Preprint
Article

Characterization of the Complete Mitochondrial Genome of Leptobotia elongata and Its Phylogenetic Implications in Cobitidae

Altmetrics

Downloads

86

Views

38

Comments

0

A peer-reviewed article of this preprint also exists.

Submitted:

25 October 2023

Posted:

31 October 2023

You are already at the latest version

Alerts
Abstract
Elongate loach (Leptobotia elongata) is an endemic fish in China. Previous studies have provided some insights into the mitochondrial genome composition, and the phylogenetic relationships of L. elongata inferred using protein-coding genes (PCGs). However, the detailed information about is limited. Therefore, in this study, we sequenced the complete mitochondrial genome of L. elongata and analyzed its structural characteristics. The PCGs and mitochondrial genome were used for selective stress analysis and genomic comparative analysis respectively. The complete mitochondrial genome of the L. elongata, together with those of 36 Cyprinidae species, was used to infer the phylogenetic relationships of the Cobitidae family through maximum likelihood (ML) reconstruction. The results showed that the genome sequence has a full length of 16,591 bp, which includes 13 PCGs, 22 transfer RNA genes (tRNA), two ribosomal RNA genes (rRNA), and two non-coding regions (CR D-loop and light chain sub-chain replication origin OL). Overall, L. elongata shared the same gene arrangement and composition of the mitochondrial genes with other teleost fishes. The Ka/Ks ratios of all mitochondrial PCGs were less than 1, indicating that all the PCGs were evolving under purifying selection. Genome comparison analyses showed a significant sequence homology of species of Leptobotia. A significant identity between L. elongata and the other 5 Leptobotia species was observed in the visualization result, except for L. mantschurica, which lacked the tRNA-Arg gene and had a shorter tRNA-Asp gene. The phylogenetic tree revealed that the Cobitidae species examined here can be grouped into two clades, with L. elongata forming a sister relationship with L. microphthalma. This study could provide additional inferences for a better understanding of the phylogenetic relationships among Cobititdae species.
Keywords: 
Subject: Biology and Life Sciences  -   Aquatic Science

1. Introduction

Elongate loach (Leptobotia elongata), belonging to Cobitidae of Cypriniformes, is indigenous to the middle and upper reaches of the Yangtze River in China [1]. It is characterized by rapid growth and exceptional ornamental value [2,3]. However, the wild population resources of L. elongata have experienced a significant decline since the 1980s due to overfishing, dam construction, destruction of feeding, and spawning grounds [4]. As a result, it has been classified as vulnerable grade (VU) in China Red Book of Endangered Animals-Fish [5].
The family Cobitidae was originally proposed by Regan [6]. In this family, extensive research focused on morphological characteristics and mitochondrial genes has been conducted for over a century [7,8,9,10,11]. Currently, many scientists tend to divide Cobitidae into three subfamilies: Nemacheilinae, Botiinae, and Cobitinae [12]. In order to maintain consistency between the phylogenetic relationship and the natural classification of Cobitidae fishes, Tang et al [13] elevated these three subfamilies to the family level, which aligns with the classification of Liu et al [9]. As the second-largest group of Cypriniformes, Cobitidae is a key element in resolving the phylogenetic relationships of Cypriniformes. Investigating the phylogenetic relationships of the L. elongate, one of the youngest species in the Cobitidae, in family Cobitidae is beneficial to resolve the taxonomic ambiguity of Cobitidae fishes. Previous studies only focused on biological characteristics [14], artificial breeding [15], embryo development, and genetic diversity [16,17]. However, its research on the phylogenetic relationships of L. elongate is limited [18]. Therefore, a reevaluation of the phylogenetic relationships of L. elongata, involving additional genes and a broader range of taxa, could provide more data for the conservation of L. elongata's wild population resources.
Mitochondrial DNA (mtDNA) is present in the cells of all eukaryotes and possesses several genetic characteristics. It is primarily inherited maternally and exhibits conservation of coding regions, rapid evolution of the control region (CR), a high mutation rate, and a relatively independent genetic transcription system [19,20]. Compared to nuclear genes, mtDNA evolves at a faster rate, allowing for a more accurate representation of phylogenetic relationships. Therefore, mtDNA is widely utilized as a molecular marker in phylogenetic studies [21,22,23]. In fish phylogeny research, genes such as cytochrome b (cytb), cytochrome oxidase (cox), and 16SrRNA are commonly employed at species-to-family level [22,24,25]. However, when investigating higher taxonomic categories, relying solely on a single mitochondrial gene may lead to misleading phylogenetic data due to limited information capacity and homogenization effects [26]. In contrast, utilizing the complete mitochondrial genome could provide a more comprehensive set of phylogenetic information [27].
In this study, we sequenced the mitochondrial genome, analyzed the structural information of L. elongata, and compared the structures and complete mitochondrial genome with some of the determined Leptobotia species. Additionally, we reconstructed phylogenetic trees using complete mitochondrial genome sequences to analyze the evolutionary relationships L. elongata in the Cobitidae family. These study might provide futher insight into the structural of L. elongate, and improve understanding of evolutionary relationships of L. elongate within the Cobitidae.

2. Materials and Methods

2.1. Sample Collection and DNA Extraction

The L. elongata sample was collected from Chengdu, Sichuan Province, China, in October 2020. The pectoral fin was collected and stored in 95% ethanol at −20°C. Genomic DNA was isolated from the pectoral fin using the phenol-chloroform method and DNA degradation and contamination were estimated using agarose gel electrophoresis.

2.2. Mitochondrial Genome Sequencing and Assembly

After qualifying the DNA sample, the DNA was mechanically fragmented using ultrasonic interruption. The fragmented DNA underwent fragment purification, end-repair, addition of A at the 3' end, connection of sequencing adapters, and selection of fragments of different sizes using agarose gel electrophoresis. PCR amplification was then performed to generate a sequencing library [18]. The qualified PCR products were sequenced on the Illumina HiSeq 2500 platform.
Prior to assembly, low-quality data, including the reads of average quality value< 5 or N content>5, were filtered using Fasta software (version 0.20.0), and the sequences linker and primer sequence were trimmed from the reads. The mitochondrial genome assembly was carried out using the following methodology. First, clean reads were assembled using SPAdes (version 3.10) [28]. Second, contigs were connected to generate scaffold sequences using SSPACE (version 2.0), and Gaps in the scaffold sequences were filled using Gapfiller (version 2.1.1) until a complete pseudo genome sequence was assembled. Lastly, the sequencing results were mapped onto the assembled pseudo genome sequence to identify and correct any incorrect bases, and the complete mitochondrial circular genome sequence was obtained by coordinate remaking.

2.3. Mitochondrial Genome Annotation and Analysis

The newly assembled sequences were annotated in the Mitos web server (http://mitos2.bioinf.uni-leipzig.de) [29] with the following parameters: E-value Exponent=5, Maximum Overlap=100, ncRNA overlap=100. The annotation results were then compared with those of closely related species. Finally, after manual correction, the final annotation results were obtained.
The secondary structure of tRNAs was obtained from the annotation results. The circular map of the mitochondrial genome was generated using OGDRAW (version 1.3.1) [30]. The relative synonymous codon usage (RSCU) values were analyzed with MEGA (version 6.0). The mitochondrial genome skew values were calculated using the following formula: ATskew=(A-T)/(AT); GCskew=(G-C)/(GC) [31]. Mafft (version 7. 310) [32] software was used for gene sequences comparative between L. elongata and six Leptobotia fishes (Leptobotia mantschurica, Leptobotia taeniops, Leptobotia microphthalma, Leptobotia rubrilabris, Leptobotia punctata, and Leptobotia pellegrini), and the evolutionary rate (Ka/Ks, ω) was calculated using KaKs_Calculator (version 2. 0) [33]. If the evolutionary rate is equal to 1, >1, or, <1, the PGCs are expected to be under no selection, positive selective constraint (purifying selection), or diversifying selection. The mitochondrial genome structure was compared between L. elongata and six Leptobotia fish species in CGVIEW server [34] with default parameters (http://stothard.afns.ualberta.ca/cgview_server/), and the alignment results were visualized using mauve software (version 2.4.0).

2.4. Phylogenetic Analyses

The phylogenetic tree was reconstructed using the complete mitochondrial genome sequences of 37 Cypriniformes species, with Myxocyprinus asiaticus and Danio rerio used as outgroups (Table 1). All the genome sequences were set to the same start points in the circular sequence. Multiple sequence alignment was performed in MAFFT software (version 7.42) with auto model, and the alignment sequences were trimmed using trimAl (version 1.4. rev15). Subsequently, the RaxML (version 8.2.0) software was used to conduct the rapid bootstrap analysis (bootstrap=1000) to construct the maximum likelihood evolution tree.

3. Result and Discussion

3.1. Mitochondrial Structural Characteristics

The complete mitochondrial genome of L. elongata was obtained through high-throughput sequencing technology, with a total length of 16,591bp (Figure 1). It consists of 37 typical animal mitochondrial genes, including 22 tRNA genes, 13 PCGs, 2 rRNA genes, and two non-coding regions (D-Loop and OL). Among the mitochondrial genes, nine genes (trnQ, trnP, trnE, nad6, trnS2, trnY, trnC, trnN, trnA) were encoded by the light (L) strand, while the remaining genes are encoded by the heavy (H) strand. The arrangement and content of the mitochondrial genome in L. elongata were similar to those reported in teleost fishes [20,35,36]. The entire base composition of the L. elongata mitochondrial genes is as follows (Table 2): 30.79 %A, 24.77 %T, 16.17 %G, and 28.27 %C, and the AT and GC percentages are 55.56 % and 44.44 % respectively, which results in a positive skew value for AT and a subtractive skew value for CG. It was suggested that the occurrence of A and C bases was more frequent in the genome. Previous studies have shown that the bias in base composition plays a crucial role in the replication and transcription of mitochondrial genomes [37].

3.2. Protein Coding Genes (PCGs)

The PCGs account for 68.89% of the total length of the L. elongata mitochondrial genome. As expected (Table 3), most PCGs started with the regular codon ATG, except for the cox1, started with GTG. Among the PCGs, there were 7 genes that shared the complete stop codon TAA. While six genes shared incomplete stop codons (TA- or T--), which existed in many teleostean as numerous studies: L. microphthalma with seven incomplete stop codons [38], Cobitis macrostigma with seven incomplete stop codons [39], Pelteobagrus fulvidraco with five incomplete stop codons [20], Parabotia kiangsiensis with three incomplete stop codons [40], etc. The presence of tRNA sequences at the 3' end of these genes is responsible for the incomplete stop codons [41], and these incomplete stop codons can be converted to TAA through post-transcriptional polyadenylation [42].
Three overlapping regions between certain PCGs: ATPase8-ATPase6, ND4-ND4L, and ND5-ND6 were also identified in this study. These overlapping regions were 4-10bp in length, with the largest overlapping occurring between ATP8 and ATP6, which was common among Cobitidae species [43]. These overlapping regions contribute to the variation in mitochondrial genome length among closely related species [44]. The relative synonymous codon usage (RSCU) values of PCGs are revealed in Table 4 and Figure 2. In the protein-coding region, a total of 2012 codons were used. According to the degeneracy of codons, serine and leucine were encoded by six codons, while the remaining amino acids were encoded by either four or two codons. In the coded passwords, CUA (leucine), AUU (isoleucine), GCC (Aminopropanoic), and GCA (Aminopropanoic) are the most common, while AAA (Lysine) and CUA (leucine) have the highest RSCU values. Therefore, PCGs preferred the codons using adenine at the third codon. The codon usage varied between different species, which was more prominent between species with further phylogenetic relationship [45]. It is relevant to gene length, mutation bias, GC composition, amino acid composition, tRNA abundance, and translational selection [46,47,48,49,50,51].

3.3. Genome Comparative Analysis

The nonsynonymous substitution ratio (Ka) and synonymous substitution ratio (Ks) were calculated to evaluate selective pressures during the evolutionary process of PCGs among Leptobotia species. It was shown that the average Ka was similar among the 6 fishes (0.0089-0.0114), with nd5 exhibiting the highest average Ka (Figure 3A; Table 5), indicating that it might be under positive selection across species. The Ks of Leptobotia microphthalma was significantly lower than the other species (Figure 3B; Table 6). There were more synonymous substitutions per synonymous sites in nd4 and atp6, exhibiting the high polymorphic nature of the genes in these fishes. nd4 has also been confirmed to be polymorphism among sharks [52] and blue-spotted maskray [45]. The Ka/Ks ratio (ω) is a means to examine molecular adaption [53,54], which could be used to estimate the evolutionary rate among Cobitidae species. In this study, the Ka/Ks ratios of all PGCs were less than 1, indicating that purifying selection possesses the leading role in the evolution of these PGCs (Figure 3C; Table 7). Therein, cox3 (0.0076) and nd4l (0.0087) were evolving under a strong purifying selection, whereas nd4 (0.0549), nd5 (0.0782), and nd2 (0.0784) were evolving under comparatively relaxed mutational constraints. Currently, selective pressure in mitochondrial PCGs has been studied poorly on other Cobitidae species [13,18,38,39,55,56,57]. While the same pattern of widespread purifying selection has been discovered in several other decapod crustaceans [58].
The comparison of the mitochondrial genome sequences between L. elongata and 6 Leptobotia species showed a significant sequence homology within the Leptobotia genus (Figure 4; Figure 5). L. elongata showed a higher identity with the other five species, except for L. microphthalma, which lacked the tRNA-Arg and a shorter tRNA-Asp, indicating that the arrangement of genes of Leptobotia species is comparatively conserved.

3.4. Ribosomal RNA and Transfer RNA Genes

The total length of rRNAs was 2638bp, with an AT skew value of 0.272 and a GC skew value of -0.095. The lengths of 12SrRNA and 16SrRNA were 955bp and 1683bp, respectively (Table 3). These rRNAs were located between tRNA-Phe and tRNA-Leu and are separated by tRNA-Val, which is consistent with the most reported teleost [59].
There were 22 tRNAs in the mitochondrial genome of L. elongata, with a total length of 1558bp. The AT content was 53.89% and AT skew value was 0.044. Each tRNA has a length of 66-76bp, with 14 encoded in the H chain and 8 encoded in the L chain. Most of the secondary structure of tRNA genes (Figure 6) in L. elongata were standard clover-shaped, except for trnS1, which lacked the DHU stem. It was very common to defect DHU stem in metazoan [42]. Additionally, there were 18 false GU pairings in the tRNA sequences of L. elongata. GU mismatch was frequently observed in teleost fishes and allowed for an expanded chemical and conformational diversity of double-stranded RNA. This diversity provided unique sites that were recognized by amino acids, contributing to a higher genetic diversity for L. elongate [60]. The base mismatch was essential for the secondary structure of tRNA and played a crucial role in the accurate translation of the genetic code. It also helped minimize errors during mRNA transcription [61].

3.5. Non-Coding Regions

Two common non-coding regions (OL and CR) were identified in the L. elongata mitogenome, the OL region was 31 bp in length and was located between tRNA-Asn and tRNA-Cys. The CR region was located between tRNA-Pro and tRNA-Phe, which is the longest no-coding region in the entire mitochondrial genome with a span distance of 926 bp. It plays a key role in the replication and transcription [62]. Similar to other vertebrates [21,63], the CR of L. elongata exhibited the highest AT content (67.39%) among all regions of the mitochondrial genome. The palindromic sequence motifs ‘tacat’ and ‘atgta’ were related to the termination of H strand replication found in the CR of L. elongata (Figure 6), which might complete the termination by forming a stable hairpin structure [64].

3.6. Phylogenetic Relationships

Based on the complete mitochondrial genome sequences of the L. elongata and 36 Cyprinidaes, the phylogenetic tree was constructed. It was shown that the entire phylogenetic tree was grouped into two major clades (Figure 7). The genus Cobitis, Pangio, Triplophysa, and Acanthocobitis formed one clade and matched the subfamily Cobitinae. The Cobitis and the Pangio were sister-lineage, the Triplophysa and the Acanthocobitis were sister-lineage, and the two sister-lineages were sister-lineages to each other. The other clade consisted of Yasuhikotakia, Sinibotia, Chromobotia, Botia, Parabotia, and Leptobotia, corresponding to the subfamily Botiinae. In the subfamily Botiinae, L. elongata was more closely related to L. microphthalma than other species.
As a diverse population, there was a controversy in the taxonomic relationship of the subfamily Cobitinae. This study exhibited a monophyly of the subfamily CobitinaeN, which consists of four clades. However, according to Liu et al [11], there were sisterhoods in many branches. Therefore, the species in Cobitinae can’t form a monophyletic group, and the classification of Cobitinae in our study is incomplete, and more taxon should be used in future studies.
It is generally considered that the subfamily Botiinae is a group with relatively clear taxonomic relationship. In this study, according to their respective genera separately, all individuals except L. mantschurica of the subfamily Botiinae were clustered into a common branch, which could be confirmed the monophyly of the subfamily Botiinae. In the previous study, the genera Botia was separated into a separate genus [7], and the genera Botia was divided into three subgenera: Sinibotia, Botia, and Hymenophysa [65]. Others did not further categorize these subgenra, but instead grouped them under the genus Botia [12,66]. In this study, subgenera Botia and subgenra sinibotia species were clustered separately and formed parallel branches with the species of other genera. Thus, the results supported that subgenera Botia and subgenra sinibotia should be raised to genus. Additionally, the phylogenetic tree showed that L. elongata and L. microphthalma formed a sister group, which together formed a sister group of other Leptobotia species. Acorrding to Li et al. [18], the L. elongata and L. mantschurica were classified as sister lineages using protein genome sequence to construct the phylogenetic tree, however, this study was analysed based on limited taxon sampling and mitochondrial genes, lacking of sufficient information of phylogenetic the L. elongata.
Slechtova et al. [67] suggested that the Leptobotia and Parabotia genera were monophyletic using only the Cytb and 12S. However, the phylogenetic tree in our study clearly showed that L. mantschurica was nested with Parabotia, indicating that Leptobotia and Parabotia genera were an unnatural group and not reciprocally monophyletic groups as previously hypothesized [13,66,68]. Additionally, there was obvious structural variation among the mitochondrial genome of L. mantschurica, compared with other Leptobotia species, proving that L. mantschurica formed a sister relationship with Parabotia fasciatus instead of Leptobotia species is reliable, which shared the same results with Tang et al. [13]. In the Parabotia species, part of the support vaule in the branch was low, suggestting that the phylogenetic relationships of these species haven‘t been solved well. Futher investigations should be performed to solve this problem.

4. Conclusion

In this study, we reported the complete mitogenome of L. elongata, the structural characteristics of the mitogenome of L. elongata were analyzed in detail, and the phylogenetic analyses of L. elongata were inferred using the complete mitogenome. The full length of the genome sequence was 16591 bp, and the arrangement of the L. elongata mitochondrial genome is similar to most teleost fishes. Almost all 13 PCGs showed the regular start codon ATG except gene cox1, which started with GTG. 6 PCGs have incomplete stop codons T--. 13PCGs were evolving under purifying selection, and the mitogenome shared the high identity with Leptobotia species. All the tRNA genes were standard clover-shaped except the lacking of DHU stem in trnS1. The phylogenetic analysis showed that L. elongata was more closely related to L. microphthalma than other species. L. mantschurica formed a sister relationship with Parabotia fasciatus, and the Leptobotia and Parabotia genera was polyphyletic. In this study, we first studied the selection pressure of complete PCGs in the L. elongata. Overall, we have a deeper understanding of the mitochondrial genome structure and phylogenetic analysis of L. elongata. However, exact information of about many Cobitidae fishes is still unkown. Extra taxon should be used for the phylogenetic research of Cobitidae in the future.

Author Contributions

Conceptualization, Z.K. and H.Y.; methodology, H.Y. and K.Z.; writing—original draft preparation, Z.K.; writing—review and editing, Z.K., M.H., H.Y., Z.L., H.L., T.J. and X.P. All authors have read and agreed to the published version of the manuscript.

Funding

This research was funded by supported by the National Training Program of Innovation and Entrepreneurship for Undergraduates of China (202110635024), and the Chongqing Alliance for Aquatic Science and Technology Innovation (CQFTIU2022-09).

Institutional Review Board Statement

The study was conducted in accordance with the Declaration of Helsinki and approved by the Animal Care and Use Committee of University of Southwest (Approval ID was 20190922, and the approval date was on 22 September 2019).

Informed Consent Statement

Not applicable.

Data Availability Statement

The datasets presented in this study were submitted to The National Center for Biotechnology Information (NCBI) database.

Acknowledgments

We would like to express our sincere thanks to Sichuan Fisheries Research Institute for their help in sample collecting.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Yuan, Y.C; Yang, H.J; Gong, S.Y; Liang, Y.Q. Threatened fishes of the world: Leptobotia elongata Bleeker, 1870 (Bottiinae). Environ Biol Fish. 2010, 87, 295–296. [Google Scholar] [CrossRef]
  2. Chen, K.G.; Wang, Z.J.; Yue, X.J. Study of the structurs of the digestive system in Leptobotia Elongata. Southwest Agric. Univ. 2002, 24, 1000–2642. [Google Scholar] [CrossRef]
  3. Tian, H.W.; Duan, X.B.; Xiong, X.; Luo, H.Q.; Liu, S.P; Chen, D.Q. Estimation of growth and population parameters of elongate loach (Leptobotia elongata) in the upper reaches of the Yangtze River. Yangtze Basin. 2013, 22, 1305–1312. [Google Scholar]
  4. Zhang, Y.; Cao, X.; Zou, Y.; Yan, Z.; Huang, Y.; Zhu, Y.; Gao, J. De novo gonad transcriptome analysis of elongate loach (Leptobotia elongata) provides novel insights into sex-related genes. Comp Biochem Phys D. 2022, 42, 100962. [Google Scholar] [CrossRef] [PubMed]
  5. Yue, P.; Chen, Y. China Red Data Book of Endangered Animals: Pisces; Science Press: Beijing, China, 1998; pp. 240–243. [Google Scholar]
  6. Regan, C.T. The classification of the teleostean fishes of the order Ostariophysi. I. Cyprinidae. Ann. Mag. Nat. Hist. 1911, 8, 13–32. [Google Scholar] [CrossRef]
  7. Hora, S.L. Classification, bionomics and evolution of homalopterid fishes. Mem. Indian. Mus. 1932, 12, 263–330. [Google Scholar]
  8. Berg, L.S. Classification of fishes both recent and fossil. Trav. Inst. Zool. Acad. Sci. USSR. 1940, 87–345. [Google Scholar] [CrossRef]
  9. Liu, H.Z.; Tzeng, C.S.; Teng, H.Y. Sequence variations in the mitochondrial DNA control region and their implications for the phylogeny of the Cypriniformes. Can. J. Zool. 2002, 80, 569–581. [Google Scholar] [CrossRef]
  10. Nalbant, T.T. Sixty million years of evolution. Part one: family Botiidae (Pisces: Ostariophysi: Cobitoidea). Trav. Mus. Hist. Nat, 44.
  11. Liu, H.Y; Cai, J.; Xie, Z.G.; Xiong, F.; Wang, Y.; Wang, Q.; Yu, J.X.; Zhai, D.D.; Xia, M.; Chen, Y.Y. DNA Barcodes for species identification and systematic evolution of cobitidae fish. Acta Agric. Univ. Jiangxiensis. 2020, 42, 766–777. [Google Scholar] [CrossRef]
  12. Liu, S.Q.; Mayden, R.L.; Zhang, J.B.; Yu, D.; Tang, Q.Y.; Deng, X.; Liu, H.Z. Phylogenetic relationships of the Cobitoidea (Teleostei: Cypriniformes) inferred from mitochondrial and nuclear genes with analyses of gene evolution. Gene. 2012, 508, 60–72. [Google Scholar] [CrossRef]
  13. Tang, Q.Y.; Liu, H.Z.; Mayden, R.; Xiong, B.X. Comparison of evolutionary rates in the mitochondrial DNA cytochrome b gene and control region and their implications for phylogeny of the Cobitoidea (Teleostei: Cypriniformes). Mol. Phylogenet.Evol. 2006, 39, 347–357. [Google Scholar] [CrossRef]
  14. Yuan, Q.; Wang, Y.; Liang, R.; Feng, J.; Kefeng, A. Field observations of the lethality characteristics of endangered and endemic fish under the stress of total dissolved gas supersaturation. River Res Appl. 2020. 38, 1156–1167. [CrossRef]
  15. Zhang, Y.H.; Ding, Y.; Zheng-Xuan, G.U.; Huang, X.X.; Wen, Y.Z.; Min, S. Artificial propagation and embryonic development observation of leptobotia elongata from Jinsha River and Yangtze River. HubeiAgric.Sci. 2018, 57, 104–107. [Google Scholar] [CrossRef]
  16. Liu, D.; Li, X.; Song, Z. No decline of genetic diversity in elongate loach (Leptobotia elongata) with a tendency to form population structure in the upper Yangtze River. Glob ecol conserv. 2020, 23, e01072. [Google Scholar] [CrossRef]
  17. Liu, D.Q.; Wu, J.Y.; Deng, L.J.; Gan, W.X.; Du, L.M.; Song, Z.B. Development of Microsatellite Markers for Leptobotia elongata (Cypriniformes: Cobitidae) Using 454 Sequencing and Cross-species Amplification. Pak J Zool. 2014, 4, 1147–1151. [Google Scholar]
  18. Li, P.; Yang, C.Z.; Tu, F.Y.; Liu, G.X. The complete mitochondrial genome of the Elongate loach Leptobotia elongata (Cypriniformes: Cobitidae). MtDNA. 2012, 23, 352–354. [Google Scholar] [CrossRef]
  19. Caccone, A.; Gentile, G.; Burns, C.E.; Sezzi, E.; Powell, J.R. Extreme difference in rate of mitochondrial and nuclear DNA evolution in a large ectotherm, Galápagos tortoises. Mol phylogenet evol. 2004, 31, 794–798. [Google Scholar] [CrossRef] [PubMed]
  20. Liu, Y.; Wu, P.D.; Zhang, D.Z.; Zhang, H.B.; Tang, B.P.; Liu, Q.N. Mitochondrial genome of the yellow catfish Pelteobagrus fulvidraco and insights into Bagridae phylogenetics. Genomics. 2018, 111, 1258–1265. [Google Scholar] [CrossRef] [PubMed]
  21. Wang, C.H.; Chen, Q.; Lu, G.Q.; Xu, J.W.; Yang, Q.L.; Li, S. Complete mitochondrial genome of the grass carp (Ctenopharyngodon idella, Teleostei): Insight into its phylogenic position within Cyprinidae. Gene. 2008, 424, 96–101. [Google Scholar] [CrossRef] [PubMed]
  22. Tang, K.L.; Agnew, M.K.; Hirt, M.V.; Sado, T.Y.; Schneider, L.M.; Freyhof, J.; Sulaiman, Z.; Swartz, E.; Vidthayanon, C.; Miya, M.; Saitoh, K.; Simons, A.M.; Wood, R.M.; Mayden, R.L. Systematics of the subfamily Danioninae (Teleostei: Cypriniformes: Cyprinidae). Mol phylogenet evol. 2010, 57, 189–214. [Google Scholar] [CrossRef]
  23. Muniyangd, N.; Raja, M.; Vikram, P. Genetic characterization of Bagarius species using cytochrome c oxidase I and cytochrome b genes. MtDNA. 2016, 27, 3781–3783. [Google Scholar] [CrossRef]
  24. Chen, D.X.; Chu, W.Y.; Liu, X.L.; Nong, X.X.; Li, Y.L.; Du, S.J.; Zhang, J. Phylogenetic studies of three sinipercid fishes (Perciformes: Sinipercidae) based on complete mitochondrial DNA sequences. MtDNA 2012, 23, 70–76. [Google Scholar] [CrossRef] [PubMed]
  25. Pavan-Kumar, A.; Raman, S.; Koringa, P.G.; Patel, N.; Chaudhari, A. Complete mitochondrial genome of threatened mahseer Tor tor (Hamilton 1822) and its phylogenetic relationship within Cyprinidae family. J. Genet. 2016, 95, 853–863. [Google Scholar] [CrossRef] [PubMed]
  26. Zou, X.H.; Ge, S. Conflicting gene trees and phylogenomics. J SYST EVOL. 2008, 46, 795–807. [Google Scholar] [CrossRef]
  27. Rieppel, O. ‘Total evidence’ in phylogenetic systematics. Biol Philos. 2009, 24, 607–622. [Google Scholar] [CrossRef]
  28. Bankevich, A.; Nurk, S.; Antipov, D.; Gurevich, A.A.; Dvorkin, M.; Kulikov, A.S.; Lesin, V.M.; Nikolenko, S.I.; Pham, S.; Prjibelski, A.D.; Pyshkin, A.V.; Sirotkin, A.V.; Vyahhi, N.; Tesler, G.; Alekeseyev, A.M.; Pevzner, P.A. SPAdes: a new genome assembly algorithm and its applications to single-cell sequencing. J Comput Biol. 2012, 19, 77–455. [Google Scholar] [CrossRef] [PubMed]
  29. 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] [PubMed]
  30. Lohse, M.; Drechsel, O.; Bock, R. OrganellarGenomeDRAW (OGDRAW): a tool for the easy generation of high-quality custom graphical maps of plastid and mitochondrial genomes. Curr Genet. 2007. 52, 74–267. [CrossRef]
  31. Perna, N.T.; Kocher, T.D. Patterns of nucleotide composition at fourfold degenerate sites of animal mitochondrial genomes. J Mol Evol. 1995, 41, 353–358. [Google Scholar] [CrossRef]
  32. Katoh, K.; Standley, D.M. MAFFT multiple sequence alignment software version 7: improvements in performance and usability. Mol Biol Evol. 2013, 30, 80–772. [Google Scholar] [CrossRef]
  33. Wang, D.P.; Zhang, Y.B.; Zhang, Z.; Zhu, J.; Yu, J. KaKs_Calculator 2.0: a toolkit incorporating gamma-series methods and sliding window strategies. Genom Proteom Bioinf. 2010, 8, 77–80. [Google Scholar] [CrossRef]
  34. Stothard, P.; Grant, R.J.; Domselaar, G.V. Visualizing and comparing circular genomes using the CGView family of tools. Brief Bioinform. 2019, 20, 1576–1582. [Google Scholar] [CrossRef] [PubMed]
  35. Zhong, L.Q.; Wang, M.H.; Li, D.; Tang, S.; Zhang, T.Q.; Bian, W.J.; Chen, X.H. Complete mitochondrial genome of Odontobutis haifengensis (Perciformes, Odontobutiae): A unique rearrangement of tRNAs and additional non-coding regions identified in the genus Odontobutis. Genomics. 2018, 110, 382–388. [Google Scholar] [CrossRef] [PubMed]
  36. Prabhu, V.R.; Singha, H.S.; Kumar, R.G.; Gopalakrishnan, A.; Nagarajan, M. Characterization of the complete mitochondrial genome of Barilius malabaricus and its phylogenetic implications. Genomics. 2020, 112, 2154–2163. [Google Scholar] [CrossRef] [PubMed]
  37. Wei, S.; Shi, M.; Chen, X.; Sharkey, M.; Achterberg, C.; Ye, G.; He, J. New views on strand asymmetry in insect mitochondrial genomes. PLoS ONE. 2010, 5, e12708. [Google Scholar] [CrossRef] [PubMed]
  38. Tian, H.; Wang, D.; Jia, X.; Duan, X.; Chen, D. The mitogenome of Leptobotia microphthalma (Teleostei, Cypriniformes: Cobitidae). MtDNA 2014, 25, 261–262. [Google Scholar] [CrossRef]
  39. Yang, N.; Li, Y.; Liu, Z.; Chen, Q.; Shen, Y. The complete mitochondrial genome of Cobitis macrostigma (Cypriniformes: Cobitidae: Cobitinae) and a phylogenetic implication for its closely related species. Biologia. 2020, 75, 393–399. [Google Scholar] [CrossRef]
  40. Ma, Q.; Zhang, T.L.; Chen, L.; Tang, Q.Y. The Complete Mitochondrial Genomes of Parabotia Kiangsiensis (Cypriniformes: Botiidae). MtDNA. 2020, 5, 3629–3631. [Google Scholar] [CrossRef] [PubMed]
  41. Ojala, D.; Montoya, J.; Attardi, G. tRNA punctuation model of RNA processing in human mitochondria. Nature, 1981, 290, 470–474. [Google Scholar] [CrossRef]
  42. Garey, J.R.; Wolstenholme, D.R. Platyhelminth mitochondrial DNA: Evidence for early evolutionary origin of a tRNAserAGN that contains a dihydrouridine arm replacement loop, and of serine-specifying AGA and AGG codons. J Mol Evol. 1989, 28, 374–387. [Google Scholar] [CrossRef]
  43. Sharma, A.; Siva, C.; Ali, S.; Sahoo, P.K.; Sarma, D. The complete mitochondrial genome of the medicinal fish, Cyprinion semiplotum: Insight into its structural features and phylogenetic implications. INT J BIOL MACROMOL. 2020, 164, 939–948. [Google Scholar] [CrossRef]
  44. Lee, Y.S.; Prakash Patil, M.; Kim, J.O.; Lee, Y.J.; Seo, Y.B.; Kim, J.K.; Suryawanshi, R.K. The Complete Mitochondrial Genome of the Fivespot Flounder, Pseudorhombus pentophthalmus (Pleuronectiformes: Paralichthyidae), from Korea and Its Phylogenetic Analysis. Fishes. 2023, 8, 150, doi103390/fishes8030150. [Google Scholar] [CrossRef]
  45. Pavan-Kumar, A.; Singh, S.; Mishra, A.; Suman, S.; Gireesh-Babu, P.; Chaudhari, A.; Shen, K.N.; Borsa, P. Characterization of mitochondrial genome of Indian Ocean blue-spotted maskray, Neotrygon indica and its phylogenetic relationship within Dasyatidae Family. Int. J. Biol. Macromol. 2022, 223, 458–467. [Google Scholar] [CrossRef]
  46. William, B. Codon distribution in vertebrate genes may be used to predict gene length. J MOL BIOL. 1987, 197, 379–388. [Google Scholar] [CrossRef]
  47. Sueoka, N. Two Aspects of DNA Base Composition: G+C Content and Translation-Coupled Deviation from Intra-Strand Rule of A=T and G=C. J MOL EVOL. 1999, 49, 49–62. [Google Scholar] [CrossRef]
  48. Akashi, H. Translational selection and yeast proteome evolution. Genetics. 2003, 164, 1291–1303. [Google Scholar] [CrossRef]
  49. Bernardi, G.; Bernardi, G. Compositional Constraints and Genome Evolution. J. Mol. Evol. 1986, 24, 1–11. [Google Scholar] [CrossRef]
  50. D'Onofrio, G.; Mouchiroud, D.; Aïssani, B.; Gautier, C.; Bernardi, G. Correlations between the compositional properties of human genes, codon usage, and amino acid composition of proteins. J Mol Evol. 1991, 32, 504–510. [Google Scholar] [CrossRef]
  51. Foster, P.G.; Jermiin, L.S.; Hickey, D.A. Nucleotide Composition Bias Affects Amino Acid Content in Proteins Coded by Animal Mitochondria. J Mol Evol. 1997, 44, 282–288. [Google Scholar] [CrossRef] [PubMed]
  52. Dudgeon, C.L.; Broderick, D.; Ovenden, J.R. IUCN classification zones concord with, but underestimate, the population genetic structure of the zebra shark Stegostoma fasciatum in the Indo-West Pacific. Mol. Ecol. 2009, 18, 248–261. [Google Scholar] [CrossRef] [PubMed]
  53. Hurst, L.D. The Ka/Ks ratio: diagnosing the form of sequence evolution. Trends Genet. 2002, 18, 486–487. [Google Scholar] [CrossRef]
  54. Yang, Z.; Bielawski, J.P. Statistical methods for detecting molecular adaptation. Trends Ecol. Evol. 2000, 15, 496–503. [Google Scholar] [CrossRef]
  55. He, Y.; Chen, Y.; Yan, J.; Page, L.M. Phylogenetic analysis and osteological comparison of the cave-dwelling spined loach, Bibarba parvoculus (Cypriniformes: Cobitidae), and its surface congener. Zool. J. Linn. Soc. 2021, 191, 1059–1074. [Google Scholar] [CrossRef]
  56. Shen, Y.J.; Wang, J.; Zhang, F.B. Complete Mitochondrial Genome of Parabotia bimaculata (Cypriniformes: Cobitidae: Botiinae), an Endemic Riverine Loach in China and Phylogenetic Analysis for Botiinae. Thalassas. 2020, 36, 387–393. [Google Scholar] [CrossRef]
  57. Yang, X.G.; Lian, Y.X.; Chen, M.M.; Li, X.Q.; Yu, D.P. Characterization and phylogenetic analysis of the complete mitochondrial genome of sun loach (Yasuhikotakia eos). MtDNA. 2021, 6, 13–14. [Google Scholar] [CrossRef]
  58. Baeza, J.A. The complete mitochondrial genome of the Caribbean spiny lobster Panulirus argus. Sci. Rep. 2018, 8, 1–10. [Google Scholar] [CrossRef]
  59. Inoue, J.G.; Miya, M.; Tsukamoto, K.; Nishida, M. Complete mitochondrial DNA sequence of the Japanese sardine Sardinops melanostictus. Fish Sci. 2010, 66, 924–932. [Google Scholar] [CrossRef]
  60. Varani, G.; McClain, W.H. The G·U wobble base pair: A fundamental building block of RNA structure crucial to RNA function in diverse biological systems. EMBO Rep. 2000, 1, 18–23. [Google Scholar] [CrossRef]
  61. Chen, L.; Lin, Y.F.; Xiao, Q.; Lin, Y.; Du, Y.; Lin, C.X.; Ward-Fear, G.; Hu, C.C.; Qu, Y.F.; Li, H. Characterization of the complete mitochondrial genome of the many-lined sun skink (Eutropis multifasciata) and comparison with other Scincomorpha species. Genomics. 2021, 113, 2526–2536. [Google Scholar] [CrossRef] [PubMed]
  62. Gao, Y.C.; Zhang, J.; Wang, Q.H.; Liu, Q.N.; Tang, B.P. The Complete Mitochondrial Genome of Box Tree Moth Cydalima perspectalis and Insights into Phylogenetics in Pyraloidea. Animals. 2023, 13, 1045. [Google Scholar] [CrossRef] [PubMed]
  63. Wang, X.; Wang, J.; He, S.; Mayden, R.L. The complete mitochondrial genome of the Chinese hook snout carp Opsariichthys bidens (Actinopterygii: Cypriniformes) and an alternative pattern of mitogenomic evolution in vertebrate. Gene. 2007, 399, 11–19. [Google Scholar] [CrossRef]
  64. Broughton, R.E.; Milam, J.E.; Roe, B.A. The Complete Sequence of the Zebrafish (Danio rerio) Mitochondrial Genome and Evolutionary Patterns in Vertebrate Mitochondrial DNA. Genome Res. 2001, 11, 1958–1967. [Google Scholar] [CrossRef]
  65. Fang, P.W. Study on Botoid Fishes of China. Sinensia 1936, 7. [Google Scholar]
  66. Tang, Q.Y. molecular phylogeny of the cobitoidea. HZAU. 2005.
  67. Slechtova, V.; Bohlen, J.; Freyhof, J.; Rab, P. Molecular phylogeny of the Southeast Asian freshwater fish family Botiidae (Teleostei: Cobitoldea) and the origin of polyploidy in their evolution. Mol. Phylogenet. Evol. 2006, 39, 529–541. [Google Scholar] [CrossRef]
  68. Tang, Q.Y.; Yu, D.; Liu, H.Z. Leptobotia zebra Should Be Revised as Sinibotia zebra (Cypriniformes: Botiidae). Zoological Research. 2008, 29, 1–9. [Google Scholar] [CrossRef]
Figure 1. Mitochondrial genome map of L. elongata.
Figure 1. Mitochondrial genome map of L. elongata.
Preprints 88733 g001
Figure 2. The relative synonymous codon usage (RSCU) in the mitogenome of L. elongata.
Figure 2. The relative synonymous codon usage (RSCU) in the mitogenome of L. elongata.
Preprints 88733 g002
Figure 3. Non-ynonymous (A) and synonymous (B) substitutional rates and the ratios of KaKs (C) of protein coding genes of L. elongata.
Figure 3. Non-ynonymous (A) and synonymous (B) substitutional rates and the ratios of KaKs (C) of protein coding genes of L. elongata.
Preprints 88733 g003
Figure 4. The comparative circle diagram of the genomestructure of Leptobotia species.
Figure 4. The comparative circle diagram of the genomestructure of Leptobotia species.
Preprints 88733 g004
Figure 5. The visualized results of the genome comparison of L.elongata.
Figure 5. The visualized results of the genome comparison of L.elongata.
Preprints 88733 g005
Figure 6. Putative secondary structure of L. elongata tRNA.
Figure 6. Putative secondary structure of L. elongata tRNA.
Preprints 88733 g006
Figure 7. Compositional features of the control region of the L. elongata mitochondrial genome. Palindromic motif sequence “TACAT’ and ‘ATGTA’ are marked in yellow and purple respectively.
Figure 7. Compositional features of the control region of the L. elongata mitochondrial genome. Palindromic motif sequence “TACAT’ and ‘ATGTA’ are marked in yellow and purple respectively.
Preprints 88733 g007
Figure 8. ML tree with boostrap values on the nodes constructed by using complete mitogenome sequences of L. elongata.
Figure 8. ML tree with boostrap values on the nodes constructed by using complete mitogenome sequences of L. elongata.
Preprints 88733 g008
Table 1. Taxonomic information and GeneBank assession numbers of all species used in phylogenetic analysis.
Table 1. Taxonomic information and GeneBank assession numbers of all species used in phylogenetic analysis.
Family Genus Species Assession
Myxocyprinae Myxocyprinus Myxocyprinus asiaticus NC_006401.1
Cyprinidae Danio Danio rerio NC_002333.2
Cobitidae Leptobotia Leptobotia microphthalma KY307846.1
Leptobotia Leptobotia elongata
Leptobotia Leptobotia pellegrini NC_031602.1
Leptobotia Leptobotia taeniops AP013304.1
Leptobotia Leptobotia rubrilabris KY307847.1
Leptobotia Leptobotia punctata MH644033.1
Leptobotia Leptobotia mantschurica AB242170.1
Parabotia Parabotia fasciata AP011437.1
Parabotia Parabotia banarescui NC_026127.1
Parabotia Parabotia kiangsiensis MT850132.1
Parabotia Parabotia lijiangensis MT323118.1
Botia Botia udomritthiruji AP011349.1
Botia Botia lohachata KP729183.1
Chromobotia Chromobotia macracanthus AB242163.1
Yasuhikotakia Yasuhikotakia morleti NC_031600.1
Yasuhikotakia modesta KY131962.1
Sinibotia Sinibotia robusta NC_027853.1
Sinibotia pulchra KT362179.1
Sinibotia reevesae NC_030322.1
Sinibotia superciliaris JX683724.1
Sinibotia superciliaris KY307848.1
Acanthocobitis Acanthocobitis botia AP012139.1
Acanthocobitis botia AP012138.1
Triplophysa Triplophysa pappenheimi NC_033972.1
Triplophysa siluroides NC_024611.1
Pangio Pangio kuhlii NC_031599.1
Pangio cuneovirgata NC_031594.1
Cobitis Cobitis lutheri NC_022717.1
Cobitis melanoleuca granoei NC_023473.1
Cobitis nalbanti MH349461.1
Cobitis elongatoides NC_023947.1
Cobitis hankugensis MN841275.1
Cobitis macrostigma MK156771.1
Cobitis sinensis NC_007229.1
Cobitis matsubarai NC_029441.1
Table 2. Nucleotide composition and skewness values of L. elongata mitogenome.
Table 2. Nucleotide composition and skewness values of L. elongata mitogenome.
Leptobotia_elongata Size (bp) A% T% G% C% A+T% G+C% AT-skew GC-skew
Mitogenome 16591 30.79 24.77 16.17 28.27 55.56 44.44 0.108 -0.272
PCGs 11430 28.56 26.77 15.55 29.13 55.33 44.67 0.032 -0.304
tRNAs 1558 28.18 25.8 23.49 22.53 53.98 46.02 0.044 0.021
rRNAs 2638 34.04 19.48 21.04 25.44 53.53 46.47 0.272 -0.095
Dloop 926 35.64 31.75 13.71 18.9 67.39 32.61 0.058 -0.159
Table 3. Summary of L. elongata mitogenome.
Table 3. Summary of L. elongata mitogenome.
Position codon
Gene stand From To size Intergenic length start stop
tRNA-phe H 1 69 69 0
12s-rRNA H 70 1024 955 0
tRNA-val H 1025 1096 72 0
16S-rRNA H 1097 2779 1683 0
tRNA-leu H 2780 2854 75 0
nd1 H 2855 3829 975 0 ATG TAA
tRNA-ile H 3838 3909 72 8
tRNA-gln L 3908 3978 71 -2
tRNA-met H 3980 4048 69 1
nd2 H 4049 5094 1046 0 ATG TA-
tRNA-trp H 5095 5163 69 0
tRNA-ala L 5166 5234 69 2
tRNA-asn L 5236 5308 73 1
OL L 5310 5340 31 1
tRNA-cys L 5339 5404 66 -2
tRNA-tyr L 5406 5476 71 1
cox1 H 5478 7028 1551 1 GTG TAA
tRNA-ser L 7030 7100 71 1
tRNA-asp H 7103 7174 72 2
cox2 H 7188 7878 691 13 ATG T--
tRNA-lys H 7879 7954 76 0
ATPase8 H 7956 8123 168 1 ATG TAA
ATPase6 H 8114 8797 684 -10 ATG TAA
cox3 H 8797 9581 785 1 ATG TA-
tRNA-gly H 9582 9653 72 0
nd3 H 9654 10002 349 0 ATG T--
tRNA-arg H 10003 10072 70 0
nd4l H 10073 10369 297 0 ATG TAA
nd4 H 10363 11744 1382 -7 ATG TA-
tRNA-his H 11745 11814 70 0
tRNA-ser H 11815 11881 67 0
tRNA-leu H 11883 11955 73 1
nd5 H 11956 13794 1839 0 ATG TAA
nd6 L 13791 14312 522 -4 ATG TAA
tRNA-glu L 14313 14381 69 0
Cytb H 14386 15526 1141 4 ATG T--
tRNA-thr H 15527 15598 72 0
tRNA-pro L 15597 15666 70 -2
CR H 15666 16591 926 0
Table 4. Relative synonymous codon usage and codon numbers of Leptobotia_elongata mi- tochondrial PCGs.
Table 4. Relative synonymous codon usage and codon numbers of Leptobotia_elongata mi- tochondrial PCGs.
Codon No. RSCU Codon No. RSCU Codon No. RSCU
UAA(*) 7 1 AAA(K) 38 1.8536 CGG(R) 5 0.5
GCA(A) 68 1.4468 AAG(K) 3 0.1464 CGU(R) 5 0.5
GCC(A) 94 2 CUA(L) 123 2.271 AGC(S) 24 1.161
GCG(A) 3 0.064 CUC(L) 57 1.0524 AGU(S) 4 0.1938
GCU(A) 23 0.4892 CUG(L) 22 0.4062 UCA(S) 50 2.4192
UGC(C) 7 1.1666 CUU(L) 53 0.9786 UCC(S) 22 1.0644
UGU(C) 5 0.8334 UUA(L) 51 0.9414 UCG(S) 1 0.0486
GAC(D) 25 1.3158 UUG(L) 19 0.351 UCU(S) 23 1.113
GAU(D) 13 0.6842 AUA(M) 60 2.0226 ACA(T) 72 1.87
GAA(E) 36 1.44 AUG(M) 28 0.9438 ACC(T) 59 1.5324
GAG(E) 14 0.56 GUG(M) 1 0.0336 ACG(T) 1 0.026
UUC(F) 63 0.9618 AAC(N) 54 1.4594 ACU(T) 22 0.5716
UUU(F) 68 1.0382 AAU(N) 20 0.5406 GUA(V) 52 1.5524
GGA(G) 48 1.4116 CCA(P) 47 1.6348 GUC(V) 23 0.6864
GGC(G) 31 0.9116 CCC(P) 47 1.6348 GUG(V) 24 0.7164
GGG(G) 39 1.1472 CCG(P) 6 0.2088 GUU(V) 35 1.0448
GGU(G) 18 0.5296 CCU(P) 15 0.5216 UGA(W) 45 1.6364
CAC(H) 42 1.7142 CAA(Q) 45 1.9148 UGG(W) 10 0.3636
CAU(H) 7 0.2858 CAG(Q) 2 0.0852 UAC(Y) 26 0.963
AUC(I) 59 0.792 CGA(R) 23 2.3 UAU(Y) 28 1.037
AUU(I) 90 1.208 CGC(R) 7 0.7
Table 5. Non-synonymous substitution rate of mitochondrial genes in Leptobotia species.
Table 5. Non-synonymous substitution rate of mitochondrial genes in Leptobotia species.
Species Leptobotia mantschurica Leptobotia taeniops Leptobotia microphthalma Leptobotia rubrilabris Leptobotia punctata Leptobotia pellegrini Average
nd4 0.0166 0.0124 0.0110 0.0146 0.0101 0.0145 0.0132
nd5 0.0203 0.0202 0.0153 0.0143 0.0150 0.0177 0.0171
nd6 0.0087 0.0060 0.0091 0.0030 0.0061 0.0091 0.0070
cytb 0.0065 0.0054 0.0041 0.0081 0.0068 0.0041 0.0058
nd1 0.0081 0.0082 0.0066 0.0114 0.0115 0.0098 0.0093
nd2 0.0265 0.0179 0.0119 0.0134 0.0178 0.0120 0.0166
cox1 0.0078 0.0029 0.0049 0.0049 0.0083 0.0078 0.0061
cox2 0.0066 0.0022 0.0036 0.0066 0.0022 0.0022 0.0039
atp8 0.0075 NA 0.0151 NA NA 0.0075 0.0100
atp6 0.0158 0.0113 0.0135 0.0135 0.0137 0.0091 0.0129
cox3 0.0095 NA NA NA 0.0019 NA 0.0057
nd3 0.0089 0.0086 0.0089 NA 0.0045 0.0089 0.0080
nd4l 0.0055 0.0055 NA NA NA NA 0.0055
Average 0.0114 0.0092 0.0094 0.0100 0.0089 0.0093
Table 6. Synonymous substitution rate of mitochondrial genes in Leptobotia species.
Table 6. Synonymous substitution rate of mitochondrial genes in Leptobotia species.
Species Leptobotia mantschurica Leptobotia taeniops Leptobotia microphthalma Leptobotia rubrilabris Leptobotia punctata Leptobotia pellegrini Average
nd4 0.3072 0.2587 0.1378 0.2807 0.2312 0.281 0.2494
nd5 0.339 0.2061 0.1508 0.2192 0.1948 0.262 0.2286
nd6 0.3892 0.2009 0.1454 0.1851 0.2136 0.2839 0.2363
cytb 0.313 0.2249 0.1558 0.1944 0.2038 0.2119 0.2173
nd1 0.2916 0.2499 0.1942 0.2269 0.2008 0.2503 0.2356
nd2 0.2375 0.2222 0.1452 0.2249 0.2176 0.2187 0.211
cox1 0.3102 0.2498 0.1397 0.1906 0.1935 0.2488 0.2221
cox2 0.1696 0.1562 0.1393 0.153 0.1376 0.1815 0.1562
atp8 0.2238 0.193 0.1062 0.3132 0.2284 0.2686 0.2222
atp6 0.3279 0.2489 0.1982 0.2665 0.2852 0.2965 0.2705
cox3 0.2624 0.1976 0.123 0.205 0.1981 0.1609 0.1912
nd3 0.351 0.2059 0.1736 0.1592 0.2136 0.241 0.2241
nd4l 0.2358 0.1912 0.1103 0.1924 0.1234 0.2873 0.1901
Avergae 0.2891 0.2158 0.1476 0.2162 0.2032 0.2456
Table 7. The ratio of the number of nonsynonymous substitutions per nonsynonymous site (Ka) to the number of synonymous substitutions per synonymous site (Ks) of mitochondrial genes in Leptobotia species.
Table 7. The ratio of the number of nonsynonymous substitutions per nonsynonymous site (Ka) to the number of synonymous substitutions per synonymous site (Ks) of mitochondrial genes in Leptobotia species.
Species Leptobotia mantschurica Leptobotia taeniops Leptobotia microphthalma Leptobotia rubrilabris Leptobotia punctata Leptobotia pellegrini Average
nd4 0.0541 0.0479 0.0802 0.0521 0.0437 0.0514 0.0549
nd5 0.0599 0.0982 0.1014 0.0651 0.0773 0.0675 0.0782
nd6 0.0224 0.0300 0.0623 0.0163 0.0283 0.0321 0.0319
cytb 0.0209 0.0239 0.0264 0.0417 0.0335 0.0193 0.0276
nd1 0.0278 0.0329 0.0338 0.0503 0.0573 0.0391 0.0402
nd2 0.1114 0.0807 0.0818 0.0596 0.0820 0.0550 0.0784
cox1 0.0251 0.0117 0.0350 0.0255 0.0430 0.0312 0.0286
cox2 0.0387 0.0140 0.0261 0.0431 0.0158 0.0121 0.0250
atp8 0.0335 0.0000 0.1420 0.0000 0.0000 0.0278 0.0339
atp6 0.0483 0.0455 0.0682 0.0508 0.0481 0.0308 0.0486
cox3 0.0361 0.0000 0.0000 0.0000 0.0096 0.0000 0.0076
nd3 0.0253 0.0420 0.0512 0.0000 0.0209 0.0369 0.0294
nd4l 0.0235 0.0286 0.0000 0.0000 0.0000 0.0000 0.0087
Average 0.0405 0.0350 0.0545 0.0311 0.0353 0.0310
Disclaimer/Publisher’s Note: The statements, opinions and data contained in all publications are solely those of the individual author(s) and contributor(s) and not of MDPI and/or the editor(s). MDPI and/or the editor(s) disclaim responsibility for any injury to people or property resulting from any ideas, methods, instructions or products referred to in the content.
Copyright: This open access article is published under a Creative Commons CC BY 4.0 license, which permit the free download, distribution, and reuse, provided that the author and preprint are cited in any reuse.
Prerpints.org logo

Preprints.org is a free preprint server supported by MDPI in Basel, Switzerland.

Subscribe

© 2024 MDPI (Basel, Switzerland) unless otherwise stated