Skip to main content
  • Research article
  • Open access
  • Published:

Intraspecific rearrangement of mitochondrial genome suggests the prevalence of the tandem duplication-random loss (TDLR) mechanism in Quasipaa boulengeri

Abstract

Background

Tandem duplication followed by random loss (TDRL) is the most frequently invoked model to explain the diversity of gene rearrangements in metazoan mitogenomes. The initial stages of gene rearrangement are difficult to observe in nature, which limits our understanding of incipient duplication events and the subsequent process of random loss. Intraspecific gene reorganizations may represent intermediate states, and if so they potentially shed light on the evolutionary dynamics of TDRL.

Results

Nucleotide sequences in a hotspot of gene-rearrangement in 28 populations of a single species of frog, Quasipaa boulengeri, provide such predicted intermediate states. Gene order and phylogenetic analyses support a single tandem duplication event and a step-by-step process of random loss. Intraspecific gene rearrangements are not commonly found through comparison of all mitochondrial DNA records of amphibians and squamate reptiles in GenBank.

Conclusions

The intraspecific variation in Q. boulengeri provides insights into the rate of partial duplications and deletions within a mitogenome, and reveals that fixation and gene-distribution in mitogenomic reorganization is likely non-adaptive.

Background

The order of mitochondrial (mt) genes in metazoans varies greatly [1, 2] and the molecular drivers that explain the underlying evolution are subject to debate [3, 4]. The most widely invoked model involves tandem duplication of mt genes followed by the random loss of one copy (TDRL) [58]. However, duplication and non-random loss may result from the transcriptional polarities of genes and their positions in the genome, as two models describe: tandem duplication and non-random loss (TDNL) [9]; and dimer-mitogenome and non-random loss (DMNL) [10]. These models attribute gene-rearrangements to clustering by common polarity. Further gene rearrangement may be a result of intra- or inter-mtDNA recombination [11, 12]. Recently, Shi et al. [13] proposed that double replications and random loss (DRRL) best explained gene rearrangements in flounders.

The TDRL model is most widely accepted, but no definitive evidence supports it. According to this model, tandem duplications involve imprecise terminations, strand slippages, and/or mispairings, which result in errors in mtDNA replication [14, 15]. The hypothesis predicts that the new, intermediate mitogenome will contain duplicated genes, one of which is subsequently randomly lost. Despite the increasing number of taxa known to have gene rearrangements, few mitogenomes exhibit intermediate states that could point to this evolutionary mechanism, even though pseudogenes and residues of tandemly duplicated sequences may provide indirect evidence for an intermediate step in genomic rearrangement [3, 6, 16].

If gene deletion occurs randomly, then populations should have mitogenomes with varying gene-orders that consist of alternative arrangements of duplicated genes. Such alternative gene arrangements have been reported in only a few closely-related lineages or species [16, 17]. These interspecific occurrences support the TDRL model, yet no information exists as to when and where the rearrangements occurred and how they subsequently dispersed within a species. Investigations at a lower (intraspecific) level may be necessary to understand the evolution of mitochondrial gene rearrangements.

Intraspecific rearrangements of mitogenomes are rarely reported in vertebrates. Many species have structural diversity in their control region (CR), but all of these involve non-coding sequences. Gene order diversity within a species is known only from asexual squamates [7, 14, 18], an amphisbaenid [19] and a bird [20]. In these cases, gene rearrangements that qualify as potential intermediate states involve either a large number of genes adjacent to the CR or the formation of pseudogenes [3, 19]. In addition, gene rearrangements in the mitogenomes of asexual squamates results from multiple independent duplications and lack the random loss of alternative states [7].

High levels of gene rearrangement have been reported from amphibians, especially among so-called modern frogs [2123]. A hotspot of gene rearrangement has been reported in the “WANCY” region (trnW, trnA, trnN, origin of light strand replication (OL), trnC, and trnY) [6, 24]. Because many amphibians have gene rearrangements in their WANCY region [17, 25, 26], reorganizations in this group facilitate testing hypotheses on how gene-rearrangement occurs. Each of the above hypotheses predicts a unique arrangement of five short tRNA genes in the WANCY region (Fig. 1), which can be compared with the results of sequencing.

Fig. 1
figure 1

Expected mitochondrial gene rearrangement under different evolutionary scenarios. a Tandem duplication–random loss model (TDRL): A, N, OL, C were tandem duplicated, followed by random loss of the redundant copies. Random loss could occur repeatly, resulting in alternative loss types [5]. b Tandem duplication and non-random loss (TDNL), or dimer-mitogenome and non-random loss (DMNL) models: a dimeric molecule was formed by two monomers linked head-to-tail, then one of the two sets of promoters lost function, and genes with the same polarity would cluster together [9, 10]. c Inter- or intra-mtDNA recombination: duplication was caused by unequal crossing over of intermolecular recombination. Redundant copies were then deleted. Intramolecular recombination could cause concerted evolution of the two copies of trnA [12]. d Double replications and random loss (DRRL) model: the CR was duplicated and translocated, then double replications of the mitogenome were successively initiated from the two CRs, leading to the duplication of the genes between the two CRs, followed by random loss [13]. Underline indicates the transcriptional direction of L-strand–encoding gene. “---” represents other coding gene. “-”, pseudogenes or noncoding sequence. Gray boxes represent the genes involved in rearrangement

Herein, we report a diversity of rearrangements in the frog Quasipaa boulengeri and the discovery of intermediate states. To test the hypotheses of mechanisms of gene rearrangement, we investigate the origins and evolution of the rearrangement by 1) determing the structure of the rearranged region for each type, 2) speculating on the steps resulting in observed gene rearrangements, and 3) placing each type on an inferred phylogeny and estimating the time at which each rearrangement arose. We supplement this with an in silico approach using mitochondrial gene orders from GenBank data. By using a custom Perl script (mtGordV0.5.pl), we explore the frequency of occurrence of intraspecific rearrangements in Amphibia and Squamata, which have high diversities in mitogenomic rearrangements.

Results and Discussion

Intraspecific rearrangements

We obtained 290 samples from 28 localities for Quasipaa boulengeri (Fig. 2 and Additional file 1: Table S1). Sequences from a region encompassing nad2 to cox1, which includes the WANCY hotspot, revealed gene-organizations atypical of vertebrates. Stable secondary structures of tRNA genes and the absence of premature stop codons in nad2 and cox1 authenticated the sequencing of mtDNA. The WANCY region in this frog differed from the typical organization by having a long noncoding sequence that ranges in size from 473 bp to 925 bp. Further, gene annotation identified different positions for the gene and its copy, even within a single population. The details of gene organization of the WANCY region for each sample were listed in Additional file 2: Table S2.

Fig. 2
figure 2

Map of sampling localities for Quasipaa boulengeri. Populations are presented as pie-diagrams with slice-size proportional to the frequency of type of mitochondrial gene rearrangement. Green: Type I; red: Type II; blue: Type III; yellow: Type IV. This map is created with ArcGIS (ESRI, http://www.esri.com/software/arcgis)

Annotation identifies four kinds (types) of gene rearrangements (Fig. 3a). The typical gene order of the trnWtrnY block is trnW, trnA, trnN, OL, trnC, and trnY (WANCY). Unlike the other types, where the OL is located before trnN (Fig. 3a), in Type I, the OL occurs after trnN, separated by an intergenic spacer (IGS or noncoding region). In Type II, two trnA occur with an IGS located between trnA1 and OL, another IGS occurs between OL and trnA2, and another IGS between trnN and trnC. The gene organization of Type III and Type IV are similar to Type II, but Type III lacks trnA1 and Type IV lacks trnA2, respectively. Except for the reorganizations of trnA, trnN, and OL, all other tRNAs and protein-coding genes have positions and lengths typical of the vertebrate mitogenome (Fig. 3a; Additional file 2: Table S2).

Fig. 3
figure 3

Diversity of intraspecific mitochondrial gene rearrangements in Quasipaa boulengeri. a Four types of gene rearrangement in the “WANCY” region. b Types of evolution and putative mechanism of gene rearrangement of the mitochondrial sequences according to the tandem duplication–random loss model (TDRL). TDRL first produces Types I and II. Type II is the intermediate state with two trnA genes. Types III and IV result from the random loss of one alternative trnA. “---”, pseudogenes or noncoding sequence. c Phylogenetic relationships and divergence times of four mitochondrial gene rearrangements in Quasipaa boulengeri. Tree topology derived from BI analyses of cox1 and cob is consistent with an ML tree. Numbers above the lines or beside the nodes are inferred divergence times (Ma) and Bayesian posterior probabilities, respectively. Types II, III, and IV form a clade, and each Type forms its own clade, except for Type III

Sequential process of TDRL

All four types of gene rearrangement in Q. boulengeri involve trnA, trnN, OL and trnC. The IGSs reveal tandem duplications in the WANCY region. These residues identify pseudogenes of trnA, trnN and trnC, whose sequences are similar to corresponding tRNAs. Additional file 3: Figure S1 shows the primary sequence of trnA and trnN for each variant. The secondary structures of these genes fold into typical stem-and-loop structures (Additional file 4: Figure S2). Each type of variant has only one trnA and trnN, except in Type II, which has two trnAs, both of which form stable clover-leaf structures. Thus, these genes are paralogs created by gene duplication. Residues are very similar to trnA and trnN, but have a loss of function owing to secondary structures or a mutation on the anticodon position.

TDRL [5, 16] best explains the gene rearrangement in Q. boulengeri, and the diversity of rearrangements rejects the alternative hypotheses of non-random loss (TDNL and DMNL), recombination, and DRRL. Genes of the same polarity do not cluster together, and the finding of alternative loss of duplicated genes, as seen in comparison of Types III and IV, contradicts non-random loss models (TDNL or DMNL). The absence of different tandem duplication junction points and no variation in the number of tandem repeats (VNTR) and this does not support the recombination model. We cannot reject the hypotheses that unequal crossing over of intermolecular recombinations were subtly inserted in front of trnA and behind trnC, but tandem duplication would essentially be a consequence of this recombination. Further, no concerted evolution rejects intramolecular recombination, because the two copies of trnA in type II differ (Additional file 3: Figure S1, and Additional file 4: Figure S2). Finally, analyses reject the hypothesis of DRRL due to the absence of two control regions in the mitogenome of Q. boulengeri [26].

The TDRL hypothesis remains the only viable explanation and our results conform to its predictions (Fig. 3b). The hypothesized duplicated region in the mitogenome of Q. boulengeri includes trnA, trnN, OL, and a partial fragment of trnC. Slipped-strand mispairing, imprecise termination or recombination have been proposed to explain mitogenomic duplications [5, 7, 15]. Regardless of the molecular mechanism, tandem duplication mutations will yield two copies each of trnA, trnN, OL, and trnC. Subsequent random loss appears to have occurred at least twice independently in the mitogenome of Q. boulengeri. First, rearrangement Type I involves the loss of OL1, trnC1, trnA2, and trnN2. Second, loss involves trnN1, trnC1, and OL2 in Type II. The retention of two copies of trnA in Type II is direct evidence for the randomness of loss because alternative losses occur in Type III and Type IV, which have the same gene order as Type II (trnA1 has been lost in Type III as compared to trnA2 in Type IV). Rather than a result of selection for one or other alternative, loss of one copy of trnA appears to have occurred by chance alone.

The sequencing cox1 and cob for 290 individuals (Additional file 1: Table S1) identifies the origin of the initial tandem duplication event and the stepwise process of random loss when viewed in terms of the phylogenetic relationships of all types of gene rearrangement. The concatenated alignment contains 1463 nucleotide positions (cox1: 626 bp; cob: 837 bp) without stop codons. Maximum likelihood (ML) and Bayesian inference (BI) reconstructions obtain similar tree topologies for the four types of rearrangement (Fig. 3c). All haplotypes cluster together by type and with moderate to strong levels of nodal support, except for Type III, which is paraphyletic. Analyses recover the group Type II + Type III + Type IV, and roots it as the sister-group of Type I. Type II forms the sister-group of Type III + Type IV and some samples of Type III unite with Type IV.

The phylogenetic analyses and gene-order strongly indicate a single tandem duplication event and stepwise random loss in Q. boulengeri. The clustering of Types II–IV suggests that they share the same primary TDRL process (Fig. 3b, c), and that Type I represents an independent random loss. The monophyly of types I, II and IV indicate a single origin of each rearrangement type. Paraphyly of Type III suggests two parallel random losses are responsible for the same gene rearrangement. The associations of Type IIIa, IIIb and IV indicate that they shared a recent common ancestor, but independently lost one duplicated copy of trnA. The possession of both copies of trnA indicates that Type II represents the intermediate state.

Although mitochondrial gene rearrangements are not uncommon among related taxa, recognized intermediate steps of gene-order rearrangements are rare, and their presence can suggest evolutionary mechanisms [7, 27]. Most intermediate states appear as pseudogenes or residues of tandemly duplicated sequences rather than as two functional gene copies [3, 8, 16]. The intermediate state of Type II, leading either to Type III or Type IV, is an example of the random loss of one trnA gene.

Alternative loss-types have been observed among closely related species, e.g. alternative losses of trnH in the anuran Babina [17, 28]. However, alternative losses have been reported rarely within a single species and even more so within a population. The finding of alternate losses in Q. boulengeri is the first observation of an intermediate state involving two functional gene copies, and, simultaneously, the loss of alternative types in a vertebrate mt genome.

The occurrence of TDRL in Q. boulengeri corresponds to the view that reorganization of the mitochondrial genome is nonadaptive [17, 29]. Our results indicate that the hotspot of gene rearrangement is adjacent to the origin of light-strand replication [6]. Homoplastic mitochondrial rearrangements are contiguous in the genome or they locate around the origin of replication [6, 30, 31]. Under these conditions, mitochondrial gene-orders appear susceptible to convergent or parallel evolution because of functional constraints or selective pressures [3234]. However, such evolution does not occur in Q. boulengeri. The gene-order and phylogenetic analyses indicate a single tandem duplication event followed by independent losses (Fig. 3), which implies that random loss and gene-order are not involved in adaptive evolution [17].

Evolution of mitogenomic rearrangement

A time-calibrated phylogeny constructed using Bayesian inference estimates the recency of mitogenomic rearrangements in Q. boulengeri (Fig. 3c). The initial diversification (Type I) dates to about 0.8 Ma and divergence among the other three types ranges from 0.4 to 0.6 Ma, suggesting that the duplication and fixation of these rearrangements can occur quite quickly. Combing interspecific rearrangement data, we summarize the rates of mitogenomic duplication and loss (Additional file 5: Table S3). This suggests that post duplication, the alternative loss types can occur in 0.2–5 Ma.

To explore how many intraspecific rearrangements existed, we gather both in silico and experimental evidence to detect the gene order in highly rearranged groups. First, our in silico approach obtains mitochondrial gene-order information (mtGordV0.5.pl;  Additional file 6: Software) from GenBank in Amphibia and Squamata, two groups with high diversities in mitogenome rearrangement. Analyses discover that intraspecific rearrangements are rare, but they exist. Two cases occur in Squamata, one in asexual squamates with multiple origins of duplication [7], and another in an amphisbaenian with alternative loss-types varying among populations [19]. No intraspecific rearrangements in Amphibia exist in data from GenBank.

Our sequences of the WANCY fragment in multiple populations of the frogs Odorrana schmackeri and Amolops mantzorum detected that this hotspot region differed from the typical vertebrate arrangement [17]. These species do not exhibit intraspecific variation in gene-order, yet evidence for variation in losses may be represented by the residues of pseudogenes (data not shown).

Intraspecific studies may provide new insights into the high incidence of rearranged mitochondrial genomes. Above the species level, rates of mitogenomic partial duplication have been found to be high, and multiple duplication events can facilitate gene-rearrangement [7, 16, 21]. However, duplications may not occur frequently within a species. The rarity of this may reflect selection or functional constraints that prevent fixation, and may shed light on the limits of intraspecific diversity of mitogenomes. It could also owe to the dearth of investigations of intraspecific mitogenomic reorganization. We predict that mitochondrial metagenomic skimming by next-generation sequencing [35, 36] will detect additional cases of intraspecific rearrangements.

Random loss within duplicated regions could occur repeatedly, and the rate of duplication excision may be relatively high. At the intraspecific level, random loss occurred independently many times both in Q. boulengeri and the lizard Bipes biporus (Fig. 3c, Additional file 5: Table S3). At the interspecific level, the sibling frog genera Babina and Odorrana share the same duplication of genes, but the pathways of deletion differed among species [17, 28].

The loss of a duplicated region and fixation could occur in short evolutionary time (0.2 Ma, Additional file 5: Table S3). Deletion of a redundant gene-copy may happen rapidly due to functional constraints and the compactness of the metazoan mitogenome, facilitating the formation of pseudogenes or the complete deletion of redundant genes [27, 37]. A functionally redundant duplicate gene copy may not persist long in a population because deleterious mutations can accumulate and cause the redundant gene to become nonfunctional [38].

Nonadaptive forces, such as genetic drift or bottlenecking, may drive the fixation and dispersal of mitogenomic reorganizations. The low effective population size of the mitogenome leaves it vulnerable to bottlenecks and genetic drift, which can drive the fixation of large-scale genomic modifications [16, 3941]. Quasipaa boulengeri resides in localized montane areas, mainly in rocky streams [42]. Its highly specific habitat may limit gene flow and result in a pattern structured by genetic drift. The upper and midstream tributaries of the Yangtze River, including some areas in Chongqing, Guizhou and Hubei, have populations containing two or more sympatric types of rearrangements (Fig. 2). This area may be the original source of the gene rearrangements, or may represent areas of secondary contact. Both scenarios explain the distribution of types. Historical demographic analyses in this area point to dispersal events for Q. boulengeri [42]. TDRL may have first occurred in this area, followed by dispersal to other places. However, the single origin of each type suggests independent fixations of alternative arrangements, in which case secondary contact could also explain the pattern.

Conclusion

The initial stages of gene rearrangement are difficult to observe in nature, which limits our understanding of the evolutionary mechanism. Intraspecific or population level investigations may represent intermediate states and fixation of initial rearrangement, and if so they potentially shed light on the evolutionary dynamics. Here, we found mitogenomic rearrangements diversity in a single frog species, Quasipaa boulengeri. Intermediate state and alternative losses types were observed in this frog, which provide direct evidence of tandem duplication and random loss model for mitochondrial gene rearrangement. The intraspecific variation in Q. boulengeri provides insights into the rate of partial duplications and deletions within a mitogenome, and reveals that fixation and gene-distribution in mitogenomic reorganization is likely non-adaptive. Our observation may shed light on the investigations of intraspecific mitogenomic reorganization.

Methods

Samples and Sequence Amplification

A total of 290 samples from 28 localities were used. Frogs were collected from 2006 to 2013, and Fig. 1 and Additional file 1: Table S1 detail the localities. Tissue samples, including liver, muscle, and tadpoles were stored in 95% ethanol at −20 °C in the Chengdu Institute of Biology, Chinese Academy of Sciences (CIB). The CIB Animal Care and Use Committee approved all procedures.

We sequenced the hotspot of gene rearrangement, from nad2 to cox1, of mtDNA for Quasipaa boulengeri, which included the WANCY region [26]. Two other fragments, cox1 and cob, were sequenced for population genetic and phylogenetic analysis. For cox1, we added published sequences (GenBank No. JX629572–JX629667) for phylogenetic analysis [43]. The sample and sequence information were provided in Additional file 1: Table S1. The PCR primers for the three fragments were those of Kurabayashi and Sumida [44] and Qing et al. [43]. To avoid Numt (nuclear copies of mtDNA genes), we designed a pair of primers to confirm amplification of the WANCY region: 5059 F-3, 5’- TTCTTTTACTTACGACTGACAT -3’; 6399R-2, 5’- ATGCCTGCGGCTAAAACTGGAAGAG-3’. PCR amplification, sub-cloning, and sequencing followed Xia et al. [17]. All newly obtained sequences were examined by checking for the presence of premature stop codons (pseudogenes).

Gene Annotation and Time-Tree Analyses

The tRNA genes were identified by using both tRNAscan-SE v.1.21 (http://lowelab.ucsc.edu/tRNAscan-SE) and MITOS (http://mitos.bioinf.uni-leipzig.de). To avoid misannotated tRNA genes, we predicted the secondary structure for each. We extracted and aligned the duplicated tRNA genes and their pseudogene residues.

We aligned sequences of each fragments using ClustalW in MEGA6 [45]. DnaSP v.5.10 [46] was used to determine DNA polymorphisms and divergences. To estimate the time-tree, we constructed phylogenies using cox1 and cob, and partitioned these genes by codon position. Six species of Quasipaa, including Q. verrucospinosa (KF199147), Q. shini (KF199148), Q. yei (KJ842105), Q. spinosa (FJ432700), Q. jiulongensis (KF199149) and Q. exilispinosa (KF199151), were chosen as outgroup taxa. The best-fit substitution model for each partition was estimated using the Akaike information criterion (AIC) implemented in PartitionFinder v1.1.1 [47]. The best model of each partition was chosen for maximum likelihood (ML) and Bayesian inference (BI) analyses, which were performed with RAxML BlackBox web-servers (http://phylobench.vital-it.ch/raxml-bb/index.php) [48] and MrBayes v.3.1 [49], respectively.

BI as implemented in BEAST2 v.2.1.2 [50] was used to obtain an ultrametric time-tree for Q. boulengeri. Each locus was assigned its own partition with unlinked substitution model but with linked clock and tree models. We assumed a substitution rate ranging from 0.65 to 1.00% per Ma for the cox1 and cob based on evolutionary rates commonly proposed for frogs [42, 51, 52]. Lacking fossil evidence, we calibrated our phylogeny using the published divergence time to the most recent common ancestor (TMRCA) between the Q. jiulongensis and Q. exilispinosa of about 9 Ma [53]. We ran BEAST for 20 million generations while logging trees every 1000 generations for a total of 20,000 trees. We determined a 10% burn-in length using Tracer v.1.5 and retained the maximum clade credibility tree using TreeAnnotator v.2.1.2.

A Perl script named mtGordV0.5.pl was written by YZ to obtain the gene-orders of mitochondrial records deposited in GenBank, based on the annotation of the sequence. Records were downloaded together as a single file, which was used as the input file of the script. For each record with more than one gene, items in the order of accession number, sequence length, species name, gene names in their original order, and total number of genes were saved in an individual line to the output file. Items were separated from each other by a tab. The script was applied to two major groups of vertebrates, amphibians and squamate reptiles. For amphibians, all 126,638 mitochondrial records were downloaded on 02 Nov 2015, and the output file contained 17,559 records. For squamates, all 110,064 records were downloaded on 25 Sep 2015, and the output file contained 21,045 records. The output files were opened using Microsoft Excel and records were aligned according to species names. The records were manually checked for intraspecific and intrageneric cases of random loss of genes after duplication. As the script did not include all variation of annotations for all mitochondrial genes, errors from missing genes were expected for a small number of records. However, when a potential case was detected, the related GenBank full records were carefully checked. More importantly, this script made such a scan possible, analyses could be conducted within a reasonable amount of time, a few days for each group in our case, and it could be applied to other groups of taxa. Regarding the speed of the script itself, the data for squamates were processed within 3 min on a ThinkPad X200 laptop computer.

Abbreviations

BI:

Bayesian inference

cob :

Cytochrome b

cox1 :

Cytochrome c oxidase subunit I

CR:

Control region

DMNL:

Dimer-mitogenome and non-random loss

DRRL:

Double replications and random loss

IGS:

Intergenic spacer

ML:

Maximum likelihood

nad2 :

NADH dehydrogenase subunit 2

OL :

Origin of light strand replication

TDNL:

Tandem duplication and non-random loss

TDRL:

Tandem duplication and random loss

WANCY:

trnW, trnA, trnN, OL, trnC, and trnY

References

  1. Boore JL. Animal mitochondrial genomes. Nucleic Acids Res. 1999;27:1767–80.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  2. Dowton M, Cameron SL, Dowavic JI, Austin AD, Whiting MF. Characterization of 67 mitochondrial tRNA gene rearrangements in the hymenoptera suggests that mitochondrial tRNA gene position is selectively neutral. Mol Biol Evol. 2009;26:1607–17. doi:10.1093/molbev/msp072.

    Article  CAS  PubMed  Google Scholar 

  3. Macey JR, Schulte JA, Larson A, Papenfuss TJ. Tandem duplication via light-strand synthesis may provide a precursor for mitochondrial genomic rearrangement. Mol Biol Evol. 1998;15:71–5.

    Article  CAS  PubMed  Google Scholar 

  4. Shao R, Barker SC, Mitani H, Takahashi M, Fukunaga M. Molecular mechanisms for the variation of mitochondrial gene content and gene arrangement among chigger mites of the genus Leptotrombidium (Acari: Acariformes). J Mol Evol. 2006;63:251–61. doi:10.1007/s00239-005-0196-y.

    Article  CAS  PubMed  Google Scholar 

  5. Boore JL. The duplication/random loss model for gene rearrangement exemplified by mitochondrial genomes of deuterostome animals. In: Sankoff D, Nadeau J, editors. Comparative Genomics. Dordrecht: Kluwer Academic Publishers; 2000. p. 133–47.

  6. San Mauro D, Gower DJ, Zardoya R, Wilkinson M. A hotspot of gene order rearrangement by tandem duplication and random loss in the vertebrate mitochondrial genome. Mol Biol Evol. 2006;23:227–34. doi:10.1093/molbev/msj025.

    Article  CAS  PubMed  Google Scholar 

  7. Fujita MK, Boore JL, Moritz C. Multiple origins and rapid evolution of duplicated mitochondrial genes in parthenogenetic geckos (Heteronotia binoei; squamata, gekkonidae). Mol Biol Evol. 2007;24:2775–86. doi:10.1093/molbev/msm212.

    Article  CAS  PubMed  Google Scholar 

  8. Shi W, Gong L, Wang SY, Miao XG, Kong XY. Tandem Duplication and Random Loss for mitogenome rearrangement in Symphurus (Teleost: Pleuronectiformes). BMC Genomics. 2015;16:355.

    Article  PubMed  PubMed Central  Google Scholar 

  9. Lavrov DV, Boore JL, Brown WM. Complete mtDNA sequences of two millipedes suggest a new model for mitochondrial gene rearrangements: duplication and nonrandom loss. Mol Biol Evol. 2002;19:163–9.

    Article  CAS  PubMed  Google Scholar 

  10. Shi W, Dong XL, Wang ZM, Miao XG, Wang SY, Kong XY. Complete mitogenome sequences of four flatfishes (Pleuronectiformes) reveal a novel gene arrangement of L-strand coding genes. BMC Evol Biol. 2013;13:173.

    Article  PubMed  PubMed Central  Google Scholar 

  11. Dowton M, Campbell NJH. Intramitochondrial recombination–is it why some mitochondrial genes sleep around? Trends Ecol Evol. 2001;16:269–71. doi:10.1016/s0169-5347(01)02182-6.

    Article  PubMed  Google Scholar 

  12. Kurabayashi A, Sumida M, Yonekawa H, Glaw F, Vences M, Hasegawa M. Phylogeny, recombination, and mechanisms of stepwise mitochondrial genome reorganization in mantellid frogs from Madagascar. Mol Biol Evol. 2008;25:874–91. doi:10.1093/molbev/msn031.

    Article  CAS  PubMed  Google Scholar 

  13. Shi W, Miao XG, Kong XY. A novel model of double replications and random loss accounts for rearrangements in the Mitogenome of Samariscus latus (Teleostei: Pleuronectiformes). BMC Genomics. 2014;15:352.

    Article  PubMed  PubMed Central  Google Scholar 

  14. Moritz C, Brown W. Tandem duplications in animal mitochondrial DNAs: variation in incidence and gene content among lizards. Proc Natl Acad Sci U S A. 1987;84:7183–7.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  15. Macey JR, Larson A, Ananjeva NB, Fang ZL, Papenfuss TJ. Two novel gene orders and the role of light-strand replication in rearrangement of the vertebrate mitochondrial genome. Mol Biol Evol. 1997;14:91–104.

    Article  CAS  PubMed  Google Scholar 

  16. Mueller RL, Boore JL. Molecular mechanisms of extensive mitochondrial gene rearrangement in plethodontid salamanders. Mol Biol Evol. 2005;22:2104–12. doi:10.1093/molbev/msi204.

    Article  CAS  PubMed  Google Scholar 

  17. Xia Y, Zheng YC, Miura I, Wong PB, Murphy RW, Zeng XM. The evolution of mitochondrial genomes in modern frogs (Neobatrachia): nonadaptive evolution of mitochondrial genome reorganization. BMC Genomics. 2014;15:691.

    Article  PubMed  PubMed Central  Google Scholar 

  18. Moritz C. Evolutionary dynamics of mitochondrial DNA duplications in parthenogenetic geckos, Heteronotia binoei. Genetics. 1991;129:221–30.

    CAS  PubMed  PubMed Central  Google Scholar 

  19. Macey JR, Papenfuss TJ, Kuehl JV, Fourcade HM, Boore JL. Phylogenetic relationships among amphisbaenian reptiles based on complete mitochondrial genomic sequences. Mol Phylogenet Evol. 2004;33:22–31. doi:10.1016/j.ympev.2004.05.003.

    Article  CAS  PubMed  Google Scholar 

  20. Sammler S, Ketmaier V, Havenstein K, Tiedemann R. Intraspecific Rearrangement of Duplicated Mitochondrial Control Regions in the Luzon Tarictic Hornbill Penelopides manillae (Aves: Bucerotidae). J Mol Evol. 2013;77:199–205. doi:10.1007/s00239-013-9591-y.

    Article  CAS  PubMed  Google Scholar 

  21. Kurabayashi A, Yoshikawa N, Sato N, Hayashi Y, Oumi S, Fujii T, et al. Complete mitochondrial DNA sequence of the endangered frog Odorrana ishikawae (family Ranidae) and unexpected diversity of mt gene arrangements in ranids. Mol Phylogenet Evol. 2010;56:543–53. doi:10.1016/j.ympev.2010.01.022.

    Article  CAS  PubMed  Google Scholar 

  22. Kurabayashi A, Sumida M. Afrobatrachian mitochondrial genomes: genome reorganization, gene rearrangement mechanisms, and evolutionary trends of duplicated and rearranged genes. BMC Genomics. 2013;14:633.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  23. Zhang P, Liang D, Mao RL, Hillis DM, Wake DB, Cannatella DC. Efficient sequencing of Anuran mtDNAs and a mitogenomic exploration of the phylogeny and evolution of frogs. Mol Biol Evol. 2013;30:1899–915. doi:10.1093/molbev/mst091.

    Article  CAS  PubMed  Google Scholar 

  24. Fonseca MM, Harris DJ. Relationship between mitochondrial gene rearrangements and stability of the origin of light strand replication. Genet Mol Biol. 2008;31:566–74.

    Article  CAS  Google Scholar 

  25. Irisarri I, Mauro DS, Abascal F, Ohler A, Vences M, Zardoya R. The origin of modern frogs (Neobatrachia) was accompanied by acceleration in mitochondrial and nuclear substitution rates. BMC Genomics. 2012;13:626.

    Article  PubMed  PubMed Central  Google Scholar 

  26. Shan X, Xia Y, Zheng YC, Zou FD, Zeng XM. The complete mitochondrial genome of Quasipaa boulengeri (Anura: Dicroglossidae). Mitochondr DNA. 2014;25:83–4. doi:10.3109/19401736.2013.782023.

    Article  CAS  Google Scholar 

  27. Kumazawa Y, Miura S, Yamada C, Hashiguchi Y. Gene rearrangements in gekkonid mitochondrial genomes with shuffling, loss, and reassignment of tRNA genes. BMC Genomics. 2014;15:930.

    Article  PubMed  PubMed Central  Google Scholar 

  28. Kakehashi R, Kurabayashi A, Oumi S, Katsuren S, Hoso M, Sumida M. Mitochondrial genomes of Japanese Babina frogs (Ranidae, Anura): unique gene arrangements and the phylogenetic position of genus Babina. Genes Genet Syst. 2013;88:59–67.

    Article  CAS  PubMed  Google Scholar 

  29. Boussau B, Brown JM, Fujita MK. Nonadaptive evolution of mitochondrial genome size. Evolution. 2011;65:2706–11.

    Article  PubMed  Google Scholar 

  30. Boore JL. The use of genome-level characters for phylogenetic reconstruction. Trends Ecol Evol. 2006;21:439–46. doi:10.1016/j.tree.2006.05.009.

    Article  PubMed  Google Scholar 

  31. Babbucci M, Basso A, Scupola A, Patarnello T, Negrisolo E. Is It an Ant or a Butterfly? Convergent Evolution in the Mitochondrial Gene Order of Hymenoptera and Lepidoptera. Genome Biol Evol. 2014;6:3326–43. doi:10.1093/gbe/evu265.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  32. Ojala D, Montoya J, Attardi G. tRNA punctuation model of RNA processing in human mitochondria. Nature. 1981;290:470–4.

    Article  CAS  PubMed  Google Scholar 

  33. Mindell DP, Sorenson MD, Dimcheff DE. Multiple independent origins of mitochondrial gene order in birds. Proc Natl Acad Sci U S A. 1998;95:10693–7.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  34. Satoh T, Sato Y, Masuyama N, Miya M, Nishida M. Transfer RNA gene arrangement and codon usage in vertebrate mitochondrial genomes: a new insight into gene order conservation. BMC Genomics. 2010;11:479.

    Article  PubMed  PubMed Central  Google Scholar 

  35. Dettai A, Hinsinger DD, Utage J, Debruyne R, Thomas M, Denys GP, et al. Fishing for barcodes in the Torrent: from COI to complete mitogenomes on NGS platforms. DNA Barcodes. 2015;3:170–86.

    Google Scholar 

  36. Crampton-Platt A, Timmermans MJTN, Gimmel ML, Kutty SN, Cockerill TD, Vun Khen C, et al. Soup to Tree: The Phylogeny of Beetles Inferred by Mitochondrial Metagenomics of a Bornean Rainforest Sample. Mol Biol Evol. 2015;32:2302–16. doi:10.1093/molbev/msv111.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  37. Selosse M-A, Albert B, Godelle B. Reducing the genome size of organelles favours gene transfer to the nucleus. Trends Ecol Evol. 2001;16:135–41. doi:10.1016/S0169-5347(00)02084-X.

    Article  PubMed  Google Scholar 

  38. Harris EE. Nonadaptive processes in primate and human evolution. Am J Phys Anthropol. 2010;143:13–45. doi:10.1002/ajpa.21439.

    Article  PubMed  Google Scholar 

  39. Lynch M, Koskella B, Schaack S. Mutation pressure and the evolution of organelle genomic architecture. Science. 2006;311:1727–30. doi:10.1126/science.1118884.

    Article  CAS  PubMed  Google Scholar 

  40. Oliveira D, Raychoudhury R, Lavrov DV, Werren JH. Rapidly evolving mitochondrial genome and directional selection in mitochondrial genes in the parasitic wasp Nasonia (Hymenoptera : Pteromalidae). Mol Biol Evol. 2008;25:2167–80. doi:10.1093/molbev/msn159.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  41. Chong RA, Mueller RL. Low metabolic rates in salamanders are correlated with weak selective constraints on mitochondrial genes. Evolution. 2013;67:894–9. doi:10.1111/j.1558-5646.2012.01830.x.

    Article  CAS  PubMed  Google Scholar 

  42. Yan F, Zhou WW, Zhao HT, Yuan ZY, Wang YY, Jiang K, et al. Geological events play a larger role than Pleistocene climatic fluctuations in driving the genetic structure of Quasipaa boulengeri (Anura: Dicroglossidae). Mol Ecol. 2013;22:1120–33. doi:10.1111/mec.12153.

    Article  PubMed  Google Scholar 

  43. Qing LY, Xia Y, Zheng YC, Zeng XM. A De Novo Case of Floating Chromosomal Polymorphisms by Translocation in Quasipaa boulengeri (Anura, Dicroglossidae). PLoS One. 2012;7:e46163. doi:10.1371/journal.pone.0046163.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  44. Kurabayashi A, Sumida M. PCR primers for the neobatrachian mitochondrial genome. Current Herpetology. 2009;28:1–11. doi:10.3105/018.028.0101.

    Article  Google Scholar 

  45. Tamura K, Stecher G, Peterson D, Filipski A, Kumar S. MEGA6: Molecular Evolutionary Genetics Analysis Version 6.0. Mol Biol Evol. 2013;30:2725–9. doi:10.1093/molbev/mst197.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  46. Librado P, Rozas J. DnaSP v5: a software for comprehensive analysis of DNA polymorphism data. Bioinformatics. 2009;25:1451–2. doi:10.1093/bioinformatics/btp187.

    Article  CAS  PubMed  Google Scholar 

  47. Lanfear R, Calcott B, Ho SYW, Guindon S. PartitionFinder: Combined Selection of Partitioning Schemes and Substitution Models for Phylogenetic Analyses. Mol Biol Evol. 2012;29:1695–701. doi:10.1093/molbev/mss020.

    Article  CAS  PubMed  Google Scholar 

  48. Stamatakis A, Hoover P, Rougemont J. A rapid bootstrap algorithm for the RAxML web servers. Syst Biol. 2008;57:758–71. doi:10.1080/10635150802429642.

    Article  PubMed  Google Scholar 

  49. Huelsenbeck JP, Ronquist F. MRBAYES: Bayesian inference of phylogenetic trees. Bioinformatics. 2001;17:754–5.

    Article  CAS  PubMed  Google Scholar 

  50. Bouckaert R, Heled J, Kühnert D, Vaughan T, Wu CH, Xie D, et al. BEAST 2: A Software Platform for Bayesian Evolutionary Analysis. PLoS Comput Biol. 2014;10:e1003537. doi:10.1371/journal.pcbi.1003537.

    Article  PubMed  PubMed Central  Google Scholar 

  51. Macey JR, Strasburg JL, Brisson JA, Vredenburg VT, Jennings M, Larson A. Molecular Phylogenetics of Western North American Frogs of the Rana boylii Species Group. Mol Phylogenet Evol. 2001;19:131–43. doi:10.1006/mpev.2000.0908.

    Article  CAS  PubMed  Google Scholar 

  52. Pröhl H, Ron S, Ryan M. Ecological and genetic divergence between two lineages of Middle American tungara frogs Physalaemus (= Engystomops) pustulosus. BMC Evol Biol. 2010;10:146.

    Article  PubMed  PubMed Central  Google Scholar 

  53. Che J, Zhou WW, Hu JS, Yan F, Papenfuss TJ, Wake DB, et al. Spiny frogs (Paini) illuminate the history of the Himalayan region and Southeast Asia. Proc Natl Acad Sci U S A. 2010;107:13765–70.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

Download references

Acknowledgments

We thank Liyan Qing, Shujun Zhang, Xiang Shan for collecting the samples. We gratefully acknowledge technical support from the laboratory of XZ.

Funding

This research was supported by the National Natural Sciences Foundation of China (NSFC–31272282, 31372181, 31401960, 31572243), by the China Postdoctoral Science Foundation (2014 M552386) and West Light Foundation of The Chinese Academy of Sciences.

Availability of data and materials

The nucleotide sequence dataset for this research article is available in the GenBank with accession numbers shown in Additional file 1: Table S1 (Accession numbers: KX571396 - KX571899).

Authors’ contributions

YX carried out molecular lab work and conceived the project and analyzed the data. YZ wrote the Perl script. YX, RWM, and XZ wrote the paper. XZ and RWM finalized the manuscript. All authors have read and approved the final manuscript.

Competing interests

The authors declare that they have no competing interests.

Consent for publication

Not applicable.

Ethics approval and consent to participate

The tissue samples collection and DNA extraction were approved under the Chengdu Institute of Biology (CIB) Animal Care and Use Committee.

Author information

Authors and Affiliations

Authors

Corresponding authors

Correspondence to Yuchi Zheng or Xiaomao Zeng.

Additional files

Additional file 1: Table S1.

Summary of sample localities and mitogenome regions sequenced for Quasipaa boulengeri. (XLS 84 kb)

Additional file 2: Table S2.

The gene organization of “WANCY” region for each sample. (XLS 77 kb)

Additional file 3: Table S3.

Time-scale of mitogenomic duplication and radom loss. (PDF 209 kb)

Additional file 4: Figure S1.

The primary sequence of trnA and trnN for each variant. (PDF 161 kb)

Additional file 5: Figure S2.

The secondary structures of of trnA and trnN fold into typical stem-and-loop structures. (DOCX 22 kb)

Additional file 6: Software.

The Perl script mtGordV0.5.pl with its readme file and example input and output files. (ZIP 5.35 kb)

Rights and permissions

Open Access This article is distributed under the terms of the Creative Commons Attribution 4.0 International License (http://creativecommons.org/licenses/by/4.0/), which permits unrestricted use, distribution, and reproduction in any medium, provided you give appropriate credit to the original author(s) and the source, provide a link to the Creative Commons license, and indicate if changes were made. The Creative Commons Public Domain Dedication waiver (http://creativecommons.org/publicdomain/zero/1.0/) applies to the data made available in this article, unless otherwise stated.

Reprints and permissions

About this article

Check for updates. Verify currency and authenticity via CrossMark

Cite this article

Xia, Y., Zheng, Y., Murphy, R.W. et al. Intraspecific rearrangement of mitochondrial genome suggests the prevalence of the tandem duplication-random loss (TDLR) mechanism in Quasipaa boulengeri . BMC Genomics 17, 965 (2016). https://doi.org/10.1186/s12864-016-3309-7

Download citation

  • Received:

  • Accepted:

  • Published:

  • DOI: https://doi.org/10.1186/s12864-016-3309-7

Keywords