Skip to main content
Advertisement
Browse Subject Areas
?

Click through the PLOS taxonomy to find articles in your field.

For more information about PLOS Subject Areas, click here.

  • Loading metrics

Mitochondrial Genomes of Two Barklice, Psococerastis albimaculata and Longivalvus hyalospilus (Psocoptera: Psocomorpha): Contrasting Rates in Mitochondrial Gene Rearrangement between Major Lineages of Psocodea

  • Hu Li ,

    Contributed equally to this work with: Hu Li, Renfu Shao

    Affiliation Department of Entomology, China Agricultural University, Beijing, China

  • Renfu Shao ,

    Contributed equally to this work with: Hu Li, Renfu Shao

    rshao@usc.edu.au (RS); caiwz@cau.edu.cn (WC)

    Affiliation GeneCology Research Centre, Faculty of Science, Education and Engineering, University of the Sunshine Coast, Maroochydore, Queensland, Australia

  • Fan Song,

    Affiliation Department of Entomology, China Agricultural University, Beijing, China

  • Xuguo Zhou,

    Affiliation Department of Entomology, University of Kentucky, Lexington, Kentucky, United States of America

  • Qianqian Yang,

    Affiliation Department of Entomology, China Agricultural University, Beijing, China

  • Zhihong Li,

    Affiliation Department of Entomology, China Agricultural University, Beijing, China

  • Wanzhi Cai

    rshao@usc.edu.au (RS); caiwz@cau.edu.cn (WC)

    Affiliation Department of Entomology, China Agricultural University, Beijing, China

Abstract

The superorder Psocodea has ∼10,000 described species in two orders: Psocoptera (barklice and booklice) and Phthiraptera (parasitic lice). One booklouse, Liposcelis bostrychophila and six species of parasitic lice have been sequenced for complete mitochondrial (mt) genomes; these seven species have the most rearranged mt genomes seen in insects. The mt genome of a barklouse, lepidopsocid sp., has also been sequenced and is much less rearranged than those of the booklouse and the parasitic lice. To further understand mt gene rearrangements in the Psocodea, we sequenced the mt genomes of two barklice, Psococerastis albimaculata and Longivalvus hyalospilus, the first representatives from the suborder Psocomorpha, which is the most species-rich suborder of the Psocodea. We found that these two barklice have the least rearranged mt genomes seen in the Psocodea to date: a protein-coding gene (nad3) and five tRNAs (trnN, trnS1, trnE, trnM and trnC) have translocated. Rearrangements of mt genes in these two barklice can be accounted for by two events of tandem duplication followed by random deletions. Phylogenetic analyses of the mt genome sequences support the view that Psocoptera is paraphyletic whereas Phthiraptera is monophyletic. The booklouse, L. bostrychophila (suborder Troctomorpha) is most closely related to the parasitic lice. The barklice (suborders Trogiomorpha and Psocomorpha) are closely related and form a monophyletic group. We conclude that mt gene rearrangement has been substantially faster in the lineage leading to the booklice and the parasitic lice than in the lineage leading to the barklice. Lifestyle change appears to be associated with the contrasting rates in mt gene rearrangements between the two lineages of the Psocodea.

Introduction

The organization of mitochondrial (mt) genome is highly conserved in insects, as in most other bilateral animals [1], [2]. With few exceptions, the mt genomes of insects consist of 13 protein-coding genes, two rRNA genes, 22 tRNA genes and a large non-coding region (also called the control region) on a single circular chromosome [1][3]. Arrangement of genes in mt genomes is usually stable; most insects known retained exactly the ancestral pattern of mt gene arrangement or have minor changes from the ancestral pattern of mt gene arrangement [1], [4][8].

One group of insects that stands out and has major changes in the organization of mt genome is the superorder Psocodea. Psocodea has ∼10,000 described species in two orders: Psocoptera (barklice and booklice) and Phthiraptera (parasitic lice) [9][12]. Complete mt genomes have been sequenced for two species of the Psocoptera and six species of the Phthiraptera [13][20]. Compared to other insects, species of the Psocodea have three unusual features in their mt genomes that have not been found in any other insects. First, all of the eight species that have been sequenced have rearranged mt genomes. The booklouse, Liposcelis bostrychophila, and the six species of parasitic lice have the most rearranged mt genomes seen in insects - they differ at nearly every gene boundary from the putative ancestor of insects. Second, the booklouse, L. bostrychophila (Psocoptera: suborder Troctomorpha) and the parasitic lice in the suborder Anoplura have multipartite mt genomes. The mt genome of the booklouse, L. bostrychophila, has two mt chromosomes; each chromosome is 7–9 kb and has 16 to 22 genes [20]. The mt genomes of the human body louse, Pediculus humanus, and the human head louse, Pediculus capitis, have 20 minichromosomes; each minichromosome is 3–4 kb in size and contains one to three genes [14], [16]. The mt genome of the human pubic louse, Pthirus pubis, has at least 14 minichromosomes; each minichromosome is 1.8–2.7 kb in size and contains one to five genes [16]. Third, in the screamer louse, Bothriometopus macrocnemis (suborder Ischnocera), all mt genes are encoded by the same strand [18]. In the pigeon louse, Campanulotes bidentatus compar (suborder Ischnocera), all mt genes except trnQ are on the same strand [17].

The order Psocoptera has three suborders: Troctomorpha, Trogiomorpha and Psocomorpha. In addition to the booklouse, L. bostrychophila (suborder Troctomorpha), the complete mt genome of a barklouse, lepidopscocid sp. (suborder Trogiomorpha), has also been sequenced. The mt genome of the lepidopsocid sp. is much less rearranged than those of the booklouse and the parasitic lice; nevertheless, eight genes including a protein-coding gene have rearranged [13]. Prior to the present study, nothing is known about the mt genomes for species in the suborder Psocomorpha, which is the largest suborder of the Psocoptera, containing 25 of the 39 extant families and ∼4,000 of the ∼5,000 described species of the Psocoptera [10], [21], [22]. To further understand mt gene rearrangements and changes in mt genome organization in the Psocodea, we sequenced the mt genomes of two barklice, Psococerastis albimaculata and Longivalvus hyalospilus, the first representatives from the suborder Psocomorpha. We found that these barklice have the least rearranged mt genomes seen in the Psocodea to date. We show that there are contrasting rates in mt gene rearrangement between the two major lineages of the Psocodea.

Materials and Methods

Ethics Statement

No specific permits were required for the insects collected for this study. The insect specimens were collected from roadside vegetation by sweeping. The field collections did not involve endangered or protected species. The species in the family of Psocidae are common insects and are not included in the “List of Protected Animals in China”.

Samples and DNA Extraction

Specimens of P. albimaculata and L. hyalospilus were collected in Kuankuoshui, Suiyang, Guizhou, China, in June 2010. Specimens were initially preserved in 95% ethanol in the field, and transferred to −20°C for long-term storage at the China Agricultural University (CAU). For each species, the genomic DNA was extracted from one male adult’s muscle tissues of the thorax using the DNeasy DNA Extraction kit (Qiagen).

PCR Amplification and Sequencing

The mt genome was amplified by PCR in overlapping fragments with universal insect mt primers [23], and species-specific primers designed from sequenced fragments (Table S1). Short PCRs (<1.5 kb) were with Taq DNA polymerase (Qiagen); the cycling conditions were: 5 min at 94°C, followed by 35 cycles of 50 s at 94°C, 50 s at 48–55°C, 1–2 min at 72°C depending on the size of amplicons, and a final elongation step at 72°C for 10 min. Long PCRs (>1.5 kb) were with Long Taq DNA polymerase (New England BioLabs); the cycling conditions were: 30 s at 95°C, followed by 40 cycles of 10 s at 95°C, 50 s at 48–55°C, 3–6 min at 68°C depending on the size of amplicons, and a final elongation step at 68°C for 10 min. The concentration and size of PCR products were measured by spectrophotometry and agarose gel electrophoresis. PCR fragments were ligated into the pGEM-T Easy Vector (Promega); the resulting plasmid DNAs were isolated using the TIANprp Midi Plasmid Kit (Qiagen). All fragments were sequenced in both directions with an ABI 3730XL Genetic Analyzer, using the BigDye Terminator Sequencing Kit (Applied Biosystems) with two vector-specific primers and internal primers for primer walking.

Assembly, Annotation and Bioinformatics Analysis

Sequence reads from the mt genome of each barklouse species were assembled into contigs with Sequencher (Gene Codes). Protein-coding genes and rRNA genes were identified by BLAST searches in GenBank and then confirmed by alignment with homologous genes from other insects. tRNA genes were identified with tRNAscan-SE v.1.21 [24]. trnR and trnS1, which could not be identified by tRNAscan-SE, were determined by sequence similarity comparison with tRNA genes of other insects. The base composition, codon usage, and nucleotide substitution were analyzed with Mega 5.0 [25]. Secondary structures of stem-loop in control region were folded using Mfold [26].

Sequence Alignment

Four species from the Psocoptera and six species from the Phthiraptera were included in our phylogenetic analyses (Table 1). These species are: 1) three barklice, P. albimaculata, L. hyalospilus and lepidopsocid sp. [13]; 2) a booklouse, L. bostrychophila [20]; 3) four chewing lice, Bothriometopus macrocnemis [18], Campanulotes bidentatus compar [17], Ibidoecus bisignatus [19], and Heterodoxus macropus [15]; 4) the human body louse, Pediculus humanus [14]; and 5) the human pubic louse, Pthirus pubis [16]. Two true bugs, Alloeorhynchus bakeri and Halyomorpha halys (Hemiptera) [27], [28], the lacewing, Chrysoperla nipponensis (Neuroptera) [29], and the ground beetle, Calosoma sp. (Coleoptera) [30], were used as outgroups.

thumbnail
Table 1. Species of insects used in the phylogenetic analyses in the present study.

https://doi.org/10.1371/journal.pone.0061685.t001

Sequences of all mt protein-coding genes and rRNA genes except nad4 were used in phylogenetic analyses; nad4 was excluded because it was not identified in the human pubic louse, P. pubis [16]. Segments of identical sequences (26–127 bp long) shared between five pairs of mt genes in the human body louse, P. humanus, and the human pubic louse, P. pubis [14], [16], were also excluded to ensure only homologous regions of the mt genes were aligned and used in subsequent phylogenetic analyses. Alignment of the nucleotide sequences of each protein-coding gene and its putative amino acid sequence was with MUSCLE [31], adjusted to preserve the reading frame. Sequences of each rRNA gene were aligned with the GUIDANCE algorithm [32], 33, adjusted to its RNA secondary structure [34]. The alignments of individual genes were concatenated after removing poorly aligned sites using Gblocks 0.91 [35].

Phylogenetic Analysis

Three alignments were used for phylogenetic analyses: 1) a concatenated nucleotide sequence alignment of protein-coding genes and two rRNA genes (PCG123R); 2) a concatenated nucleotide sequence alignment of the first and the second codon positions of protein-coding genes and two rRNA genes (PCG12R); and 3) a concatenated amino acid sequence alignment of protein-coding genes (AA). Partitioned ML and Bayesian analyses were run with PCG123R, PCG12R and AA matrix, using RAxML 7.0.3 [36] and MrBayes 3.2.1 [37]. The best-fit model for the amino acid sequence alignment was determined with ProtTest [38], and the jModelTest 0.1.1 [39] was used for the nucleotide sequence of each gene, according to the Akaike Information Criterion (AIC). For the ML analyses, GTRMIX option for nucleotide sequence and MtREV model for amino acid sequence were used to optimize the topology. For the combined dataset, 1,000 independent runs from random starting trees were performed to find the highest scoring replicate. Node support was calculated by acquiring bootstrap values from heuristic searches of 1000 resampled datasets, using the rapid bootstrap feature of RAxML [40].

For Bayesian analyses, the most appropriate substitution model was GTR+I+G model for each protein-coding gene, 1st and 2nd codon positions of protein-coding genes, and rrnL gene; GTR+G model for rrnS gene; and MtREV model for amino acid sequence of each protein-coding genes. Two simultaneous runs of 10 million generations were conducted for the matrix and trees were sampled every 1,000 generations, with the first 25% discarded as burn-in. Stationarity was considered to be reached when the average standard deviation of split frequencies was below 0.01 [41].

Results and Discussion

Mitochondrial Genomes of the Barklice, Psococerastis Albimaculata and Longivalvus Hyalospilus

The mt genome of P. albimaculata contains the entire set of 37 genes (13 protein-coding genes, 22 tRNA genes, and two rRNA genes; Figure 1 and Table S2) and a putative control region that are usually present in animal mt genomes [1], [3]. We found the same set of mt genes in L. hyalospilus, except the control region and the adjacent trnM gene due to our unsuccess to amplify this region (Figure 1 and Table S3). These two barklice have the same arrangement of mt genes to each other but differs from the putative ancestor of insects: a protein-coding gene (nad3) and five tRNA genes (trnN, trnS1, trnE, trnM and trnC) have translocated. Genes are encoded by both strands in the mt genomes of these two barklice: 14 genes on one strand whereas the rest on the other strand (Table S2 and S3). Thirteen pairs of adjacent genes in the mt genomes of these two barklice overlap by 1 to 16 bp. All of the protein-coding genes of the two barklice start with ATN codons and stop with TAA/TAG codons or truncated codons TA or T.

thumbnail
Figure 1. Mitochondrial genomes of the barklice, Psococerastis albimaculata and Longivalvus hyalospilus.

Circular maps were drawn with CGView [42]. Arrows indicate the orientation of gene transcription. Protein-coding genes are shown as blue arrows, rRNA genes as purple arrows, tRNA genes as brown arrows and large non-coding regions (>100 bp) as grey rectangle. Abbreviations of gene names are: atp6 and atp8 for ATP synthase subunits 6 and 8, cox13 for cytochrome oxidase subunits 1–3, cytb for cytochrome b, nad16 and nad4L for NADH dehydrogenase subunits 1–6 and 4L, rrnL and rrnS for large and small rRNA subunits. tRNA genes are indicated with their one-letter corresponding amino acids; the two tRNA genes for leucine and serine have different anticodons: L1 (anticodon TAG), L2 (TAA), S1 (TCT) and S2 (TGA). The GC content is plotted using a black sliding window, as the deviation from the average GC content of the entire sequence. GC-skew is plotted as the deviation from the average GC-skew of the entire sequence. The inner cycle indicates the location of genes in the mt genome.

https://doi.org/10.1371/journal.pone.0061685.g001

The nucleotide compositions of the mt genomes of the two barklice are biased toward A and T. The nucleotide skew statistics for the entire majority strand indicate moderate A skew and obvious C skew, and the coding strand of protein-coding genes display a moderate T skew and slight C skew (Table 2). The A+T-richness of mt genomes of these two barklice is reflected further in the codon usage (Table S4). The overall ratio of G+C-rich codons to A+T-rich codons is 0.13 in the two barklice and there is a strong bias to A+T at the third codon positions of the protein-coding genes (Table 3).

thumbnail
Table 2. Codon usage of the protein-coding genes in the mitochondrial genomes of Psococerastis albimaculata and Longivalvus hyalospilus.

https://doi.org/10.1371/journal.pone.0061685.t002

thumbnail
Table 3. Nucleotide composition at each codon position of the protein-coding genes in the mitochondrial genomes of Psococerastis albimaculata and Longivalvus hyalospilus.

https://doi.org/10.1371/journal.pone.0061685.t003

The multiple alignment of 21 tRNA genes (excluded trnM) extends over 1,366 positions and contains 1,229 conserved nucleotides (90.0%) between the two barklice. Nucleotides at the stems and anticodon loops were conserved; variations were largely restricted to the loop of TψC and DHU arm (Figure 2). The nucleotide sequences of the mt genomes of two barklice, P. albimaculata and L. hyalospilus, have been deposited in GenBank under accession numbers JQ910989 and JQ910986.

thumbnail
Figure 2. Inferred secondary structure of the mitochondrial tRNAs of the barklice, Psococerastis albimaculata and Longivalvus hyalospilus.

tRNAs are labeled with the abbreviations of their corresponding amino acids. Lines indicate Watson-Crick bonds; dots indicate GU bonds. Green circle with red letter inside highlight variations of nucleotides in tRNA between two barklice.

https://doi.org/10.1371/journal.pone.0061685.g002

The Control Region in the Mitochondrial Genome of the Barklouse, Psococerastis albimaculata

The putative control region (911 bp) was flanked by rrnS and trnM in the mt genome of P. albimaculata. The control region is highly AT-rich (86.9%; majority strand) and can be divided into five parts (Figure 3A): 1) 59 bp leading sequence; 2) 180 bp tandem repeat sequences consisting of five 36 bp repeat units (Figure 3B); 3) 448 bp A+T-rich sequences (A+T 88.2%); 4) 180 bp tandem repeated sequences consisting of nine 20 bp repeat units (Figure 3B); and 5) 44 bp stem-loop structure (Figure 3C). Stem-loops are common in the control regions of animal mt genomes and may have roles in the initiation of gene transcription and DNA replication [43][47]. The pattern of two tandem repeated sequences with an A+T-rich sequence in between is also present in the control region of the other barklouse, lepidopsocid-RS (Trogiomorpha), but not in the booklouse, L. bostrychophila (Troctomorpha), nor in any parasitic lice.

thumbnail
Figure 3. The control region in the mitochondrial genome of the barklouse, Psococerastis albimaculata.

A) Structural elements in the control region; B) sequences of the tandem-repeat units; C) predicated stem-loop structures.

https://doi.org/10.1371/journal.pone.0061685.g003

Outside the control region, there are 146 bp non-coding sequences in 13 intergenic regions of P. albimaculata, and 178 bp non-coding sequences in 12 intergenic regions in L. hyalospilus. Most non-coding sequences are scatted in short runs (1–16 bp in P. albimaculata and 1–26 bp in L. hyalospilus). However, two of these non-coding sequences are longer in length and locate in the same intergenic regions in the two barklice: 1) between trnQ and nad2 (29 bp in P. albimaculata and 38 bp in L. hyalospilus); and 2) between nad5 and nad3 (71 bp in P. albimaculata and 73 bp in L. hyalospilus). These two non-coding sequences are in the regions where gene rearrangement occurred and are likely generated from events of tandem duplication followed by random deletions (See below).

Phylogenetic Relationships among Major Lineages of the Psocodea Inferred from Mitochondrial Genome Sequences

The Psocoptera (booklice and barklice) and the Phthiraptera (parasitic lice) have traditionally been recognized as two separate orders [12], [48], [49]. The Psocoptera (booklice and barklice) are free-living insects and consist of over 5,000 species with a world-wide distribution, and are divided into three suborders: Trogiomorpha, Troctomorpha and Psocomorpha [10], [22]. Members of the order Phthiraptera (∼4,900 species) are wingless insects, parasitic on birds and mammals. There are four recognized suborders in the Phthiraptera: Amblycera, Ischnocera, Anoplura and Rhynchophthirina [9], [11], [50]. Both morphological and molecular analyses indicate a close relationship between parasitic lice (Phthiraptera) and booklice (family Liposcelididae); the order Psocoptera is therefore paraphyletic [9], [12], [20], [51][55]. The monophyly of the Phthiraptera, however, has been controversial. A number of shared morphological and physiological characters support the monophyly of the Phthiraptera [9], [48], [49]. Analyses of mt 12S and 16S rDNA and nuclear 18S rDNA sequences, however, indicate that the parasitic lice are paraphyletic: the suborder Amblycera is more closely related to the booklouse family Liposcelididae than to the other three suborders of the parasitic lice [12], [51], [54].

We tested the phylogenetic relationships among the major lineages of the Psocodea with the mt genome sequences of P. albimaculata and L. hyalospilus, and eight other species of the Psocodea. We used three different datasets: 1) a concatenated nucleotide sequence alignment of protein-coding genes and two rRNA genes (PCG123R); 2) a concatenated nucleotide sequence alignment of the first and the second codon positions of protein-coding genes and two rRNA genes (PCG12R); and 3) a concatenated amino acid sequence alignment of protein-coding genes (AA). All three datasets were run with partitioned ML and Bayesian analyses based on best-fit models. We recovered two major clades in the Psocodea with strong support regardless the dataset and the method we used: 1) species of barklice in the suborders Psocomorpha and Trogiomorpha were clustered together (Clade A, Figure 4); and 2) the booklouse, L. bostrychophila (suborder Troctomorpha), formed a clade with the parasitic lice (Clade B, Figure 4). The parasitic lice (Phthiraptera) are monophyletic with strong support; within the parasitic lice, the suborder Ischnocera, however, is paraphyletic.

thumbnail
Figure 4. Phylogenetic relationships among major lineages of the Psocodea inferred from mitochondrial genome sequences.

Numbers close to the branching points are ML bootstrap support values (above) and Bayesian posterior probabilities (below) in percentages; only support above 50% is shown. Numbers from left to right are from PCG123R, PCG12R and AA alignments respectively.

https://doi.org/10.1371/journal.pone.0061685.g004

Contrasting Rates in Mitochondrial Gene Rearrangement between Two Major Lineages of the Psocodea

Seven species from the Clade B above have been sequenced for complete mt genomes previously, i.e., the booklouse, L bostrychophila (suborder Troctomorpha) [20] and six parasitic lice [14][19]. All of these seven species have extremely rearranged mt genomes, having little similarity in gene arrangement with each other, nor with any other insects. Both translocations and inversions occurred in these species relative to the ancestral gene arrangement of insects. Only four ancestral gene boundaries of insects (trnG-nad3, trrnL1-rrnL, nad4-nad4L, and atp8-atp6) were found in the Clade B species, of which only atp8-atp6 was retained by all of the seven species in the Clade B. Apparently, mt gene rearrangement has been occurring much more often in the Clade B species than in other insects.

In contrast, species from the Clade A investigated to date have much less rearranged mt genomes and retained most of the ancestral gene arrangements of insects [13]. In particular, the two barklice we sequenced in the present study, P. albimaculata and L. hyalospilus, have the least rearranged mt genomes seen in the Psocodea: a protein-coding gene (nad3) and five tRNAs (trnN, trnS1, trnE, trnM and trnC) have translocated relative to that of the ancestor of insects (Figure 5). Gene rearrangement in these two barklice occurred in the two “hot spot” regions for gene rearrangement [56]: 1) between cox3 and trnH, and 2) between CR and trnY (Figure 5). Two events of tandem duplication followed by random deletions could account for, straightforwardly, the rearrangement of mt genes observed in P. albimaculata and L. hyalospilus (Figure 6). The number of mt genes that have rearranged in P. albimaculata and L. hyalospilus is even less than in the other barklouse, lepidopsocid-RS (cox2 and seven tRNAs rearranged) [13]. Together, 20 gene boundaries (Figure 5, seven gene blocks: A, B, D, E, F, G, and H) are shared and 32 ancestral gene boundaries are retained in these three barklice.

thumbnail
Figure 5. Comparison of mitochondrial gene arrangement between barklice, booklice and the hypothetical ancestor of insects.

Abbreviations of species names are: Pa for P. albimaculata, Lh for L. hyalospilus, Dy for Drosophila yakuba, Le for lepidopsocid, and Lb for L. bostrychophila. Abbreviations of gene names follow Figure 1. Genes are transcribed from left to right except those underlined, which have the opposite transcriptional orientation. Pa and Lh have the same set of mt genes and gene arrangement, except the control region and the adjacent trnM gene due to our unsuccess to amplify this region in Lh. Shared gene boundaries are indicated by brackets and letter codes: A) trnY-cox1-trnL2; B) from trnK to cox3; C) atp8-atp6; D) trnF-nad5; E) trnS1- trnE; F) from trnH to cytb; G) from nad1 to CR; and H) trnQ-nad2. Two ancestral gene blocks highlighted by the orange; broken lines indicate the inferred blocks that had duplicated in tandem. The circling arrows indicate inversions of mt genes relative to the hypothetical ancestor of insects.

https://doi.org/10.1371/journal.pone.0061685.g005

thumbnail
Figure 6. Inferred tandem-duplication-and-random-deletion events that account for the mitochondrial gene rearrangements in the barklice, Psococerastis albimaculata and Longivalvus hyalospilus.

A) genes between cox3 and nad4; B) genes between CR and cox1. Two longer non-coding sequences are highlighted by green color.

https://doi.org/10.1371/journal.pone.0061685.g006

Tandem duplication followed by random deletion model has been proposed to account for mt gene rearrangement [8], [57][59]. This model can explain the mt gene rearrangements observed in the two barklice, P. albimaculata and L. hyalospilus (Figure 6). The rearranged mt gene blocks of these two barklice, one from trnG to trnH and another from trnM to trnY can be generated by a single event of tandem duplication of the ancestral gene block from trnG to trnH (Figure 6A) and from trnI to trnY (Figure 6B), respectively, and followed by random deletion of excess genes. The non-coding sequences present in the two rearranged gene boundaries (between trnQ and nad2, and between nad5 and nad3) are likely traces of the random deletion. The mt gene rearrangement in the other barklouse lepidopsocid-RS (suborder Trogiomorpha) is more complicated than in P. albimaculata and L. hyalospilus; cox2 gene and four tRNA genes have translocated from long distance and cannot be accounted for alone by tandem duplication followed by random deletion model. The mt gene rearrangements in the booklice and parasitic lice are much more complicated than in the barklice; multiple mechanisms and frequent rearrangement events are likely involved [16], [60], [61].

What caused the substantial difference between the two clades of the Psocodea in the rates of mt gene rearrangement? An obvious difference between the two clades is the lifestyle. The barklice in the Clade A are entirely free-living insects, which often feed on fungal spores; whereas Clade B (booklice and parasitic lice) is a mixture of short-term commensal and parasitic. The booklouse, L. bostrychophila, is mainly an inhabitant of households and a major pest to stored grains world-wide; moreover, there are many records of various species of booklice in the plumage of birds and the pelage of mammals, as well as in their nests [62], [63]. This association is believed to be a short-term commensal (non-parasitic) relationship, which may have given rise to a parasitic and permanent association [64]. All parasitic lice (Phthiraptera) feed on the skin, skin debris or blood of their vertebrate hosts and spend their entire life cycle on the body of the host [12]. The lifestyle change in the Clade B appears to be associated with an increased rate of mt gene rearrangement. However, why they are linked and exactly what biological and lifestyle factors contributed to the contrasting rates in mt gene rearrangement between the two major lineages of the Psocodea are not yet clear to us and remains to be determined.

Supporting Information

Table S1.

Primers used in the present study.

https://doi.org/10.1371/journal.pone.0061685.s001

(DOC)

Table S2.

Genes in the mitochondrial genome of the barklouse, Psococerastis albimaculata.

https://doi.org/10.1371/journal.pone.0061685.s002

(DOC)

Table S3.

Genes in the mitochondrial genome of the barklouse, Longivalvus hyalospilus.

https://doi.org/10.1371/journal.pone.0061685.s003

(DOC)

Table S4.

Codon usage in the protein-coding genes of the mitochondrial genomes of the barklice, Psococerastis albimaculata and Longivalvus hyalospilus.

https://doi.org/10.1371/journal.pone.0061685.s004

(DOC)

Acknowledgments

We thank Prof. Fasheng Li and Dr. Luxi Liu (China Agricultural University) for their help in the insect identification.

Author Contributions

Conceived and designed the experiments: HL WZC. Performed the experiments: HL FS. Analyzed the data: HL RS FS. Contributed reagents/materials/analysis tools: RS WZC. Wrote the paper: HL RS XGZ QQY ZHL WZC.

References

  1. 1. Boore JL (1999) Animal mitochondrial genomes. Nucleic Acids Res 27: 1767–1780.
  2. 2. Lavrov DV (2007) Key transitions in animal evolution: a mitochondrial DNA perspective. Integr Comp Biol 47: 734–743.
  3. 3. Wolstenholme DR (1992) Animal mitochondrial DNA: structure and evolution. Int Rev Cytol 141: 173–216.
  4. 4. Dowton M, Castro LR, Austin AD (2002) Mitochondrial gene rearrangements as phylogenetic characters in the invertebrates: the examination of genome ‘morphology’. Invertebr Syst 16: 345–356.
  5. 5. Shao R, Barker SC (2007) Mitochondrial genomes of parasitic arthropods: implications for studies of population genetics and evolution. Parasitology 134: 153–167.
  6. 6. Cameron SL, Sullivan J, Song H, Miller KB, Whiting MF (2009) A mitochondrial genome phylogeny of the Neuropterida (lace-wings, alderflies and snakeflies) and their relationship to the other holometabolous insect orders. Zool Scr 38: 575–590.
  7. 7. Beckenbach AT (2012) Mitochondrial genome sequences of Nematocera (lower Diptera): evidence of rearrangement following a complete genome duplication in a winter crane fly. Genome Biol Evol 4: 89–101.
  8. 8. Li H, Liu H, Shi AM, Štys P, Zhou XG, et al. (2012) The complete mitochondrial genome and novel gene arrangement of the unique-headed bug Stenopirates sp. (Hemiptera: Enicocephalidae). PLoS ONE 7: e29419.
  9. 9. Lyal CHC (1985) Phylogeny and classification of the Psocodea, with particular reference to the lice (Psocodea: Phthiraptera). Syst Entomol 10: 145–165.
  10. 10. Lienhard C, Smithers CN (2002) Psocoptera: world catalogue and bibliography. Geneva: Muséum d’Histoire Naturelle Press.
  11. 11. Barker SC, Whiting M, Johnson KP, Murrell A (2003) Phylogeny of the lice (Insecta: Phthiraptera) inferred from small subunit rRNA. Zool Scr 32: 407–414.
  12. 12. Johnson KP, Yoshizawa K, Smith VS (2004) Multiple origins of parasitism in lice. Proc R Soc Lond B 271: 1771–1776.
  13. 13. Shao R, Dowton M, Murrell A, Barker SC (2003) Rates of gene rearrangement and nucleotide substitution are correlated in the mitochondrial genomes of insects. Mol Biol Evol 20: 1612–1619.
  14. 14. Shao R, Kirkness EF, Barker SC (2009) The single mitochondrial chromosome typical of animals has evolved into 18 minichromosomes in the human body louse, Pediculus humanus. Genome Res 19: 904–912.
  15. 15. Shao R, Campbell NJH, Barker SC (2011) Numerous gene rearrangements in the mitochondrial genome of the wallaby louse, Heterodoxus macropus (Phthiraptera). Mol Biol Evol 18: 858–865.
  16. 16. Shao R, Zhu XQ, Barker SC, Herd K (2012) Evolution of extensively fragmented mitochondrial genomes in the lice of humans. Genome Biol Evol 4: 1088–1101.
  17. 17. Covacin C, Shao R, Cameron S, Barker SC (2006) Extraordinary number of gene rearrangements in the mitochondrial genomes of lice (Phthiraptera: Insecta). Insect Mol Biol 15: 63–68.
  18. 18. Cameron SL, Johnson KP, Whiting MF (2007) The mitochondrial genome of the screamer louse Bothriometopus (Phthiraptera: Ischnocera): effects of extensive gene rearrangements on the evolution of the genome. J Mol Evol 65: 589–604.
  19. 19. Cameron SL, Yoshizawa K, Mizukoshi A, Whiting MF, Johnson KP (2011) Mitochondrial genome deletions and minicircles are common in lice (Insecta: Phthiraptera). BMC Genomics 12: 394.
  20. 20. Wei DD, Shao R, Yuan ML, Dou W, Barker SC, et al. (2012) The multipartite mitochondrial genome of Liposcelis bostrychophila: insights into the evolution of mitochondrial genomes in bilateral animals. PLoS ONE 7: e33973.
  21. 21. Yoshizawa K (2002) Phylogeny and higher classification of suborder Psocomorpha (Insecta: Psocodea: ‘Psocoptera’). Zool J Linn Soc 136: 371–400.
  22. 22. Johnson KP, Mockford EL (2003) Molecular systematics of Psocomorpha (Psocoptera). Syst Entomol 28: 409–440.
  23. 23. Simon C, Buckley TR, Frati F, Stewart JB, Beckenbach AT (2006) Incorporating molecular evolution into phylogenetic analysis, and a new compilation of conserved polymerase chain reaction primers for animal mitochondrial DNA. Annu Rev Ecol Evol Syst 37: 545–579.
  24. 24. Lowe TM, Eddy SR (1997) tRNAscan–SE: a program for improved detection of transfer RNA genes in genomic sequence. Nucleic Acids Res 25: 955–964.
  25. 25. Tamura K, Peterson D, Peterson N, Stecher G, Nei M, et al. (2011) MEGA5: Molecular evolutionary genetics analysis using maximum likelihood, evolutionary distance, and maximum parsimony methods. Mol Biol Evol 28: 2731–2739.
  26. 26. Zuker M (2003) Mfold web server for nucleic acid folding and hybridization prediction. Nucleic Acids Res 31: 3406–3415.
  27. 27. Lee W, Kang J, Jung C, Hoelmer K, Lee SH, et al. (2009) Complete mitochondrial genome of brown marmorated stink bug Halyomorpha halys (Hemiptera: Pentatomidae), and phylogenetic relationships of hemipteran suborders. Mol Cells 28: 155–165.
  28. 28. Li H, Liu HY, Cao LM, Shi AM, Yang HL, et al. (2012) The complete mitochondrial genome of the damsel bug Alloeorhynchus bakeri (Hemiptera: Nabidae). Int J Biol Sci 8: 93–107.
  29. 29. Haruyama N, Mochizuki A, Sato Y, Naka H, Nomura M (2011) Complete mitochondrial genomes of two green lacewings, Chrysoperla nipponensis (Okamoto, 1914) and Apochrysa matsumurae Okamoto, 1912 (Neuroptera: Chrysopidae). Mol Biol Rep 38: 3367–3373.
  30. 30. Song H, Sheffield NC, Cameron SL, Miller KB, Whiting MF (2010) When phylogenetic assumptions are violated: base compositional heterogeneity and among-site rate variation in beetle mitochondrial phylogenomics. Syst Entomol 35: 429–448.
  31. 31. Edgar RC (2004) MUSCLE: multiple sequence alignment with high accuracy and high throughput. Nucleic Acids Res 32: 1792–1797.
  32. 32. Penn O (2010) GUIDANCE: a web server for assessing alignment confidence scores. Nucleic Acids Res 38: W23–W28.
  33. 33. Penn O, Privman E, Landan G, Graur D, Pupko T (2010) An alignment confidence score capturing robustness to guide-tree uncertainty. Mol Biol Evol 27: 1759–1767.
  34. 34. Li H, Liu H, Song F, Shi A, Zhou X, et al. (2012) Comparative mitogenomic analysis of damsel bugs representing three tribes in the family Nabidae (Insecta: Hemiptera). PLoS ONE 7: e45925.
  35. 35. Castresana J (2000) Selection of conserved blocks from multiple alignments for their use in phylogenetic analysis. Mol Biol Evol 17: 540–552.
  36. 36. Stamatakis A (2006) RAxML-VI-HPC: Maximum likelihood-based phylogenetic analyses with thousands of taxa and mixed models. Bioinformatics 22: 2688–2690.
  37. 37. Ronquist F, Huelsenbeck JP (2003) MrBayes 3: Bayesian phylogenetic inference under mixed models. Bioinformatics 19: 1572–1574.
  38. 38. Abascal F, Zardoya R, Posada D (2007) ProtTest: Selection of best-fit models of protein evolution. Bioinformatics 24: 1104–1105.
  39. 39. Posada D (2008) jModelTest: Phylogenetic model averaging. Mol Biol Evol 25: 1253–1256.
  40. 40. Stamatakis A, Hoover P, Rougemont J (2008) A rapid bootstrap algorithm for the RAxML Web servers. Syst Biol 57: 758–771.
  41. 41. Huelsenbeck JP, Ronquist F, Nielsen R, Bollback JP (2001) Bayesian inference of phylogeny and its impact on evolutionary biology. Science 294: 2310–2314.
  42. 42. Stothard P, Wishart DS (2005) Circular genome visualization and exploration using CGView. Bioinformatics 21: 537–539.
  43. 43. Clayton DA (1982) Replication of animal mitochondrial DNA. Cell 28: 693–705.
  44. 44. Clayton DA (1991) Replication and transcription of vertebrate mitochondrial DNA. Annu Rev Cell Biol 7: 453–478.
  45. 45. Clary DO, Wolstenholme DR (1987) Drosophila mitochondrial DNA: Conserved sequences in the A+T-rich region and supporting evidence for a secondary structure model of the small ribosomal RNA. J Mol Evol 25: 116–125.
  46. 46. Zhang DX, Szymura JM, Hewitt GM (1995) Evolution and structural conservation of the control region of insect mitochondrial DNA. J Mol Evol 40: 382–391.
  47. 47. Li H, Gao JY, Liu HY, Liu H, Liang AP, et al. (2011) The architecture and complete sequence of mitochondrial genome of an assassin bug Agriosphodrus dohrni (Hemiptera: Reduviidae). Int J Biol Sci 7: 792–804.
  48. 48. Hennig W (1966) Phylogenetic systematics. Urbana: University of Illinois Press.
  49. 49. Jamieson BGM, Dallai R, Afzelius BA (1999) Insects: Their spermatozoa and phylogeny. Enfield: Science Publisher.
  50. 50. Clay T (1970) The Amblycera (Phthiraptera: Insecta). Bull Br Mus (Nat Hist) Entomol 25: 73–98.
  51. 51. Yoshizawa K, Johnson KP (2003) Phylogenetic position of Phthiraptera (Insecta: Paraneoptera) and elevated rate of evolution in mitochondrial 12S and 16S rDNA. Mol Phylogenet Evol 29: 102–114.
  52. 52. Yoshizawa K, Johnson KP (2006) Morphology of male genitalia in lice and their relatives and phylogenetic implications. Syst Entomol 31: 350–361.
  53. 53. Yoshizawa K, Johnson KP (2010) How stable is the ‘Polyphyly of Lice’ hypothesis (Insecta: Psocodea)?: A comparison of phylogenetic signal in multiple genes. Mol Phylogenet Evol 55: 939–951.
  54. 54. Murrell A, Barker SC (2005) Multiple origins of parasitism in lice: phylogenetic analysis of SSU rDNA indicates that the Phthiraptera and Psocoptera are not monophyletic. Parasitol Res 97: 274–280.
  55. 55. Yoshizawa K, Lienhard C (2010) In search of the sister group of the true lice: a systematic review of booklice and their relatives, with an updated checklist of Liposcelididae (Insecta: Psocodea). Arthropod Syst Phyl 68: 181–195.
  56. 56. Dowton M, Austin AD (1999) Evolutionary dynamics of a mitochondrial rearrangement ‘Hot Spot’ in the Hymenoptera. Mol Biol Evol 16: 298–309.
  57. 57. Moritz C, Brown WM (1986) Tandem duplication of D-loop and ribosomal RNA sequences in lizard mitochondrial DNA. Science 308: 1425–1427.
  58. 58. Boore JL (2000) The duplication/random loss model for gene rearrangement exemplified by mitochondrial genomes of deuterostome animal. In: Sankoff D, Nadeau JH, editor. Comparative genomics. Dordrecht: Kluwer Academic Press. p.133–147.
  59. 59. Shao R, Barker SC, Mitani H, Takahashi M, Fukunaga M (2006) Molecular mechanisms for the variation of mitochondrial gene content and gene arrangement among chigger mites of the genus Leptotrombidium (Acari: Acariformes). J Mol Evol 63: 251–261.
  60. 60. Dowton M, Campbell NJH (2001) Intramitochondrial recombination – is it why some mitochondrial genes sleep around? Trends Ecol Evol 16: 269–271.
  61. 61. Kraytsberg Y, Schwartz M, Brown TA, Ebralidse K, Kunz WS, et al. (2004) Recombination of human mitochondrial DNA. Science 304: 981.
  62. 62. Pearlman JV (1960) Some African Psocoptera found on rats. Entomologist 93: 246–250.
  63. 63. Mockford EL (1967) Some Psocoptera from the plumage of birds. Proc Entomol Soc Wash 69: 307–309.
  64. 64. Hopkins GHE (1949) The host associations of the lice of mammals. Proc Zool Soc Lond 119: 387–604.