Next Article in Journal
Management of Thermal Injuries in Donkeys: A Case Report
Previous Article in Journal
Effects of Dog-Based Animal-Assisted Interventions in Prison Population: A Systematic Review
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Complete Mitochondrial Genome of Three Species of the Genus Microtus (Arvicolinae, Rodentia)

1
Departamento de Biología Experimental, Área de Genética, Universidad de Jaén, Paraje de las Lagunillas s/n, 23071 Jaén, Spain
2
Department of Ecology, Faculty of Science, Charles University, 12844 Prague, Czech Republic
*
Author to whom correspondence should be addressed.
Animals 2020, 10(11), 2130; https://doi.org/10.3390/ani10112130
Submission received: 27 October 2020 / Revised: 9 November 2020 / Accepted: 14 November 2020 / Published: 16 November 2020
(This article belongs to the Section Animal Genetics and Genomics)

Abstract

:

Simple Summary

The knowledge and analysis of mitochondrial genomes can be useful to answer some interesting questions about the biology of some species. This is the case of vole species that are characterized by unusual genetic features related to sex chromosomes and variation of their karyotypes. In this work, we describe the mitochondrial genomes of three emblematic vole species, demonstrating that they are highly conserved in the organization of their protein coding, transfer RNA and ribosomal RNA genes. In addition, we performed a detailed analysis of their control regions identifying several domains and conserved boxes related to mitochondrial DNA regulation. Finally, a phylogenetic analysis with the mitochondrial DNA confirmed the established phylogeny of the analyzed voles.

Abstract

The 65 species of the genus Microtus have unusual sex-related genetic features and a high rate of karyotype variation. However, only nine complete mitogenomes for these species are currently available. We describe the complete mitogenome sequences of three Microtus, which vary in length from 16,295 bp to 16,331 bp, contain 13 protein-coding genes (PCGs), two ribosomal RNA genes, 22 transfer RNA genes and a control region. The length of the 13 PCGs and the coded proteins is the same in all three species, and the start and stop codons are conserved. The non-coding regions include the L-strand origin of replication, with the same sequence of 35 bp, and the control region, which varies between 896 bp and 930 bp in length. The control region includes three domains (Domains I, II and III) with extended termination-associated sequences (ETAS-1 and ETAS-2) in Domain I. Domain II and Domain III include five (CSB-B, C, D, E and F) and three (CSB-1, CSB-2, and CSB-3) conserved sequence blocks, respectively. Phylogenetic reconstructions using the mitochondrial genomes of all the available Microtus species and one representative species from another genus of the Arvicolinae subfamily reproduced the established phylogenetic relationships for all the Arvicolinae genera that were analyzed.

1. Introduction

The rodent subfamily Arvicolinae (lemmings, muskrats and voles) includes numerous species that have been classified into 21 genera [1], of which the genus Microtus is the most species-rich (65 extant species). This genus underwent fast recent radiation about 1.2–2 million years ago (MYA) [2] in what was one of the most explosive speciation events ever recorded in vertebrates [1].
Some species from the genus Microtus are characterized by a series of unusual sex-related genetic features [3,4,5,6,7,8,9,10]. Moreover, the genus Microtus shows one of the highest rates of karyotype variation of all mammals [11], with highly variable chromosome numbers ranging from 2n = 18 in M. oregoni [12] to 2n = 62 in M. duodecimcostatus and M. lusitanicus [13]. Despite this variability in their karyotypes, the majority of Microtus species have similar external morphologies and therefore are hard to distinguish using traditional taxonomic characteristics [14,15,16,17]. In addition, the genus Microtus underwent a recent rapid radiation resulting in the emergence of numerous cryptic species with unclear phylogenetic relationships [18,19,20,21]. Although some of these cryptic species can be distinguished using mitochondrial and/or nuclear markers [22,23], Microtus phylogeny is not clearly established and many systematic relationships within this genus are still unresolved [20,24].
Phylogenetic reconstructions have been carried out using karyotype banding, chromosome painting and repeated DNA sequence analysis [3,8,25,26]. Despite lacking the reliability to solve intra-specific relationships [27], mitochondrial markers COI and NADH are nevertheless regarded as very precise tools for deciphering interspecific relationships between Arvicolinae species [16,17,24,28]. Several nuclear genetic markers such as the growth hormone receptor (GRH) including exon 10 [27] have been employed to solve intra-specific relationships in the genus Microtus. An interesting phylogenomic approach employing partial mitogenomes and genotyping-by-sequencing single nucleotide polymorphism (SNP) data from seven Microtus species has also been recently described [29].
Traditionally, complete mitogenomes have been sequenced by PCR amplification of overlapping fragments [30,31]. However, next-generation sequencing (NGS) now allows us to assemble much more easily the whole mitochondrial genome due to its high-copy portion in relation to the nuclear DNA [32,33,34,35,36,37]. Nevertheless, although complete mitochondrial genomes are very useful for phylogenetic reconstructions, the number of species with fully described mitogenomes is still low [38,39]. Only eight complete mitogenomes are currently available for Microtus species, and therefore, in light of the fast radiation processes taking place in this genus, the description of mitogenomes from other Microtus species is potentially of great interest. In this paper, we report the sequencing and characterization by next-generation sequencing (NGS) data of the complete mitogenome of three further Microtus species, M. cabrerae, M. chrotorrhinus and M. thomasi.

2. Materials and Methods

2.1. DNA Extraction, Sequencing and Mitogenome Assembly

A single male specimen of M. cabrerae (Siles, Spain (TOPC-0902)) and one of M. thomasi (Edessa, Greece, (R231)) were used for DNA extraction, while for M. chrotorrhinus (North America), an established cell line from a male individual was employed [4,40]. All capture and sacrifice protocols were approved by the Junta de Andalucía Ethics Committee for Animal Experimentation (code: 04/09/2018/130). Tissue samples from the specimens were stored in 100% ethanol at −20 °C. The total DNA was extracted from tissues and from cultured cells using the Gentra Puregene Tissue Kit (Qiagen, Hilden, Germany). The extracted genomic DNA of M. cabrerae was sequenced using the Illumina technology in Macrogen (Europe) and M. thomasi and M. chrotorrhinus DNAs in Novogene (Hong Kong). Briefly, for genome sequencing, we used approximately 3 µg of genomic DNA for the construction of a library of 750-bp-long fragments for M. cabrerae and 350-bp-long fragments for the other two species. These libraries were sequenced using the Illumina® Hiseq™ 2000 (San Diego, CA, USA) platform with the 2 × 100 nt paired-end option. Two Gbp of sequences were obtained for each species. Graph-based clustering analysis [41,42] was performed using RepeatExplorer implemented within the Galaxy environment (http://repeatexplorer.org/). Sequence clusters corresponding to mtDNA were selected and assembled. We also used the Geneious v.R7.1 software [43] to assemble and annotate the mitogenomes. Both assembled versions of each species’ mitogenome were aligned and revised manually using the complete M. arvalis mitogenome as reference [38] (accession number: MG948434). In addition, the sequence of the control region (D-loop) of the mitochondrial genomes was confirmed by PCR amplification and sequencing using the primer pair Pro+ (5′-ACCATCAGCACCCAAAGCTG-3′) and Phe- (5′-AAGCATTTTCAGTGCTTTGCTT-3′) [35,44].

2.2. Annotation and Sequence Analysis

The annotation of these three mitogenomes was performed using web-based services MITOS (http://mitos.bioinf.uni-leipzig.de/help.py) [45] and tRNA scan-SE (http://lowelab.ucsc.edu/tRNAscan-SE/) [46]. The annotations of the protein-coding genes (PCGs), transfer RNAs (tRNAs) and rRNA genes were refined by manual comparison, according to Markova et al. [47], with the M. arvalis mitogenome [38]. The base composition was estimated using the Bioedit program (v7.0.9.0) (http://www.mbio.ncsu.edu/BioEdit/bioedit.html), and codon usage was analyzed using MEGA version 10 [48]. The circularized drawing of the mitogenome was performed with the OrganellarGenomeDRAW tools (http://ogdraw.mpimp-golm.mpg.de/) [49].

2.3. Phylogenetic Analysis

For the phylogenetic analyses, in addition to the sequenced mitogenomes described above, we also included nine complete mitochondrial genomes from Microtus species available in GenBank and from other representative species of the Arvicolinae subfamily [1,38,50,51,52,53,54,55,56,57,58,59,60,61,62]. As outgroup, we used a species of the genus Akodon (A. montensis) [63] from the subfamily Sigmodontinae.
Complete mitogenomes were aligned using ClustalW, and the phylogenetic relationships were reconstructed using the Bayesian inference (BI) implemented in MrBayes v. 3.1 [64]. Runs of two million generations were conducted. Trees were sampled every 1000 generations with a burn-in of 25%. The best-fit nucleotide substitution model with the lowest BIC (Bayesian Information Criterion) value was chosen (GTR + G + I) using MEGA version X [48].

3. Results and discussion

3.1. Gene Organization

The complete mitogenomes of M. cabrerae (MN058077), M. chrotorrhinus (MN058078) and M. thomasi (MN058079) analyzed were 16,331 bp, 16,297 bp and 16,295 bp in length, respectively. These values are similar to the mitogenomes of other species from this genus, which range between 16,283 bp (M. rossiaemeridionalis) and 16,312 bp (M. kikuchii) [1,60]. They were also comparable in size to those from other species of the Arvicolinae subfamily such as Proedromys liangshanensis (16,296 bp) [57] and Neodon forresti (16,397 bp; GenBank accession number: KU891252.1). All our results confirm that the mitochondrial genomes in the Arvicolinae subfamily are very similar in size.
The mitogenomes from M. cabrerae, M. chrotorrhinus and M. thomasi include a control region (D-loop) and a conserved set of 37 vertebrate mitochondrial genes, with 13 protein-coding genes (PCGs), 22 tRNA genes and two rRNA genes (12S rRNA and 16S rRNA) (Table 1). As expected, the organization and structures of these three mitogenomes were identical to those described for other Microtus and mammal species (Figure 1). Hence, twelve PCGs, 14 tRNAs and two rRNAs are located on the heavy strand, while Nd6 and eight tRNAs are found on the light strand. The D-loop is emplaced between the tRNA-Pro and tRNA-Phe genes [37,38,57]. The percentage of identity observed in pairwise comparing of these three complete mitogenomes varied between 86.34% (M. thomasi–M. cabrerae pair-wise comparison) and 87.55% (M. thomasi–M. chrotorrhinus pair-wise comparison). These results fall into the range of the identity values that we calculated for comparisons between the available Microtus mitogenomes (85.50–98.75%).

3.2. Nucleotide Composition

We identified a bias towards A and T nucleotides, which is commonly reported in mitogenome sequences in mammals [35,37]. Hence, the A+T compositions of the H-strands are 58.19%, 59.27% and 59.77% in the M. cabrerae, M. chrotorrhinus and M. thomasi mitogenomes, respectively. The 13 mitochondrial PCGs are AT-biased, with an A+T content ranging from 54.34–56.63% for Cox3 to 64.71–66.67% for the Atp8 gene. The control region, the two rRNA genes and the 22 tRNAs are also AT-biased in all three species (Table 2).

3.3. Protein-Coding Genes and Codon Usage

The 13 mitochondrial PCGs from the three analyzed species are 11,390 bp in length (11,358 bp in codons and 32 bp in stop codons) and encode 3786 amino acids (Table 2). The three described mitogenomes also contain some overlapping nucleotides and gaps between PCGs or between PCGs and tRNAs (Table 1). The longest overlap of 43 bp is located between the Atp8 and Atp6 genes (Table 1).
Twelve mitochondrial PCGs use exactly the same start codon for translation initiation in all the three species: GTG for Nd1, ATT for Nd2 and Nd3, ATG for the other nine PCGs. Only the Nd5 gene has variation in the start codon: ATA in M. cabrerae and M. thomasi, and ATT in M. chrotorrhinus. Similarly, 12 PCGs genes use exactly the same five stop codons for translation termination in all three species, two incomplete (T-- for Nd1, Cox3 and Nd4; TA- for Atp6) and the three other complete (TAG for Nd6, TAA for the rest of PCGs). However, Nd5 uses TAA in M. cabrerae and M. thomasi, and TAG in M. chrotorrhinus (Table 2).
The most abundant start and stop codons were ATG and TAA, respectively, a finding that agrees with previous work on other mammal mitogenomes [30,31,35,37,65,66]. Incomplete stop codons (T-- or TA-), like those used in three PCGs (Nd1, Cox3, and Nd4), are commonly observed in metazoan mitogenomes. They might be further completed by poly-adenylation of the 3′-end of the mRNA occurring after transcription, giving rise to the complete functional TAA stop codon [67,68].
The length of the 13 PCGs is the same in all three species and, consequently, in the coded proteins as well. The percentages of nucleotide identity range between 82.03%, observed when comparing the M. cabrereae and M. thomasi Nd2 gene sequences, and 89.71%, obtained when comparing the M. chrotorrhinus and M. thomasi Atp8 gene sequences (Table 2).
Due to the A+T richness of the Microtus mitogenomes, a strong bias toward A+T-rich codons in the codon usage of the PCGs was observed (Table 3). The most frequently used codons are CTA (Leu), ATC (Ile), ATA (Met), ATT (Ile), TTC (Phe), ACA (Thr), AAC (Asn), TCA (Ser), CCA (Pro) and GCC (Ala) (Table 3). Accordingly, leucine (15.1–15.5%), isoleucine (8.9–9.7%), threonine (8.2–8.3%), serine (7.5–7.7%), alanine (6.8–6.9%), phenylalanine (6.2–6.3%), methionine (5.7–6.0%) and glycine (5.6–5.7%) are the most common amino acids found in the mitochondria protein sequences, comprising approximately 65% of the coded amino acids. The high proportion of A+T in PCGs appears to be a shared feature in mammal species [35,66]. For any amino acid, the relative synonymous codon usage value (RSCU) is equivalent to the number of times that a codon appears in a gene in relation to the number of expected occurrences under an assumption of equal codon usage. The six codons with the highest RSCU values described in the PCGs from Microtus are as follows: CTA(L) (2.58–2.62), TCA(S) (2.40–2.55), CGA(R) (2.38–2.50), ACA(T) (1.97–2.01), CCA(P) (1.95–2.22) and GCC(A) (1.81–2.05), three of which are A+T rich (Table 3).

3.4. rRNAs, tRNA Genes and Non-Coding Regions

The tRNA-Val is located between the rRNA genes 12S and 16S. rRNAs genes appeared flanked by tRNA-Phe and tRNA-Leu(UUR) (Table 1; Figure 1). The combined sizes of the two rRNA genes in M. cabrerae, M. chrotorrhinus and M. thomasi are 2518 bp, 2509 bp and 2515 bp, respectively. The combined sizes of the 22 tRNA genes in M. cabrerae, M. chrotorrhinus and M. thomasi are 1488 bp, 1500 bp and 1489 bp, respectively. The length of the tRNA genes ranged between 56 bp for tRNA-Hys in M. cabrerae to 75 bp for tRNA-Leu(UUR) in the three mitogenomes (Table 1).
Non-coding regions are important during replication and for the maintenance of the mitogenomes. These included the L-strand origin of replication (OL), intergenic spacers and the control region [69]. The three mitogenomes from Microtus species analyzed here have identical ORs of 35 bp. This region is located between tRNA-Asn and tRNA-Cys in the WANCY region, which refers to a cluster of five tRNA genes (tRNA-Trp, tRNA-Ala, tRNA-Asn, tRNA-Cys and tRNA-Tyr). The same organization is present in other Microtus and Arvicolinae species, as well as in most mammal species [31,37,38,57]. Hence, the OR from the mitogenomes of other Microtus species varies in length from 34 to 40 bp, while in most Arvicolinae species it is 34 bp in length and is highly conserved (Figure 2). Intergenic spacers were also found in the mitogenomes, with sizes in the range 1–11 bp (Table 1).
The D-loops in M. cabrerae, M. chrotorrhinus and M. thomasi are 930 bp, 903 bp and 896 bp in length, respectively, and are located between tRNA-Pro and tRNA-Phe (Table 1; Figure 1). In other Arvicolinae species, the length of this region ranges between 657 bp (M. ochrogaster) and 1089 bp (Myodes rufocanus). Control regions are the most clearly differentiated regions in the mitogenomes of the three Microtus species analyzed, with a nucleotide similarity of 80.67–82.09%. Similar values are observed when comparing the D-loop regions of different Microtus species (77.1–97.3%) and those from other Arvicolinae species (64.8–97.8%).
The control region includes three domains (Domains I, II and III). In Domain I, the extended termination-associated sequences (ETAS-1 and ETAS-2) were identified (Figure 3). The ETAS-1 sequence is better conserved than the ETAS-2 sequences in these species. Thus, pair-wise comparisons of ETAS-1 sequences showed a similarity of 89.83–96.6%, while the ETAS-2 sequence similarity was 67.3–76.9%.
The conserved sequence blocks CSB-1, CSB-2 and CSB-3 [70,71] were identified within Domain III. CSB-1 is the best-conserved block, having similarity values of 66.7–84% when comparing the three Microtus species with each other. CSB-2 and CSB-3 are less well conserved, with similarities of 52.6–68.4% and 54.1–68.0%, respectively (Figure 3). No repetitive DNA sequences were found to be present between CSB1 and CSB2 on the Microtus D-loop, as occurs in other mammal species [35]. Five other conserved sequence blocks (CSB-B, C, D, E and F) were identified in central Domain II [72], all of which are well conserved with nucleotide similarities of 78.95–100% (Figure 3).

3.5. Phylogenetic Analysis

The phylogenetic positions of the three analyzed species were assessed using Bayesian inference (Figure 4). All Microtus species are clustered in a well-supported clade that also includes Neodon and Lasiopodomys species. The Alexandromys species are grouped in the same branch with genera Neodon and Lasiopodomys, which have been considered as subgenera of Microtus genus, although phylogenetic relationships are not clearly established [20,24]. Proedromys is also close to this Microtus clade since it appears grouped as part of a well-supported node (posterior probability values = 1). The Microtus-Proedromys group is also associated with the clade of Myodes and Eothenomys (posterior probability values = 0.93). Finally, Dicrostonyx and Ondatra show a basal position. These results agree with the previously established phylogenetic relationships for these genera [73,74,75,76].
A number of conclusions can be drawn for the Microtus species analyzed. The two species from the Terricola group included here, M. (Terricola) thomasi and M. (Terricola) subterraneus, are closely associated with M. arvalis and M. rossiaemeridionalis. The close phylogenetic relationship between the Terricola and Microtus subgenera has been previously reported [20,29]. Although the two subgenera Aulacomys and Pedomys are not resolved the three North American species, M. chrotorrhinus, M. ochrogaster and M. richardsoni are grouped together. These Nearctic species fall within the phylogeny of Microtus, in line with previous studies [20], rather than being basal, as recently reported [29]. M. cabrerae and M. agrestis are grouped, which could support their inclusion in the subgenus Agricola as has been previously proposed [20]. Ostensibly, M. cabrerae and M. agrestis share certain unusual genetic features, including the presence of giant sex chromosomes, which could be regarded as additional proof of their phylogenetic proximity. However, several studies have clearly demonstrated that these enlarged sex chromosomes arose and evolved independently in the genus Microtus [4], and hence, their presence is not a robust criterion for the inclusion of M. cabrerae and M. agrestis in the same subgenus. A previous mitochondrial phylogenetic reconstruction obtained similar results, with M. cabrerae and M. agrestis clustered in the same clade, but the level of genetic divergence indicated that both species could be considered as members of two different subgenera (Agricola and Iberomys) [29]. The genus Iberomys, which is based on the description of archaic morphological characters, has been proposed, with only the species Microtus cabrerae [77,78]; however, no support is obtained for this genus [20].
The phylogenetic results obtained here with mitochondrial data need to be validated by the use of other nuclear markers because of the limitation inherent to mitogenomes, maternal inheritance, accelerated rates of substitution, introgression, effective population size and neutrality [79]. However, sequencing and characterization of mitogenomes from other species of the genus Microtus and closely related taxa from the subfamily Arvicolinae will help our understanding of the phylogenic relationships of this rodent species group and hence resolve some of the issues that remain open.

4. Conclusions

The complete mitogenomes of three Microtus species are described. Our results demonstrated that these mitogenomes have the organization and characteristics of the described mitogenomes of voles and mammalian species and contain 13 protein-coding genes (PCGs), two ribosomal RNA genes, 22 transfer RNA genes and a control region. We identified the conserved domain and the sequence-conserved blocks of the region control. Phylogenetic reconstructions reproduced the established phylogenetic relationships for all the Arvicolinae genera that were analyzed. Our results could be useful in future studies about the identifications and phylogeny of Arvicolidae species, especially of the genus Microtus.

Author Contributions

Conceptualization: A.S., T.P. and P.L.; methodology, L.L. and G.A.-M.; formal analysis and validation: M.R., J.A.M., L.L. and G.A.-M.; writing—original draft preparation, A.S., P.L. and L.L.; writing—review and editing: M.R., J.A.M. and T.P. All authors have read and agreed to the published version of the manuscript.

Funding

This work was funded by the Consejería de Innovación, Ciencia y Empresa of the Junta de Andalucía (Group: RNM924), and by the Universidad de Jaén (as part of the program Plan de Apoyo a la Investigación 2019-2020, Acción 1). Michail Rovatsos was supported by Charles University Research Centre program (204069).

Acknowledgments

The authors would like to express their gratitude to the Consejería de Medio Ambiente y Ordenación del Territorio de la Junta de Andalucía for the permits for using a specimen of M. cabrerae from Jerez Zoo and to William S. Modi and Polina L. Perelman for the cell line of Microtus chrotorrhinus.

Conflicts of Interest

The authors declare that they have no conflict of interest.

References

  1. Triant, D.A.; DeWoody, J.A. Accelerated molecular evolution in Microtus (Rodentia) as assessed via complete mitochondrial genome sequences. Genetica 2006, 128, 95–108. [Google Scholar] [CrossRef] [PubMed]
  2. Fink, S.; Fischer, M.C.; Excoffier, L.; Heckel, G. Genomic scans support repetitive continental colonization events during the rapid radiation of voles (Rodentia: Microtus): The utility of AFLPs versus mitochondrial and nuclear sequence markers. Syst. Biol. 2010, 59, 548–572. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  3. Sitnikova, N.A.; Romanenko, S.A.; O’Brien, P.C.; Perelman, P.L.; Fu, B.; Rubtsova, N.V.; Serdukova, N.A.; Golenishchev, F.N.; Trifonov, V.A.; Ferguson-Smith, M.A.; et al. Chromosomal evolution of Arvicolinae (Cricetidae, Rodentia). I. The genome homology of tundra vole, field vole, mouse and golden hamster revealed by comparative chromosome painting. Chromosome Res. 2007, 15, 447–456. [Google Scholar] [CrossRef] [PubMed]
  4. Lamelas, L.; Arroyo, M.; Fernández, F.J.; Marchal, J.A.; Sánchez, A. Structural and evolutionary relationships in the giant sex chromosomes of three Microtus species. Genes 2018, 9, 27. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  5. Ohno, S.; Stenius, C.; Christian, L. The X0 as the normal female of the creeping vole (Microtus oregoni). In Chromosomes Today; Darlington, C.D., Lewis, K.R., Eds.; Oliver and Boyd: Edinburgh, United Kingdom, 1966; Volume 1, pp. 182–187.oregoni). In Chromosomes Today; Darlington, C.D., Lewis, K.R., Eds.; Oliver and Boyd: Edinburgh, UK, 1966; Volume 1, pp. 182–187. [Google Scholar]
  6. Burgos, M.; Jiménez, R.; Díaz de la Guardia, R. XY females in Microtus cabrerae (Rodentia, Microtidae): A case of possibly Y-linked sex reversal. Cytogenet. Cell Genet. 1988, 49, 275–277. [Google Scholar] [CrossRef] [PubMed]
  7. Marchal, J.A.; Acosta, M.J.; Bullejos, M.; Díaz de la Guardia, R.; Sánchez, A. Origin and spread of the SRY gene on the X and Y chromosomes of the rodent Microtus cabrerae: Role of L1 elements. Genomics 2008, 91, 142–151. [Google Scholar] [CrossRef] [Green Version]
  8. Acosta, M.J.; Marchal, J.A.; Fernández-Espartero, C.; Romero-Fernández, I.; Rovatsos, M.T.; Giagia-Athanasopoulou, E.B.; Gornung, E.; Castiglia, R.; Sánchez, A. Characterization of the satellite DNA Msat-160 from species of Terricola (Microtus) and Arvicola (Rodentia, Arvicolinae). Genetica 2010, 138, 1085–1098. [Google Scholar] [CrossRef]
  9. Borodin, P.M.; Basheva, E.A.; Golenischev, F.N.; Dashkevich, O.A.; Kartavtseva, I.N.; Lisachov, A.P.; Torgasheva, A.A. Parallel occurrence of asynaptic sex chromosomes in gray voles (Microtus Schrank, 1798). Paleontol. J. 2013, 47, 1035–1040. [Google Scholar] [CrossRef]
  10. Rovatsos, M.T.; Marchal, J.A.; Romero-Fernández, I.; Arroyo, M.; Athanasopoulou, E.B.; Sánchez, A. Extensive sex chromosome polymorphism of Microtus thomasi/Microtus atticus species complex associated with cryptic chromosomal rearrangements and independent accumulation of heterochromatin. Cytogenet. Genome Res. 2017, 151, 198–207. [Google Scholar] [CrossRef]
  11. Maruyama, T.; Imai, H.T. Evolutionary rate of the mammalian karyotype. J. Theor. Biol. 1981, 90, 111–121. [Google Scholar] [CrossRef]
  12. Libbus, B.L.; Johnson, L.A. The creeping vole, Microtus oregoni: Karyotype and sex-chromosome differences between two geographical populations. Cytogenet. Cell Genet. 1988, 47, 181–184. [Google Scholar] [CrossRef] [PubMed]
  13. Gornung, E.; Castiglia, R.; Rovatsos, M.; Marchal, J.A.; Díaz de la Guardia-Quiles, R.; Sánchez, A. Comparative cytogenetic study of two sister species of Iberian ground voles, Microtus (Terricola) duodecimcostatus and M. (T.) lusitanicus (Rodentia, Cricetidae). Cytogenet. Genome Res. 2011, 132, 144–150. [Google Scholar] [CrossRef] [PubMed]
  14. Baker, R.J.; Hamilton, M.J.; Van Den Bussche, R.A.; Wiggins, L.E.; Sugg, D.W.; Smith, M.H.; Lomakin, M.D.; Gaschak, S.P.; Bundova, E.G.; Ruoenskaya, G.A.; et al. Small mammals from the most radioactive sites near the Chernobyl nuclear power plant. J. Mammal. 1996, 77, 155–170. [Google Scholar] [CrossRef] [Green Version]
  15. Fink, S.; Excoffier, L.; Heckel, G. Mitochondrial gene diversity in the common vole Microtus arvalis shaped by historical divergence and local adaptations. Mol. Ecol. 2004, 13, 3501–3514. [Google Scholar] [CrossRef] [PubMed]
  16. Jaarola, M.; Searle, J.B. Phylogeography of field voles (Microtus agrestis) in Eurasia inferred from mitochondrial DNA sequences. Mol. Ecol. 2002, 11, 2613–2621. [Google Scholar] [CrossRef]
  17. Bannikova, A.A.; Lebedev, V.; Lissovsky, A.; Matrosova, V.; Abramson, N.I.; Obolenskaya, E.V.; Tesakov, A.S. Molecular phylogeny and evolution of the Asian lineage of vole genus Microtus (Rodentia: Arvicolinae) inferred from mitochondrial cytochrome b sequence. Biol. J. Linn. Soc. 2010, 99, 595–613. [Google Scholar] [CrossRef] [Green Version]
  18. Chaline, J.; Graf, J.D. Phylogeny of the Arvicolidae (Rodentia): Biochemical and paleontological evidence. J. Mammal. 1988, 69, 22–33. [Google Scholar] [CrossRef]
  19. Modi, W.S. Phylogenetic history of LINE-1 among arvicolid rodents. Mol. Biol. Evol. 1996, 13, 633–641. [Google Scholar] [CrossRef]
  20. Jaarola, M.; Martinkova, N.; Gunduz, I.; Brunhoff, C.; Zima, J.; Nadachowski, A.; Amori, G.; Bulatova, N.S.; Chondropoulos, B.; Fraguedakis-Tsolis, S.; et al. Molecular phylogeny of the speciose vole genus Microtus (Arvicolinae, Rodentia) inferred from mitochondrial DNA sequences. Mol. Phylogenet. Evol. 2004, 33, 647–663. [Google Scholar] [CrossRef]
  21. Fletcher, N.K.; Acevedo, P.; Herman, J.S.; Paupério, J.; Alves, P.C.; Searle, J.B. Glacial cycles drive rapid divergence of cryptic field vole species. Ecol. Evol. 2019, 9, 14101–14113. [Google Scholar] [CrossRef] [Green Version]
  22. De Woody, J.A. Nucleotide variation in the p53 tumor suppressor gene of voles from Chernobyl, Ukraine. Mutat. Res. 1999, 439, 25–36. [Google Scholar] [CrossRef]
  23. Belfiore, N.M.; Hoffman, F.G.; Baker, R.J.; DeWoody, J.A. The use of nuclear and mitochondrial single nucleotide polymorphisms to identify cryptic species. Mol. Ecol. 2003, 12, 2011–2017. [Google Scholar] [CrossRef] [PubMed]
  24. Conroy, C.J.; Cook, J.A. Molecular systematics of a holarctic rodent (Microtus: Muridae). J. Mamm. 2000, 81, 344–359. [Google Scholar] [CrossRef] [Green Version]
  25. Shevchenko, A.T.; Mazurok, N.A.; Slobodyanyuk, S.Y.; Zakian, S.M. Comparative analysis of the MSAT-160 repeats in four species of common vole (Microtus, Arvicolidae). Chromosome Res. 2002, 10, 117–126. [Google Scholar] [CrossRef] [PubMed]
  26. Romanenko, S.A.; Sitnikova, N.A.; Serdukova, N.A.; Perelman, P.L.; Rubtsova, N.V.; Bakloushinskaya, I.Y.; Lyapunova, E.A.; Just, W.; Ferguson-Smith, M.A.; Yang, F.; et al. Chromosomal evolution of Arvicolinae (Cricetidae, Rodentia). II. The genome homology of two mole voles (genus Ellobius), the field vole and golden hamster revealed by comparative chromosome painting. Chromosome Res. 2007, 15, 891–897. [Google Scholar] [CrossRef]
  27. Galewski, T.; Tilak, M.K.; Sanchez, S.; Chevret, P.; Paradis, E.; Douzery, E. The evolutionary radiation of Arvicolinae rodents (voles and lemmings): Relative contribution of nuclear and mitochondrial DNA phylogenies. BMC Evol. Biol. 2006, 6, 80. [Google Scholar] [CrossRef] [Green Version]
  28. Conroy, C.J.; Cook, J.A. MtDNA evidence for repeated pulses of speciation within arvicoline and murid rodents. J. Mammal. Evol. 1999, 6, 221–245. [Google Scholar] [CrossRef]
  29. Barbosa, S.; Paupério, J.; Pavlova, S.V.; Alves, P.C.; Searle, J.B. The Microtus voles: Resolving the phylogeny of one of the most speciose mammalian genera using genomics. Mol. Phylogenet. Evol. 2018, 125, 85–92. [Google Scholar] [CrossRef]
  30. Cabria, M.T.; Rubines, J.; Gómez-Moliner, B.; Zardoya, R. On the phylogenetic position of a rare Iberian endemic mammal, the Pyrenean desman (Galemys pyrenaicus). Gene 2006, 375, 1–13. [Google Scholar] [CrossRef]
  31. Kim, J.Y.; Park, Y.C. Gene organization and characterization of the complete mitogenome of Hypsugo alaschanicus (Chiroptera: Vespertilionidae). Genet. Mol. Res. 2015, 14, 16325–16331. [Google Scholar] [CrossRef]
  32. Ye, F.; Samuels, D.C.; Clark, T.; Guo, Y. High-throughput sequencing in mitochondrial DNA research. Mitochondrion 2014, 17, 157–163. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  33. Pita, S.; Panzera, F.; Vela, J.; Mora, P.; Palomeque, T.; Lorite, P. Complete mitochondrial genome of Triatoma infestans (Hemiptera, Reduviidae, Triatominae), main vector of Chagas disease. Infect. Genet. Evol. 2017, 54, 158–163. [Google Scholar] [CrossRef] [PubMed]
  34. Fernández-Pérez, J.; Nantón, A.; Ruiz-Ruano, F.J.; Camacho, J.P.M.; Méndez, J. First complete female mitochondrial genome in four bivalve species genus Donax and their phylogenetic relationships within the Veneroida order. PLoS ONE 2017, 8, 12. [Google Scholar] [CrossRef] [PubMed]
  35. Gutiérrez, J.; Lamelas, L.; Aleix-Mata, G.; Arroyo, M.; Marchal, J.A.; Palomeque, T.; Lorite, P.; Sánchez, A. Complete mitochondrial genome of the Iberian Mole Talpa occidentalis (Talpidae, Insectivora) and comparison with Talpa europaea. Genetica 2018, 146, 415–423. [Google Scholar] [CrossRef] [PubMed]
  36. Aleix-Mata, G.; Ruiz-Ruano, F.J.; Pérez, J.M.; Sarasa, M.; Sánchez, A. Complete mitochondrial genome of the Western Capercaillie Tetrao urogallus (Phasianidae, Tetraoninae). Zootaxa 2019, 4550, 585–593. [Google Scholar] [CrossRef]
  37. Aleix-Mata, G.; Gutiérrez, J.; Ruiz-Ruano, F.J.; Lorite, P.; Marchal, J.A.; Sánchez, A. The complete mitochondrial genome of Talpa aquitania (Talpidae; Insectivora), a mole species endemic to northern Spain and southern France. Mol. Biol. Rep. 2020, 47, 2387–2403. [Google Scholar] [CrossRef]
  38. Folkertsma, R.; Westbury, M.V.; Eccard, J.A.; Hofreiter, M. The complete mitochondrial genome of the common vole, Microtus arvalis (Rodentia: Arvicolinae). Mitochondrial DNA Part B 2018, 3, 446–447. [Google Scholar] [CrossRef]
  39. Zhu, L.; Qi, Z.; Wen, Y.C.; Min, J.Z.; Song, Q.K. The complete mitochondrial genome of Microtus fortis pelliceus (Arvicolinae, Rodentia) from China and its phylogenetic analysis. Mitochondrial DNA Part B 2019, 4, 2039–2041. [Google Scholar] [CrossRef] [Green Version]
  40. Modi, W.S. Nucleotide sequence and genomic organization of a tandem satellite array from the rock vole Microtus chrotorrhinus (Rodentia), Mamm. Genome 1992, 3, 226–232. [Google Scholar]
  41. Novák, P.; Neumann, P.; Macas, J. Graph-based clustering and characterization of repetitive sequences in next-generation sequencing data. BMC Bioinform. 2010, 11, 378. [Google Scholar] [CrossRef] [Green Version]
  42. Novák, P.; Neumann, P.; Pech, J.; Steinhaisl, J.; Macas, J. RepeatExplorer: A Galaxy-based web server for genome-wide characterization of eukaryotic repetitive elements from next-generation sequence reads. Bioinformatics 2013, 29, 792–793. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  43. Kearse, M.; Moir, R.; Wilson, A.; Stones-Havas, S.; Cheung, M.; Sturrock, S.; Buxton, S.; Cooper, A.; Markowitz, S.; Duran, C.; et al. Geneious Basic: An integrated and extendable desktop software platform for the organization and analysis of sequence data. Bioinformatics 2012, 28, 1647–1649. [Google Scholar] [CrossRef] [PubMed]
  44. Haring, E.; Herzig-Straschil, B.; Spitzenberger, F. Phylogenetic analysis of Alpine voles of the Microtus multiplex complex using the mitochondrial control region. J. Zool. Evol. Res. 2000, 38, 231–238. [Google Scholar] [CrossRef]
  45. 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]
  46. Lowe, T.M.; Chan, P.P. tRNAscan-SE On-line: Search and contextual analysis of transfer RNA genes. Nucl. Acids Res. 2016, 44, W54–W57. [Google Scholar] [CrossRef]
  47. Marková, S.; Filipi, K.; Searle, J.B.; KotlÍk, P. Mapping 3′ transcript ends in the bank vole (Clethrionomys glareolus) mitochondrial genome with RNA-Seq. BMC Genom. 2015, 16, 870. [Google Scholar] [CrossRef] [Green Version]
  48. Kumar, S.; Stecher, G.; Li, M.; Knyaz, C.; Tamura, K. MEGA X: Molecular evolutionary genetics analysis across computing platforms. Mol. Biol. Evol. 2018, 35, 1547–1549. [Google Scholar] [CrossRef]
  49. Lohse, M.; Drechsel, O.; Kahlau, S.; Bock, R. OrganellarGenomeDRAW-A suite of tools for generating physical maps of plastid and mitochondrial genomes and visualizing expression data sets. Nucleic Acids Res. 2013, 41, W575–W581. [Google Scholar] [CrossRef]
  50. Alqahtani, F.; Duckett, D.; Pirro, S.; Mandoiu, I.I. Complete mitochondrial genome of the water vole, Microtus richardsoni (Cricetidae, Rodentia). Mitochondrial DNA Part B 2020, 5, 2498–2499. [Google Scholar] [CrossRef]
  51. Bendová, K.; Marková, S.; Searle, J.B.; Kotlík, P. The complete mitochondrial genome of the bank vole Clethrionomys glareolus (Rodentia: Arvicolinae). Mitochondrial DNA A DNA Mapp. Seq. Anal. 2016, 27, 111–112. [Google Scholar] [CrossRef]
  52. Bondareva, O.V.; Abramson, N.I. The complete mitochondrial genome of the common pine vole Terricola subterraneus (Arvicolinae, Rodentia). Mitochondrial DNA Part B 2019, 4, 3925–3926. [Google Scholar] [CrossRef] [Green Version]
  53. Cao, W.; Xia, Y.; Dang, X.; Xu, Q. The first complete mitochondrial genome of the Microtus ochrogaster. Mitochondrial DNA Part A 2016, 27, 3682–3683. [Google Scholar] [CrossRef] [PubMed]
  54. Chen, S.; Chen, G.; Wei, H.; Wang, Q. Complete mitochondrial genome of the Père David’s Vole, Eothenomys melanogaster (Rodentia: Arvicolinae). Mitochondrial DNA A DNA Mapp. Seq. Anal. 2016, 27, 2496–2497. [Google Scholar] [CrossRef]
  55. Cong, H.; Kong, L.; Liu, Z.; Wu, Y.; Motokawa, M.; Harada, M.; Li, Y. Complete mitochondrial genome of the mandarin vole Lasiopodomys mandarinus (Rodentia: Cricetidae). Mitochondrial DNA A DNA Mapp. Seq. Anal. 2016, 27, 760–761. [Google Scholar] [CrossRef]
  56. Fedorov, V.B.; Goropashnaya, A.V. Complete mitochondrial genomes of the North American collared lemmings Dicrostonyx groenlandicus Traill, 1823 and Dicrostonyx hudsonius Pallas, 1778 (Rodentia: Arvicolinae). Mitochondrial DNA B Resour. 2016, 1, 878–879. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  57. Hao, H.; Liu, S.; Zhang, X.; Chen, W.; Song, Z.; Peng, H.; Liu, Y.; Yue, B. Complete mitochondrial genome of a new vole Proedromys liangshanensis (Rodentia: Cricetidae) and phylogenetic analysis with related species: Are there implications for the validity of the genus Proedromys? Mitochondrial DNA 2011, 22, 28–34. [Google Scholar] [CrossRef]
  58. Jiang, J.Q.; Wu, S.X.; Chen, J.J.; Liu, C.Z. Characterization of the complete mitochondrial genome of short-tailed field vole, Microtus agrestis. Mitochondrial DNA Part B 2018, 3, 845–846. [Google Scholar] [CrossRef] [Green Version]
  59. Jiang, X.; Gao, J.; Ni, L.; Hu, J.; Li, K.; Sun, F.; Xie, J.; Bo, X.; Gao, C.; Xiao, J.; et al. The complete mitochondrial genome of Microtus fortis calamorum (Arvicolinae, Rodentia) and its phylogenetic analysis. Gene 2012, 498, 288–295. [Google Scholar] [CrossRef]
  60. Lin, Y.H.; Waddell, P.J.; Penny, D. Pika and vole mitochondrial genomes increase support for both rodent monophyly and glires. Gene 2002, 294, 119–129. [Google Scholar] [CrossRef]
  61. Zhang, Q.; He, T.; Wei, H.; Li, F.; Feng, Y.; Zong, H.; Chen, S. Characterization of the complete mitochondrial genome and phylogenetic relationship of Neodon sikimensis (Rodentia: Arvicolinae). Mitochondrial DNA Part B 2016, 1, 445–446. [Google Scholar] [CrossRef] [Green Version]
  62. Zhao, H.; Qi, X.; Li, C. Complete mitochondrial genome of the muskrat (Ondatra zibethicus) and its unique phylogenetic position estimated in Cricetidae. Mitochondrial DNA Part B 2018, 3, 296–298. [Google Scholar] [CrossRef] [Green Version]
  63. Vilela, J.F.; Mello, B.; Voloch, C.M.; Schrago, C.G. Sigmodontine rodents diversified in South American prior to the complete rise of the Panamanian Isthmus. J. Zoolog. Syst. Evol. Res. 2014, 52, 249–256. [Google Scholar] [CrossRef]
  64. Ronquist, F.; Huelsenbeck, J.P. MrBayes 3: Bayesian phylogenetic inference under mixed models. Bioinformatics 2003, 19, 1572–1574. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  65. Xu, Y.; Huang, X.; Hu, Y.; Tu, F. Description of the mitogenome of Gansu mole (Scapanulus oweni). Mitochondrial DNA Part A DNA Mapp. Seq. Anal. 2016, 27, 2083–2084. [Google Scholar] [CrossRef]
  66. Kim, N.H.; Lim, S.J.; Chae, H.M.; Park, Y.C. Complete mitochondrial genome of the Amur hedgehog Erinaceus amurensis (Erinaceidae) and higher phylogeny of the family Erinaceidae. Genet. Mol. Res. 2017, 16. [Google Scholar] [CrossRef] [PubMed]
  67. Ojala, D.; Montoya, J.; Attardi, G. tRNA punctuation model of RNA processing in human mitochondria. Nature 1981, 290, 470–474. [Google Scholar] [CrossRef]
  68. Mouchaty, S.K.; Gullberg, A.; Janke, A.; Arnason, U. The phylogenetic position of the Talpidae within Eutheria based on analysis of complete mitochondrial sequences. Mol. Biol. Evol. 2000, 17, 60–67. [Google Scholar] [CrossRef] [Green Version]
  69. Fernández-Silva, P.; Enriquez, J.A.; Montoya, J. Replication and transcription of mammalian mitochondrial DNA. Exp. Physiol. 2003, 88, 41–56. [Google Scholar] [CrossRef]
  70. Walberg, M.W.; Clayton, D.A. Sequence and properties of the human KB cell and mouse L cell D-loop regions of mitochondrial DNA. Nucleic Acids Res. 1981, 9, 5411–5421. [Google Scholar] [CrossRef]
  71. Sbisà, E.; Tanzariello, F.; Reyes, A.; Pesole, G.; Saccone, C. Mammalian mitochondrial D-loop region structural analysis: Identification of new conserved sequences and their functional and evolutionary implications. Gene 1997, 205, 125–140. [Google Scholar] [CrossRef]
  72. Ketmaier, V.; Bernardini, C. Structure of the mitochondrial control region of the Eurasian otter (Lutra lutra; Carnivora, Mustelidae): Patterns of genetic heterogeneity and implications for conservation of the species in Italy. J. Hered. 2005, 96, 318–328. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  73. Buzan Elena, V.; Krystufek, B.; Hänfling, B.; Hutchinson, W.F. Mitochondrial phylogeny of Arvicolinae using comprehensive taxonomic sampling yields new insights. Biol. J. Linn. Soc. 2008, 94, 825–835. [Google Scholar] [CrossRef] [Green Version]
  74. Abramson, N.I.; Lebedev, V.S.; Tesakov, A.S.; Bannikova, A.A. Supraspecies relationships in the subfamily Arvicolinae (Rodentia, Cricetidae): An unexpected result of nuclear gene analysis. Mol. Biol. 2009, 43, 834–846. [Google Scholar] [CrossRef]
  75. Jin, Z.M.; Yu, C.W.; Liu, Z. Sequencing and analysis of the complete mitochondrial genome of the northern red-backed vole (Myodes rutilus) from China. Mitochondrial DNA Part B 2019, 4, 1575–1577. [Google Scholar]
  76. Chen, W.C.; Hao, H.B.; Sun, Z.Y.; Liu, Y.; Liu, S.Y.; Yue, B.S. Phylogenetic position of the genus Proedromys (Arvicolinae, Rodentia): Evidence from nuclear and mitochondrial DNA, Biochem. Syst. Ecol. 2012, 42, 59–68. [Google Scholar] [CrossRef]
  77. Gromov, I.M.; Polyakov, I.Y. Fauna of the USSR, Voles (Microtinae). In Mammals; Brill Publishing Company: Leiden, Germany, 1992; Volume 3, p. 725. [Google Scholar]
  78. Cuenca-Bescós, G.; López-García, J.M.; Galindo-Pellicena, M.A.; García-Perea, R.; Gisbert, J.; Rofes, J.; Ventura, J. Pleistocene history of Iberomys, an endangered endemic rodent from southwestern Europe. Integr. Zool. 2014, 9, 481–497. [Google Scholar] [CrossRef]
  79. Rubinoff, D.; Holland, B.S. Between Two Extremes: Mitochondrial DNA is neither the Panacea nor the Nemesis of Phylogenetic and Taxonomic Inference. Syst. Biol. 2005, 54, 952–961. [Google Scholar] [CrossRef]
Figure 1. Representative map of the mitochondrial genome of the three analyzed Microtus species. Genes encoded by the heavy strand are shown outside the circle, while those encoded by the light strand are shown inside.
Figure 1. Representative map of the mitochondrial genome of the three analyzed Microtus species. Genes encoded by the heavy strand are shown outside the circle, while those encoded by the light strand are shown inside.
Animals 10 02130 g001
Figure 2. Nucleotide sequence alignment of the L-strand origin of replication (OL) of species from the subfamily Arvicolinae. Dots indicate identity and dashes denote deletions.
Figure 2. Nucleotide sequence alignment of the L-strand origin of replication (OL) of species from the subfamily Arvicolinae. Dots indicate identity and dashes denote deletions.
Animals 10 02130 g002
Figure 3. Nucleotide sequence alignment of the control region of Microtus cabrerae (M. ca), M. chrotorrhinus (M. ch) and M. thomasi (M. th). Dots indicate the identity of the nucleotides and dashes the indels. All the conserved sequences identified are included in boxes: the two conserved extended termination-associated sequences (ETAS1 and ETAS2) in Domain I, five conserved blocks (CSB-F, E, D, C, B) in the central conserved Domain II (underlined) and three conserved blocks (CSB1, CSB-2 and CSB-3) in Domain III. Dots indicate identity and dashes denote deletions.
Figure 3. Nucleotide sequence alignment of the control region of Microtus cabrerae (M. ca), M. chrotorrhinus (M. ch) and M. thomasi (M. th). Dots indicate the identity of the nucleotides and dashes the indels. All the conserved sequences identified are included in boxes: the two conserved extended termination-associated sequences (ETAS1 and ETAS2) in Domain I, five conserved blocks (CSB-F, E, D, C, B) in the central conserved Domain II (underlined) and three conserved blocks (CSB1, CSB-2 and CSB-3) in Domain III. Dots indicate identity and dashes denote deletions.
Animals 10 02130 g003
Figure 4. Bayesian Inference tree using the complete mitogenome sequences of ten Microtus species available in GenBank and seven representative species of genera from the subfamily Arvicolinae, and the mitogenome sequence of Akodon montensis (subfamily Sigmodontinae) used as an outgroup. Bayesian posterior probabilities are indicated at each node.
Figure 4. Bayesian Inference tree using the complete mitogenome sequences of ten Microtus species available in GenBank and seven representative species of genera from the subfamily Arvicolinae, and the mitogenome sequence of Akodon montensis (subfamily Sigmodontinae) used as an outgroup. Bayesian posterior probabilities are indicated at each node.
Animals 10 02130 g004
Table 1. Gene organization of the three Microtus mitogenomes (M.ca: M. cabrerae; M.ch: M. chrotorrhinus; M.th: M. thomasi).
Table 1. Gene organization of the three Microtus mitogenomes (M.ca: M. cabrerae; M.ch: M. chrotorrhinus; M.th: M. thomasi).
GeneNucleotide PositionsSize (bp) 1Strand 2AnticodonIntergenic Nucleotide 1
M.caM.chM.thM.caM.chM.thM.ca M.ch M.th
tRNAPhe1–6666HTTC
12S rRNA69–101869–101569–1016950947948H 2
tRNAVal1019–10891018–10851016–1086717070HGTA
16S rRNA1088–26551084–26451085–2651156815621567H −2
tRNALeu (uur)2656–27302647–27212653–272775HTTA 11
Nd12731–36852722–36762728–3682955H
tRNAIle3686–37533677–37453683–3751686969HATC
tRNAGln3751–38223743–38143749–382072LCAA−3
tRNAMet3824–38923817–38853822–389069HATG121
Nd23893–49273886–49203891–49251035H
tRNATrp4929–49954922–49884927–499367HTGA1
tRNAAla4997–50654990–50584995–506369LGCA1
tRNAAsn5067–51365061–51305066–513570LAAC122
Rep_origin5137–51715131–51655136–5170 35 H
tRNACys5169–52365163–52305168–5235686968LTGC−3
tRNATyr5237–53035231–52975236–530267LTAC
Cox15305–68495299–68435304–68481545H 1
tRNASer (ucn)6847–69156841–69096846–691469LTCA−3
tRNAAsp6919–69866913–69806918–698568HGAC3
Cox26988–76716982–76656987–7670684H 1
tRNALys7675–77377669–77327674–7737636464HAAA3
Atp87738–79417733–79367738–7941204H
Atp67899–85787894–85737899–8578 680 H −43
Cox38579–93628574–93578579–9362784H
tRNAGly9363–94309358–94259363–943068HGGA
Nd39431–97789426–97739431–9778348H
tRNAArg9780–98459775–98419780–9846666767HCGA1
Nd4L9849–101459844–101409850–10146297H 32
Nd410139–1151610134–1151110140–115171378H −7
tRNAHis11517–1157211512–1157911527–11585566859HCAC 9
tRNASer11584–1164211580–1163811586–1164459HAGC11
tRNALeu11642–1171111638–1170711644–1171370HCTA−1
Nd511712–1352311708–1351911714–135251812H
Nd613523–1404413516–1404013522–14046525L −4
tRNAGlu14045–1411314041–1410914047–1411569LGAA
CytB14119–1526114115–1525714121–152631143H 5
tRNAThr15264–1533315260–1532615266–15331706766HACA2
tRNAPro15334–1540115327–1539415332–1539968LCCA
D-loop15402–1633115395–1629715400–16295930903896H
1 The corresponding three values are only indicated when they are different. 2 Strand: H: heavy and L: light.
Table 2. Data about the 13 PCG of the three Microtus mitogenomes (M.ca: M. cabrerae; M.ch: M. chrotorrhinus; M.th: M. thomasi).
Table 2. Data about the 13 PCG of the three Microtus mitogenomes (M.ca: M. cabrerae; M.ch: M. chrotorrhinus; M.th: M. thomasi).
GeneGene Length (bp)% IdentityA + T Content (%)Start/StopcodonsProtein Length (aa)
M.ca/M.chM.ca/M.thM.ch/M.thM.caM.chM.thM.ca/M.ch/M.th
Nd195584.5085.8686.3956.9657.4957.70GTG/T--318
Nd2103582.2282.0383.9659.3261.8461.93ATT/TAA344
Cox1154586.2186.1586.2855.3457.8656.50ATG/TAA514
Cox268486.1186.1185.3856.2957.4658.33ATG/TAA227
Atp820485.7887.7589.7164.7166.6765.69ATG/TAA67
Atp668086.9384.5886.7856.3956.9859.77ATG/TA-226
Cox378485.3384.4485.8454.3456.6356.12ATG/T--261
Nd334883.3383.3384.4858.6257.7659.48ATT/TAA115
Nd4L29782.8382.8382.4957.9155.8959.60ATG/TAA98
Nd4137882.1580.4883.7457.8458.3560.16ATG/T--459
Nd5181283.1782.3485.1056.2957.7358.44ATA/TAA ATT/TAG ATA/TAA603
Nd652585.1485.5287.6260.7663.0560.76ATG/TAG174
CytB114387.6687.7588.1056.5256.2658.53ATG/TAA380
Total11,390 3786
Table 3. Codon usage of mitochondrial genomes protein-coding genes (PCG) of the three Microtus species.
Table 3. Codon usage of mitochondrial genomes protein-coding genes (PCG) of the three Microtus species.
Codonn%RSCUCodonn%RSCU
UUU(F)69–921.8–2.40.58–0.79UAU(Y)53–561.4–1.50.87–0.91
UUC(F)142–1693.7–4.51.21–1.42UAC(Y)65–691.7–1.81.09–1.13
UUA(L)84–1052.2–2.80.86–1.10UAA(*)11–120.33.38–3.69
UUG(L)15–200.4–0.50.15–0.20UAG(*)1–20–0.10.31–0.62
CUU(L)82–902.2–2.40.85–0.92CAU(H)13–300.3–0.80.26–0.61
CUC(L)79–1002.1–2.60.82–1.02CAC(H)68–871.8–2.31.39–1.74
CUA(L)251–2546.6–6.72.58–2.62CAA(Q)66–671.7–1.81.72–1.74
CUG(L)35–460.9–1.20.36–0.47CAG(Q)10–110.30.26–0.28
AUU(I)142–1783.7–4.70.77–1.00AAU(N)35–430.9–1.10.43–0.52
AUC(I)178–2294.7–6.01.00–1.24AAC(N)122–1303.2–3.41.48–1.57
AUA(M)164–1734.3–4.61.48–1.59AAA(K)83–942.2–2.51.66–1.86
AUG(M)45–601.2–1.60.41–0.53AAG(K)7–170.2–0.40.14–34
GUU(V)22–330.6–0.90.54–0.73GAU(D)20–260.5–0.70.53–0.70
GUC(V)46–501.2–1.31.01–1.22GAC(D)48–551.3–1.41.30–1.47
GUA(V)73–861.9–2.31.74–1.82GAA(E)65–831.7–2.21.38–1.75
GUG(V)19–240.5–0.60.46–0.53GAG(E)12–290.3–0.80.25–0.62
UCU(S)37–3810.76–0.80UGU(C)6–80.20.38–0.50
UCC(S)62–761.6–2.01.30–1.56UGC(C)24–260.6–0.71.50–1.63
UCA(S)117–1223.1–3.22.40–2.55UGA(W)75–962–2.51.47–1.88
UCG(S)6–120.2–0.30.13–0.25UGG(W)6–270.2–0.70.12–0.53
CCU(P)33–500.9–1.30.64–1.01CGU(R)4–90.1–0.20.25–0.56
CCC(P)34–560.9–1.50.68–1.09CGC(R)10–190.3–0.50.63–1.19
CCA(P)99–1142.6–3.01.95–2.22CGA(R)38–401–1.12.38–2.50
CCG(P)2–170.1–0.40.04–0.33CGG(R)3–50.10.19–31
ACU(T)55–581.4–1.50.69–0.75AGU(S)12–170.3–0.40.25–0.36
ACC(T)87–942.3–2.51.12–1.19AGC(S)36–430.9–1.10.75–0.88
ACA(T)153–1594.0–4.21.97–2.01AGA(*)000
ACG(T)7–120.2–0.30.09–0.15AGG(*)000
GCU(A)39–521.0–1.40.60–0.80GGU(G)37–411–1.10.69–0.75
GCC(A)116–1343.1–3.51.81–2.05GGC(G)50–651.3–1.70.93–1.19
GCA(A)78–822.1–2.21.19–1.28GGA(G)58–861.5–2.31.06–1.61
GCG(A)8–110.2–0.30.12–0.17GGG(G)41–541.1–1.40.75–0.99
RSCU: relative synonymous codon usage. * Termination codon.
Publisher’s Note: MDPI stays neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Share and Cite

MDPI and ACS Style

Lamelas, L.; Aleix-Mata, G.; Rovatsos, M.; Marchal, J.A.; Palomeque, T.; Lorite, P.; Sánchez, A. Complete Mitochondrial Genome of Three Species of the Genus Microtus (Arvicolinae, Rodentia). Animals 2020, 10, 2130. https://doi.org/10.3390/ani10112130

AMA Style

Lamelas L, Aleix-Mata G, Rovatsos M, Marchal JA, Palomeque T, Lorite P, Sánchez A. Complete Mitochondrial Genome of Three Species of the Genus Microtus (Arvicolinae, Rodentia). Animals. 2020; 10(11):2130. https://doi.org/10.3390/ani10112130

Chicago/Turabian Style

Lamelas, Luz, Gaël Aleix-Mata, Michail Rovatsos, Juan Alberto Marchal, Teresa Palomeque, Pedro Lorite, and Antonio Sánchez. 2020. "Complete Mitochondrial Genome of Three Species of the Genus Microtus (Arvicolinae, Rodentia)" Animals 10, no. 11: 2130. https://doi.org/10.3390/ani10112130

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