliginosa (Lv et al., 2021). The cox1 gene in the S. atilius mitogenome used CGA as the starting codon, consistent with other known insects (Yang et al., 2019). However, the starting codon of cox1 was not always uniform, for example, TTG in Episymploce splendens (Yan et al., 2022) and ATA in Anastatus fulloi (Yi et al., 2022). The typical termination codons TAA was employed in 12 PCGs, which was common among metazoans (Yan et al., 2022; Yi et al., 2022), with one exception of an incomplete stop codon T for nad5. This exception was commonly observed in arthropod mitogenomes (Huang et al., 2022; Yi et al., 2022), and might be attributed to post-transcriptional modification during the mRNA maturation process (Boore, 1999; Lv et al., 2021).
The four most frequently used codons, UUU (Phe), AUU (Ile), UUA (Leu), and AAU (Asn) were observed in the S. atilius mitogenome, which was similar to other insect mitogenomes, such as those of Ephemeroptera (Li et al., 2021), Coleoptera (Zeng et al., 2021) and Hemiptera (Nguyen et al., 2020). Meanwhile, the RSCU analysis of the PCGs also indicated that A and U were the components that contributed to the high A+T bias of the full mitogenome. The nucleotide diversity (Pi) and the ratio of Ka/Ks of the PCGs among 43 Tephritoidea species
Table IV. List of taxa used for phylogenetic analysis.
Superfamily |
Family |
Genus |
Species |
GenBank accession |
Length (bp) |
Tephritoidea |
Tephritidae |
Acrotaeniostola |
Acrotaeniostola dissimilis |
MH900079 |
15,384 |
Anastrepha |
Anastrepha fraterculus |
KX926433 |
16,739 |
||
Bactrocera biguttula |
MK293875 |
15,829 |
|||
Bactrocera carambolae |
EF014414 |
15,915 |
|||
Bactrocera correcta |
JX456552 |
15,936 |
|||
Bactrocera dorsalis |
DQ845759 |
15,915 |
|||
Bactrocera limbifera |
MG566056 |
15,860 |
|||
Bactrocera melastomatos |
KT881557 |
15,945 |
|||
Bactrocera oleae |
GU108463 |
15,821 |
|||
Bactrocera ritsemai |
MF668132 |
15,927 |
|||
Bactrocera rubigina |
MN714223 |
15,285 |
|||
Bactrocera ruiliensis |
MN477221 |
15,870 |
|||
Bactrocera tuberculata |
MT196006 |
15,273 |
|||
Bactrocera zonata |
KP296150 |
15,935 |
|||
Ceratitis |
Ceratitis capitata |
AJ242872 |
15,980 |
||
Ceratitis fasciventris |
KY436396 |
16,017 |
|||
Ceratitis quilicii |
MT998948 |
16,035 |
|||
Ceratitis rosa |
MT997010 |
16,047 |
|||
Dacus |
Dacus bivittatus |
MG962404 |
15,833 |
||
Dacus conopsoides |
MH351199 |
15,852 |
|||
Dacus longicornis |
NC_032690 |
16,253 |
|||
Dacus trimacula |
MK940811 |
15,851 |
|||
Euleia |
Euleia heraclei |
MT410819 |
15,514 |
||
Felderimyia |
Felderimyia fuscipennis |
MT702879 |
16.536 |
||
Neoceratitis |
Neoceratitis asiatica |
MF434829 |
15,481 |
||
Procecidochares |
Procecidochares utilis |
KC355248 |
15,922 |
||
Sphaeniscus |
Sphaeniscus atilius |
OQ909100 |
16,285 |
||
Zeugodacus |
Zeugodacus caudatus |
KT625491 |
15,866 |
||
Zeugodacus cilifer |
MT702880 |
15,843 |
|||
Zeugodacus cucurbitae |
JN635562 |
15,825 |
|||
Zeugodacus depressus |
KY131831 |
15,832 |
|||
Zeugodacus diaphorus |
KT159730 |
15,890 |
|||
Zeugodacus proprediaphora |
MN688227 |
15,829 |
|||
Zeugodacus scutellatus |
KP722192 |
15,915 |
|||
Zeugodacus tau |
KP711431 |
15,687 |
|||
Platystomatidae |
Rivellia |
Rivellia syngenesiae |
MT410799 |
17,835 |
|
Lauxanioidea |
Lauxaniidae |
Pachycerina |
Pachycerina decemlineata |
NC_034923 |
16,286 |
Sciomyzoidea |
Sciomyzidae |
Sciomyza |
Sciomyza simplex |
MT410781 |
16,553 |
were calculated. The results showed that nad6 exhibited the most diverse nucleotide variability among all PCGs, while nad5, nad1, and cox3 exhibited a relatively low variation rate and were the most conserved genes. The overall ratios of Ka/Ks for most PCG genes were significantly less than 1, which suggests that these PCGs were under purifying selection. The cox1 gene had the lowest Ka/Ks ratio (Ka/Ks=0.06), indicating that this gene had a relatively slow evolutionary rate (Hurst, 2002). This phenomenon occurred in almost all animals (Xiao et al., 2019), and had been subjected to species identification and evolutionary analysis in various arthropod species in Tephritoidea, as well as like in other insects (Demari-Silva et al., 2015; Zhou et al., 2020).
As in other insects, most tRNA genes in S. atilius could be folded into the typical clover-leaf secondary structure, while trnS1 lacked the dihydrouridine (DHU) arm and trnT lacked the TΨC arm. This feature of trnS1 had been observed in many other insect mitogenomes (Huang et al., 2022; Lv et al., 2021; Yuan et al., 2021; Yi et al., 2022; Soumia et al., 2022). However, the feature of trnT was less found in other insects. Other tRNAs, trnA also lacked the TΨC arm as described in Chrysodeixis acuta (Soumia et al., 2022). So, we speculated that this phenomenon could cause by the loss of gene in the process of evolution. We have added reason in the discussion. In addition, 14 wobble base pairs (G-U) and 5 pairs of U-U base mismatches in the tRNA genes of the S. atilius mitogenome were observed. Previous reports had suggested that wobble and mismatched pairs, which commonly occurred in insect tRNAs, were usually corrected through the editing process and sustain the transport function (Varani and McClain, 2000; Lavrov et al., 2000). The length, location, and base composition of the two rRNA genes were similar to those of other Tephritoidea insects, such as Bactrocera arecae (Yong et al., 2015) and Bactrocera biguttula (Teixeira et al., 2019).
In addition, the S. atilius mitogenome contained one large non-coding regions of which, the one located between rrnS and trnI was supposed to act as the origin of genome replication and gene transcription (Wolstenholme, 1992; Boore, 1999). The control region of insect mitogenomes ranged in size from tens of to several thousands of base pairs (Zhang et al., 1995; Lewis et al., 1995; Inohira et al., 1997). The control region was a source of length variation in the mitogenomes (Zhang et al., 1995; Lewis et al., 1995; Inohira et al., 1997). In the S. atilius mitogenome, 2143 bp in length was sequenced, which was more than the longest control region of 1,141 bp in Tephritidae insects (Mu et al., 2022). The control region was located between rrnS and trnI genes and also had a higher AT content (88.25%), compared to other Tephritidae insects, such as Bactrocera arecae (86.0%) (Yong et al., 2015), Bactrocera melastomatos (89.0%) (Yong et al., 2016), and Ceratitis fasciventris (90.24%) (Drosopoulou et al., 2017).
Phylogenetic analyses were based on the nucleotide sequences of the 13 PCGs used ML and BI methods to construct phylogenetic trees from the mitogenomes of 38 species of Diptera to elaborate phylogenetic relationship. The same phylogenetic relationships of S. atilius and other Tephritidae species were produced by two methods. In the present study, the phylogenetic relationships within Tephritidae could be presented as follows: ((Bactrocera, Dacus, Zeugodacus), Felderimyia, Anastrepha), (Acrotaeniostola, (Neoceratitis, Ceratitis), Euleia, Rivellia), (Procecidochares, (Tephritis, Sphaenisscus))))), in concordance with the findings of previous phylogenetic studies (Drosopoulou et al., 2019; Teixeira et al., 2019; Jia et al., 2019; Yang et al., 2020). The results showed that four genera (Felderimyia, Zeugodacus, Dacus and Bactrocera) were closely related and formed a sister group. Dacus and Zeugodacus constituted a sister to Bactrocera. This pattern had been demonstrated in the previous studies (Krosch et al., 2012; Virgilio et al., 2015; Jiang et al., 2016; San et al., 2018), which further be verified by our results. Besides, S. atilius and T. femoralis were closely related, and both of them formed the sister group of P. utilis. In this study, the phylogenetic placement of S. atilius was firstly investigated. Although the phylogenetic analyses on S. atilius were still limited, this result contributed to a molecular basis for the classification and phylogeny of S. atilius within Tephritidae.
CONCLUSION
In conclusion, the mitogenome of S. atilius, was the first one from the genus Sphaenisscus. This mitogenome showed high conservation in terms of gene size, organization, AT bias and secondary structures of tRNAs. The phylogenetic placement S. atilius was investigated and clarified, revealing that S. atilius and T. femoralis were closely related, and both of them formed the sister group of P. utilis. These results provided a framework for further studies of the phylogenetics and evolution of S. atilius.
Declarations
Funding
This study was supported by Special funds for Henan Provinces Scientific and Technological Development Guided by the Central Government (Z20221341063), Natural Science Foundation of Henan Province (No. 212300410229), Key Project for University Excellent Young Talents of Henan Province (No. 2020GGJS260), the Project of Science and Technology Innovation Team (No. XNKJTD-007 and KJCXTD-202001). The funders had no role in study design, data collection and analysis, decision to publish, or preparation of the manuscript.
Statement of conflict of interest
The authors declare that there is no conflict of interest.
REFERENCES
Aluja, M. and Mangan, R.L., 2008. Fruit fly (Diptera: Tephritidae) host status determination: Critical conceptual, methodological, and regulatory considerations. Annu. Rev. Ent., 53: 473-502. https://doi.org/10.1146/annurev.ento.53.103106.093350
Aluja, M. and Norrbom, A.L., 2000. Fruit flies (Tephritidae): phylogeny and evolution of behavior. CRC Press, Boca Raton. https://doi.org/10.1201/9781420074468
Bernt, M., Donath, A., Juhling, F., Externbrink, F., Florentz, C., Fritzsch, G., Putz, J., Middendorf, M. and Stadler, P.F., 2013. MITOS: Improved de novo metazoan mitochondrial genome annotation. Mol. Phylogenet. Evol., 69: 313-319. https://doi.org/10.1016/j.ympev.2012.08.023
Boore, J.L., 1999. Animal mitochondrial genomes. Nucl. Acids Res., 27: 1767-1780. https://doi.org/10.1093/nar/27.8.1767
Cameron, S.L., 2014. Insect mitochondrial genomics: Implications for evolution and phylogeny. Annu. Rev. Ent., 59: 95-117. https://doi.org/10.1146/annurev-ento-011613-162007
Capella-Gutierrez, S., Silla-Martinez, J.M. and Gabaldon, T., 2009. TrimAl: A tool for automated alignment trimming in large-scale phylogenetic analyses. Bioinformatics, 25: 1972-1973. https://doi.org/10.1093/bioinformatics/btp348
Cheng, X.F., Zhang, L.P., Yu, D.N., Storey, K.B. and Zhang, J.Y., 2016. The complete mitochondrial genomes of four cockroaches (Insecta: Blattodea) and phylogenetic analyses within cockroaches. Gene, 586: 115-122. https://doi.org/10.1016/j.gene.2016.03.057
Demari-Silva, B., Foster, P.G., de Oliveira, T.M., Bergo, E.S., Sanabani, S.S., Pessôa, R. and Sallum, M.A.M., 2015. Mitochondrial genomes and comparative analyses of Culex camposi, Culex coronator, Culex usquatus and Culex usquatissimus (Diptera:Culicidae), members of the coronator group. BMC Genomics, 16: 831. https://doi.org/10.1186/s12864-015-1951-0
Dohm, P., Kovac, D., Freidberg, A., Rull, J. and Aluja, M., 2014. Basic biology and host use patterns of tephritid flies (Phytalmiinae: Acanthonevrini, Dacinae: Gastrozonini) breeding in bamboo (Poaceae: Bambusoidea). Ann. entomol. Soc. Am., 107: 184-203. https://doi.org/10.1603/AN13083
Drosopoulou, E., Pantelidou, C., Gariou-Papalexiou, A., Augustinos, A.A., Chartomatsidou, T., Kyritsis, G.A., Bourtzis, K., Mavragani-Tsipidou, P. and Zacharopoulou, A., 2017. The chromosomes and the mitogenome of Ceratitis fasciventris (Diptera: Tephritidae): Two genetic approaches towards the Ceratitis FAR species complex resolution. Sci. Rep., 7: 4877. https://doi.org/10.1038/s41598-017-05132-3
Drosopoulou, E., Syllas, A., Goutakoli, P., Zisiadis, G.A., Konstantinou, T., Pangea, D., Sentis, G., Sauers-Muller, A., Wee, S.L., Augustinos, A.A., Zacharopoulou, A. and Bourtzis, K., 2019. Tauhe complete mitochondrial genome of Bactrocera carambolae (Diptera: Tephritidae): Genome description and phylogenetic implications. Insects, 10: 429. https://doi.org/10.3390/insects10120429
Du, C., Zhang, L., Lu, T., Ma, J., Zeng, C.,Yue, B.S. and Zhang, X.Y., 2017. Mitochondrial genomes of blister beetles (Coleoptera, Meloidae) and two large intergenic spacers in Hycleus genera. BMC Genomics, 18: 698. https://doi.org/10.1186/s12864-017-4102-y
Du, Z., Wu, Y., Chen, Z., Cao, L., Ishikawa, T., Kamitani, S., Sota, T.J., Song, F., Tian, L., Cai, W.Z. and Li, H., 2021. Global phylogeography and invasion history of the spotted lanternfly revealed by mitochondrial phylogenomics. Evol. Appl., 14: 915-930. https://doi.org/10.1111/eva.13170
Fu, Y.T., Suleman, Yao, C.Q., Wang H.M., Wang W. and Liu, G.H., 2022. A novel mitochondrial genome fragmentation pattern in the buffalo louse Haematopinus tuberculatus (Psocodea: Haematopinidae). Int. J. mol. Sci., 23: 13092. https://doi.org/10.3390/ijms232113092
Gong, R.Y., Guo, X., Ma, J.N., Song, X.H., Shen, Y.M., Geng, F.N., Price, M.G., Zhang, X.Y. and Yue, B.S., 2018. Complete mitochondrial genome of Periplaneta brunnea (Blattodea: Blattidae) and phylogenetic analyses within Blattodea. J. Asia Pac. Ent., 21: 885-895. https://doi.org/10.1016/j.aspen.2018.05.006
Greiner, S., Lehwark P. and Bock, R., 2019. Organellar genome DRAW (OGDRAW) version 1.3.1: expanded toolkit for the graphical visualization of organellar genomes. Nucl. Acids Res., 47: W59-W64. https://doi.org/10.1093/nar/gkz238
Han, H.Y. and Kwon, Y.J., 2010. A list of North Korean tephritoid species (Diptera: Tephritoidea) deposited in the Hungarian natural history museum. Anim. Syst. Evol. Diver., 26: 251-260. https://doi.org/10.5635/KJSZ.2010.26.3.251
Hoang, D.T., Chernomor, O., von Haeseler, A., Minh, B.Q. and Vinh, L.S., 2018. UFBoot2: Improving the ultrafast bootstrap approximation. Mol. Biol. Evol., 35: 518-522. https://doi.org/10.1093/molbev/msx281
Huang, X., Chen, B., Wei, Z. and Shi, A., 2022. First report of complete mitochondrial genome in the tribes Coomaniellini and Dicercini (Coleoptera: Buprestidae) and phylogenetic implications. Genes (Basel), 13: 1074. https://doi.org/10.3390/genes13061074
Hurst, L.D., 2002. The Ka/Ks ratio: diagnosing the form of sequence evolution. Trends Genet., 18: 486. https://doi.org/10.1016/S0168-9525(02)02722-1
Inohira, K., Hara, T. and Matsuura, E.T., 1997. Nucleotide sequence divergence in the A+T-rich region of mitochondrial DNA in Drosophila simulans and Drosophila mauritiana. Mol. Biol. Evol., 14: 814-822. https://doi.org/10.1093/oxfordjournals.molbev.a025822
Jia, P.F., Liu, J.H. and Dan, W.L., 2019. Complete mitochondrial genome of the bamboo-shoot fruit fly, Acrotaeniostola dissimilis (Diptera: Tephritidae) and its phylogenetic relationship within family Tephritidae. Mitochondrial DNA B Resour., 5: 106-107. https://doi.org/10.1080/23802359.2019.1694455
Jiang, F., Pan, X.B., Li, X.K., Yu, Y.X., Zhang, J.H., Jiang, H.S., Dou, L.D. and Zhu, S.F., 2016. The first complete mitochondrial genome of Dacus longicornis (Diptera: Tephritidae) using next-generation sequencing and mitochondrial genome phylogeny of Dacini tribe. Sci. Rep., 6: 36426. https://doi.org/10.1038/srep36426
Kalyaanamoorthy, S., Minh, B.Q.,Wong, T.K.F., von Haeseler, A. and Jermiin, L.S., 2017. ModelFinder: fast model selection for accurate phylogenetic estimates. Nat. Methods, 14: 587-589. https://doi.org/10.1038/nmeth.4285
Katoh, K. and Standley, D.M., 2013. MAFFT multiple sequence alignment software version 7: improvements in performance and usability. Mol. Biol. Evol. 30: 772-780. https://doi.org/10.1093/molbev/mst010
Korneyev, V.A., 1999. Phylogenetic relationships among higher groups of Tephritidae. In: Fruit flies (Tephritidae) (ed. M.A.N. Aluja). Crc Press, Boca Raton. pp. 73-113. https://doi.org/10.1201/9781420074468.sec2
Krosch, M.N., Schutze, M.K., Armstrong, K.F., Graham, G.C., Yeates, D.K. and Clarke, A.R., 2012. A molecular phylogeny for the tribe Dacini (Diptera: Tephritidae): Systematic and biogeographic implications. Mol. Phylogenet. Evol., 64: 513-523. https://doi.org/10.1016/j.ympev.2012.05.006
Kumar, S., Stecher, G. and Tamura, K., 2016. MEGA7: Molecular evolutionary genetics analysis version 7.0 for bigger datasets. Mol. Biol. Evol., 33: 1870-1874. https://doi.org/10.1093/molbev/msw054
Lavrov, D.V., Brown, W.M. and Boore, J.L., 2000. A novel type of RNA editing occurs in the mitochondrial tRNAs of the centipede, Lithobius forficatus. Proc. natl. Acad. Sci. USA, 97: 13738-13742. https://doi.org/10.1073/pnas.250402997
Lewis, D.L., Farr, C.L. and Kaguni L.S., 1995. Drosophila melanogaster mitochondrial DNA: completion of the nucleotide sequence and evolutionary comparisons. Insect mol. Biol., 4: 263-278. https://doi.org/10.1111/j.1365-2583.1995.tb00032.x
Li, R., Ma, Z. and Zhou, C., 2021. The first two complete mitochondrial genomes of Neoephemeridae (Ephemeroptera): Comparative analysis and phylogenetic implication for Furcatergalia. Genes (Basel), 12: 1875. https://doi.org/10.3390/genes12121875
Librado, P. and Rozas, J., 2009. DnaSP v5: A software for comprehensive analysis of DNA polymorphism data. Bioinformatics, 25: 1451-1452. https://doi.org/10.1093/bioinformatics/btp187
Lv, S.S., Zhang, Y.J., Gong, N. and Chen, X.S., 2021. Characterization and phylogenetic analysis of the mitochondrial genome sequence of Nisia fuliginosa (Hemiptera: Fulgoroidea: Meenoplidae). J. Insect Sci., 21: 8. https://doi.org/10.1093/jisesa/ieab050
Mazzon, L., Whitmore, D., Cerretti, P. and Korneyev, V.A., 2021. New and confirmed records of fruit flies (Diptera, Tephritidae) from Italy. Biodivers. Data J., 9: e69351. https://doi.org/10.3897/BDJ.9.e69351
Minh, B.Q., Schmidt, H.A., Chernomor, O., Schrempf, D., Woodhams, M.D., von Haeseler, A. and Lanfear, R., 2020. IQ-TREE 2: New models and efficient methods for phylogenetic inference in the genomic era. Mol. Biol. Evol., 37: 1530-1534. https://doi.org/10.1093/molbev/msaa015
Mu, Y.L., Zhang, C.H., Zhang, Y.J., Yang, L. and Chen, X.S., 2022. Characterizing the complete mitochondrial genome of Arma custos and Picromerus lewisi (Hemiptera: Pentatomidae: Asopinae) and conducting phylogenetic analysis. J. Insect Sci., 22: 6. https://doi.org/10.1093/jisesa/ieab105
Nguyen, D.T., Wu, B., Xiao, S. and Hao, W., 2020. Evolution of a record-setting AT-rich genome: Indel mutation, recombination, and substitution bias. Genome Biol. Evol., 12: 2344-2354. https://doi.org/10.1093/gbe/evaa202
Ojala, D., Montoya, J. and Attardi, G., 1981. tRNA punctuation model of RNA processing in human mitochondria. Nature, 29: 470-474. https://doi.org/10.1038/290470a0
Pape, T., Bickel, D. and Meier, R., 2009. Diptera diversity: Status, challenges and tools. Brill, Leiden. https://doi.org/10.1163/ej.9789004148970.I-459
Patel, R.K. and Jain, M., 2012. NGS QC Toolkit: A toolkit for quality control of next generation sequencing data. PLoS One, 7: e30619. https://doi.org/10.1371/journal.pone.0030619
Peng, Y., Leung, H.C., Yiu, S.M. and Chin, F.Y., 2012. IDBA-UD: A de novo assembler for single-cell and metagenomic sequencing data with highly uneven depth. Bioinformatics, 28: 1420-1428. https://doi.org/10.1093/bioinformatics/bts174
Perna, N.T. and Kocher, T.D., 1995. Patterns of nucleotide composition at fourfold degenerate sites of animal mitochondrial genomes. J. mol. Evol., 41: 353-358. https://doi.org/10.1007/BF01215182
San Jose, M., Doorenweerd, C., Leblanc, L., Barr, N., Geib, S. and Rubinoff, D., 2018. Incongruence between molecules and morphology: A seven-gene phylogeny of Dacini fruit flies paves the way for reclassification (Diptera: Tephritidae). Mol. Phylogenet. Evol., 121: 139-149. https://doi.org/10.1016/j.ympev.2017.12.001
Siderhurst, M.S. and Jang, E.B., 2010. Cucumber volatile blend attractive to female melon fly, Bactrocera cucurbitae (Coquillett). J. chem. Ecol., 36: 699-708. https://doi.org/10.1007/s10886-010-9804-4
Song, S.N., Tang, P., Wei, S.J. and Chen, X.X., 2016. Comparative and phylogenetic analysis of the mitochondrial genomes in basal hymenopterans. Sci. Rep., 6: 20972. https://doi.org/10.1038/srep20972
Soumia, P.S, Shirsat, D.V., Krishna, R., Pandi G, G.P., Choudhary, J.S., Naaz, N., Karuppaiah, V., Gedam, P.A., Anandhan, S. and Singh, M., 2022. Unfolding the mitochondrial genome structure of green semilooper (Chrysodeixis acuta Walker): an emerging pest of onion (Allium cepa L.). PLoS One, 17: e0273635. https://doi.org/10.1371/journal.pone.0273635
Soumia, P.S., Shirsat, D.V., Krishna, R., G, G.P.P., Choudhary, J.S., Naaz, N.V.K., Gedam, P.A. and Singh, M., 2022. Unfolding the mitochondrial genome structure of green semilooper (Chrysodeixis acuta Walker): An emerging pest of onion (Allium cepa L.). PLoS One, 17: e0273635. https://doi.org/10.1371/journal.pone.0273635
Teixeira da Costa, L., Powell, C., van Noort, S., Costa, C., Sinno, M., Caleca, V., Rhode, C., Kennedy, R.J., van Staden, M. and van Asch, B., 2019. The complete mitochondrial genome of Bactrocera biguttula (Bezzi) (Diptera: Tephritidae) and phylogenetic relationships with other Dacini. Int. J. Biol. Macromol., 126: 130-140. https://doi.org/10.1016/j.ijbiomac.2018.12.186
Vaidya, G., Lohman, D.J. and Meier, R., 2011. Sequence Matrix: Concatenation software for the fast assembly of multi-gene datasets with character set and codon information. Cladistics, 27: 171-180. https://doi.org/10.1111/j.1096-0031.2010.00329.x
Varani, G. and McClain, W.H., 2000. The G x U wobble base pair. A fundamental building block of RNA structure crucial to RNA function in diverse biological systems. EMBO Rep., 1: 18-23. https://doi.org/10.1093/embo-reports/kvd001
Virgilio, M., Jordaens, K., Verwimp, C., Whit,e I.M. and De Meyer, M., 2015. Higher phylogeny of frugivorous flies (Diptera, Tephritidae, Dacini): Localised partition conflicts and a novel generic classification. Mol. Phylogenet. Evol., 85: 171-179. https://doi.org/10.1016/j.ympev.2015.01.007
Wang, C.Y., Zhao, M., Wang, S.J., Xu, H.L., Yang, Y.M., Liu, L.N. and Feng, Y., 2021. The complete mitochondrial genome of Lepidotrigona flavibasis (Hymenoptera: Meliponini) and high gene rearrangement in Lepidotrigona mitogenomes. J. Insect Sci., 21: 10. https://doi.org/10.1093/jisesa/ieab038
Wang, T., Cai, B., Wang, X.N. and Ren, Y.L., 2022. Mitochondrial genome of Dacus haikouensis Wang and Cheng 2002 using next-generation sequencing from China and its phylogenetic implication. Mitochond.DNA B Resour., 7: 317-319. https://doi.org/10.1080/23802359.2021.2017365
White, I.M. and Elson-Harris, M.M., 1992. Fruit flies of economic significance: Their identification and bionomics. CAB International in association with ACIAR, Wallingford, Oxon, UK. pp. 601.
Wilson, A.J. and Xu, J., 2012. Mitochondrial inheritance: diverse patterns and mechanisms with an emphasis on fungi. Mycology, 3: 158-166.
Wolstenholme, D.R., 1992. Animal mitochondrial DNA: Structure and evolution. Int. Rev. Cytol., 141: 173-216. https://doi.org/10.1016/S0074-7696(08)62066-5
Xiao, L.F., Zhang, S.D., Long, C.P., Guo, Q.Y., Xu, J.S., Dai, X.H. and Wang, J.G., 2019. Complete mitogenome of a leaf-mining buprestid beetle, Trachys auricollis, and its phylogenetic implications. Genes (Basel), 10: 992. https://doi.org/10.3390/genes10120992
Yan, L., Hou, Z.Z., Ma, J.N., Wang, H.M., Gao, J., Zeng, C.J., Chen, Q., Yu, B.S. and Zhang, X.Y., 2022. Complete mitochondrial genome of Episymploce splendens (Blattodea: Ectobiidae): A large intergenic spacer and lacking of two tRNA genes. PLoS One, 17: e0268064. https://doi.org/10.1371/journal.pone.0268064
Yang, M., Zhang, H., Song, L., Shi, Y. and Liu, X., 2019. The complete mitochondrial genome of Mahanta tanyae compared with other zygaenoid moths (Lepidoptera: Zygaenoidea). J. Asia-Pac. Entomol., 22: 513-521. https://doi.org/10.1016/j.aspen.2019.03.010
Yang, M.J., Liu, J.H., Wan, X.S., Zhang, Q.L., Fu, D.Y., Wang, X.B., Dan, W.L. and Zhou, X.H., 2020. Complete mitochondrial genome of the black-winged fly, Felderimyia fuscipennis (Diptera: Tephritidae) and its phylogenetic relationship within family Tephritidae. Mitochond. DNA B Resour., 5: 3638-3639. https://doi.org/10.1080/23802359.2020.1831984
Yi, J.Q., Wu, H., Liu, J.B., Li, J.H., Lu, Y.L., Zhang, Y.F., Cheng, Y.J., Guo, Y., Li, D.S. and An, Y.X., 2022. Novel gene rearrangement in the mitochondrial genome of Anastatus fulloi (Hymenoptera Chalcidoidea) and phylogenetic implications for Chalcidoidea. Sci. Rep., 12: 1351. https://doi.org/10.1038/s41598-022-05419-0
Yong, H.S., Song, S.L., Lim, P.E., Chan, K.G., Chow, W.L. and Eamsobhana, P., 2015. Complete mitochondrial genome of Bactrocera arecae (Insecta: Tephritidae) by next-generation sequencing and molecular phylogeny of Dacini tribe. Sci. Rep., 5: 15155. https://doi.org/10.1038/srep15155
Yong, H.S., Song, S.L., Lim, P.E., Eamsobhana, P. and Suana, I.W., 2016. Complete mitochondrial genome of three Bactrocera fruit flies of Subgenus Bactrocera (Diptera: Tephritidae) and their phylogenetic implications. PLoS One, 11: e0148201. https://doi.org/10.1371/journal.pone.0148201
Yuan, L.L., Ge, X.Y., Xie, G.L., Liu, H.Y. and Yang, Y.X., 2021. First complete mitochondrial genome of Melyridae (Coleoptera, Cleroidea): Genome description and phylogenetic implications. Insects, 12: 87. https://doi.org/10.3390/insects12020087
Zeng, L.Y., Pang, Y.T., Feng, S.Q., Wang, Y.N., Stejskal, V., Aulicky, R., Zhang, S.F. and Li, Z.H., 2021. Comparative mitochondrial genomics of five Dermestid beetles (Coleoptera: Dermestidae) and its implications for phylogeny. Genomics, 113: 927-934. https://doi.org/10.1016/j.ygeno.2020.10.026
Zhang, B., Nardi, F., Hull-Sanders, H., Wan, X. and Liu, Y., 2014. The complete nucleotide sequence of the mitochondrial genome of Bactrocera minax (Diptera: Tephritidae). PLoS One, 9: e100558. https://doi.org/10.1371/journal.pone.0100558
Zhang, D.X. and Hewitt, G.M., 1997. Insect mitochondrial control region: A review of its structure, evolution and usefulness in evolutionary studies. Biochem. Syst. Ecol., 25: 99-120. https://doi.org/10.1016/S0305-1978(96)00042-7
Zhang, D.X., Szymura, J.M. and Hewitt, G.M., 1995. Evolution and structural conservation of the control region of insect mitochondrial DNA. J. mol. Evol., 40: 382-391. https://doi.org/10.1007/BF00164024
Zhou, X.H., Liu, J.H., Zhang, Q.L., Wan, X.S., Fu, D.Y., Wang, X.B., Dan, W.L. and Yang, M.J., 2020. Complete mitochondrial genome of melon fly, Zeugodacus cucurbitae (Diptera: Tephritidae) from Kunming, Southwest China and the phylogeny within subfamily Dacinae. Mitochond. DNA B Resour., 5: 2828-2829. https://doi.org/10.1080/23802359.2020.1790318
To share on other social networks, click on any share button. What are these?