Next Article in Journal
Transcriptomic Profiling of Paulownia fortunei (Seem.) Hemsl. Roots in Response to Chromium and Copper Stress
Previous Article in Journal
Genetic Diversity and Metabolic Profile of Tibetan Medicinal Plant Saussurea obvallata
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

The First Mitochondrial Genome of Family Xylococcidae (Hemiptera, Coccomorpha) and Its Phylogenetic Implications

The Key Laboratory for Silviculture and Conservation of Ministry of Education, Beijing Forestry University, Beijing 100083, China
*
Author to whom correspondence should be addressed.
These authors contributed equally to this work.
Genes 2025, 16(5), 594; https://doi.org/10.3390/genes16050594
Submission received: 7 April 2025 / Revised: 13 May 2025 / Accepted: 14 May 2025 / Published: 17 May 2025
(This article belongs to the Section Animal Genetics and Genomics)

Abstract

:
Background: The mitochondrial genome has been used for understanding higher-level phylogenetic relationships within Coccomorpha. we sequenced and analyzed the first mitochondrial genome of Xylococcus castanopsis Wu & Huang, 2017 to elucidate its genomic features and phylogenetic position. Methods: The complete mitogenome was assembled using NOVOPlasty and annotated with MITOS. We analyzed genome organization, codon usage, and tRNA structures. Phylogenetic relationships were reconstructed using 13 protein-coding genes from 19 scale insect species with Bayesian Inference and Maximum Likelihood methods. Result: The mitochondrial genome is 16,363 bp in size and contains the typical 37 mitochondrial genes, with an A + T content of 89.2%. All protein-coding genes start with the ATN and end with TAA/TAG or a single T- residue. Sixteen tRNAs exhibit the typical cloverleaf structure, while the remaining six lack either the dihydrouridine (DHU) or TΨC (T) arm. Gene rearrangements occur only in individual tRNAs and transpositions between the gene clusters trnS2-ND1 and trnL1-rrnL-trnV-rrnS. Phylogenetic analysis consistently place Xylococcidae as a sister group to all scale insects except Matsucoccidae. Conclusions: This study provides the first complete mitogenome for Xylococcidae, revealing characteristic gene rearrangements. Phylogenetic reconstruction resolves the phylogenetic position of Xylococcidae as a distinct lineage sister to all scale insects except Matsucoccidae, providing critical evolutionary insights.

1. Introduction

Scale insects (Hemiptera: Coccomorpha) are small and live hidden, causing plant weakening and pathogen transmission through feeding on phloem sap, which in severe cases can even lead to plant death [1]. Currently, there are more than 8500 described species in 1238 genera, belonging to 35 extant and 21 extinct families [2]. They were divided into two informal groups: archaeococcoids and neococcoids. The archaeococcoids typically feature abdominal spiracles, at least in females. However, neococcoids lack the abdominal spiracles and multifaceted eyes [3,4].
The taxonomic status of the subfamily Xylococcinae has undergone multiple revisions. It was considered a part of the family Margarodidae sensu lato by various authors, such as Morrison [5], who recognized it as the tribe Xylococcini. Zahradnik elevated Xylococcinae to family rank [6], while Koteja restricted Xylococcidae to two genera [7]: Xylococcus and Xylococculus. Therefore, the relationships within Margarodidae sensu lato, particularly regarding the systematic position of Xylococcidae, remain unresolved. These conflicting classifications demonstrate the persistent difficulties in establishing a satisfactory system for Coccomorpha. For consistency, we follow Koteja’s [7,8] classification, which treats Xylococcidae as a distinct family.
The family Xylococcidae currently includes four genera: Baisococcus Koteja, 1989; Mesophthirus Gao, Shih, Rasnitsyn & Ren, 2019; Xylococculus Morrison, 1927; and Xylococcus Lôw, 1882. However, only seven extant species remain within Xylococcus (Xylococcus castanopsis Wu & Huang, 2017; Xylococcus filiferus Löw, 1882; Xylococcus japonicus Oguma, 1926; Xylococcus quercicola Danzig, 1980); Xylococculus (Xylococculus betulae (Pergande, 1898); Xylococculus macrocarpae (Coleman, 1908); and Xylococculus quercus (Ehrhorn, 1900), while the other two genera are extinct [2]. In China, only X. castanopsis has been recorded [9]. Remaining extant species are distributed in North America, Northeast Asia and Europe [2,10].
Mitochondria, essential organelles present in nearly all eukaryotic cells, are crucial for energy production [11]. Mitochondrial genomes of insects are now widely used in studies on species identification, population genetic structure, biogeography, and phylogenetics [12,13]. However, the current research on mitochondrial genomes of scale insects remains relatively sparse. Only a handful of cocciologists have published studies on their mitochondrial genomes, with the studies primarily focusing on their structure and characteristics and exploring phylogenetic relationships based on existing mitochondrial genomes [14,15,16,17]. Despite the high diversity of scale insects, only mitogenomes of 24 species have been recorded in NCBI.
Whether based on phylogenetic trees reconstructed from the morphological characteristics of adult males or those combining molecular data with morphological traits, the position of Xylococcidae remains unresolved, and its mitochondrial genome has yet to be documented. In this paper, the mitogenome of X. castanopsis was sequenced and analyzed in detail. We then performed a phylogeny analysis of Coccomorpha mitogenomes and tried to explore the systematic position of Xylococcidae. This contributes to a better understanding of the phylogenetic relationships within Coccomorpha.

2. Materials and Methods

2.1. Sampling, Genomic DNA Extraction and Sequencing

The sample for DNA extraction was collected from the bark of Castanea henryi (Skan) Rehd. et Wils. (Fagaceae) from Guangdong Province, China, and was preserved in 95% ethanol under −20 °C at the Beijing Forestry University, Beijing, China (BFUC). The total genomic DNA was extracted using the TIANamp Genomic DNA Kit, following the manufacturer’s protocol. Voucher specimens are deposited at Beijing Forestry University, Beijing, China (BFUC). The COX1 sequences were amplified using the primers C1-1554F and C1-2342R [18] and sequenced via the Sanger method on an ABI 3730xl sequencer to provide seed sequences for subsequent mitochondrial genome assembly. We sequenced the genome of X. castanopsis using next-generation sequencing on the Illuminia Hiseq 2500 platform at Berry Genomics, Beijing, China. The sequencing achieved a depth of 6× with 150 bp paired-end reads. A single-species Illumina TruSeq library was prepared from total genomic DNA, featuring an average insert size of 150 bp. The sequence data of phylogenetic analysis were obtained from the National Center for Biotechnology Information at https://www.ncbi.nlm.nih.gov (refer to Table 1).

2.2. Mitogenome Assembly and Annotation

Following the removal of adapters and low-quality reads, we obtained approximately 12 GB of clean data and assembled it for reference-guided assembly using NOVOPlasty v4.3.3 [19], with the mitochondrial genome of Matsucoccus matsumurae as the reference and the COX1 sequence of X. castanopsis serving as a bait. This process successfully yielded a mitogenome of 16,363 bp. Then, annotated using the MITOS v2.1.9 on the Galaxy platform, the accurate boundaries of protein-coding genes (PCGs) were confirmed by comparison with other scale insects using MEGA 7.0.26 [20,21]. Six tRNA genes (trnA, trnQ, trnL1, trnS2, trnW, and trnV) were not identified by MITOS v2.1.9; however, they were successfully detected using the manual method and ARWEN v1.2 [22]. The rrnS was detected by MITOS v2.1.9, and the rrnL was determined by the upstream and downstream of tRNAs and alignment with the other scale insects.

2.3. Bioinformatic Analysis

Base composition, codon usage of protein-coding genes (PCGs), and relative synonymous codon usage (RSCU) values were calculated using MEGA v7.0.26. Figures were then generated using the ggplot2 package in R v4.3.2 [23,24]. Base composition skew values were computed as follows: AT skew = (A − T)/(A + T) and GC skew = (G − C)/(G + C) [25].

2.4. Phylogenetic Analysis

The phylogenetic analysis included a total of 31 species, comprising 19 species from the Coccomorpha, 4 species from the Aphidomorpha, 3 species from the Aleyrodomorpha, 2 species from the Psyllomorpha, and 3 species of order Thysanoptera. Phylogenetic relationship was inferred using Maximum Likelihood (ML) and Bayesian Inference (BI). A dataset comprising protein-coding genes from 31 species was concatenated in Phylosuite v1.2.2 for phylogenetic analysis [26]. Thirteen sequences were aligned in batches with MAFFT v7.505 using the ‘G-INS-i (accurate)’ strategy and codon alignment mode [27]. Ambiguously aligned fragments of 13 alignments were removed in batches using Gblocks v0.91b [28]. Finally, the 13 protein-coding genes were concatenated in Phylosuite. Before constructing Maximum Likelihood tree and Bayesian Inference tree, the protein-coding dataset was partitioned schemes for evolutionary models were selected using PartitionFinder v2.1.1 [29]. For the Maximum Likelihood tree, the best partitioning scheme and evolutionary models for the 13 protein-coding genes were selected using the greedy algorithm combined with the AICc criterion. However, for the Bayesian tree, the greedy algorithm was applied in conjunction with the BIC criterion. Maximum likelihood phylogenies were inferred using IQ-TREE v2.2.0 under the edge-linked partition model for 5000 ultrafast bootstraps, as well as the Shimodaira–Hasegawa-like approximate likelihood ratio test [30,31,32]. Bayesian Inference phylogenies were inferred using MrBayes v3.2.6 under the partition model (2 parallel runs, 2,000,000 generations) [33], in which the initial 25% of sampled data were discarded as burn-in.

3. Results

3.1. General Features and Nucleotide Composition

The length of the mitogenome sequence of X. castanopsis was 16,363 bp, containing 13 PCGs, 22 tRNAs and 2 rRNAs, amongst which 9 PCGs and 15 tRNAs were encoded on the forward strand (+) and 13 genes were encoded by the reverse strand (-) (Figure 1 and Table 2). Ten regions of gene overlap between adjacent genes were detected in the mitogenome, ranging from −1 to −14 bp, and 20 intergenic regions, ranging from 1 to 94 bp. The overall base composition of the mitogenome sequence was 42.9% T, 46.3% A, 7.7% C and 3.1% G, with a strong bias towards A + T (89.2%) (Table 3).

3.2. Protein-Coding Genes

The 13 protein-coding genes (PCGs) span a total of 10,747 bp, comprising 65.7% of the complete mitogenome length of X. castanopsis, with a high A + T content of 88.8% (Table 3). The AT skew values for the 13 PCGs range from −0.20 in ND4 to 0.12 in ATP8, while the GC skew values vary from −0.78 in ATP8 to 0.73 in ND4L. The start codons for all protein-coding genes were ATN, while the stop codons varied: COX2 ended with T-, ND5 ended with TAG, and the remaining genes all terminated with TAA. The relative synonymous codon usage (RSCU) was depicted in Figure 2. The usage of codons for protein-coding genes is presented in descending order of frequency as follows: AAT, ATT, TTT, TAT and ATA (Table 4).

3.3. tRNA Genes and rRNA Genes

The tRNAs of X. castanopsis had a length ranging from 47 bp (trnA) to 73 bp (trnH). Among 22 tRNAs, 16 tRNAs (trnC, trnD, trnE, trnG, trnH, trnI, trnK, trnL1, trnL2, trnM, trnN, trnQ, trnR, trnS2, trnT, and trnW) exhibited a complete cloverleaf structure, while 3 tRNAs (trnA, trnS1, and trnY) lacked the dihydrouridine (DHU) arm, and 3 other tRNAs (trnF, trnP, and trnV) lacked the TψC (T) arm (Figure 3).
The two rRNAs were both on the reverse strand in the mitogenome of X. castanopsis. The rrnL gene was 1,340 bp with 90.1% A + T content. The rrnS gene was 668 bp with 88.6% A + T content.

3.4. Gene Rearrangements Analysis

The mitochondrial genome of X. castanopsis exhibits significant gene rearrangements when compared to the ancestral sequence and other archeaococcoids (Figure 4). The genes trnS2-ND1 and trnL1-rrnL-trnV-rrnS have undergone transposition (T), and the positions of trnI-trnQ and trnM have been transposed (T). Additionally, the gene trnV has undergone inversion (I), and the rearrangement involving trnY and trnW-trnC is an inverted transposition (iT).

3.5. Phylogenetic Analysis of Coccomorpha

The phylogenetic trees reconstructed using the Maximum Likelihood and Bayesian Inference methods, based on the dataset of 13 protein-coding genes, were identical and had high support (Figure 5). Four infraorders (Aleyrodomorpha, Psyllomorpha, Aphidomorpha and Coccomorpha) were recovered as monophyletic with strong support. Aleyrodomorpha was sister to other infraorders. The phylogenetic relationship of Aleyrodomorpha + (Psyllomorpha + (Aphidomorpha + Coccomorpha))) was found with high statistical support (bootstrap value = 100%, posterior possibility = 1).
Within Coccomorpha, the Matsucoccidae was a sister to other coccoid families analyzed here. Xylococcidae formed a sister group with all other scale insects excluding Matsucoccidae. The two species of Monophlebidae formed separate clades, strongly supporting their monophyly (bootstrap value = 98%, posterior possibility = 1). Putoidae was found to be the sister group to all neococcids. The Pseudococcidae was grouped with the remaining neococcids. Among the remaining neococcids, the phylogenetic relationship of (Eriococcidae + (Kerriidae + (Cerococcidae + (Aclerdidae + Coccidae)))) was found with high statistical support (bootstrap value = 100%, posterior possibility = 1).

4. Discussion

Rearrangements in mitochondrial genomes are common in scale insects [14,15,16,17,18]. The Matsucoccidae exhibits gene arrangements that are similar to the ancestral sequence, with only individual tRNA transpositions (T). However, due to morphological similarities between xylococcids and matsucoccids, they are often placed by cocciologists as tribe Xylococcini and Matsucoccini within Margarodidae sensu lato, or alternatively [5,34], genera such as Xylococcus and Matsucoccus are classified under Xylococcidae [6]. Our results show that, compared to the ancestral arrangement, Matsucoccidae exhibits tRNA transpositions, while Xylococcidae exhibits more rearrangements, such as tRNA transpositions (T), inversions (I), inverted transpositions (iT), and transpositions (T) between the gene clusters trnS2-ND1 and trnL1-rrnL-trnV-rrnS.
The Coccomorpha is divided into two informal groups: archaeococcids and neococcoids. Archaeococcids are characterized by the presence of abdominal spiracles, and adult males possess compound eyes or a row of unicorneal eyes encircling the head, while the neococcoids lack the abdominal spiracles and compound eyes (except for Kermesidae and some Coccidae) [4,35]. Within Coccomorpha, Matsucoccidae is considered to be more morphologically similar to Orthezia and Aphis (Aphidomorpha) than to Xylococcidae [36]. However, Koteja proposed a potential relationship between Matsucoccidae and Xylococcidae based on mouthpart morphology, while suggesting a less likely affinity with Ortheziidae [7]. Due to the lack of data on Ortheziidae, our results recognize Matsucoccidae as the sister with other scale insects, showing closer phylogenetic proximity to Xylococcidae, consistent with previous findings [34,37,38].
Although the Xylococcidae and Matsucoccidae were included in Margarodidae sensu lato, the morphology of adult males and females of Xylococcidae is distinct from that of Matsucoccidae. The pores are present in the atrium of the abdominal spiracles of Xylococcidae, while absent in those of Matsucoccidae. Additionally, bitubular ducts are absent in Xylococcidae but present in Matsucoccidae [39]. In males, Matsucoccidae possess only a cluster of cupolae at the wing base, whereas Xylococcidae exhibit both a row of cupolae along the subcostal complex and a cluster at the wing base [8]. Morphological analysis of adult males reveals close phylogenetic relationships between Matsucoccidae and Xylococcidae, evidenced by shared derived characteristics including (1) the presence of a dorsal waxy tuft, (2) a group of very small setae on the dorsal surface of antennal pedicel and coxal bases, and (3) the development of alar folds [40,41]. In our study, Xylococcidae was grouped with all other scale insects except for Matsucoccidae, with strong support (98% bootstrap support and one posterior probability). This mainly supports the phylogenetic result obtained by Hodgson and Hardy based on male morphological characteristics [38]. Monophlebidae has consistently shown strong monophyly, whether based on molecular or morphological data [37,42,43]. However, there is still no satisfactory classification system [44]. Our results suggest that Monophlebidae forms a clade. It forms a sister with all other scale insects.

Author Contributions

Conceptualization, H.X. and S.W.; methodology, Y.L. and H.X.; software, X.Z. and Y.L.; validation, H.X.; data curation, X.Z.; writing—original draft preparation, X.Z. and Y.L.; writing—review and editing, H.X. and S.W.; supervision, H.X. and S.W.; project administration, S.W.; funding acquisition, S.W. All authors have read and agreed to the published version of the manuscript.

Funding

This project was supported by the National Natural Science Foundation of China (No. 32270476).

Institutional Review Board Statement

Not applicable.

Informed Consent Statement

Not applicable.

Data Availability Statement

The data that support the findings of this study are openly available in the National Center for Biotechnology Information at https://www.ncbi.nlm.nih.gov/nuccore (accessed on 6 March 2025), reference number PV227316.

Acknowledgments

We thank the editor and anonymous reviewers for valuable comments and improvement of the manuscript.

Conflicts of Interest

The authors declare no conflicts of interest.

References

  1. Gullan, P.J.; Martin, J.H. Sternorrhyncha: (Jumping plant-lice, whiteflies, aphids, and scale insects). In Encyclopedia of Insects; Resh, V.H., Cardé, R.T., Eds.; Elsevier: San Diego, CA, USA, 2009; pp. 957–967. [Google Scholar]
  2. García Morales, M.; Denno, B.D.; Miller, D.R.; Miller, G.L.; Ben-Dov, Y.; Hardy, N.B. ScaleNet: A Literature-Based Model of Scale Insect Biology and Systematics. Database. Available online: http://scalenet.info (accessed on 3 September 2024).
  3. Borchsenius, N.S. On the evolution and phylogenetic interrelations of Coccoidea (Insecta, Homoptera). Zool. Zhurnal 1958, 37, 765–780, (In Russian with English Summary). [Google Scholar]
  4. Williams, D.J.; Gullan, P.J.; Miller, D.R.; Matile-Ferrero, D.; Han, S.I. A study of the scale insect genera Puto ignoret (Hemiptera: Sternorrhyncha: Coccoidea: Putoidae) and Ceroputo Sulc (Pseudococcidae) with a comparison to Phenacoccus Cockerell (Pseudococcidae). Zootaxa 2011, 2802, 1–22. [Google Scholar] [CrossRef]
  5. Morrison, H. A classification of the higher groups and genera of the coccid family Margarodidae. USDA Tech. Bull. 1928, 52, 62–67. [Google Scholar]
  6. Zahradnik, J. Cervci—Coccinea. Scale insects—Coccinea. Bestimmungstabellen Tschechoslow. Fauna 1959, 3, 527–552. [Google Scholar]
  7. Koteja, J. On the phylogeny and classification of the scale insects (Homoptera, Coccinea) (discussion based on the morphology of the mouthparts). Acta Zool. Cracoviensia 1974, 19, 267–325. [Google Scholar]
  8. Koteja, J. Scale insects (Homoptera: Coccinea) a day after. In Studies on Hemipteran Phylogeny; Thomas Say Publications in Entomology; Entomological Society of America Lanham: Lanham, MD, USA, 1996; 244p. [Google Scholar]
  9. Wu, S.A.; Huang, S.B.; Dong, Q.-G. First records of the family Xylococcidae (Hemiptera: Coccomorpha) in China, with description of a new species. Zootaxa 2017, 4312, 547–556. [Google Scholar] [CrossRef]
  10. Ben-Dov, Y. An updated checklist of the scale insects (Hemiptera: Coccoidea) of the Margarodidae sensu lato group. Zootaxa 2011, 2859, 1–62. [Google Scholar] [CrossRef]
  11. Nass, M.M.; Nass, S. Intramitochondrial fibers with DNA characteristics: I. Fixation and electron staining reactions. J. Cell Biol. 1963, 19, 593–611. [Google Scholar] [CrossRef]
  12. Beard, C.B.; Hamm, D.M.; Collins, F.H. The mitochondrial genome of the mosquito Anopheles gambiae: DNA sequence, genome organization, and comparisons with mitochondrial sequences of other insects. Insect Mol. Biol. 1993, 2, 103–124. [Google Scholar] [CrossRef]
  13. Manchekar, M.; Scissum-Gunn, K.; Song, D.; Khazi, F.; McLean, S.L.; Nielsen, B.L. DNA recombination activity in soybean mitochondria. J. Mol. Biol. 2006, 356, 288–299. [Google Scholar] [CrossRef]
  14. Hu, K.; Yu, S.; Zhang, N.; Tian, M.; Ban, Q.; Fan, Z.; Qiu, J. The first complete mitochondrial genome of Matsucoccidae (Hemiptera, Coccoidea) and implications for its phylogenetic position. Biodivers. Data J. 2022, 10, e94915. [Google Scholar] [CrossRef]
  15. Hou, Y.F.; Wei, J.F.; Zhao, T.Y.; Li, C.F.; Wang, F. First complete mitochondrial genome of the tribe Coccini (Hemiptera, Coccomorpha, Coccidae) and its phylogenetic implications. ZooKeys 2023, 1180, 333–354. [Google Scholar] [CrossRef] [PubMed]
  16. Lu, C.C.; Huang, X.; Deng, J. Mitochondrial genomes of soft scales (Hemiptera: Coccidae): Features, structures and significance. BMC Genom. 2023, 24, 37. [Google Scholar] [CrossRef]
  17. Xu, H.; Liu, X.; Wang, P.; Li, H.; Wu, S.A. Phylogenetic Implications of Mitogenomic Sequences and Gene Rearrangements of Scale Insects (Hemiptera and Coccoidea). Insects 2023, 14, 257. [Google Scholar] [CrossRef] [PubMed]
  18. Deng, J.; Lu, C.C.; Huang, X.L. The first mitochondrial genome of scale insects (Hemiptera: Coccoidea). Mitochondrial DNA B Resour. 2019, 4, 2094–2095. [Google Scholar] [CrossRef]
  19. Dierckxsens, N.; Mardulyn, P.; Smits, G. NOVOPlasty: De novo assembly of organelle genomes from whole genome data. Nucleic Acids Res. 2017, 45, e18. [Google Scholar]
  20. Donath, A.; Jühling, F.; Al-Arab, M.; Bernhart, S.H.; Reinhardt, F.; Stadler, P.F.; Middendorf, M.; Bernt, M. Improved Annotation of Protein-Coding Genes Boundaries in Metazoan Mitochondrial Genomes. Nucleic Acids Res. 2019, 47, 10543–10552. [Google Scholar] [CrossRef] [PubMed]
  21. 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]
  22. Laslett, D.; Canback, B. ARWEN: A program to detect tRNA genes in metazoan mitochondrial nucleotide sequences. Bioinformatics 2008, 24, 172–175. [Google Scholar] [CrossRef]
  23. Wickham, H. Ggplot2: Elegant Graphics for Data Analysis; Springer: New York, NY, USA, 2016. [Google Scholar]
  24. R Core Team. R: A Language and Environment for Statistical Computing; MSOR Connections; R Core Team: Vienna, Austria, 2014; p. 1. [Google Scholar]
  25. 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]
  26. Zhang, D.; Gao, F.; Jakovlić, 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] [PubMed]
  27. Katoh, K.; Standley, D.M. MAFFT multiple sequence alignment software version 7: Improvements in performance and usability. Mol. Biol. Evol. 2013, 30, 772–780. [Google Scholar] [CrossRef]
  28. 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] [PubMed]
  29. Lanfear, R.; Frandsen, P.B.; Wright, A.M.; Senfeld, T.; Calcott, B. PartitionFinder 2: New methods for selecting partitioned models of evolution for molecular and morphological phylogenetic analyses. Mol. Biol. Evol. 2017, 34, 772–773. [Google Scholar] [CrossRef]
  30. 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. 2015, 32, 268–274. [Google Scholar] [CrossRef] [PubMed]
  31. Minh, B.Q.; Nguyen, M.A.; von Haeseler, A. Ultrafast approximation for phylogenetic bootstrap. Mol. Biol. Evol. 2013, 30, 1188–1195. [Google Scholar] [CrossRef]
  32. Guindon, S.; Dufayard, J.F.; Lefort, V.; Anisimova, M.; Hordijk, W.; Gascuel, O. New algorithms and methods to estimate maximum-likelihood phylogenies: Assessing the performance of PhyML 3.0. Syst. Biol. 2010, 59, 307–321. [Google Scholar] [CrossRef]
  33. Ronquist, F.; Teslenko, M.; van der Mark, P.; Ayres, D.L.; Darling, A.; Höhna, S.; Larget, B.; Liu, L.; Suchard, M.A.; Huelsenbeck, J.P. MrBayes 3.2: Efficient Bayesian phylogenetic inference and model choice across a large model space. Syst. Biol. 2012, 61, 539–542. [Google Scholar] [CrossRef]
  34. Gavrilov-Zimin, I.A. Ontogenesis, morphology and higher classification of archaeocococcids (Homoptera: Coccinea: Orthezoiidea). Zoosyst. Ross. 2018, 2, 1–264. [Google Scholar]
  35. Hodgson, C.J. A review of neococcid scale insects (Hemiptera: Sternorrhyncha: Coccomorpha) based on the morphology of the adult males. Zootaxa 2020, 4765, 1–264. [Google Scholar] [CrossRef]
  36. Beardsley, J.W. External morphology of the adult male of Matsucoccus bisetosus. Ann. Entomol. Soc. Am. 1968, 61, 1449–1459. [Google Scholar] [CrossRef]
  37. Hodgson, C.J.; Hardy, N.B. The phylogeny of the superfamily Coccoidea (Hemiptera: Sternorrhyncha) based on the morphology of extant and extinct macropterous males. Syst. Entomol. 2013, 38, 794–804. [Google Scholar] [CrossRef]
  38. Vea, I.M.; Grimaldi, D.A. Diverse new scale insects (Hemiptera: Coccoidea) in amber from the Cretaceous and Eocene with a phylogenetic framework for fossil Coccoidea. Am. Mus. Novit. 2015, 3823, 1–15. [Google Scholar] [CrossRef]
  39. Kondo, T.; Watson, G.W. Encyclopedia of Scale Insect Pests. Encyclopedia of Scale Insect Pests; CABI: Wallingford, UK, 2022; 608p. [Google Scholar]
  40. Hodgson, C.J.; Foldi, I. A review of the Margarodidae sensu Morrison (Hemiptera: Coccoidea) and some related taxa based on the morphology of adult males. Zootaxa 2006, 1263, 1–250. [Google Scholar] [CrossRef]
  41. Koteja, J. Xylococcidae and related groups (Hemiptera: Coccinea) from Baltic amber. Pr. Muz. Ziemi 2008, 49, 19–56. [Google Scholar]
  42. Gullan, P.J.; Cook, L.G. Phylogeny and higher classification of the scale insects (Hemiptera: Sternorrhyncha: Coccoidea). Zootaxa 2007, 1668, 413–425. [Google Scholar] [CrossRef]
  43. Song, N.; Wang, M.; Tang, H.; Dang, Z. A phylogenetic analysis of scale insects (Hemiptera, Coccoidea) based on genomic and transcriptomic data. J. Syst. Evol. 2024, 63, 693–707. [Google Scholar] [CrossRef]
  44. Li, J.N.; Xu, H.; Wu, S.A. A new genus and species of giant mealybugs (Hemiptera: Coccomorpha: Monophlebidae) from eastern China. Zootaxa 2023, 5254, 434–442. [Google Scholar] [CrossRef]
Figure 1. Circular maps of the mitogenomes of X. castanopsis.
Figure 1. Circular maps of the mitogenomes of X. castanopsis.
Genes 16 00594 g001
Figure 2. The relative synonymous codon usage (RSCU) in the mitogenome of X. castanopsis.
Figure 2. The relative synonymous codon usage (RSCU) in the mitogenome of X. castanopsis.
Genes 16 00594 g002
Figure 3. Secondary structures of the transfer RNA genes (tRNAs) in X. castanopsis mitogenome.
Figure 3. Secondary structures of the transfer RNA genes (tRNAs) in X. castanopsis mitogenome.
Genes 16 00594 g003
Figure 4. Gene orders in ancestor and scale insects’ mitogenomes. The order was divided into four relatively conserved gene regions colored with yellow, blue, green and red.
Figure 4. Gene orders in ancestor and scale insects’ mitogenomes. The order was divided into four relatively conserved gene regions colored with yellow, blue, green and red.
Genes 16 00594 g004
Figure 5. Phylogenetic trees inferred from mitogenomes of Coccomorpha under ML and BI. The numbers at nodes indicate Maximum Likelihood bootstrap values (left) and Bayesian posterior probabilities (right).
Figure 5. Phylogenetic trees inferred from mitogenomes of Coccomorpha under ML and BI. The numbers at nodes indicate Maximum Likelihood bootstrap values (left) and Bayesian posterior probabilities (right).
Genes 16 00594 g005
Table 1. Species involved in the phylogenetic analysis.
Table 1. Species involved in the phylogenetic analysis.
OrderFamilySpeciesNCBI Accession Number
ThysanopteraAeolothripidaeAeolothrips albicinctusPP932666
ThripidaeThrips palmiMN127983
Megalurothrips usitatusON815612
HemipteraAleyrodidaeAleyrodes shizuokensisMT880225
Paraleyrodes mineiPP727237
Trialeurodes vaporariorumAY521265
PsyllidaePsyllopsis discrepansOR608273
Trioza erytreaeMG989238
AphididaeAphis citricidusMK540501
Aphis glycinesMK111111
Aphis gossypiiOR449278
Myzus persicaeKU236024
MatsucoccidaeMatsucoccus matsumuraePP103290
MonophlebidaeIcerya purchasiOX7316821
Coronaproctus castanopsisSRR26067561
PutoidaePuto sinensisPQ727056
XylococcidaeXylococcus castanopsisPV227316
PseudococcidaeParacoccus marginatusPP060471
Phenacoccus acerisSRR14087799
Phenacoccus manihotiMZ958983
EriococcidaeAcanthococcus coriaceusOP351525
Apiomorpha munitaOP351523
KerriidaeAlbotachardina sinensisOP351521
CerococcidaeAntecerococcus theydoniOP351522
AclerdidaeNipponaclerda biwakoensisOP351526
Aclerda takahashiiMW839575
CoccidaeDiensmococcus koreanusMW302211
Saissetia coffeaeMN863803
Parasaissetia nigraOK040656
Ceroplastes japonicusOP351524
Ceroplastes floridensisOK040657
Table 2. Mitochondrial genome characteristics of X. castanopsis.
Table 2. Mitochondrial genome characteristics of X. castanopsis.
GeneLocationSizeStrandStart CodonStop CodonAnticodonIntergenic Length
trnS21273–134472+--CCT-
ND11439–2356918-ATATAA-94
trnM2424–249370+--CAT67
trnI2500–256566+--GAT6
trnQ2564–261754---TTG−2
ND22617–3564948+ATTTAA-−1
trnY3565–362561+--GTA0
trnW3631–370361---TCA5
trnC3705–377369---GCA1
COX13775–53041530+ATGTAA-1
trnL25300–536566+--TAA−5
COX25369–6038670+ATAT--3
trnK6039–610466+--TTT0
trnD6114–617865+--GTC9
ATP86179–6328150+ATATAA-0
ATP66318–6947630+ATATAA-−11
COX36937–7683750+ATGTAA-−14
trnG7729–779264+--TCC45
ND37790–8143354+ATATAA-−3
trnA8157–820347+--AGC13
trnR8259–831254+--TCG55
trnN8312–837564+--GTT−1
trnS18377–843960+--TCT1
trnE8447–851266+--TTC7
trnF8512–856857---GAA−1
ND58574–10,2051632-ATTTAG-5
trnH10,227–10,29973---GTG21
ND410,303–11,6101308-ATGTAA-3
ND4L11,604–11,897294-ATTTAA-−7
trnT11,887–11,95569+--TGT−11
trnP11,959–12,01759---TGG3
ND612,021–12,485465+ATATAA-3
cytb12,493–13,5791086+ATATAA-7
trnL113,603–13,66967---TAA24
rrnL13,700–15,0091340----0
trnV15,010–15,07162+--AAC0
rrnS15,072–15,739668----0
“+/-” indicates whether the gene is located on the forward or reverse strand, respectively. The minus signs (-) appearing in the “Start codon”, “Stop codon”, and “Anticodon” columns simply denote blank entries or absence of relevant data for those particular fields.
Table 3. The nucleotide composition of the mitochondrial genome.
Table 3. The nucleotide composition of the mitochondrial genome.
Gene% T% C% A% G% A + TAT SkewGC Skew
ATP647.07.543.02.590.0−0.04−0.49
ATP841.35.352.70.794.00.12−0.78
COX143.910.538.67.182.5−0.06−0.19
COX242.48.145.24.387.60.03−0.31
COX349.36.541.23.090.4−0.09−0.37
cytb45.38.941.04.886.3−0.05−0.30
ND151.73.336.38.788.0−0.180.45
ND249.46.842.81.092.2−0.07−0.74
ND344.65.446.93.191.50.02−0.27
ND454.03.035.77.389.7−0.200.42
ND4L53.71.038.86.592.5−0.160.73
ND553.73.036.46.990.1−0.190.39
ND648.84.745.21.394.0−0.04−0.57
PCGs48.86.040.05.288.8−0.10−0.07
rrnL48.12.642.17.290.1−0.070.47
rrnS45.13.143.68.288.6−0.020.45
2 rRNA47.02.842.67.689.6−0.050.46
22 tRNA44.34.845.95.090.20.020.01
Total42.97.746.33.189.20.04−0.43
Table 4. The utilization of codons of the mitogenome of X. castanopsis.
Table 4. The utilization of codons of the mitogenome of X. castanopsis.
Amino AcidCodonCountRSCUAmino AcidCodonCountRSCU
PheUUU(F)4471.77SerUCU(S)361.33
UUC(F)580.23 UCC(S)130.48
LeuUUA(L)2354.62 UCA(S)702.59
UUG(L)250.49 UCG(S)80.3
Leu (c)CUU(L)210.41Ser (s)AGU(S)240.89
CUC(L)50.1 AGC(S)50.19
CUA(L)180.35 AGA(S)381.41
CUG(L)10.02 AGG(S)220.81
IleAUU(I)4781.83ThrACU(T)381.69
AUC(I)440.17 ACC(T)170.76
MetAUA(M)2431.81 ACA(T)271.2
AUG(M)250.19 ACG(T)80.36
ValGUU(V)183AlaGCU(A)72.8
GUC(V)10.17 GCC(A)20.8
GUA(V)50.83 GCA(A)10.4
GUG(V)00 GCG(A)00
TyrUAU(Y)3701.89CysUGU(C)141.47
UAC(Y)220.11 UGC(C)50.53
*UAA(*)1701.68TrpUGA(W)681.77
UAG(*)320.32 UGG(W)90.23
HisCAU(H)311.88ArgCGU(R)51.82
CAC(H)20.12 CGC(R)00
GlnCAA(Q)232 CGA(R)62.18
CAG(Q)00 CGG(R)00
AsnAAU(N)4831.89ProCCU(P)241.96
AAC(N)290.11 CCC(P)20.16
LysAAA(K)1771.71 CCA(P)231.88
AAG(K)300.29 CCG(P)00
AspGAU(D)382GlyGGU(G)80.64
GAC(D)00 GGC(G)00
GluGAA(E)271.86 GGA(G)383.04
GAG(E)20.14 GGG(G)40.32
The “*” denotes stop codon.
Disclaimer/Publisher’s Note: The statements, opinions and data contained in all publications are solely those of the individual author(s) and contributor(s) and not of MDPI and/or the editor(s). MDPI and/or the editor(s) disclaim responsibility for any injury to people or property resulting from any ideas, methods, instructions or products referred to in the content.

Share and Cite

MDPI and ACS Style

Zhao, X.; Li, Y.; Xu, H.; Wu, S. The First Mitochondrial Genome of Family Xylococcidae (Hemiptera, Coccomorpha) and Its Phylogenetic Implications. Genes 2025, 16, 594. https://doi.org/10.3390/genes16050594

AMA Style

Zhao X, Li Y, Xu H, Wu S. The First Mitochondrial Genome of Family Xylococcidae (Hemiptera, Coccomorpha) and Its Phylogenetic Implications. Genes. 2025; 16(5):594. https://doi.org/10.3390/genes16050594

Chicago/Turabian Style

Zhao, Xiaxia, Yuang Li, Han Xu, and Sanan Wu. 2025. "The First Mitochondrial Genome of Family Xylococcidae (Hemiptera, Coccomorpha) and Its Phylogenetic Implications" Genes 16, no. 5: 594. https://doi.org/10.3390/genes16050594

APA Style

Zhao, X., Li, Y., Xu, H., & Wu, S. (2025). The First Mitochondrial Genome of Family Xylococcidae (Hemiptera, Coccomorpha) and Its Phylogenetic Implications. Genes, 16(5), 594. https://doi.org/10.3390/genes16050594

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