Next Article in Journal
Increase in Cd Tolerance through Seed-Borne Endophytic Fungus Epichloë gansuensis Affected Root Exudates and Rhizosphere Bacterial Community of Achnatherum inebrians
Previous Article in Journal
PsnWRKY70 Negatively Regulates NaHCO3 Tolerance in Populus
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

A Novel Mitochondrial Genome Fragmentation Pattern in the Buffalo Louse Haematopinus tuberculatus (Psocodea: Haematopinidae)

1
Research Center for Parasites & Vectors, College of Veterinary Medicine, Hunan Agricultural University, Changsha 410128, China
2
Department of Zoology, University of Swabi, Swabi 23561, Pakistan
3
Department of Biomedical Sciences and One Health Center for Zoonoses and Tropical Veterinary Medicine, Ross University School of Veterinary Medicine, Basseterre P.O. Box 334, Saint Kitts and Nevis
4
The Centre for Bioinnovation, School of Science and Engineering, University of the Sunshine Coast, Sippy Downs, QLD 4556, Australia
*
Author to whom correspondence should be addressed.
Int. J. Mol. Sci. 2022, 23(21), 13092; https://0-doi-org.brum.beds.ac.uk/10.3390/ijms232113092
Submission received: 24 August 2022 / Revised: 8 October 2022 / Accepted: 25 October 2022 / Published: 28 October 2022
(This article belongs to the Section Molecular Genetics and Genomics)

Abstract

:
Sucking lice are obligate ectoparasites of mammalian hosts, causing serious public health problems and economic losses worldwide. It is well known that sucking lice have fragmented mitochondrial (mt) genomes, but many remain undetermined. To better understand patterns of mt genome fragmentation in the sucking lice, we sequenced the mt genome of the buffalo louse Haematopinus tuberculatus using next-generation sequencing (NGS). The mt genome of H. tuberculatus has ten circular minichromosomes containing a total of 37 genes. Each minichromosome is 2.9–5.0 kb long and carries one to eight genes plus one large non-coding region. The number of mt minichromosomes of H. tuberculatus (ten) is different from those of congeneric species (horse louse H. asini, domestic pig louse H. suis and wild pig louse H. apri) and other sucking lice. Two events (gene translocation and merger of mt minichromosome) are observed in Haematopinus. Compared to other studies, our phylogeny generated from mt genome datasets showed a different topology, suggesting that inclusion of data other than mt genomes would be required to resolve phylogeny of sucking lice. To our knowledge, this is the first report of a ten mt minichromosomes genome in sucking lice, which opens a new outlook into unexplored mt genome fragmentation patterns in sucking lice.

1. Introduction

The sucking lice (Psocodea: Anoplura) are obligate ectoparasites of eutherian mammals. There are approximately 540 known species in 15 families [1]. The single genus family Haematopinidae contains 21 described species, which are important ectoparasites of domestic animals that cause significant economic losses [2,3]. In addition, Haematopinus species are vectors of several pathogens, such as African swine fever virus [4], swinepox virus [5], classical swine fever virus [6] and Anaplasma spp. [7].
Metazoan mitochondrial (mt) genomes are usually circular DNA molecules of 13–20 kb with 36–37 genes including 12–13 protein-coding genes, two rRNA genes and 22 tRNA genes [8,9,10]. However, mt genomes of eutherian mammalian lice and some avian lice exhibit diverse fragmentation patterns. An example of an extremely fragmented mt genome is the human body louse Pediculus humanus humanus with 20 mt minichromosomes [11]. To date, the mt genomes of 21 sucking lice species (12 complete mt genomes and 9 incomplete mt genomes) have been sequenced, all are extensively fragmented with different numbers of minichromosomes [11,12,13,14,15,16,17,18,19,20,21,22]. Often, mt gene arrangement and composition are stable among members of a louse genus [12,17,19]; however, substantial variation in mt karyotype, the number of mt minichromosomes, gene arrangement and gene content has been also reported among congeneric lice, including sucking lice. In primate lice, the human louse P. humanus and chimpanzee louse P. schaeffi have 20 and 18 minichromosomes, respectively [12,17]. Macaque louse Pedicinus obtusus and colobus louse P. badii have 12 and 14 minichromosomes, respectively [19]. The goat louse Bovicola caprae of 13 minichromosomes is different from cattle louse B. bovis and sheep louse B. ovis (12 minichromosomes) [23]. Furthermore, variation in congeneric avian lice has also been found in pigeon lice, Columbicola columbae, C. macrourae, C. passerinae 1 and C. passerinae 2 have 15, 16, 17 and 17 minichromosomes, respectively [24]. Conversely while the horse louse H. asini, pig louse H. suis and wild pig louse H. apri all have nine mt minichromosomes [13,16], gene content and gene order of three minichromosomes in H. asini differ from those of H. suis and H. apri. Interestingly, the mt genomes of both H. suis and H. apri also have tRNA pseudogenes [13]. Based on these findings, we hypothesize that various mt genome fragmentation patterns exist in the genus Haematopinus. However, this hypothesis is built on only three Haematopinus species [13,16], thus, there is a need to obtain more mt genomes to test this hypothesis.
To further explore mt genome evolution in Haematopinus, we used next-generation sequencing (NGS) on H. tuberculatus. We found that the mt genome of H. tuberculatus is fragmented into ten circular minichromosomes. We analyzed mt genome fragmentation pattern and phylogeny, as well as variation in mt minichromosome composition and recombination with in the genus Haematopinus. Our results are invaluable in understanding the evolution of fragmented mt genomes in the sucking lice.

2. Results and Discussion

2.1. General Features of the mt Genome of the Buffalo Louse H. tuberculatus

Sequencing the H. tuberculatus genome produced 3.4 Gb of Illumina short-read sequence data, a total of 6,710,412 × 2 raw reads. After quality filtration, 3,841,215 × 2 clean reads were suitable for assembly of the mt genome. Assembling these sequence-reads into contigs, identified all 37 mt genes including 13 protein-coding genes, 22 tRNA genes and two rRNA genes, typical of bilateral animals. There are ten minichromosomes (Figure 1; Table 1); each minichromosome is 2.9–5.0 kb in size and consists of a coding region and one non-coding regions (NCR) (Table 1). The coding regions have 1–8 genes each and vary in size from 67 bp to 2627 bp (Table 1). All genes are in identical orientation relative to the transcription origin except trnT, nad1 and trnQ (Figure 1). The raw data (BioProject accession number: PRJNA883441) and nucleotide sequences (GenBank accession numbers: ON416547-56) of H. tuberculatus have been deposited in the NCBI database.
The NCR is composed largely of motifs conserved between different minichromosomes [24], and this region includes the D-loop which is involved in DNA replication and the initiation of transcription [25]. We assembled the full-length NCRs for all mt minichromosomes of H. tuberculatus, which ranged from 2280 bp (trnH-nad5-trnF-nad6 minichromosome) to 2901 bp (trnR-nad4L minichromosome) in size (Table 1). The longest NCR (2901 bp) in the buffalo louse H. tuberculatus is shorter than that in the horse louse H. anisi (3264 bp) [16], while it is longer than those of other sucking lice. As in most parvorder Anoplura, there is a GC-rich motif (70 bp, 55.7% C and G) downstream of the 3′-end of the coding region in each NCR. Remarkably, the AT-rich motif (54 bp, 94.4% A and T) is in the middle of the NCR, rather than upstream of the 5′-end of coding region, differing from other parvorder Anoplura [11,12,13].

2.2. Numbers of Minichromosomes among Parvorder Anoplura

All sucking lice sequenced to date have fragmented mt genome with variable numbers of, i.e., 9, 11, 12, 14, 18 or 20 minichromosomes. All mt genes have been identified in each of the 13 complete mt genomes, each circular minichromosome comprises one coding region and one NCR (Figure 2). An additional nine incomplete mt genomes of sucking lice are shown in Figure S1. Previous studies have suggested that fragmented mt minichromosomes are under strong selection to remain functional, and the related function may be affected along with an increased number of mt minichromosomes [26,27]. In the present study, we identified a novel pattern in the mt genome of H. tuberculatus with ten minichromosomes. Previous studies have indicated that the number of mt minichromosomes is evolutionarily unstable across Anoplura, even between congeneric species [12,17,19]. The substantial variation in the number of mt minichromosomes among Anoplura suggests that the process of mt genome fragmentation is a continuous process.

2.3. Variation in mt Minichromosomal Composition among Haematopinus Lice

H. suis, H. apri and H. asini each have nine mt minichromosomes [13,16]; however, H. tuberculatus has ten. The distribution of genes across the nine minichromosomes is identical between H. suis and H. apri [13]. Six minichromosomes in H. tuberculatus have the same gene content and gene order as their counterparts in H. asini, H. suis and H. apri, but the remaining differs [13,16]. In H. suis/apri, one minichromosome carries four genes, i.e., trnR-nad4L-nad6-trnM, which are found on three separated minichromosomes in H. tuberculatus (Figure 3) and H. asini [16]. In H. tuberculatus and H. asini, one minichromosome has four genes, trnH-nad5-trnF-nad6 (Figure 3), however, in H. suis and H. apri, the corresponding minichromosome has only three genes, trnH-nad5-trnF [13]. In H. tuberculatus, H. suis and H. apri, one minichromosome has two genes, rrnS-trnC (Figure 3). In contrast, this minichromosome in H. asini has four genes, trnR-nad4L-rrnS-trnC. In H. tuberculatus and H. asini, trnM occurs on its own minichromosome (Figure 3), however, in H. suis and H. apri, trnM is along with trnR-nad4L and nad6. These results clearly show the substantial variation in mt karyotype among Haematopinus species. Several previous studies compared mt genomes between the lice within the same genus, and showed substantial variation in mt karyotypes [14,19,20,21,22,23,24]. Taken together, these studies indicate that intra-genus variation in mt minichromosome composition is common in lice.

2.4. Recombination of mt Minichromosomes in the Haematopinus Lice

Recombination has been proposed as a possible mechanism contributing to the evolution of mt genome fragmentation across animal clades [27]. Long identical nucleotide sequences ranging from 14 to 133 bp are shared between mt genes, providing evidence for recombination between mt minichromosomes in sucking lice [11,12,13,14,15,16,17]. Similarly, seven stretches of identical nucleotide sequences, 7 to 32 bp long, were found between five pairs of mt genes in the buffalo louse (Table 2). trnL1 and trnL2 share three stretches of identical sequences of 7, 11 and 25 bp long. Meanwhile, in pig lice, the two genes share 9, 10 and 16 bp long identical sequences, whereas in horse louse these two genes share 15 bp long identical sequences with one another (Table 2). rrnL and rrnS share two stretches of identical sequences, 10 and 33 bp long, in H. asini; however, in H. suis, H. apri and H. tuberculatus, these two genes share only one stretch, 9 to 11 bp long of identical sequence, suggesting that recombination is occasional (Table 2). Previous studies found that recombination among tRNA genes could affect tRNA secondary structures [12,16]. Among Haematopinus, in addition to the pair of tRNA genes mentioned above, trnP and trnT share longer identical sequences than expected in H. suis (26 bp), H. apri (26 bp) and H. asini (27 bp), respectively. Nevertheless, in H. tuberculatus, the two genes only share 7 bp long identical sequences as other sucking lice do (Table 2). Among protein-coding genes, atp8 and atp6 in H. tuberculatus share a 32 bp identical sequence, three to four times more than expected by chance (Table 2), nad4 and cox1 share a 19 bp identical sequence in H. tuberculatus, and share 12 and 18 bp identical sequences in H. suis, but do not share longer-than expected identical sequences in H. asini and H. apri. Meanwhile, nad4 and cytb share 17 bp in H. tuberculatus; and 20 bp in H. asini, but no longer-than expected identical sequences are seen in H. suis and H. apri, nor in other sucking lice (Table 2). There is a 14 bp identical sequence shared by nad4L and trnV genes in H. tuberculatus, which is approximately twice as in other sucking lice. These results indicate that recombination is a likely cause of shared identical sequences between mt genes in Haematopinus lice.
Gene translocation between mt minichromosomes has been reported in the horse louse H. asini [16] and in the shrew louse P. reclinate [21], indicating that it is common in sucking lice. In the present study, translocations in Haematopinus lice can be also accounted for by two events of recombination. Firstly, in H. suis/apri, a nad6 moved from the trnR-nad4L-nad6-trnM minichromosome to the trnH-nad5-trnF to generate a new minichromosome trnH-nad5-trnF-nad6 in H. tuberculatus (Figure 1) and H. asini [16]. Second, in H. suis/apri, trnR-nad4L transferred from the minichromosome that contained trnR-nad4L-nad6-trnM, while in H. tuberculatus, trnR-nad4L moved to rrnS-trnC to generate a minichromosome, trnR-nad4L-rrnS-trnC in H. asini (Figure 1). These results suggest that recombination resulted in gene translocation between mt minichromosomes.
Previous studies have showed that merging and splitting occur between the minichromosomes of sucking lice [18,19,22]. Specifically, mergers but no split have been previously observed in Haematopinus spp. Sequenced [18]. Two mergers occurred in H. tuberculatus in the current study. First, the ancestral minichromosomes, trnK-nad4 and atp8-atp6-trnN, merged into one minichromosome, trnK-nad4-atp8-atp6-trnN (Figure 4). Second, nad2 and trnI-cox1-trnL2 merged into nad2-trnI-cox1-trnL2 (Figure 4). These data suggest that mt minichromosome merging is common in H. tuberculatus.

2.5. Phylogenetic Relationships

In the present study, the monophyly of Haematopinus (Haematopinidae), Polyplax (Polyplacidae) and Hoplopleura (Hoplopleuridae) was strongly supported by Bayesian inference (BI) analysis (Bpp = 0.9) and maximum likelihood (ML) analysis (Bv = 100) (Figure 5). The family Haematopinidae was sister to a clade of the families Polyplacidae + Hoplopleuridae to the exclusion of the families Pediculidae, Pthiridae and Pedicinidae with strong BI support (Bpp = 1.0) and moderate ML support (Bv = 49) (Figure 5). These results are consistent with those observed in the previous studies using nuclear genomic sequences [28,29]. In employing mt genomic datasets, however, several studies have indicated that the family Haematopinidae and families Pediculidae + Pthiridae + Pedicinidae were more closely related than to the families Polyplacidae and Hoplopleuridae with strong BI support (Bpp = 1.0), but weak support in ML analyses [19,21,30]. The mt genome is a valuable genetic marker for phylogenetic and evolutionary studies of different organisms because of its lacking of recombination, low mutation rate, and matrilineal inheritance [31,32,33]. However, recombination is found frequently in the fragmented mt genomes of Anoplura lice [27]. Recombination in mt genomes has substantial effects on phylogenetic and evolutionary studies that utilize mt genes [34,35,36]. The traditional methods for phylogenetic analysis are based on the assumption that mtDNA does not recombine; ignoring the occurrence of recombination can lead to incorrect phylogenetic reconstruction and positive selection analyses [37,38]. Collectively, our data along with others suggest that the deeper relationships among families within the parvorder Anoplura are challenge to resolve due to occurrence of recombination. Consequently, inclusion of data other than mt genomes would be greatly helpful in order to resolve phylogeny of sucking lice.

3. Materials and Methods

3.1. Sample Collection and DNA Extraction

Adult lice H. tuberculatus were collected from naturally infected buffalo Bubalus bubalis in Khyber Pakhtunkhwa province, Pakistan. They were identified to species morphologically [1], and stored in 100% (v/v) ethanol at −40 °C after five washes in physiological saline. Total genomic DNA was extracted from ten individual lice (five females and five males) using the DNeasy Tissue Kit (Promega, Madison, USA) according to the manufacturer’s protocol. The molecular identity of each sucking louse as H. tuberculatus was further verified by PCR-based sequencing of regions in the mt cox1 and rrnS genes as previously described [20]. The cox1 gene sequences of H. tuberculatus were 100% identical to that of H. tuberculatus (GenBank accession no: EU375757) from Bubalus bubalis in the United Kingdom.

3.2. Sequencing and Assembling

DNA concentration of each sample was determined using the Qubit system (Thermo Fisher Scientific, Waltham, MA, USA). Total DNA sequencing was performed by Novogene Bioinformatics Technology Co., Ltd. (Tianjing, China) using the Illumina HiSeq2500 platform (Illumina, San Diego, CA, USA) to produce 2 × 250 bp paired-end reads and raw data were recorded in FASTQ format. The raw reads were filtered to remove containing adaptor sequences and low-quality reads (the ‘N’ percent of one end > 5%, average quality score Q < 20 and length < 75 bp after trimming) using Trimmomatic v.0.32 [39]. The mt cox1 and rrnS sequences of H. tuberculatus were used as the initial references to de novo assembled the clean reads using Geneious Prime 2020 (www.geneious.com, accessed on 1 November 2021). The assembly parameters were: minimum overlap identity 99%, maximum 3% gaps per read, maximum gap 5 bp and minimum overlap 150 bp. A circular minichromosome was identified if both ends of a contig overlapped. Previous studies [18,19] showed that the NCR are highly conserved among the mt minichromosomes of a sucking louse. The conserved NCR sequences were identified between the mt cox1 and rrnS minichromosomes and were used as references to align the clean read sequence dataset. This allowed us to extract sequence reads derived from the two ends of the coding regions of all other mt minichromosomes. We then assembled all minichromosomes individually in full length using the same method stated above for mt cox1 and rrnS minichromosome assembly.

3.3. Verification of mt Minichromosomes

The size and circular organization of each mt minichromosome of H. tuberculatus were verified by long PCR using specific primers (Table S1), which were designed from the coding region of each minichromosome using the Primer Premier 5.0 (Premier Biosoft Interpairs, Palo Alto, CA, USA). The forward primer and reverse primer of each pair were next to each other with a small gap or no gap in between. PCR with these primers amplified each circular minichromosome in full or near full size if it had a circular organization (Figure S2). These positive amplicons were also sequenced with Illumina HiSeq2500 platform as described above. To obtain full-length and accurate sequences of the NCR of the all minichromosomes, we have re-assembled the NCR of each mt minichromosome using these obtained sequences according to the same method.

3.4. Annotation and Visualization

Genes were predicted with MITOS web server (http://mitos.bioinf.uni-leipzig.de/index.py, accessed on 5 November 2021) [40] and manually curated. Sequences of each protein-coding gene were then aligned against the corresponding gene of H. suis [13] and H. asini [16] using the MAFFT 7.263 software [41] to further identify gene boundaries. The location of protein-coding genes was further confirmed in ORFfinder (https://0-www-ncbi-nlm-nih-gov.brum.beds.ac.uk/orffinder/, accessed on 5 November 2021). Amino acid sequences of each protein-coding genes were inferred using MEGA 11 [42], and deduced amino acid sequences were used in BLAST searches of the protein database of GenBank. tRNA genes were identified using the program tRNAscan-SE [43] and ARWEN [44], and rRNA genes were identified with BLAST searches of the NCBI database and in comparison with alignments from H. suis [13] and H. asini [16]. The circular map of H. tuberculatus mt genome was illustrated using Microsoft PowerPoint v.2021.

3.5. Phylogenetic Analysis

Amino acid sequences inferred from the nucleotide sequences of 11 mt protein-coding genes common (nad2 and nad5 excluded because these genes are unidentified in H. kitti and H. elephantis) for all sucking lice (Table 3), using the elephant louse species, H. elephantis (GenBank: KF933032-41) as an outgroup [45]. The deduced amino acid sequences were aligned individually using MAFFT 7.122 and concatenated to form a single dataset; ambiguously aligned regions were excluded using Gblocks 0.91b using default parameters [46].
Phylogenetic analyses were conducted using two methods: BI and ML. BI was carried out using MrBayes 3.2.6 [47]. The most suitable model (MtArt) of evolution was selected by ProtTest 3.4 [48] at the default setting based on the Akaike information criterion (AIC). As MtArt model is a very recent addition to the models commonly used, we could not implement it in the current version of MrBayes, which used the best scoring alternative model MtREV. Four independent Markov chains (three heated and one cold) were run simultaneously for 1,000,000 metropolis coupled MCMC generations, sampling a tree every 100 generations. The first 2500 trees represented burn-in, and the remaining trees were tested for stability of likelihood values and used to compute Bayesian posterior probabilities (Bpp). We assumed that stationarity had been reached when the estimated sample size (ESS) was greater than 100, the potential scale reduction factor (PSRF) approached 1.0 and the average standard deviation of split frequencies (ASDSF) was < 0.01. ML was conducted with IQ-TREE v.2.1.3 [49]. The “Auto” option was set under the best evolutionary models, and the ML trees were constructed using an ultrafast bootstrap approximation approach with 10,000 replicates. The Bootstrap value (Bv) was calculated using 100 bootstrap replicates. Phylogenetic trees were drawn using FigTree v.1.42.

4. Conclusions

The newly-described mt genome of H. tuberculatus presented here has a novel mt genome fragmentation pattern, differing from other three Haematopinus lice, proved our hypothesis. Our findings indicate that recombination plays a major role in generating the variation in the composition of mt minichromosomes among Haematopinus lice. Compared to other studies, our phylogeny generated from mt genome datasets showed a different topology. Therefore, inclusion of data other than mt genomes would be required to resolve phylogeny of sucking lice. This is the first report of a mt genome with ten mt minichromosomes in sucking lice, which opened new outlook into unexplored fragmentation pattern in their mt genomes. Our results would encourage further investigation on mt genome fragmentation pattern in parasitic lice and other insects.

Supplementary Materials

The supporting information can be downloaded at: https://0-www-mdpi-com.brum.beds.ac.uk/article/10.3390/ijms232113092/s1.

Author Contributions

G.-H.L., Y.-T.F. and C.Y.: study conception and design, manuscript revision. Y.-T.F.: experiments. Y.-T.F. and G.-H.L.: data analysis. S., H.-M.W. and W.W.: study implementation and manuscript preparation. Suleman: resources. All authors have read and agreed to the published version of the manuscript.

Funding

This study was supported in part, by the National Natural Science Foundation of China (Grant No. 32172884) and the Training Programme for Excellent Young Innovators of Changsha (Grant No. KQ2106044).

Institutional Review Board Statement

Not applicable.

Informed Consent Statement

Not applicable.

Data Availability Statement

The fragmented mitochondrial genome sequences of H. tuberculatus from buffalo have been deposited in the GenBank database under the accession numbers ON416547–56.

Conflicts of Interest

Authors declare no conflict of interest for this article.

References

  1. Kim, K.C.; Ludwig, H.W. The family classification of the Anoplura. Syst. Entomol. 1978, 3, 249–284. [Google Scholar] [CrossRef]
  2. Durden, L.A.; Musser, G.G. The sucking lice (Insecta, Anoplura) of the world: A taxonomic checklist with records of mammalian hosts and geographical distributions. Bull. Amer. Mus. Nat. Hist. 1994, 218, 1–90. [Google Scholar] [CrossRef]
  3. Scofield, A.; Campos, K.F.; Silva, A.M.M.; Oliveira, C.H.B.; Barbosa, J.D.; Góes-Cavalcante, G. Infestation by Haematopinus quadripertusus on cattle in São Domingos do Capim, state of Pará, Brazil. Rev. Bras. Parasitol. Vet. 2012, 21, 315–318. [Google Scholar] [CrossRef] [Green Version]
  4. Saegerman, C.; Bonnet, S.; Bouhsira, E.; De Regge, N.; Fite, J.; Etoré, F.; Garigliany, M.M.; Jori, F.; Lempereur, L.; Le Potier, M.F.; et al. An expert opinion assessment of blood-feeding arthropods based on their capacity to transmit African swine fever virus in Metropolitan France. Transbound. Emerg. Dis. 2021, 68, 1190–1204. [Google Scholar] [CrossRef] [PubMed]
  5. Thibault, S.; Drolet, R.; Alain, R.; Dea, S. A sporadic skin disorder in nursing piglets. Swine Health Prod. 1998, 6, 276–278. [Google Scholar]
  6. Wall, R.; Shearer, D. Lice (Phthiraptera). In Veterinary Ectoparasites: Biology, Pathology and Control, 2nd ed.; Blackwell Science Ltd.: Oxford, UK, 2008; pp. 162–178. [Google Scholar]
  7. Da-Silva, A.; Lopes, L.; Diaz, J.; Tonin, A.; Stefani, L.; Araújo, D. Lice outbreak in buffaloes: Evidence of Anaplasma marginale transmission by sucking lice Haematopinus tuberculatus. J. Parasitol. 2013, 99, 546–547. [Google Scholar] [CrossRef]
  8. Wolstenholme, D.R. Genetic novelties in mitochondrial genomes of multicellular animals. Curr. Opin. Genet. Dev. 1992, 2, 919–925. [Google Scholar] [CrossRef]
  9. Boore, J.L. Animal mitochondrial genomes. Nucleic Acids Res. 1999, 27, 1767–1870. [Google Scholar] [CrossRef] [Green Version]
  10. Lavrov, D.V. Key transitions in animal evolution: A mitochondrial DNA perspective. Integr. Comp. Biol. 2007, 47, 734–743. [Google Scholar] [CrossRef] [Green Version]
  11. Shao, R.; Kirkness, E.F.; Barker, S.C. The single mitochondrial chromosome typical of animals has evolved into 18 minichromosomes in the human body louse, Pediculus humanus. Genome Res. 2009, 19, 904–912. [Google Scholar] [CrossRef] [Green Version]
  12. Shao, R.; Zhu, X.Q.; Barker, S.C.; Herd, K. Evolution of extensively fragmented mitochondrial genomes in the lice of humans. Genome Biol. Evol. 2012, 4, 1088–1101. [Google Scholar] [CrossRef] [PubMed]
  13. Jiang, H.; Barker, S.C.; Shao, R. Substantial variation in the extent of mitochondrial genome fragmentation among blood-sucking lice of mammals. Genome Biol. Evol. 2013, 5, 1298–1308. [Google Scholar] [CrossRef]
  14. Dong, W.G.; Song, S.; Guo, X.G.; Jin, D.C.; Yang, Q.Q.; Barker, S.C.; Shao, R. Fragmented mitochondrial genomes are present in both major clades of the blood-sucking lice (suborder Anoplura): Evidence from two Hoplopleura rodent lice (family Hoplopleuridae). BMC Genom. 2014, 15, 751. [Google Scholar] [CrossRef]
  15. Dong, W.G.; Song, S.; Jin, D.C.; Guo, X.G.; Shao, R. Fragmented mitochondrial genomes of the rat lice, Polyplax asiatica and Polyplax spinulosa: Intra-genus variation in fragmentation pattern and a possible link between the extent of fragmentation and the length of life cycle. BMC Genom. 2014, 15, 44. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  16. Song, S.D.; Barker, S.C.; Shao, R. Variation in mitochondrial minichromosome composition between blood-sucking lice of the genus Haematopinus that infest horses and pigs. Parasit. Vectors 2014, 7, 144. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  17. Herd, K.E.; Barker, S.C.; Shao, R. The mitochondrial genome of the chimpanzee louse, Pediculus schaeffi: Insights into the process of mitochondrial genome fragmentation in the blood-sucking lice of great apes. BMC. Genom. 2015, 16, 661. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  18. Shao, R.; Li, H.; Barker, S.C.; Song, S. The mitochondrial genome of the guanaco louse, Microthoracius praelongiceps: Insights into the ancestral mitochondrial karyotype of sucking lice (Anoplura, Insecta). Genome. Biol. Evol. 2017, 9, 431–445. [Google Scholar] [CrossRef] [Green Version]
  19. Fu, Y.T.; Dong, Y.; Wang, W.; Nie, Y.; Liu, G.H.; Shao, R. Fragmented mitochondrial genomes evolved in opposite directions between closely related macaque louse Pedicinus obtusus and colobus louse Pedicinus badii. Genomics 2020, 112, 4924–4933. [Google Scholar] [CrossRef]
  20. Fu, Y.T.; Nie, Y.; Duan, D.Y.; Liu, G.H. Variation of mitochondrial minichromosome composition in Hoplopleura lice (Phthiraptera: Hoplopleuridae) from rats. Parasit. Vectors 2020, 13, 506. [Google Scholar] [CrossRef]
  21. Dong, W.G.; Dong, Y.; Guo, X.G.; Shao, R. Frequent tRNA gene translocation towards the boundaries with control regions contributes to the highly dynamic mitochondrial genome organization of the parasitic lice of mammals. BMC. Genom. 2021, 22, 598. [Google Scholar] [CrossRef]
  22. Dong, Y.; Zhao, M.; Shao, R. Fragmented mitochondrial genomes of seal lice (family Echinophthiriidae) and gorilla louse (family Pthiridae): Frequent minichromosomal splits and a host switch of lice between seals. BMC Genom. 2022, 23, 283. [Google Scholar] [CrossRef]
  23. Song, F.; Li, H.; Liu, G.H.; 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]
  24. Sweet, A.D.; Johnson, K.P.; Cameron, S.L. Mitochondrial genomes of Columbicola feather lice are highly fragmented, indicating repeated evolution of minicircle-type genomes in parasitic lice. PeerJ. 2020, 8, e8759. [Google Scholar] [CrossRef] [Green Version]
  25. Clayton, D.A. Replication of animal mitochondrial DNA. Cell 1982, 28, 693–705. [Google Scholar] [CrossRef]
  26. Rand, D.M. The units of selection of mitochondrial DNA. Annu. Rev. Ecol. Evol. Syst. 2001, 32, 415–448. [Google Scholar] [CrossRef]
  27. Feng, S.; Pozzi, A.; Stejskal, V.; Opit, G.; Yang, Q.; Shao, R.; Dowling, D.K.; Li, Z. Fragmentation in mitochondrial genomes in relation to elevated sequence divergence and extreme rearrangements. BMC. Biol. 2022, 20, 7. [Google Scholar] [CrossRef] [PubMed]
  28. Johnson, K.P.; Nguyen, N.P.; Sweet, A.D.; Boyd, B.M.; Warnow, T.; Allen, J.M. Simultaneous radiation of bird and mammal lice following the K-Pg boundary. Biol. Lett. 2018, 14, 20180141. [Google Scholar] [CrossRef] [Green Version]
  29. de Moya, R.S.; Yoshizawa, K.; Walden, K.K.O.; Sweet, A.D.; Dietrich, C.H.; Kevin, P.J. Phylogenomics of parasitic and nonparasitic lice (Insecta: Psocodea): Combining sequence data and exploring compositional bias solutions in next generation data sets. Syst. Biol. 2021, 70, 719–738. [Google Scholar] [CrossRef]
  30. Deng, Y.P.; Yi, J.N.; Fu, Y.T.; Nie, Y.; Zhang, Y.; Liu, G.H. Comparative analyses of the mitochondrial genomes of the cattle tick Rhipicephalus microplus clades A and B from China. Parasitol. Res. 2022, 121, 1789–1797. [Google Scholar] [CrossRef]
  31. Fu, Y.T.; Zhang, Y.; Xun, Y.; Liu, G.H.; Suleman; Zhao, Y. Characterization of the complete mitochondrial genomes of six horseflies (Diptera: Tabanidae). Infect. Genet. Evol. 2021, 95, 105054. [Google Scholar] [CrossRef]
  32. Deng, Y.P.; Zhang, X.L.; Li, L.Y.; Yang, T.; Liu, G.H.; Fu, Y.T. Characterization of the complete mitochondrial genome of the swine kidney worm Stephanurus dentatus (Nematoda: Syngamidae) and phylogenetic implications. Vet. Parasitol. 2021, 295, 109475. [Google Scholar] [CrossRef] [PubMed]
  33. Light, J.E.; Smith, V.S.; Allen, J.M.; Durden, L.A.; Reed, D.L. Evolutionary history of mammalian sucking lice (Phthiraptera: Anoplura). BMC Evol. Biol. 2010, 10, 292. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  34. Zbawicka, M.; Wenne, R.; Burzyński, A. Mitogenomics of recombinant mitochondrial genomes of Baltic Sea Mytilus mussels. Mol. Genet. Genom. 2014, 289, 1275–1287. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  35. Schierup, M.H.; Hein, J. Consequences of recombination on traditional phylogenetic analysis. Genetics 2000, 156, 879–891. [Google Scholar] [CrossRef]
  36. Zheng, C.; Nie, L.; Wang, J.; Zhou, H.; Hou, H.; Wang, H.; Liu, J. Recombination and evolution of duplicate control regions in the mitochondrial genome of the Asian big-headed turtle, Platysternon megacephalum. PLoS ONE 2013, 8, e82854. [Google Scholar] [CrossRef] [Green Version]
  37. Rokas, A.; Ladoukakis, E.; Zouros, E. Animal mitochondrial DNA recombination revisited. Trends Ecol. Evol. 2003, 18, 411–417. [Google Scholar] [CrossRef]
  38. Slate, J.; Gemmell, N.J. Eve ‘n’ Steve: Recombination of human mitochondrial DNA. Trends Ecol. Evol. 2004, 19, 561–563. [Google Scholar] [CrossRef]
  39. Bolger, A.M.; Lohse, M.; Usadel, B. Trimmomatic: A flexible trimmer for Illumina sequence data. Bioinformatics 2014, 30, 2114–2120. [Google Scholar] [CrossRef] [Green Version]
  40. Bernt, M.; Donath, A.; Jühling, F.; Externbrink, F.; Florentz, C.; Fritzsch, G.; Pütz, J.; Middendorf, M.; Stadler, P.F. MITOS: Improved de novo metazoan mitochondrial genome annotation. Mol. Phylogenet. Evol. 2013, 69, 313–319. [Google Scholar] [CrossRef]
  41. Katoh, K.; Standley, D.M. A simple method to control over-alignment in the MAFFT multiple sequence alignment program. Bioinformatics 2016, 32, 1933–1942. [Google Scholar] [CrossRef] [Green Version]
  42. Tamura, K.; Stecher, G.; Kumar, S. MEGA11: Molecular evolutionary genetics analysis version 11. Mol. Biol. Evol. 2021, 38, 3022–3027. [Google Scholar] [CrossRef] [PubMed]
  43. Lowe, T.M.; Chan, P.P. tRNAscan-SE on-line: Integrating search and context for analysis of transfer RNA genes. Nucleic Acids Res. 2016, 44, W54–W57. [Google Scholar] [CrossRef] [PubMed]
  44. Laslett, D.; Canbäck, B. ARWEN: A program to detect tRNA genes in metazoan mitochondrial nucleotide sequences. Bioinformatics 2008, 24, 172–175. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  45. Shao, R.; Barker, S.C.; Li, H.; Song, S.; Poudel, S.; Su, Y. Fragmented mitochondrial genomes in two suborders of parasitic lice of eutherian mammals (Anoplura and Rhynchophthirina, Insecta). Sci. Rep. 2015, 5, 17389. [Google Scholar] [CrossRef] [Green Version]
  46. 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]
  47. 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] [Green Version]
  48. Darriba, D.; Taboada, G.L.; Doallo, R.; Posada, D. ProtTest 3: Fast selection of best-fit models of protein evolution. Bioinformatics 2011, 27, 1164–1165. [Google Scholar] [CrossRef] [Green Version]
  49. 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]
Figure 1. The complete mitochondrial genome of the buffalo louse, Haematopinus tuberculatus. Each minichromosome has a coding region and a non-coding region (NCR, in black). The names and transcript orientation of genes are indicated in the coding region and the minichromosomes are in alphabetical order of protein-coding genes and rRNA genes. Abbreviations: atp6 and atp8, ATP synthase F0 subunits 6 and 8; cob, cytochrome b; cox1–3, cytochrome c oxidase subunits 1–3; nad1–6 and nad4L, NADH dehydrogenase subunits 1–6 and 4L; rrnS and rrnL, small and large subunits of ribosomal RNA. tRNA genes are indicated with their single-letter abbreviations of the corresponding amino acids.
Figure 1. The complete mitochondrial genome of the buffalo louse, Haematopinus tuberculatus. Each minichromosome has a coding region and a non-coding region (NCR, in black). The names and transcript orientation of genes are indicated in the coding region and the minichromosomes are in alphabetical order of protein-coding genes and rRNA genes. Abbreviations: atp6 and atp8, ATP synthase F0 subunits 6 and 8; cob, cytochrome b; cox1–3, cytochrome c oxidase subunits 1–3; nad1–6 and nad4L, NADH dehydrogenase subunits 1–6 and 4L; rrnS and rrnL, small and large subunits of ribosomal RNA. tRNA genes are indicated with their single-letter abbreviations of the corresponding amino acids.
Ijms 23 13092 g001
Figure 2. Numbers of mitochondrial minichromosomes of 12 sucking lice which all genes were identified in the mt genome.
Figure 2. Numbers of mitochondrial minichromosomes of 12 sucking lice which all genes were identified in the mt genome.
Ijms 23 13092 g002
Figure 3. The differences among all minichromosomes of four Haematopinus lice. (a) Ten circular minichromosomes of Haematopinus tuberculatus; (b) Nine circular minichromosomes of H. asini; (c) nine circular minichromosomes of H. suis. (d) Nine circular minichromosomes of H. apri. See Figure 1 legend for gene name abbreviation, pV indicates pseudo trnV. * indicates the identical minichromosomes identified among four Haematopinus lice.
Figure 3. The differences among all minichromosomes of four Haematopinus lice. (a) Ten circular minichromosomes of Haematopinus tuberculatus; (b) Nine circular minichromosomes of H. asini; (c) nine circular minichromosomes of H. suis. (d) Nine circular minichromosomes of H. apri. See Figure 1 legend for gene name abbreviation, pV indicates pseudo trnV. * indicates the identical minichromosomes identified among four Haematopinus lice.
Ijms 23 13092 g003
Figure 4. The ancestral mitochondrial minichromosomes of sucking lice that merges in Haematopinus tuberculatus. Gene name and transcription orientation are indicated in the coding region; non-coding regions (NCR) are in black. See Figure 1 legend for gene name abbreviation. Ancestral mt minichromosomes are inferred by Shao et al., 2017.
Figure 4. The ancestral mitochondrial minichromosomes of sucking lice that merges in Haematopinus tuberculatus. Gene name and transcription orientation are indicated in the coding region; non-coding regions (NCR) are in black. See Figure 1 legend for gene name abbreviation. Ancestral mt minichromosomes are inferred by Shao et al., 2017.
Ijms 23 13092 g004
Figure 5. Phylogenetic relationships among 17 species of the parvorder Anopluran lice inferred by Bayesian inference method (BI) and maximum likelihood (ML) of deduced amino acid sequences of eight mitochondrial proteins using MrBayes and IQ-Tree. The elephant louse, Haematomyzus elephantis, was used as the outgroup. Posterior probability values (Bpp) and bootstrap values (Bv) are indicated at nodes.
Figure 5. Phylogenetic relationships among 17 species of the parvorder Anopluran lice inferred by Bayesian inference method (BI) and maximum likelihood (ML) of deduced amino acid sequences of eight mitochondrial proteins using MrBayes and IQ-Tree. The elephant louse, Haematomyzus elephantis, was used as the outgroup. Posterior probability values (Bpp) and bootstrap values (Bv) are indicated at nodes.
Ijms 23 13092 g005
Table 1. Mitochondrial minichromosomes of the buffalo louse Haematopinus tuberculatus, determined by next-generation sequencing using Illumina.
Table 1. Mitochondrial minichromosomes of the buffalo louse Haematopinus tuberculatus, determined by next-generation sequencing using Illumina.
MinichromosomeSize (bp)Size of Coding Region (bp)Size of Non-Coding Region (bp)
trnK-nad4-atp8-atp6-trnN467322812392
trnE-cytb-trnV401312162797
nad2-trnI-cox1-trnL2501926272392
trnD-trnY-cox2-trnS1-trnS2-trnP-cox3-trnA340018821518
trnQ (−) -nad1 (−) -trnT (−) -trnG-nad3-trnW407915062573
trnR-nad4L43693444025
trnH-nad5-trnF-nad645082,2282280
rrnS-trnC35867932793
trnL1-rrnL388212112671
trnM2966672899
Total40,49514,15526,340
Note: minus (−) indicates the mt genes have the opposite orientation of transcription relative to the non-coding region.
Table 2. Long stretches of identical sequence shared between mitochondrial genes in the buffalo louse, Haematopinus tuberculatus.
Table 2. Long stretches of identical sequence shared between mitochondrial genes in the buffalo louse, Haematopinus tuberculatus.
Long Stretches of Identical Sequence Shared (bp)
Pairs of GenesBuffalo
Louse
Horse
Louse
Pig LicePrimate LiceGuanaco LouseRodent Lice
HatHaasHas HaapPhc PhhPtpPesPebPeoMipHok HoaHospPoaPosPor
cox1nad4191112, 181113, 1813, 18111411111014 1013181011
nad4cytb172010 1011 11151214131111 109111211
atp8atp63299 910 10811119119 109889
nad4LtrnV1488 87 76771068 66767
rrnLrrnS1110, 3310 911 1110121011109 139101010
trnL1trnL27, 11, 25159, 10, 169, 10, 1632, 3332, 3332, 3532, 348, 14, 3232, 327, 10, 277 N/A 828 6,11, 2510, 25
trnTtrnP72726267 7687767 76678
Note: Abbreviations of species names are: Hat, Haematopinus tuberculatus (buffalo louse); Haas, Haematopinus asini (horse louse); Has, Haematopinus suis (domestic pig louse); Haap, Haematopinus apri (wild pig louse); Phc, Pediculus humanus capitis (human head louse); Phh, Pediculus humanus humanus (human body louse); Ptp, Pthirus pubis (human pubic louse); Pes, Pediculus schaeffi (chimpanzee louse); Peb, Pedicinus badii (monkey louse); Peo, Pedicinus obtutas (monkey louse); Mip, Microthoracius praelongiceps (guanaco louse); Hok, Hoplopleura kitti (rat louse); Hoa, Hoplopleura akanezumi (mouse louse); Hosp, Hoplopleura sp. (rat louse); Poa, Polyplax asiatica (rat louse); Pos, Polyplax spinulosa (rat louse); Por, Polyplax reclinata (shrew louse); N/A, not available. Stretches of shared identical sequences longer than expected by chance are in bold.
Table 3. The sucking lice included in phylogenetic analyses in this study.
Table 3. The sucking lice included in phylogenetic analyses in this study.
SpeciesHostGenBank Accession NumberReference
Haematopinus apriWild pigKC814611-19[13]
Haematopinus asiniHorseKF939318, KF939322, KF939324,
KF939326, KJ434034-38
[16]
Haematopinus suisDomestic pigKC814602-10[13]
Hoplopleura kittiRatKJ648933-43[14]
Hoplopleura sp.RatMT792483-94[20]
Microthoracius praelongicepsGuanacoKX090378-KX090389[18]
Pediculus humanus capitisHumanJX080388-407[12]
Pediculus humanus humanusHumanFJ499473-90[11]
Pediculus schaeffiChimpanzeeKC241882-97, KR706168-69 [17]
Pedicinus badiiMonkeyMT721726-37[19]
Pedicinus obtutasMonkeyMT792495–506[20]
Pthirus pubisHumanJQ976018, MT721740,
HM241895-8, EU219987-95
[12]
[19]
Polyplax asiaticaRatKF647751-61[15]
Polyplax reclinataShrewMW291451-61[21]
Polyplax spinulosaRatKF647762-72[15]
Haematopinus tuberculatusBuffaloOP574152-61 Present study
Publisher’s Note: MDPI stays neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Share and Cite

MDPI and ACS Style

Fu, Y.-T.; Suleman; Yao, C.; Wang, H.-M.; Wang, W.; Liu, G.-H. A Novel Mitochondrial Genome Fragmentation Pattern in the Buffalo Louse Haematopinus tuberculatus (Psocodea: Haematopinidae). Int. J. Mol. Sci. 2022, 23, 13092. https://0-doi-org.brum.beds.ac.uk/10.3390/ijms232113092

AMA Style

Fu Y-T, Suleman, Yao C, Wang H-M, Wang W, Liu G-H. A Novel Mitochondrial Genome Fragmentation Pattern in the Buffalo Louse Haematopinus tuberculatus (Psocodea: Haematopinidae). International Journal of Molecular Sciences. 2022; 23(21):13092. https://0-doi-org.brum.beds.ac.uk/10.3390/ijms232113092

Chicago/Turabian Style

Fu, Yi-Tian, Suleman, Chaoqun Yao, Hui-Mei Wang, Wei Wang, and Guo-Hua Liu. 2022. "A Novel Mitochondrial Genome Fragmentation Pattern in the Buffalo Louse Haematopinus tuberculatus (Psocodea: Haematopinidae)" International Journal of Molecular Sciences 23, no. 21: 13092. https://0-doi-org.brum.beds.ac.uk/10.3390/ijms232113092

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