The first complete mitochondrial genome of the Mariana Trench Freyastera benthophila (Asteroidea: Brisingida: Brisingidae) allows insights into the deep-sea adaptive evolution of Brisingida.
Journal: 2018/December - Ecology and Evolution
ISSN: 2045-7758
Abstract:
Starfish (phylum Echinodermata) are ecologically important and diverse members of marine ecosystems in all of the world's oceans, from the shallow water to the hadal zone. The deep sea is recognized as an extremely harsh environment on earth. In this study, we present the mitochondrial genome sequence of Mariana Trench starfish Freyastera benthophila, and this study is the first to explore in detail the mitochondrial genome of a deep-sea member of the order Brisingida. Similar to other starfish, it contained 13 protein-coding genes, two ribosomal RNA genes, and 22 transfer RNA genes (duplication of two tRNAs: trnL and trnS). Twenty-two of these genes are encoded on the positive strand, while the other 15 are encoded on the negative strand. The gene arrangement was identical to those of sequenced starfish. Phylogenetic analysis showed the deep-sea Brisingida as a sister taxon to the traditional members of the Asteriidae. Positive selection analysis indicated that five residues (8 N and 16 I in atp8, 47 D and 196 V in nad2, 599 N in nad5) were positively selected sites with high posterior probabilities. Compared these features with shallow sea starfish, we predict that variation specifically in atp8, nad2, and nad5 may play an important role in F. benthophila's adaptation to deep-sea environment.
Relations:
Content
Citations
(3)
Chemicals
(3)
Genes
(3)
Organisms
(2)
Processes
(2)
Similar articles
Articles by the same authors
Discussion board
Ecology and Evolution. Oct/31/2018; 8(22): 10673-10686
Published online Oct/30/2018

The first complete mitochondrial genome of the Mariana Trench Freyastera benthophila (Asteroidea: Brisingida: Brisingidae) allows insights into the deep‐sea adaptive evolution of Brisingida

Abstract

Starfish (phylum Echinodermata) are ecologically important and diverse members of marine ecosystems in all of the world's oceans, from the shallow water to the hadal zone. The deep sea is recognized as an extremely harsh environment on earth. In this study, we present the mitochondrial genome sequence of Mariana Trench starfish Freyastera benthophila, and this study is the first to explore in detail the mitochondrial genome of a deep‐sea member of the order Brisingida. Similar to other starfish, it contained 13 protein‐coding genes, two ribosomalRNAgenes, and 22 transferRNAgenes (duplication of twotRNAs: trnL and trnS). Twenty‐two of these genes are encoded on the positive strand, while the other 15 are encoded on the negative strand. The gene arrangement was identical to those of sequenced starfish. Phylogenetic analysis showed the deep‐sea Brisingida as a sister taxon to the traditional members of the Asteriidae. Positive selection analysis indicated that five residues (8 N and 16 I in atp8, 47 D and 196 V in nad2, 599 N in nad5) were positively selected sites with high posterior probabilities. Compared these features with shallow sea starfish, we predict that variation specifically in atp8, nad2, and nad5 may play an important role in F. benthophila's adaptation to deep‐sea environment.

1INTRODUCTION

The class Asteroidea (sea stars and starfish) is one of the most familiar and diverse groups of the phylum Echinodermata with a long paleontological history, including nearly 1,800 species grouped into 35 families (Clark & Downey, 1992; Matsubara, Komatsu, & Wada, 2004). They are present in all of the world's oceans and occur from intertidal to abyssal, and they are most diverse in the Indo‐Pacific and tropical Atlantic regions (Mah & Blake, 2012). To date, the phylogenetic relationships of these starfish have not yet been fully resolved (Knott & Wray, 2000).

Because of its maternal inheritance, and low frequency of gene recombination, mitochondrial genes (e.g., COI) are widely used for phylogenetic analysis (Boore, 1999). Compared to one gene, complete mitogenomes include more genetic information and usually could obtain more accurate phylogenetic relationship and therefore have become more popular in recent years (Fan, Hu, Wen, & Zhang, 2011; Shen, Ma, Ren, & Zhao, 2009; Shen et al., 2017). Up to now, the complete mitochondrial genomes have been reported in many marine organisms, such as sea cucumber (Fan et al., 2011; Perseke et al., 2010; Scouras, Beckenbach, Arndt, & Smith, 2004; Sun, Qi, & Kong, 2010), sea urchin (Cantatore, Roberti, Rainaldi, Gadaleta, & Saccone, 1989; De Giorgi, Martiradonna, Lanave, & Saccone, 1996; Qureshi & Jacobs, 1993), brittle star (Perseke et al., 2008, 2010; Scouras et al., 2004; Smith, Arndt, Gorski, & Fajber, 1993), sea lily (Perseke et al., 2008; Scouras & Smith, 2006), shellfish (Plazzi, Ribani, & Passamonti, 2013; Ren, Liu, Jiang, Guo, & Liu, 2010), and crab (Liu & Cui, 2010; Yang & Yang, 2008).

Animal mitogenome is typically always circular molecule, except for some classes of cnidarians (Bridge, Cunningham, Schierwater, Desalle, & Buss, 1992). It contains 37 genes in general: 13 protein‐coding genes (PCGs) (cytochrome c oxidase subunits I–III [cox1cox3], NADH dehydrogenase subunits 1–6 and 4L [nad16, nad4L], ATP synthase subunits 6 and 8 [atp6, atp8], apocytochrome b [cob]), two ribosomal RNAs (12S and 16S), and 22 transfer RNAs (tRNAs). All 13 protein‐coding genes play key roles in oxygen usage and energy metabolism (Boore, 1999; Xu et al., 2007). Because variation in mitochondrial protein‐coding genes that involved in oxidative phosphorylation can directly influence metabolic performance, an increasing number of researches related to adaptive evolution of these genes have been reported (Maliarchuk, 2011; Xu et al., 2007; Yu, Wang, Ting, & Zhang, 2011; Zhou, Shen, Irwin, Shen, & Zhang, 2014). Despite strong functional constraints, mitochondrial DNA may be subject to positive directional selection in response to pressures from extreme harsh environment (Tomasco & Lessa, 2011). Indeed, mtDNA analyses have demonstrated the existence of adaptive evolution in the ATP synthase genes of the sea anemones, alvinocaridid shrimp, and galliform birds (Sun, Hui, Wang, & Sha, 2018; Zhang, Zhang, Wang, Zhang, & Lin, 2017; Zhou et al., 2014); the NADH dehydrogenase genes of sea anemones, alvinocaridid shrimp, Tibetan horses, and Chinese snub‐nosed monkeys (Sun et al., 2018; Xu et al., 2007; Yu et al., 2011; Zhang et al., 2017); the cytochrome b gene of cetaceans and alpacas (da Fonseca, Johnson, O'Brien, Ramos, & Antunes, 2008); and the cytochrome c oxidase genes of Tibetan antelope and anthropoids (Adkins & Honeycutt, 1994; Luo et al., 2008).

Here, we report the complete mitogenome of the starfish Freyastera benthophila, which collected from Mariana Trench at 5,463 m depth. Freyastera benthophila exist in abyssal and is mainly distributed in the southern Pacific, eastern Pacific off California, mid‐Atlantic (between Azores and Spain), the Bay of Bengal, and Biscay, ranging from 4,250 to 5,000 m depth (Downey, 1986). The mitogenome features, organization, codon usage, and gene arrangement information were presented. The phylogenetic relationships between F. benthophila and 19 other species from Echinodermata were analyzed. To infer the deep‐sea adaptive evolution, positive selection analysis of mitochondrial genes was also performed.

2MATERIALS AND METHODS

2.1Sample collection and DNA extraction

The specimen was collected at Mariana Trench, in June 2016 (10°51.0971′N, 141°57.2705′E, at 5,463 m depth). The collection was accomplished by deep‐sea human occupied vehicle (HOV) “Jiao Long” during an expedition. The specimen was preserved in 95% ethanol. Total genomic DNA was extracted from ethanol‐fixed tissue with tissue DNA kit (Omega Bio‐Tek) and stored at −20°C.

2.2PCR amplification and DNA sequencing

The universal metazoan primers for mtDNA were used in PCR. Three fragments cox3, cob, and 16S were successfully amplified with the primers cox3F + cox3R (Boore, Macey, & Medina, 2005), cobF424 + cobR876 (Boore & Brown, 2000), and 16SarL + 16SbrH (Boore et al., 2005). In addition, partial sequences of cox1, cox2, nad4L, atp6, nad4, and nad5 were amplified with the degenerate primers from conserved regions of other starfish in GenBank. The remaining gaps were amplified with the species‐specific primers designed according to the obtained sequences. Finally, the whole mitogenome was amplified based on eight pairs of primers (Table 1).

Table 1

Primers used for amplifying and sequencing the mitogenome of Freyastera benthophila

NameSequence (5′–3′)RegionLocation
W1‐FCCGCAAGAGTCGAAAGAGcox13473–3490
W1‐RTCAAGGAGTCGTGGCATTcox25448–5465
W2‐FTAGCTCTTTCCCGAACACnad4L4993–5010
W2‐RGAACCGTAAACACTATCTGCTcox37233–7253
W3‐FGACTCGCAGCTAATCTTACAatp66466–6485
W3‐RCAAGACCGTATCCACCTAACnad48630–8649
W4‐FCCCTCCTTCCAACCCTCATCnad48287–8306
W4‐RCACCCATCTTTCGTAGGTCTTGTnad510735–10757
W5‐FCCACCGCTACTTCTCAACATnad510545–10564
W5‐RTAGAGCGAAGGATTGCATAGcob12860–12879
W6‐FCCACCTATTCTTCCTTCACCcob12614–12633
W6‐RGCATAATCATTTGCCTCTTA16S15183–15202
W7‐FAGCTCGATAGGGTCTTCTCGTC16S15063–15084
W7‐RGCAGTGGCATTGTTGACTTTGAnad12040–2061
W8‐FAGCTAACGGCTGAAACAATCnad11957–1976
W8‐RTTCCTGCGTAATGGGCTAcox13902–3919
John Wiley & Sons, Ltd

PCRs were performed with a gradient machine (Applied Biosystems Inc.). The cycling was set up with an initial denature step at 94°C for 5 min, followed by 35 cycles (94°C for 30 s, 45–55°C for 1 min, 72°C 1–3 min), and a final extension was executed at 72°C for 10 min. LA‐PCR was carried out in a 20 μl reaction volume containing 12.6 μl ddH2O, 2 μl 10 × LA‐PCR buffer (Mg2+ plus, Takara), 3.2 μl dNTP mix (2.5 mM each), 0.5 μl each primer (10 μM), 0.2 μl LA Taq DNA polymerase (5 U/μl, Takara), and 1 μl DNA template (50 ng/μl). PCR products were electrophoresed on a 1.0% agarose gel and purified with gel extraction kit (Omega Bio‐Tek) and sequenced with ABI 3730x1 DNA analyzer (Applied Biosystems Inc.).

2.3Gene annotation and Sequence analysis

Raw sequencing reads were first processed using Phred with the quality score 20 and assembled in Phrap with default parameters (Ewing & Green, 1998; Ewing, Hillier, Wendl, & Green, 1998). Then, all assemblies and sequence quality were verified manually in Consed (Gordon, Abajian, & Green, 1998). DOGMA (Wyman, Jansen, & Boore, 2004), ORFfinder (http://www.ncbi.nlm.nih.gov/projects/gorf/orfig.cgi), and BLAST (http://blast.ncbi.nlm.nih.gov/Blast.cgi) were used to identify protein‐encoding genes and rRNA genes. The tRNA genes were identified by tRNAscan‐SE 1.21 (Lowe & Eddy, 1997) and ARWEN 1.2.3.c (Laslett & Canbäck, 2008). Secondary structures for tRNAs were drawn using MITOS Web server (Bernt et al., 2013). Codon usage analysis was estimated with CodonW 1.4.4 (Peden, 1999). The mitochondrial gene map was drawn with GenomeVx (Conant & Wolfe, 2008).

2.4Phylogenetic analysis

Twenty echinoderm mt genomes including the one obtained in this study were used for phylogenetic analysis. All complete mtDNA sequences vailable in GenBank are listed in Table 2. Crinoidea is generally considered as the earliest diverged group of echinoderms (Scouras & Smith, 2006). In this study, Antedon mediterranea (Crinoidea) was rooted as the out‐group. The amino acid sequence from each of 13 protein‐coding genes was aligned separately using Clustal ×2.0 (Larkin et al., 2007), and then, the relatively poor homologous sequence was eliminated. The aligned amino acid sequences were concatenated into a single dataset. The phylogenetic reconstruction approach was performed using neighbor joining (NJ) and maximum likelihood (ML) with MEGA 5.0 (Tamura et al., 2011). The assessment of node reliability was performed using 1,000 bootstrap replicates.

Table 2
List of taxa used in the phylogenetic analysis
CrinoideaOphiuroideaEchinoideaHolothuroideaAsteroidea
TaxonClassificationAccession numberReferences
Antedom mediterraneaCrinoidea; Articulata; Comatulida; AntedonidaePerseke et al. (2008)
Ophiocomina nigraOphiuroidea; Ophiuridea; Ophiurida; Ophiurina; Gnathophiurina; OphiocomidaePerseke et al. (2010)
Astrospartus mediterraneusOphiuroidea; Ophiuridea; Euryalida; GorgonocephalidaePerseke et al. (2010)
Strongylocentrotus purpuratusEchinoidea; Euechinoidea; Echinacea; Echinoida; StrongylocentrotidaeQureshi and Jacobs (1993)
Echinocardium cordatumEchinoidea; Euechinoidea; Atelostomata; Spatangoida; LoveniidaePerseke et al. (Unpublished)
Paracentrotus lividusEchinoidea; Euechinoidea; Echinacea; Echinoida; EchinidaeCantatore et al. (1989)
Cucumaria miniataHolothuroidea; Dendrochirotacea; Dendrochirotida; CucumariidaeScouras et al. (2004)
Apostichopus japonicusHolothuroidea; Aspidochirotacea; Aspidochirotida; StichopodidaeSun et al. (2010)
Holothuria forskaliHolothuroidea; Aspidochirotacea; Aspidochirotida; HolothuriidaePerseke et al. (2010)
Parastichopus nigripunctatusHolothuroidea; Aspidochirotacea; Aspidochirotida; StichopodidaeSasaki and Hamaguchi (Unpublished)
Stichopus horrensHolothuroidea; Aspidochirotacea; Aspidochirotida; StichopodidaeFan et al. (2011)
Freyastera benthophilaAsteroidea; Forcipulatacea; Brisingida; BrisingidaeThis study
Aphelasterias japonicaAsteroidea; Forcipulatacea; Forcipulatida; AsteriidaeTang et al. (2014)
Pisaster ochraceusAsteroidea; Forcipulatacea; Forcipulatida; AsteriidaeSmith, Banfield, Doteval, Gorski, and Kowbel (1990)
Asterias amurensisAsteroidea; Forcipulatacea; Forcipulatida; AsteriidaeMatsubara et al. (2005)
Astropecten polyacanthusAsteroidea; Valvatacea; Paxillosida; AstropectinidaeMatsubara et al. (2005)
Luidia quinariaAsteroidea; Valvatacea; Paxillosida; LuidiidaeMatsubara et al. (2005)
Acanthaster brevispinusAsteroidea; Valvatacea; Valvatida; AcanthasteridaeYasuda et al. (2006)
Acanthaster planciAsteroidea; Valvatacea; Valvatida; AcanthasteridaeYasuda et al. (2006)
Patiria pectiniferaAsteroidea; Valvatacea; Valvatida; AsterinidaeAsakawa et al. (1995)
John Wiley & Sons, Ltd

2.5Positive selection analysis

To evaluate the variation in selective pressure between deep‐sea F. benthophila and other eight shallow sea starfish, we used a codon‐based likelihood approach implemented in the CODEML program of the pamlX package (Xu & Yang, 2013; Yang, 2007). All models correct the transition/transversion rate and codon usage biases (F3 × 4). The branch model tests were used to analyze the difference of selective pressure between the deep‐sea and shallow sea starfish. The “one‐ratio” model (model 0), “free‐ratio” model (model 1), and “two‐ratio” model were used in the combined dataset of 13 protein‐coding genes. Considering that positive selection may occur in some amino acids during the evolution of a protein, we used two branch site models (A and A null). Bayes empirical Bayes (Yang, Wong, & Nielsen, 2005) analysis was used to calculate the posterior probabilities of a specific codon site.

3RESULTS AND DISCUSSION

3.1General features

The mitogenome of the F. benthophila is a 16,175‐bp circular molecule (Figure 1) with a nucleotide composition of 34.70% A, 21.13% C, 10.65% G, and 33.53% T bases. The genome has an overall A + T content of 68.23%, which appears to be high for Asteroidea. Among the eight species in Asteroidea, the lowest A + T content is 56.34% in Acanthaster planci (Table 3). Freyastera benthophila has the smallest complete mitogenome found in Asteroidea thus far. The size of Asteroidea mitogenomes ranged from 16,524 bp in Luidia quinaria to 16,175 bp in F. benthophila (Table 3). The synteny and identity level between F. benthophila and each of the other seven starfish mitogenomes is shown in Figure 2. The lack of similarity between F. benthophila and L. quinaria is the most obvious feature in the plot.

Figure 1

Mitochondrial gene map of Freyastera benthophila. All of 37 genes are encoded on the both strands. Genes for proteins andrRNAs are shown with standard abbreviation. Genes fortRNAs are designated by a single letter for the corresponding amino acid with two leucinetRNAs and two serinetRNAs differentiated by numerals

Table 3
Genomic characteristics of Asteroidea mtDNAs
Freyastera benthophilaAsterias amurensisAstropecten polyacanthusLuidia quinariaAphelasterias japonicaAcanthaster brevispinusAcanthaster planciPatiria pectinifera
Entire genome length (bp)16,17516,42716,30416,52416,21516,25416,23416,260
Entire genome A + T%68.2365.4564.0062.9864.3256.3756.3461.27
Protein‐coding gene length (bp)11,50611,48811,53911,50611,50411,48811,49111,501
Protein‐coding gene A + T%67.1564.4562.4861.0262.9555.9555.6460.12
12S gene length (bp)891893901884900928929897
12S gene A + T%65.6661.7062.0459.1662.2253.4554.0458.86
16S gene length (bp)1,6021,6201,6291,7511,6021,5451,5491,531
16S gene A + T%72.2870.1969.1269.6769.6655.3456.0466.49
tRNA length (bp)1,5571,5801,5611,5631,5661,5461,5501,585
tRNA A + T%60.6567.2267.0166.5466.9961.4561.3564.35
Largest NCR length (bp)284483395402281551531445
Largest NCR A + T%67.2563.9864.5670.1562.6353.9054.9959.78
nad21062 (ATG/TAG)1062 (GTG/TAA)1068 (GTG/TAA)1068 (GTG/TAA)1062 (ATG/TAG)1065 (ATG/TAG)1065 (ATG/TAG)1065 (ATG/TAA)
nad1972 (ATG/TAG)978 (GTG/TAA)976 (ATG/T‐)978 (GTG/TAG)978 (GTG/TAA)981 (GTG/TAG)981 (GTG/TAG)981 (GTG/TAG)
cox11557 (ATG/TAA)1551 (ATG/TAA)1554 (ATG/TAA)1554 (ATG/TAA)1552 (ATG/T‐)1553 (ATG/TA‐)1553 (ATG/TA‐)1554 (ATG/TAA)
nad4L297 (ATG/TAA)288 (ATG/TAA)297 (ATT/TAA)297 (ATT/TAA)297 (ATC/TAA)297 (ATT/TAA)297 (ATT/TAA)297 (ATT/TAA)
cox2690 (ATG/TAA)690 (ATG/TAA)688 (ATG/T‐)693 (ATG/TAG)690 (ATG/TAA)688 (ATG/T‐)688 (ATG/T‐)688 (ATG/T‐)
atp8168 (ATG/TAA)168 (ATG/TAA)168 (ATG/TAA)168 (ATG/TAA)168 (ATG/TAA)165 (ATG/TAA)165 (ATG/TAA)165 (ATG/TAA)
atp6693 (ATG/TAA)693 (ATG/TAA)693 (ATG/TAA)693 (ATG/TAA)693 (ATG/TAA)693 (ATG/TAA)693 (ATG/TAA)693 (ATG/TAA)
cox3783 (TAG/TAA)780 (ATG/TAA)783 (ATG/TAA)783 (ATG/TAA)783 (ATG/TAA)783 (ATG/TAA)783 (ATG/TAA)783 (ATG/TAA)
nad3351 (ATG/TAA)351 (ATG/TAA)351 (ATT/TAA)351 (ATT/TAG)351 (ATG/TAA)351 (ATT/TAA)351 (ATT/TAA)333 (ATT/TAG)
nad41386 (ATG/TAA)1383 (ATG/TAA)1380 (ATG/TAA)1380 (ATG/TAA)1383 (ATG/TAA)1383 (ATG/TAG)1383 (ATG/TAA)1383 (ATG/TAA)
nad51920 (ATG/TAA)1917 (ATG/TAA)1905 (ATG/TAA)1911 (ATG/TAA)1920 (ATG/TAA)1902 (ATG/TAA)1902 (ATG/TAA)1932 (GTG/TAA)
nad6489 (ATG/TAG)489 (ATG/TAG)492 (ATG/TAA)492 (ATG/TAA)489 (ATG/TAG)489 (ATG/TAG)489 (ATG/TAG)489 (ATG/TAA)
cob1138 (ATG/T‐)1138 (ATG/T‐)1138 (ATG/T‐)1138 (ATG/T‐)1138 (ATG/T‐)1138 (ATG/T‐)1138 (ATG/T‐)1138 (ATG/T‐)
ReferenceThis studyMatsubara et al. (2005)Matsubara et al. (2005)Matsubara et al. (2005)Tang et al. (2014)Yasuda et al. (2006)Yasuda et al. (2006)Asakawa et al. (1995)
John Wiley & Sons, Ltd
Figure 2

The synteny and identity level of Freyastera benthophila mitogenome against each of the other seven starfish mitogenomes. Ideograms and ribbons represent the similarity pairwise blastn searches. In F. benthophila ideogram, the 13 coding genes are colored in blue, control regions are colored in black, andrRNAs are colored in green. The figure was produced using Circoletto (Darzentas, 2010).FB(F. benthophila),AB(Acanthaster brevispinus),AP1 (Acanthaster planci),AJ(Aphelasterias japonica),AA(Asterias amurensis),AP2 (Astropecten polyacanthus),LQ(Luidia quinaria),PP(Patiria pectinifera)

The genome encodes 37 genes including 13 protein‐coding genes (PCGs), two rRNA genes, and 22 tRNA genes (duplication of two tRNAs: trnL and trnS) on both strands. Fifteen of the genes are encoded on the negative strand, while the other 22 are encoded on the positive strand. A total of 22 noncoding regions were found, with the largest continuous region (284 bp, A + T = 67.25%) located between trnT and 16S. Due to its AT richness, we predict that this part is mitochondrial control region. Furthermore, we found four overlaps: trnC/trnV, trnA/trnL1, atp8/atp6, and cox3/trnS2. Table 4 presents a summary of the organization of F. benthophila mitogenome. The complete mitochondrial DNA sequence has been deposited in GenBank (Accession Number: MG563681).

Table 4

Gene content of the Freyastera benthophila mitogenome

GeneLocationSizeCodonIntergenic nucleotide (bp)Strand
StartEndNucleotide (bp)Amino acidStartStop
nad2110621,062353ATGTAG0L
trnI10641132691L
nad111472118972323ATGTAG14L
trnL221192191730L
trnG221622846924L
trnY22862356711L
trnD23572426700H
trnM24282500731L
trnV25092579718H
trnC2578264871−2L
trnW26512721712L
trnA273528047013H
trnL12804287572−1L
trnN28762947720L
trnQ29513022723H
trnP30233093710L
cox1313046861,557518ATGTAA36H
trnR46874757710H
nad4L4758505429798ATGTAA0H
cox250565745690229ATGTAA1H
trnK57475821751H
atp85824599116855ATGTAA2H
atp659766668693230ATGTAA−16H
cox366737455783260ATGTAA4H
trnS27454752471−2L
nad375497899351116ATGTAA24H
nad4791192961,386461ATGTAA11H
trnH9448951770151H
trnS195199586681H
nad59587115061,920639ATGTAA0H
nad61152412012489162ATGTAG17L
cob12027131641,138379ATGT‐14H
trnF1316513235710H
12S13236141268910H
trnE1412714194680H
trnT1419514263690H
16S14548161491,602284L
John Wiley & Sons, Ltd

3.2Protein‐coding genes

With regard to PCGs, nine (cox1cox3, nad3nad5, nad4L, cob, atp6, and atp8) are encoded by the positive strand, and the remaining three (nad1, nad2, and nad6) are encoded by the negative strand. These features have been observed in all Asteroidea mitogenomes published so far. Thirteen PCGs initiate with the standard start codon ATG. Most of PCGs terminate with the stop codon TAA (9 of 13), and three genes terminate with the stop codon TAG. Incomplete termination codon T is used by cob. However, mitogenomes often use a variety of nonstandard initiation codons (Wolstenholme, 1992). Nonstandard initiation codon GTG and incomplete termination codon TA are also used in other starfish (Table 3). The lengths of PCGs are 11,506 bp, and the A + T content is 67.15% higher than that of other Asteroidea species (Table 3).

The codon usage of F. benthophila is shown in Figure 3. Among PCGs, leucine (15.85%) and cysteine (0.99%) are the most and the least frequently used amino acids, respectively. Codons, UUA (leucine 6.67%) and ACG (threonine 0.08%), are the most and the least frequently used, respectively. We predict that the richness of A and T occurrence frequency of the mitogenome caused the corresponding amino acid bias to some extent. It is obvious that the A + T content of the third codon position (74.10%) is higher than that of the first (63.43%) and second positions (63.67%).

Figure 3

Codon usage in Freyastera benthophila. All codons for amino acids have been classified. Each amino acid is designated by a single letter for the corresponding codon. x‐axis and y‐axis represent the used times of each codon

3.3Ribosomal RNA and transfer RNA genes

Boundaries of both the small and the large ribosomal genes were determined by BLAST and DOGMA. The 16S and 12S genes of F. benthophila are 1,602 bp (A + T = 72.28%) and 891 bp (A + T = 65.66%) in length, respectively. These lengths are typical for Asteroidea, whereas the AT contents are higher than those of other starfish (Table 3).

We analyzed the entire mitogenome sequence of F. benthophila and successfully identified 22 tRNA genes based on their potential secondary structures using the tRNAscan‐SE, ARWEN, and MITOS Web server (Table 4, Supporting Information Figure S1). The length of these tRNA genes ranged from 68 bp (trnS1 and trnE) to 75 bp (trnK). Twenty‐one of these genes displayed a common cloverleaf secondary structure, and the remaining one lacked a DHU arm from trnS1. The D‐stem absence has been found in many other starfish, such as Acanthaster brevispinus, Acanthaster planci, Aphelasterias japonica, Asterias amurensis, L. quinaria, and Patiria pectinifera (Asakawa, Himeno, Miura, & Watanabe, 1995; Matsubara et al., 2005; Tang et al., 2014; Yasuda et al., 2006).

3.4Gene arrangement

Mitochondrial gene arrangement has been demonstrated to be an effective means to solve the deep phylogenetic studies (Boore, 1999; Boore & Brown, 1998). In recent years, some research on mt gene arrangement of echinoderms has been reported (Arndt & Smith, 1998; Perseke et al., 2008, 2010; Scouras et al., 2004).

In this study, mitochondrial gene order of echinoderm was compared among species within classes Asteroidea, Echinoidea, Holothuroidea, Ophiuroidea, and Crinoidea (Figure 4). We expected that the mt gene order of starfish may reveal some phylogenetically information. However, the gene component and gene order of eight species of Asteroidea are completely identical to each other. This phenomenon also happened in the class Echinoidea. We obtained 27 complete mt genomes of Echinoidea from NCBI genebank, and the gene component and gene order of 27 species of Echinoidea are also completely identical to each other. Because Strongylocentrotus purpuratus has been considered as a model for developmental and systems biology, we took S. purpuratus as a representative for Echinoidea in Figure 4 (Sodergren et al., 2006). However, mitochondrial gene order has undergone significant changes in the classes of Holothuroidea, Ophiuroidea, and Crinoidea. Scouras et al. (2004) suggested that it is difficult to resolve the echinoderm phylogeny using the mitochondrial gene rearrangement.

Figure 4
Comparison of mitochondrial gene arrangement in Echinodermata. The bars show identical gene blocks. The noncoding regions are not presented, and gene segments are not drawn to scale

It is interesting that mt gene order of the species in the classes of Asteroidea and Echinoidea is completely identical to each other. If the tRNA is not considered, gene order of PCGs in species within the class Holothuroidea is also the same. This raises the questions: As these species are distributed throughout the world's oceans, why had the mt gene order not been changed and how do they evolve over time. More studies of mt genome species are needed to further investigate whether this pattern is common among starfish, sea urchins, and sea cucumbers.

3.5Phylogenetic analysis

The gene order and transcriptional orientation of the eight Asteroidea species are completely identical to each other, so the mt genome structures would not provide the phylogenetic information. Thus, we performed the phylogenetic analysis using all amino acids of mt protein‐coding genes (Figure 5). Almost all the phylogenetic relationships are supported with high values (NJ/ML bootstraps 99–100). Acanthaster brevispinus is first clustered with A. planci and then united with P. pectinifera; meanwhile, Astropecten polyacanthus and L. quinaria formed a clade. And these five starfish formed the Valvatacea clade. Then, A. japonica is first clustered with Pisaster ochraceus and then united with A. amurensis. Finally, F. benthophila with these three species formed a Forcipulatacea clade. Blake (1987) recognized that Brisingida and Forcipulatida are the two orders within the Forcipulatacea and suggested that they were the most primitive asteroids (Blake, 1987, 1988). Mah and Foltz (2011) described that the largest clade within the Forcipulatacea is formed by the Brisingida and Asteriidae, which forms a clade of deep‐sea and Southern Hemisphere taxa. In the present study, the results supported the deep‐sea Brisingida as a sister taxon to the traditional members of the Asteriidae, and the branch support values are higher than those in previous studies (Glover et al., 2016; Mah & Foltz, 2011). However, the number of Brisingida species with complete mitogenome is still limited, and more mitogenomes and analysis are necessary to determine the phylogenetic relationship among members of Brisingida.

Figure 5

Phylogenetic trees based on the concatenated amino acids of 13 protein‐coding genes. The branch length is determined withNJanalysis. Antedon mediterranea was used as out‐group.NJ(left number) andML(right number) bootstrap values are given for each branch. The red dot highlights the species sequenced in this study

3.6Positive selection analysis

We examined the potential positive selection in Brisingida lineage because of the colonization of deep‐sea environments which may affect the function of mitochondrial genes. The results of selective pressure analyses are shown in Table 5. When the ω ratios for the 13 concatenated mitochondrial protein‐coding genes were tested between the deep‐sea F. benthophila and other eight shallow sea starfish, we failed to find a significant difference in their ω ratios, which may be due to the large bias of sample sizes (p >0.05) (Table 5). In addition, in the analyses of individual genes, we found five residues with high posterior probabilities in the atp8 (8 N, 16 I), nad2 (47 D, 196 V), and nad5 (599 N), respectively (Table 5). Similar results have been observed in deep‐sea animals, and the authors concluded that it may be related to the adaptation to environment (Sun et al., 2018; Zhang et al., 2017). Under the deep‐sea extreme environment, survival may require a modified and adapted energy metabolism (Sun et al., 2018).

Table 5
Selective pressure analyses of the mitochondrial genes of starfish
TreesBranch modelModel compared2△lnLLRT p‐value
ModellnLEstimates of parameters
NJ/MLModel 1−77,922.08746Model 1 versus Model 0127.224180.00000
Two ratio−77,985.69322ω0 = 0.06446
ω1 = 0.06371
Two ratio versus Model 00.012660.91027
Model 0−77,985.69955ω = 0.06438
GeneBranch site model (BSM)Model compared2△lnLLRT p‐valuePositive sites
ModellnLEstimates of parameters
atp8Model A−1,269.41181Site class012a2bModel A versus Model A null1.828420.176328 N 0.996**
Proportion0.538650.213950.177070.0703316 I 0.971*
Background ω0.082401.000000.082401.00000
Foreground ω0.082401.00000151.29990151.29990
Model A null−1,270.32602
nad2Model A−7,398.27984Site class012a2bModel A versus Model A null2.543160.1107747 D 0.969*
Proportion0.837560.109930.046410.00609196 V 0.959*
Background ω0.042531.000000.042531.00000
Foreground ω0.042531.00000127.13451127.13451
Model A null−7,399.55142
nad5Model A−12,858.89872Site class012a2bModel A versus Model A null0.341800.55879599 N 0.961*
Proportion0.761890.186030.041850.01022
Background ω0.043241.000000.043241.00000
Foreground ω0.043241.000001.527241.52724
Model A null−12,859.06962
*Posterior probability >95%; **Posterior probability >99%.John Wiley & Sons, Ltd

Because ATP synthase directly produces ATP, variation in ATPase protein sequence should influence ATP production (Mishmar et al., 2003; Wallace, 2007). Amino acid variations have been widely reported in the ATPase proteins (da Fonseca et al., 2008; Mishmar et al., 2003; Zhang et al., 2017; Zhou et al., 2014). Nad2, nad4, and nad5 are suggested to act as proton‐pumping devices (Brandt, 2006; da Fonseca et al., 2008); thus, mutations in these proteins should influence metabolic efficiency (da Fonseca et al., 2008; Hassanin, Ropiquet, Couloux, & Cruaud, 2009; Zhang et al., 2017). Therefore, we predict that mitochondrial protein‐coding genes, specifically atp8, nad2, and nad5, may play an important role in F. benthophila's adaptation to deep‐sea environment.

4CONCLUSIONS

In this study, we determined the mitogenome of the deep‐sea member F. benthophila, which is 16,175 bp in length and encodes 37 genes including 13 PCGs, two rRNA genes, and 22 tRNA genes on the both strands. We described the mitogenome features, codon usage, gene arrangement, phylogenetic analysis, and positive selection of the starfish F. benthophila. This study is the first determination of the mitogenome of a deep‐sea member of the order Brisingida and may shed light on the adaptive evolution of Brisingida species to the deep‐sea environment.

CONFLICT OF INTEREST

The authors declare no conflict of interest.

AUTHOR CONTRIBUTIONS

Haibin Zhang and Wendan Mu designed the study. Haibin Zhang contributed to the project coordination and collected the samples. Wendan Mu conducted the sequence analyses and drafted the manuscript. Haibin Zhang and Jun Liu helped to draft the manuscript. All authors read and approved the final manuscript.

DATA ACCESSIBILITY

The complete mitochondrial DNA sequence has been deposited in GenBank (Accession Number: MG563681).

Supporting information

ECE3-8-10673-s001.pdfClick here for additional data file.

ACKNOWLEDGEMENTS

The authors thank the captains and crews of the R/V Xiangyanghong 09 and the pilots of HOV “Jiao Long” for their technical support. This study was supported by The National Key R & D Program of China (2017YFC0306600), Strategic Priority Research Program of the Chinese Academy of Sciences (CAS) (XDB06010104), National Natural Science Foundation of China (41576127), Knowledge Innovation Program of CAS (SIDSSE–201401), Hundred Talents Program of CAS (SIDSSE–BR–201401), and Major scientific and technological projects of Hainan Province (ZDKJ2016009).

References

  • 1.Adkins, R., &Honeycutt, R.(1994). Evolution of the primate cytochrome c oxidase subunit II gene. Journal of Molecular Evolution, 38, 215231.[PubMed][Google Scholar]
  • 2.Arndt, A., &Smith, M. J.(1998). Mitochondrial gene rearrangement in the sea cucumber genus Cucumaria. Molecular Biology and Evolution, 15, 10091016. [PubMed][Google Scholar]
  • 3.Asakawa, S.,Himeno, H.,Miura, K., &Watanabe, K.(1995). Nucleotide sequence and gene organization of the starfish Asterina Pectinifera mitochondrial genome. Genetics, 140, 10471060.[PubMed][Google Scholar]
  • 4.Bernt, M.,Donath, A.,Jühling, F.,Externbrink, F.,Florentz, C.,Fritzsch, G., …Stadler, P. F.(2013). MITOS: Improved de novo metazoan mitochondrial genome annotation. Molecular Phylogenetics and Evolution, 69(2), 313319. [PubMed][Google Scholar]
  • 5.Blake, D. B.(1987). A classification and phylogeny of post‐Palaeozoic sea stars (Asteroidea: Echinodermata). Journal of Natural History, 21(2), 481528. [PubMed][Google Scholar]
  • 6.Blake, D. B.(1988).Paxillosidans are not primitive asteroids: A hypothesis based on functional considerationsIn BurkeR. D. (Ed.), Echinoderm biology (pp. 309314). Rotterdam, The Netherlands: Balkema.
  • 7.Boore, J. L.(1999). Animal mitochondrial genomes. Nucleic Acids Research, 27(8), 17671780. [PubMed][Google Scholar]
  • 8.Boore, J. L., &Brown, W. M.(1998). Big trees from little genomes: Mitochondrial gene order as a phylogenetic tool. Current Opinion in Genetics & Development, 8, 668674. [PubMed][Google Scholar]
  • 9.Boore, J. L., &Brown, W. M.(2000). Mitochondrial genomes of Galathealinum, Helobdella, and Platynereis: Sequence and gene arrangement comparisons indicate that Pogonophora is not a phylum and Annelida and Arthropoda are not sister taxa. Molecular Biology and Evolution, 17(1), 87106. [PubMed][Google Scholar]
  • 10.Boore, J. L.,Macey, J. R., &Medina, M.(2005). Sequencing and comparing whole mitochondrial genomes of animals. Methods in Enzymology, 395, 311348. [PubMed][Google Scholar]
  • 11.Brandt, U.(2006). Energy converting NADH: Quinone oxidoreductase (complex I). Annual Review of Biochemistry, 75, 6992. [PubMed][Google Scholar]
  • 12.Bridge, D.,Cunningham, C. W.,Schierwater, B.,Desalle, R., &Buss, L. W.(1992). Class‐level relationships in the phylum Cnidaria: Evidence from mitochondrial genome structure. Proceedings of the National Academy of Sciences of the United States of America, 89(18), 87508753. [PubMed][Google Scholar]
  • 13.Cantatore, P.,Roberti, M.,Rainaldi, G.,Gadaleta, M. N., &Saccone, C.(1989). The complete nucleotide sequence, gene organization, and genetic code of the mitochondrial genome of Paracentrotus lividus. Journal of Biological Chemistry, 264(19), 1096510975.[PubMed][Google Scholar]
  • 14.Clark, A. M., &Downey, M. E.(1992). Starfishes of the Atlantic. London, UK: Chapman and Hall.
  • 15.Conant, G. C., &Wolfe, K. H.(2008). GenomeVx: Simple web‐based creation of editable circular chromosome maps. Bioinformatics, 24(6), 861862. [PubMed][Google Scholar]
  • 16.da Fonseca, R. R.,Johnson, W. E.,O'Brien, S. J.,Ramos, M. J., &Antunes, A.(2008). The adaptive evolution of the mammalian mitochondrial genome. BMC Genomics, 9, 119[PubMed][Google Scholar]
  • 17.Darzentas, N.(2010). Circoletto: Visualizing sequence similarity with Circos. Bioinformatics, 26(20), 26202621. [PubMed][Google Scholar]
  • 18.De Giorgi, C.,Martiradonna, A.,Lanave, C., &Saccone, C.(1996). Complete sequence of the mitochondrial DNA in the sea urchin Arbacia lixula: Conserved features of the echinoid mitochondrial genome. Molecular Phylogenetics and Evolution, 5(2), 323332.[PubMed][Google Scholar]
  • 19.Downey, M. E.(1986). Revision of the Atlantic Brisingida (Echinodermata: Asteroidea), with description of a new genus and family. Washington, DC: Smithsonian Institution Press.
  • 20.Ewing, B., &Green, P.(1998). Base‐calling of automated sequencer traces using phred. II. Error probabilities. Genome Research, 8(3), 186194. [PubMed][Google Scholar]
  • 21.Ewing, B.,Hillier, L. D.,Wendl, M. C., &Green, P.(1998). Base‐calling of automated sequencer traces using phred. I. Accuracy assessment. Genome Research, 8, 175185. [PubMed][Google Scholar]
  • 22.Fan, S. G.,Hu, C. Q.,Wen, J., &Zhang, L. P.(2011). Characterization of mitochondrial genome of sea cucumber Stichopus horrens: A novel gene arrangement in Holothuroidea. Science China Life Sciences, 54(5), 434441. [PubMed][Google Scholar]
  • 23.Glover, A. G.,Wiklund, H.,Rabone, M.,Amon, D. J.,Smith, C. R.,O'Hare, T., …Dahlgren, T. G.(2016). Abyssal fauna of the UK‐1 polymetallic nodule exploration claim, Clarion‐Clipperton Zone, central Pacific Ocean: Echinodermata. Biodiversity Data Journal, 4, e9277.[Google Scholar]
  • 24.Gordon, D.,Abajian, C., &Green, P.(1998). Consed: A graphical tool for sequence finishing. Genome Research, 8, 195202. [PubMed][Google Scholar]
  • 25.Hassanin, A.,Ropiquet, A.,Couloux, A., &Cruaud, C.(2009). Evolution of the mitochondrial genome in mammals living at high altitude: New insights from a study of the tribe Caprini (Bovidae, Antilopinae). Journal of Molecular Evolution, 68, 293310. [PubMed][Google Scholar]
  • 26.Knott, K. E., &Wray, G. A.(2000). Controversy and consensus in asteroid systematics: New insights to ordinal and familial relationships. American Zoologist, 40, 382392.[Google Scholar]
  • 27.Larkin, M. A.,Blackshields, G.,Brown, N. P.,Chenna, R.,Mcgettigan, P. A.,Mcwilliam, H., …Lopez, R.(2007). Clustal W and Clustal X version 2.0. Bioinformatics, 23(21), 29472948. [PubMed][Google Scholar]
  • 28.Laslett, D., &Canbäck, B.(2008). ARWEN: A program to detect tRNA genes in metazoan mitochondrial nucleotide sequences. Bioinformatics, 24(2), 172175. [PubMed][Google Scholar]
  • 29.Liu, Y., &Cui, Z.(2010). Complete mitochondrial genome of the Asian paddle crab Charybdis japonica (Crustacea: Decapoda: Portunidae): Gene rearrangement of the marine brachyurans and phylogenetic considerations of the decapods. Molecular Biology Reports, 37, 25592569. [PubMed][Google Scholar]
  • 30.Lowe, T. M., &Eddy, S. R.(1997). tRNAscan‐SE: A program for improved detection of transfer RNA genes in genomic sequence. Nucleic Acids Research, 25(5), 955964. [PubMed][Google Scholar]
  • 31.Luo, Y.,Gao, W.,Gao, Y.,Tang, S.,Huang, Q.,Tan, X., …Huang, T.(2008). Mitochondrial genome analysis of Ochotona curzoniae and implication of cytochrome c oxidase in hypoxic adaptation. Mitochondrion, 8(5–6), 352357. [PubMed][Google Scholar]
  • 32.Mah, C. L., &Blake, D. B.(2012). Global diversity and phylogeny of the Asteroidea (Echinodermata). PLoS ONE, 7(4), e35644[PubMed][Google Scholar]
  • 33.Mah, C., &Foltz, D.(2011). Molecular phylogeny of the Forcipulatacea (Asteroidea: Echinodermata): Systematics and biogeography. Zoological Journal of the Linnean Society, 162, 646660. [PubMed][Google Scholar]
  • 34.Maliarchuk, B. A.(2011). Adaptive evolution of the Homo mitochondrial genome. Molecular Biology, 45(5), 845850.[Google Scholar]
  • 35.Matsubara, M.,Komatsu, M.,Araki, T.,Asakawa, S.,Yokobori, S.,Watanabe, K., &Wada, H.(2005). The phylogenetic status of Paxillosida (Asteroidea) based on complete mitochondrial DNA sequences. Molecular Phylogenetics and Evolution, 36(3), 598605. [PubMed][Google Scholar]
  • 36.Matsubara, M.,Komatsu, M., &Wada, H.(2004). Close relationship between Asterina and Solasteridae (Asteroidea) supported by both nuclear and mitochondrial gene molecular phylogenies. Zoological Science, 21(7), 785793. [PubMed][Google Scholar]
  • 37.Mishmar, D.,Ruiz‐Pesini, E.,Golik, P.,Macaulay, V.,Clark, A. G.,Hosseini, S., …Wallace, D. C.(2003). Natural selection shaped regional mtDNA variation in humans. Proceedings of the National Academy of Sciences of the United States of America, 100(1), 171176. [PubMed][Google Scholar]
  • 38.Peden, J. F.(1999). Analysis of codon usage. Ph.D. Thesis, University of Nottingham.
  • 39.Perseke, M.,Bernhard, D.,Fritzsch, G.,Brümmer, F.,Stadler, P. F., &Schlegel, M.(2010). Mitochondrial genome evolution in Ophiuroidea, Echinoidea, and Holothuroidea: Insights in phylogenetic relationships of Echinodermata. Molecular Phylogenetics and Evolution, 56(1), 201211. [PubMed][Google Scholar]
  • 40.Perseke, M.,Fritzsch, G.,Ramsch, K.,Bernt, M.,Merkle, D.,Middendorf, M., …Schlegel, M.(2008). Evolution of mitochondrial gene orders in echinoderms. Molecular Phylogenetics and Evolution, 47(2), 855864. [PubMed][Google Scholar]
  • 41.Plazzi, F.,Ribani, A., &Passamonti, M.(2013). The complete mitochondrial genome of Solemya velum (Mollusca: Bivalvia) and its relationships with Conchifera. BMC Genomics, 14, 409[PubMed][Google Scholar]
  • 42.Qureshi, S. A., &Jacobs, H. T.(1993). Two distinct, sequence‐specific DNA‐binding proteins interact independently with the major replication pause region of sea urchin mtDNA. Nucleic Acids Research, 21(12), 28012808. [PubMed][Google Scholar]
  • 43.Ren, J.,Liu, X.,Jiang, F.,Guo, X., &Liu, B.(2010). Unusual conservation of mitochondrial gene order in Crassostrea oysters: Evidence for recent speciation in Asia. BMC Evolutionary Biology, 10, 394[PubMed][Google Scholar]
  • 44.Scouras, A.,Beckenbach, K.,Arndt, A., &Smith, M. J.(2004). Complete mitochondrial genome DNA sequence for two ophiuroids and a holothuroid: The utility of protein gene sequence and gene maps in the analyses of deep deuterostome phylogeny. Molecular Phylogenetics and Evolution, 31(1), 5065. [PubMed][Google Scholar]
  • 45.Scouras, A., &Smith, M. J.(2006). The complete mitochondrial genomes of the sea lily Gymnocrinus richeri and the feather star Phanogenia gracilis: Signature nucleotide bias and unique nad4L gene rearrangement within crinoids. Molecular Phylogenetics and Evolution, 39(2), 323334. [PubMed][Google Scholar]
  • 46.Shen, Y.,Kou, Q.,Zhong, Z.,Li, X.,He, L.,He, S., &Gan, X.(2017). The first complete mitogenome of the South China deep‐sea giant isopod Bathynomus sp. (Crustacea: Isopoda: Cirolanidae) allows insights into the early mitogenomic evolution of isopods. Ecology and Evolution, 7(6), 18691881. [PubMed][Google Scholar]
  • 47.Shen, X.,Ma, X.,Ren, J., &Zhao, F.(2009). A close phylogenetic relationship between Sipuncula and Annelida evidenced from the complete mitochondrial genome sequence of Phascolosoma esculenta. BMC Genomics, 10, 136[PubMed][Google Scholar]
  • 48.Smith, M. J.,Arndt, A.,Gorski, S., &Fajber, E.(1993). The phylogeny of echinoderm classes based on mitochondrial gene arrangements. Journal of Molecular Evolution, 36(6), 545554. [PubMed][Google Scholar]
  • 49.Smith, M. J.,Banfield, D. K.,Doteval, K.,Gorski, S., &Kowbel, D. J.(1990). Nucleotide sequence of nine protein‐coding genes and 22 tRNAs in the mitochondrial DNA of the sea star Pisaster ochraceus. Journal of Molecular Evolution, 31(3), 195204. [PubMed][Google Scholar]
  • 50.Sodergren, E.,Weinstock, G. M.,Davidson, E. H.,Cameron, R. A.,Gibbs, R. A.,Angerer, R. C., …Wright, R.(2006). The Genome of the Sea Urchin Strongylocentrotus purpuratus. Science, 314(5801), 941952.[PubMed][Google Scholar]
  • 51.Sun, S.,Hui, M.,Wang, M., &Sha, Z.(2018). The complete mitochondrial genome of the alvinocaridid shrimp Shinkaicaris leurokolos (Decapoda, Caridea): Insight into the mitochondrial genetic basis of deep‐sea hydrothermal vent adaptation in the shrimp. Comparative Biochemistry and Physiology Part D Genomics Proteomics, 25, 4252. [PubMed][Google Scholar]
  • 52.Sun, X. J.,Qi, L., &Kong, L. F.(2010). Comparative mitochondrial genomics within sea cucumber (Apostichopus japonicus): Provide new insights into relationships among color variants. Aquaculture, 309(1–4), 280285. [PubMed][Google Scholar]
  • 53.Tamura, K.,Peterson, D.,Peterson, N.,Stecher, G.,Nei, M., &Kumar, S.(2011). MEGA5: Molecular evolutionary genetics analysis using maximum likelihood, evolutionary distance, and maximum parsimony methods. Molecular Biology and Evolution, 28(10), 27312739. [PubMed][Google Scholar]
  • 54.Tang, M.,Tan, M.,Meng, G.,Yang, S.,Su, X.,Liu, S., …Zhang, A.(2014). Multiplex sequencing of pooled mitochondrial genomes‐a crucial step toward biodiversity analysis using mito‐metagenomics. Nucleic Acids Research, 42(22), e166[PubMed][Google Scholar]
  • 55.Tomasco, I. H., &Lessa, E. P.(2011). The evolution of mitochondrial genomes in subterranean caviomorph rodents: Adaptation against a background of purifying selection. Molecular Phylogenetics and Evolution, 61, 6470. [PubMed][Google Scholar]
  • 56.Wallace, D. C.(2007). Why do we still have a maternally inherited mitochondrial DNA? Insights from evolutionary medicine. Annual Review of Biochemistry, 76, 781821. [PubMed][Google Scholar]
  • 57.Wolstenholme, D. R.(1992). Animal mitochondrial DNA: Structure and evolution. International Review of Cytology, 141(6), 173216. [PubMed][Google Scholar]
  • 58.Wyman, S. K.,Jansen, R. K., &Boore, J. L.(2004). Automatic annotation of organellar genomes with DOGMA. Bioinformatics, 20(17), 32523255. [PubMed][Google Scholar]
  • 59.Xu, S.,Luosang, J.,Hua, S.,He, J.,Ciren, A.,Wang, W., …Zheng, X.(2007). High altitude adaptation and phylogenetic analysis of Tibetan horse based on the mitochondrial genome. Journal of Genetics and Genomics, 34(8), 720729. [PubMed][Google Scholar]
  • 60.Xu, B., &Yang, Z.(2013). PAMLX: A graphical user interface for PAML. Molecular Biology and Evolution, 30(12), 27232724. [PubMed][Google Scholar]
  • 61.Yang, Z.(2007). PAML 4: Phylogenetic analysis by maximum likelihood. Molecular Biology and Evolution, 24(8), 15861591. [PubMed][Google Scholar]
  • 62.Yang, Z.,Wong, W. S., &Nielsen, R.(2005). Bayes empirical Bayes inference of amino acid sites under positive selection. Molecular Biology and Evolution, 22(4), 11071118. [PubMed][Google Scholar]
  • 63.Yang, J.‐S., &Yang, W.‐J.(2008). The complete mitochondrial genome sequence of the hydrothermal vent galatheid crab Shinkaia crosnieri (Crustacea: Decapoda: Anomura): A novel arrangement and incomplete tRNA suite. BMC Genomics, 9, 257[PubMed][Google Scholar]
  • 64.Yasuda, N.,Hamaguchi, M.,Sasaki, M.,Nagai, S.,Saba, M., &Nadaoka, K.(2006). Complete mitochondrial genome sequences for Crown‐of‐thorns starfish Acanthaster planci and Acanthaster brevispinus. BMC Genomics, 7(1), 17[PubMed][Google Scholar]
  • 65.Yu, L.,Wang, X.,Ting, N., &Zhang, Y.(2011). Mitogenomic analysis of Chinese snub‐nosed monkeys: Evidence of positive selection in NADH dehydrogenase genes in high‐altitude adaptation. Mitochondrion, 11(3), 497503. [PubMed][Google Scholar]
  • 66.Zhang, B.,Zhang, Y. H.,Wang, X.,Zhang, H. X., &Lin, Q.(2017). The mitochondrial genome of a sea anemone Bolocera sp. exhibits novel genetic structures potentially involved in adaptation to the deep‐sea environment. Ecology and Evolution, 7(13), 49514962. [PubMed][Google Scholar]
  • 67.Zhou, T.,Shen, X.,Irwin, D. M.,Shen, Y., &Zhang, Y.(2014). Mitogenomic analyses propose positive selection in mitochondrial genes for high‐altitude adaptation in galliform birds. Mitochondrion, 18, 7075. [PubMed][Google Scholar]
Collaboration tool especially designed for Life Science professionals.Drag-and-drop any entity to your messages.