Quick viewing(Text Mode)

Evidence for Adaptive Selection in the Mitogenome of a Mesoparasitic Monogenean Flatworm Enterogyrus Malmbergi

Evidence for Adaptive Selection in the Mitogenome of a Mesoparasitic Monogenean Flatworm Enterogyrus Malmbergi

G C A T T A C G G C A T genes

Article Evidence for Adaptive Selection in the Mitogenome of a Mesoparasitic Monogenean Enterogyrus malmbergi

Dong Zhang 1,2,3 , Hong Zou 1,2, Shan G. Wu 1,2,3, Ming Li 1,2,3, Ivan Jakovli´c 4 , Jin Zhang 4, Rong Chen 4, Wen X. Li 1,2,3,* and Gui T. Wang 1,2,3 1 State Key Laboratory of Freshwater Ecology and Biotechnology, Institute of Hydrobiology, Chinese Academy of Sciences, Wuhan 430072, China; [email protected] (D.Z.); [email protected] (H.Z.); [email protected] (S.G.W.); [email protected] (M.L.); [email protected] (G.T.W.) 2 Key Laboratory of Aquaculture Disease Control, Ministry of Agriculture, Institute of Hydrobiology, Chinese Academy of Sciences, Wuhan 430072, China 3 University of Chinese Academy of Sciences, Beijing 100000, China 4 Bio-Transduction Lab, Wuhan 430075, China; [email protected] (I.J.); [email protected] (J.Z.); [email protected] (R.C.) * Correspondence: [email protected]

 Received: 19 September 2019; Accepted: 29 October 2019; Published: 30 October 2019 

Abstract: Whereas a majority of monogenean flatworms are ectoparasitic, i.e., parasitize on external surfaces (mainly gills) of their fish hosts, Enterogyrus species (subfamily Ancyrocephalinae) are mesoparasitic, i.e., parasitize in the stomach of the host. As there are numerous drastic differences between these two environments (including lower oxygen availability), we hypothesized that this life-history innovation might have produced adaptive pressures on the energy metabolism, which is partially encoded by the mitochondrial genome (OXPHOS). To test this hypothesis, we sequenced mitochondrial genomes of two Ancyrocephalinae species: mesoparasitic E. malmbergi and ectoparasitic Ancyrocephalus mogurndae. The mitogenomic architecture of E. malmbergi is mostly standard for monogeneans, but that of A. mogurndae exhibits some unique features: missing trnL2 gene, very low AT content (60%), a non-canonical start codon of the nad2 gene, and exceptionally long tandem-repeats in the non-coding region (253 bp). Phylogenetic analyses produced paraphyletic Ancyrocephalinae (with embedded Dactylogyrinae), but with low support values. Selective pressure (PAML and HYPHY) and protein structure analyses all found evidence for adaptive evolution in cox2 and cox3 genes of the mesoparasitic E. malmbergi. These findings tentatively support our hypothesis of adaptive evolution driven by life-history innovations in the mitogenome of this species. However, as only one stomach-inhabiting mesoparasitic monogenean was available for this analysis, our findings should be corroborated on a larger number of mesoparasitic monogeneans and by physiological studies.

Keywords: Ancyrocephalus mogurndae; positive selection; ; ; flatworms; replication initiation; mtDNA

1. Background In contrast to a vast majority (more than 95% in estimation [1]) of other monogenean flatworms, which parasitize on external body surfaces, such as fins, skins, and gills of fish hosts, species belonging to the mesoparasitic genus Enterogyrus (Ancyrocephalinae, Dactylogyridae) parasitize in the digestive system (stomach) of their fish hosts [2,3]. This genus currently contains eight known species [4]. They insert the sharp curved terminal ends of their anchors into the epithelial tissue (stomach wall) [3,5]; thus, causing wounds and chronic morbidity to the host [5]. Their peculiar mesoparasitic lifestyle enables

Genes 2019, 10, 863; doi:10.3390/genes10110863 www.mdpi.com/journal/genes Genes 2019, 10, 863 2 of 16 them to survive environmental changes associated with long inter-continental marine migrations (such as variations in salinity), which results in their vicariant global distribution, unlike that of ectoparasitic monogeneans [6,7]. However, it remains unclear which physiological and genetic adaptations facilitated this migration from gills and fins (plesiomorphic state for monogeneans [8]) to the drastically different, and arguably, inhospitable environment of the stomach, characterized by high acidity, absence of light, and lower oxygen availability, etc. Mitochondrion is the energy factory in metazoans, by encoding the genes that make up four of the five subunits of the oxidative phosphorylation enzymatic complex (OXPHOS): Complex I (nad1–6), III (cytb), IV (cox1–3), and V (atp6 and atp8) (Complex II is completely encoded by the nuclear genes) [9]. Due to the abundance of mitochondria in cells, (mostly) maternal inheritance, absence of introns, small genomic size (in metazoans), and an increasingly large set of available orthologous sequences, mitochondrial genomes (mitogenomes) have become a popular tool in population genetics [10], phylogenetics [11,12], and diagnostics [13] studies. An additional reason for their popularity in these types of studies is the fact that they were long considered to be an almost neutral marker, evolving under a strong purifying selection [14–16]. However, there is growing evidence for episodic positive selection in mitogenomes of some taxa, mostly identifiable as non-synonymous substitutions in second and first codon positions of OXPHOS genes [9,17–19]. In particular, radical changes in physicochemical properties of amino acids are usually seen as evidence for the signature of positive selection [17,19]. These amino acid (AA) substitutions can increase or decrease the aerobic capacity in response to the environmental adaption [18]. For example, Scott et al. [19] identified a mutation in the cox3 gene of bar-headed goose that may have resulted in adaptations in mitochondrial enzyme kinetics and O2 transport capacity, and thereby enabled this species to fly at exceptionally high altitudes; Almeida et al. [9] found adaptive mitogenomic changes to the deep sea environment in cephalopods; Guo et al. [20] detected hypoxia adaptation in the mitogenomes of anchialine cave shrimps; and Yu et al. [21] found evidence of positive selection in the mitochondrial NADH dehydrogenase genes of Chinese snub-nosed monkeys that live in high-altitudes. The energy metabolism, crucial for any living organism, is affected by the environmental conditions [22]. The examples above show that environmental changes (especially the concentration of oxygen) can increase the fitness of an organism carrying a beneficial mutation, which can be detected as signals of adaptive selection in mitochondrial OXPHOS genes. Adult helminths are generally considered to be facultative aerobes, and their oxygen consumption is in direct positive correlation with the availability of oxygen [23]. Hypoxic environment limits the efficiency of metabolic pathways that rely on acetyl-CoA as a substrate and decreases the energy output of mitochondria [24]. For example, some endoparasites of mammals adapt to the low oxygen availability in the digestive tract (which is about 1/4 of the air concentration) by switching to the anaerobic glycolytic phosphoenolpyruvate carboxykinase (PEPCK)-succinate pathway [25]. Although much less is known about metabolic adaptations of fish parasites, it is reasonable to assume that a switch from the ectoparasitic lifestyle (the gills environment being characterized by a relative abundance of oxygen and almost neutral PH) to the mesoparasitic lifestyle (the stomach environment being characterized by much lower availability of oxygen and highly acidic PH) must have been accompanied by comparable metabolic adaptations. Therefore, as the ancestor of the stomach-dwelling parasites from the genus Enterogyrus was almost certainly an ectoparasite (gills of fins) [8], this genus presents a good model to study signatures of selection in proteins encoded by the mitochondrial genome. So far, such studies were hindered by the unavailability of mitogenomic data for the mesoparasitic monogeneans. Here, we sequenced and characterized two complete mitochondrial genomes belonging to two species from the subfamily Ancyrocephalinae with different life-histories, ectoparasitic (gills) Ancyrocephalus mogurndae (Yamaguti, 1940), and mesoparasitic (stomach) E. malmbergi Bilong Bilong, 1988, and searched for the evidence of adaptive selection signals in the latter mitogenome. Genes 2019, 10, 863 3 of 16

2. Materials and Methods

2.1. Specimen Collection and Identification A. mogurndae was obtained from the gills of Siniperca chuatsi (Basilewsky, 1855) (Centrarchiformes: Sinipercidae) bought at a local market in the Wuhan city, Hubei Province on 17th of January 2017. E. malmbergi was obtained from the stomach of Oreochromis niloticus (Linnaeus, 1758) (Cichliformes: Cichlidae), caught in a freshwater pond in the Sun Yat-sen University (Guangzhou city, Guangdong province). Morphological identification of the latter species was conducted as described previously [3], so further details are not given here, whereas A. mogurndae was identified morphologically by the hard parts of the haptor and reproductive organs, as described in Wu et al. [26]. Additionally, to confirm the taxonomic identity of A. mogurndae, its 28S rRNA gene was amplified using a primer pair designed before [27]: C1 (50-ACCCGCTGAATTTAAGCAT-30) and D2 (50-TGGTCCGTGTTTCAAGAC-30). It shares a very high identity of 99.74% with conspecific homologs available in GenBank: 779/781 identical bp with AY841871. All sampled and identified parasites were first washed in 0.6% saline and then stored in 100% ethanol at 4 ◦C.

2.2. DNA Extraction, Amplification, and Sequencing Due to a small size of these parasites, to ensure a sufficient amount of DNA for amplification and sequencing, we used two kinds of genomic DNA: mixture DNA (20 parasite specimens) and individual DNA (a single parasite specimen). Both were extracted using TIANamp MicroDNA Kit (Tiangen Biotech, Beijing, China). The mixture DNA was first used to amplify the whole mitogenome. First, we selected 14 monogenean mitogenomes from the GenBank, aligned them using ClustalX [28], and designed degenerate primer pairs (Supplementary Table S1) matching the generally conserved regions of mitochondrial genes. On the basis of these obtained fragments, specific primers were designed using Primer Premier 5 [29], and the remaining mitogenome was amplified and sequenced in several PCR steps (Supplementary Table S1). PCR products were sequenced bi-directionally using both degenerate and specific primers mentioned above on an ABI 3730 automatic sequencer using the Sanger method. All obtained fragments were BLASTed [30] to confirm that the amplicon was the actual target sequence. To address the possibility of intraspecific sequence variation present in the mixture DNA, we inspected all chromatograms for double peaks or any other sign of the existence of two different sequences, and then we used individual DNA and long-range PCR to further verify the obtained sequences. If we found two different sequences, we used the DNA extracted from a single specimen to assemble the mitogenomic sequence, thereby ensuring that the entire sequence belongs to a single specimen.

2.3. Sequence Annotation and Analyses Both mitogenomes were assembled and annotated following the procedure described before [31–34] using DNAstar v7.1 software [35], and MITOS [36] and ARWEN [37] web tools. The annotated genome was recorded in a Word document, and PhyloSuite software [38] was then used to parse and extract the annotation information, as well as create GenBank submission files and organization tables for mitogenomes. Data for the comparative genomics analyses were also extracted and processed using PhyloSuite. Tandem Repeats Finder [39] was employed to find tandem repeats in the long non-coding regions (LNCR), and their secondary structures were predicted by the Mfold software [40]. Genetic distances (identity) among mitogenomic sequences were calculated with the “DistanceCalculator” function in Biopython using the “identity” model.

2.4. Phylogenetic Analyses Phylogenetic analyses were conducted using the two newly sequenced ancyrocephaline mitogenomes and 28 monogenean mitogenomes available in the GenBank (22/June/2019). Among these species, six polyopisthocotylean monogeneans were used as outgroups (Supplementary Table S2). Amino acid alignment of 12 protein-coding genes was used for the phylogenetic analyses. Best-fit Genes 2019, 10, 863 4 of 16 homogeneous model (mtZOA+F+I+G4) was calculated with ModelFinder [41], integrated in PhyloSuite. On the basis of the selected model, phylogenetic analyses were performed using two different algorithms: maximum likelihood (ML) and Bayesian inference (BI). ML analysis was conducted using RAxML [42] using an ML+rapid bootstrap (BS) algorithm with 1000 replicates. Bayesian inference analysis with the empirical MTZOA model was conducted using PhyloBayes (PB) MPI 1.5a [43]. For each analysis, two MCMC chains were run after the removal of invariable sites from the alignment, and the analysis was stopped when the conditions considered to indicate a good run (PhyloBayes manual) were reached: maxdiff <0.1 and minimum effective size >300. Phylograms were visualized and annotated using iTOL [44], with the help of several dataset files generated by PhyloSuite.

2.5. Selection Analyses Based on the tree reconstructed above, mutation pressures acting on the 12 mt OXPHOS genes were tested by comparing the nonsynonymous/synonymous substitution ratios (ω = dN/dS), through which positive (ω > 1), neutral (ω = 1), and purifying (ω < 1) selection can be distinguished. To avoid mutational saturation, which can blur the evolutionary signal at higher taxonomic levels, only monopisthocotyleans monogeneans, excluding Gyrodactylidae, were included in the selection analysis (15 species in total). First, we invoked the CODEML package in PAML 4.7 [45] to conduct a branch-site evolutionary analysis of the 12 PCGs. The modified branch-site model A allowed ω to vary, both among sites in the protein and across branches on the tree. After specifying foreground-lineages (E. malmbergi), the null model fixes ω to 1, whereas the alternative model assumes that sequences may have experienced positive selection (ω > 1). The significance of differences between the two models was assessed using likelihood ratio tests (LRTs). Subsequently, positively selected sites of genes with p < 0.05 in LRTs were further evaluated using the Bayes Empirical Bayes (BEB) [45] analysis with posterior probabilities 0.95. A mixed-effects model of evolution (MEME) and fixed-effect likelihood ≥ (FEL) methods implemented in HYPHY [46] were also used to infer the positively selected sites on the same foreground-lineage. Sites were considered as candidates under positive selection when they met the following conditions: β+ > α, significant likelihood ratio test (p < 0.05) in MEME, and p < 0.05 in the likelihood ratio test of FEL. The optimal 3D structures of proteins were predicted using I-TASSER [47] on the basis of the confidence score (C-score) value. Superimposition, visualization, and manipulation of the 3D structures were conducted using PYMOL [48], and the prediction of protein structure and function using the PredictProtein server [49].

3. Results

3.1. Genome Organization and Base Composition The complete mitochondrial genomes of E. malmbergi (MN095193) and A. mogurndae (MN095192) are circular molecules of 14,760 bp and 14,107 bp, respectively (Figure1). The E. malmbergi mitogenome contains the standard [50] 36 flatworm mitochondrial genes (12 protein-encoding genes (PCGs; atp8 is absent), 22 tRNA genes, and 2 rRNA genes), whereas we managed to identify only 35 genes in the mitogenome of A. mogurndae, from which the trnL2 gene appears to be missing (Table1 and Figure1). The PCGs encoded by both mitogenomes mostly used canonical start and stop codons (start: ATG and GTG; stop: TAA and TAG), but the nad2 gene of A. mogurndae uses non-canonical codons: the abbreviated version of TAA (T–) as the stop codon (otherwise relatively common, but unique in our dataset), and it proved difficult to determine the initial codon (Table1 and Supplementary Table S3). The architecture and similarity of orthologous sequences of the two newly sequenced mitogenomes are summarized in Table1. The average sequence similarity of PCGs between the two studied mitogenomes ranged from 43.14% (nad4L) to 66.57% (cytb) (Table1). The gene order of E. malmbergi (Figure1) was identical to that of E. johnii, Tetrancistrum nebulosi, Cichlidogyrus sclerosus Genes 2019, 10, 863 5 of 16 and Paragyrodactylus variegatus, whereas the gene order of A. mogurndae was almost identical, with the exception of a transposition of trnK and the missing trnL2. Genes 2019, 10, x FOR PEER REVIEW 5 of 15

FigureFigure 1. Visual 1. Visual representation representation of the of circular the circular mitochondrial mitochondrial genomes genomes of A. of mogurndae A. mogurndaeand E.and malmbergi E. . Protein-codingmalmbergi. Protein genes- arecoding red, genes tRNAs are arered, yellow, tRNAs are rRNAs yellow, are rRNAs green, are and green, non-coding and non-coding regions regions are grey. are grey. Table 1. Comparison of the annotated mitochondrial genomes of Ancyrocephalus mogurndae and EnterogyrusTable 1. malmbergi Comparison. of the annotated mitochondrial genomes of Ancyrocephalus mogurndae and Enterogyrus malmbergi. Position Intergenic Codon Gene Position Size Intergenic Codon Anti-codonAnti- Identity Gene Size Nucleotides Identity FromFrom To To Nucleotides Start Start Stop Stop codon EE.. malmbergi /A/A.. mogurndae mogurndae cox1cox1 1/11/1 1551/1572 1551/1572 1551/1572 1551/1572 ATG/ATGGTG/GTG TAG/TAG TAG/TAG 66.5466.54 trnTtrnT 1579/15581579/1558 1646/1627 1646/1627 68/70 68 /7027/− 2715/ 15 TGT/TGTTGT/TGT 57.1457.14 − rrnLrrnL 1647/16281647/1628 2599/2571 2599/2571 953/944 953/ 944 66.07 66.07 trnCtrnC 2600/25722600/2572 2665/2638 2665/2638 66/67 66 /67 GCA/GCAGCA/GCA 5050 rrnSrrnS 2666/26392666/2639 3408/3357 3408/3357 743/719 743/ 719 60.64 60.64 cox2cox2 3409/33793409/3379 4005/3951 4005/3951 597/573 597/ 5730/21 0 /21 ATG/ATGATG/ATG TAA/TAA TAA/TAA 55.0255.02 trnEtrnE 4001/39474001/3947 4067/4014 4067/4014 67/68 67 /68 −5/−55 / 5 TTC/TTCTTC/TTC 47.8947.89 − − nad6nad6 4071/40224071/4022 4517/4468 4517/4468 447/447 447/ 4473/7 3/7 ATG/GTGATG/GTG TAA/TAG TAA/TAG 50.5650.56 trnYtrnY 4519/44764519/4476 4586/4542 4586/4542 68/67 68 /671/7 1/7 GTA/GTAGTA/GTA 47.8347.83 trnL1trnL1 4585/46494585/4649 4650/4709 4650/4709 66/61 66 /61 −2/142 /14 TAG/TAGTAG/TAG 44.9344.93 − trnS2trnS2 4651/45704651/4570 4716/4634 4716/4634 66/65 66 /650/27 0 /27 TGA/TGATGA/TGA 67.1667.16 trnL2trnL2 4727/4727-/ -4794/ 4794- /-68/- 68/-10/0 10 /0 TAA/ TAA-/ - trnRtrnR 4802/49394802/4939 4865/5003 4865/5003 64/65 64 /657/229 7/229 TCG/TCGTCG/TCG 64.6264.62 nad5nad5 4867/50094867/5009 6420/6568 6420/6568 1554/1560 1554/1560 1/5 1/5 ATG/ATGATG/ATG TAG/TAG TAG/TAG 50.3550.35 trnGtrnG 7404/79997404/7999 7469/8065 7469/8065 66/67 66 /67983/1430 983/1430 TCC/TCCTCC/TCC 56.7256.72 cox3cox3 7470/80707470/8070 8144/8723 8144/8723 675/654 675/ 6540/4 0/4 ATG/ATGATG/ATG TAA/TAG TAA/TAG 50.8150.81 trnHtrnH 8125/87198125/8719 8189/8782 8189/8782 65/64 65 /64 −20/−205 / 5 GTG/GTGGTG/GTG 69.2369.23 − − cytbcytb 8193/87868193/8786 9269/9871 9269/9871 1077/1086 1077/1086 3/3 3/3 ATG/ATGATG/ATG TAA/TAG TAA/TAG 66.5766.57 nad4Lnad4L 9278/98629278/9862 9529/10107 9529/10107 252/246 252/ 2468/−10 8/ 10 ATG/ATGATG/ATG TAG/TAG TAG/TAG 43.1443.14 − nad4nad4 9502/100959502/10095 10716/11309 10716/11309 1215/1215 1215/1215 −28/−2813/ 13 ATG/ATGATG/ATG TAG/TAG TAG/TAG 51.1851.18 − − trnQtrnQ 10717/1131110717/11311 10783/11372 10783/11372 67/62 67 /620/1 0/1 TTG/TTGTTG/TTG 64.7164.71 trnFtrnF 10782/1137210782/11372 10846/11437 10846/11437 65/66 65 /66 −2/−12 / 1 GAA/GAAGAA/GAA 68.6668.66 − − trnMtrnM 10839/1143010839/11430 10904/11495 10904/11495 66/66 66 /66 −8/−88 / 8 CAT/CATCAT/CAT 62.6962.69 − − atp6atp6 10908/1150510908/11505 11417/12017 11417/12017 510/513 510/ 5133/9 3/9 ATG/ATGATG/ATG TAA/TAG TAA/TAG 56.4856.48 nad2nad2 11422/1202511422/12025 12243/12850 12243/12850 822/826 822/ 8264/7 4/7 ATG/ATAATG/ATA TAG/TAGT/ T 47.32 47.32 trnVtrnV 12244/1285112244/12851 12308/12916 12308/12916 65/66 65 /66 TAC/TACTAC/TAC 65.1565.15 trnAtrnA 12316/1292112316/12921 12377/12982 12377/12982 62/62 62 /627/4 7/4 TGC/TGCTGC/TGC 74.1974.19 trnDtrnD 12385/1298312385/12983 12447/13048 12447/13048 63/66 63 /667/0 7/0 GTC/GTCGTC/GTC 43.9443.94 nad1nad1 12453/1304912453/13049 13340/13933 13340/13933 888/885 888/ 8855/0 5/0 ATG/ATGATG/ATG TAG/TAG TAG/TAG 63.1863.18 trnNtrnN 13346/1400613346/14006 13405/14073 13405/14073 60/68 60 /685/2 5/2 GTT/GTTGTT/GTT 6060 trnPtrnP 13421/1407613421/14076 13487/14141 13487/14141 67/66 67 /6615/2 15 /2 TGG/TGGTGG/TGG 62.6962.69 trnI 13487/13937 13554/14003 68/67 1/3 GAT/GAT 82.35 trnI 13487/13937 13554/14003 68/67 −1/3− GAT/GAT 82.35 trnKtrnK 13558/1456813558/14568 13622/14634 13622/14634 65/67 65 /673/7 3/7 CTT/CTTCTT/CTT 54.4154.41 nad3nad3 13624/1421613624/14216 13971/14560 13971/14560 348/345 348/ 3451/74 1 /74 ATG/ATGATG/ATG TAG/TAG TAG/TAG 51.1351.13 trnS1 13971/14634 14029/14691 59/58 1/ 1 GCT/TCT 67.8 trnS1 13971/14634 14029/14691 59/58 −1/−1 − GCT/TCT 67.8 trnWtrnW 14042/1469914042/14699 14104/14760 14104/14760 63/62 63 /6212/7 12 /7 TCA/TCATCA/TCA 61.961.9

Genes 2019, 10, 863 6 of 16

Genes 2019, 10, x FOR PEER REVIEW 6 of 15 3.2. Non-Coding Regions 3.2. Non-Coding Regions The long non-coding region (LNCR) was located between nad5 and trnG genes in both mitogenomes (FigureThe2). long The twonon-coding segments region were (LNCR) 983 bp wasand 1430 located bp between in size in nad5E. malmbergiand trnG andgenesA. in mogurndae, both respectively.mitogenomes The (Figure LNCR 2). ofTheE. two malmbergi segmentscontained were 983 two bp and highly 1430 repetitive bp in size regions in E. malmbergi (HRR): HRR1and A. was composedmogurndae, of respectively. four successive The LNCR tandem of repeatsE. malmbergi (TRs), co wherentained repeat two highly units repetitive 1–3 were regions identical (HRR): (122 bp), whereasHRR1 was unit composed 4 lost one of nucleotidefour successive at the tandem end; HRR2 repeats was (TRs), comprised where repeat of 23 units successively 1–3 were short identical TRs, in (122 bp), whereas unit 4 lost one nucleotide at the end; HRR2 was comprised of 23 successively short which units 1–22 were identical (16 bp), whereas the unit 23 was 3 bp shorter at the 30 end (Figure2). Similarly,TRs, in which we found units one 1–22 HRR were in identical the LNCR (16 of bp),A. mogurndae whereas the; it containedunit 23 was five 3 bp successive shorter at long the TRs, 3′ end where repeat(Figure units 2). Similarly, 1–3 were we identical found (253one HRR bp), whereasin the LNCR unit of 4 wasA. mogurndae longer, with; it contained one nucleotide five successive insertion at long TRs, where repeat units 1–3 were identical (253 bp), whereas unit 4 was longer, with one the 102nd position, and unit 5 exhibited one nucleotide deletion at the 253rd position (Figure2). The nucleotide insertion at the 102nd position, and unit 5 exhibited one nucleotide deletion at the 253rd consensus repeat patterns of the HRR1 in E. malmbergi and the HRR in A. mogurndae are capable of position (Figure 2). The consensus repeat patterns of the HRR1 in E. malmbergi and the HRR in A. forming stem-loop structures (Figure2). mogurndae are capable of forming stem-loop structures (Figure 2).

FigureFigure 2.2. StemStem-loop-loop structures structures of the of theconsensus consensus repeat repeat pattern pattern in repetitive in repetitive regions regions(RRs) of (RRs)the non of- the non-codingcoding region region (NCR) (NCR) of A of. mogurndaeA. mogurndae andand E. malmbergiE. malmbergi. Thermodynamic. Thermodynamic energy energy value valuess (dG) (dG) are are shownshown nextnext toto thethe secondary s structures.tructures. Only Only the the NCR NCR containing containing RRs RRs was was shown. shown.

3.3.3.3. PhylogenyPhylogeny andand SelectionSelection Analyses Analyses BIBI andand ML produced produced an an identical identical topology topology (Figure (Figure 3), wherein3), wherein A. mogurndaeA. mogurndae was resolvedwas resolved as the as thebasal basal species species within within the Dactylogyridae, the Dactylogyridae, followed followed by E. malmbergi by E. malmbergi. The rem. Theaining remaining dactylogyrids dactylogyrids were weredivided divided into twointo sister two sister-clades:-clades: (E. johni (E., johniDactylogyrus, lamellatus lamellatus) and (T.) and nebulosi (T. nebulosi, (C. mbirizei,(C., mbirizei (C. , (C.sclerosus sclerosus, C. ,halliC.))). halli However,))). However, the support the support values within values the within Dactylogyridae the Dactylogyridae clade in the ML clade tree in were the ML treevery were low very(Figure low 3). (Figure The subfamily3). The subfamilyAncyrocephalinae Ancyrocephalinae was rendered was paraphyletic rendered by paraphyletic the embedded by the Dactylogyrinae species, D. lamellatus. embedded Dactylogyrinae species, D. lamellatus.

Genes 2019, 10, 863 7 of 16

Genes 2019, 10, x FOR PEER REVIEW 7 of 15

FigureFigure 3. A3. A phylogram phylogram reconstructed reconstructed using using mitogenomes mitogenomes of of 30 30 monogeneans monogeneans(amino (amino acidacid sequencessequences of of all PCGs) and the mtZOA model. Scale bar corresponds to the estimated number of substitutions all PCGs) and the mtZOA model. Scale bar corresponds to the estimated number of substitutions per per site. Statistical support values of Bayesian analyses and maximum likelihood are shown on the site. Statistical support values of Bayesian analyses and maximum likelihood are shown on the nodes nodes (left/right, respectively). Taxonomic families and orders are shown in different colors. (left/right, respectively). Taxonomic families and orders are shown in different colors. Dactylogyrinae Dactylogyrinae and Ancyrocephalinae subfamilies are also displayed. and Ancyrocephalinae subfamilies are also displayed. With the mesoparasitic E. malmbergi selected as the test branch within the phylogenetic tree With the mesoparasitic E. malmbergi selected as the test branch within the phylogenetic tree described above (only the topology of monopisthocotylean monogeneans, excluding gyrodactylids, described above (only the topology of monopisthocotylean monogeneans, excluding gyrodactylids, was used), branch-site specific analyses in PAML, and MEME and FEL analyses in HYPHY, all was used), branch-site specific analyses in PAML, and MEME and FEL analyses in HYPHY, all detected detected evidence for positive selection in this branch (Figure 3, Table 2, and Supplementary Table evidence for positive selection in this branch (Figure3, Table2, and Supplementary Table S4). Under S4). Under the branch-site model (BSM) test, cox1, cox2, cox3, nad1, and nad5 were identified as genes theunder branch-site positive model selection (BSM) (p < test,0.05; cox1Supplementary, cox2, cox3 Table, nad1, S5and). However,nad5 were BEB identified analyses only as genes identified under positiveseveral selection codons in (p cox2< 0.05; and Supplementarycox3 genes as undergoing Table S5). positive However, selection BEB ( analysesp-value < only0.05). identified In comparison, several codonsthe MEME in cox2 methodand cox3 identifiedgenes asepis undergoingodic diversifying positive positive selection selection (p-value (p-value< 0.05). < 0.05) In in comparison, eleven PCGs the MEME(all PCGs method except identified nad1) (Supplementary episodic diversifying Table S4 positive). The FEL selection method (p -value found evidence< 0.05) in of eleven pervasive PCGs (allpositive PCGs exceptdiversifyingnad1 )selection (Supplementary in all 12 PCGs Table except S4). for The nad3 FEL, which method only found had 30 evidence sites under of n pervasiveegative positiveselection diversifying (at p ≤ 0.05; selection Supplementary in all 12 Table PCGs S4 except). Intriguingly, for nad3, whichnad1 was only special had 30again, sites with under only negative one selectionsite under (at p pervasive0.05; Supplementary positive diversifying Table S4). Intriguingly, selection andnad1 137was sites special under again, negative with only selection one site ≤ under(Supplementary pervasive positive Table S4 diversifying). selection and 137 sites under negative selection (Supplementary Table S4).As positively selected sites are more reliable when they are supported by two or more methods [51,52],As positively only cox2 and selected cox3 genes sites were are moreused in reliable subsequent when analyses they areof positive supported selection by twoon proteins. or more methodsPositively [51 ,selected52], only sitescox2 thatand werecox3 consistentlygenes were identified used in subsequent by three methods analyses (BSM, of positive MEME, selection and FEL) on proteins.were mapped Positively onto selected the predicted sites that3D structures were consistently of Cox2 and identified Cox3 proteins by three of E. methods malmbergi (BSM, (Figure MEME, 4). andPredictProtein FEL) were mapped analysis onto showed the predictedthat these 3Dpositively structures selected of Cox2 sites and were Cox3 located proteins in, or of closeE. malmbergi to, the (Figurefunctional4). PredictProtein regions: 13 positively analysis showedselected thatsites these were positivelylocated within selected a helical sites transmembrane were located in, region, or close and one positively selected site (codon 171 in cox2) was located within a protein-binding region to, the functional regions: 13 positively selected sites were located within a helical transmembrane (Figure 4, Table 2). region, and one positively selected site (codon 171 in cox2) was located within a protein-binding region (Figure4, Table2).

Genes 2019, 10, 863 8 of 16

Table 2. Positively selected sites of cox2 and cox3 genes identified by PAML, and HYPHY.

PAML Branch-Site HYPHY Model (p-Value < 0.05, PredictProtein (https: GENES AA Positions MEME Posterior Probabilities FEL //www.predictprotein.org//) p < 0.05 & 95%) p < 0.05 + ≥ β > α cox2 2   5   Protein binding region 6   Protein binding region 7  8  Protein binding region 9   Protein binding region 11  Helical transmembrane region 15   Helical transmembrane region 18    Helical transmembrane region 19    Helical transmembrane region 20   Helical transmembrane region 28  Helical transmembrane region 31   Helical transmembrane region 32  Helical transmembrane region 39   44   45   Protein binding region 46  47   49  Helical transmembrane region 51  Helical transmembrane region 53   Helical transmembrane region 54  Helical transmembrane region 57  Helical transmembrane region 60   Helical transmembrane region 61  Helical transmembrane region 62    Helical transmembrane region 64   Helical transmembrane region 66    Helical transmembrane region 68    Helical transmembrane region 69  Helical transmembrane region 71  72   73   74   79   Protein binding region 80   Protein binding region 81  Protein binding region 82   88    91  99  117  118  119  126   128  129  133    136  148   152  154  165    170  171    Protein binding region 177    Genes 2019, 10, 863 9 of 16

Table 2. Cont.

PAML Branch-Site HYPHY Model (p-Value < 0.05, PredictProtein (https: GENES AA Positions MEME Posterior Probabilities FEL //www.predictprotein.org//) p < 0.05 & 95%) p < 0.05 + ≥ β > α cox3 2   Helical transmembrane region 3    Helical transmembrane region 8   Helical transmembrane region 11  Helical transmembrane region 15   Helical transmembrane region 21  25  Helical transmembrane region 40   Helical transmembrane region 42    44   47   49  50    Protein binding region; 52  Polynucelotide-binding region 53   55    Helical transmembrane region 56    Helical transmembrane region 57    Helical transmembrane region 60  Helical transmembrane region 66   Helical transmembrane region 71  Helical transmembrane region 73   Helical transmembrane region 74  Helical transmembrane region 76    Helical transmembrane region 77   Protein binding region 82   83  85  90  91  94   Helical transmembrane region 96  Helical transmembrane region 97    Helical transmembrane region 101   Helical transmembrane region 104   Helical transmembrane region 111    Helical transmembrane region 112   Helical transmembrane region 115   130   Helical transmembrane region 131  Helical transmembrane region 132   Helical transmembrane region 139   Helical transmembrane region 147  149    150  154  155  156    157   160   Helical transmembrane region 163   Helical transmembrane region 164  Helical transmembrane region 165   Helical transmembrane region 171   Helical transmembrane region 175   Helical transmembrane region 179   Helical transmembrane region 180   Helical transmembrane region 184   185   Genes 2019, 10, 863 10 of 16

Table 2. Cont.

PAML Branch-Site HYPHY Model (p-Value < 0.05, PredictProtein (https: GENES AA Positions MEME Posterior Probabilities FEL //www.predictprotein.org//) p < 0.05 & 95%) p < 0.05 + ≥ β > α 187   Protein binding region 189  Protein binding region 191  195   196   197   199   Helical transmembrane region 216  Helical transmembrane region 217   Helical transmembrane region 218   Helical transmembrane region Helical transmembrane region; 220   Protein binding region 221    Helical transmembrane region

Genes 2019, 10, x FOR PEER REVIEW 8 of 15

Figure 4. Distribution of positively selected sites in the three-dimensional (3D) structures of Cox2 and Figure 4. Distribution of positively selected sites in the three-dimensional (3D) structures of Cox2 and Cox3 proteins. The positively selected sites are shown in red. Cox3 proteins. The positively selected sites are shown in red.

4. DiscussionTable 2. Positively selected sites of cox2 and cox3 genes identified by PAML, and HYPHY.

HYPHY Mitogenomic architecturePAML of E. Branch malmbergi-Site is mostly standard for monogeneans,PredictProtein but the mitogenome of A. mogurndae possessesAA Model some ( intriguingp-Value < 0.05, features: the absenceMEME of trnL2 gene, low AT content (the GENES FEL Positions Posterior Probabilities ≥ p < 0.05 & (https://www.predictprotein.o lowest among the Monogenea; Supplementaryp Table < 0.05 S1, Supplementary Figure S1), exceptionally 95%) β+ > α long tandem-repeats within the NCR, and non-canonical start/stop codonsrg//) of the nad2 gene. On the basis ofcox2 results reported2 in other related species [12, 53], as a working hypothesis, we propose ATA 5   Protein binding region as the initial codon of nad2 in A. mogurndae. The gene order shared between E. malmbergi and other 6   Protein binding region four monogeneans 7 were previously identified as the plesiomorphic gene order for the subphylum Neodermata [54].8 This finding further confirms the hypothesis that identicalProtein gene binding orders region found in phylogenetically distant9 lineages are an indicator of shared ancestry [55]. TRsProtein with highbinding repeat region numbers and large size have11 been reported before in the subclass Helical [transmembrane32,56], but with region 253 bp, the TR in A. mogurndae15 is the longest reported so far. Along with a few previousHelical transmembrane studies [32 region,56], these 18    Helical transmembrane region 19    Helical transmembrane region 20   Helical transmembrane region 28  Helical transmembrane region 31   Helical transmembrane region 32  Helical transmembrane region 39   44   45   Protein binding region 46  47   49  Helical transmembrane region 51  Helical transmembrane region 53   Helical transmembrane region 54  Helical transmembrane region 57  Helical transmembrane region 60   Helical transmembrane region 61  Helical transmembrane region 62    Helical transmembrane region 64   Helical transmembrane region 66    Helical transmembrane region 68    Helical transmembrane region

Genes 2019, 10, 863 11 of 16

findings consistently reject the hypothesis that monopisthocotylean monogeneans possess fewer and smaller (in size) TRs in the LNCR than polyopisthocotylean monogeneans [57]. Since the presence of tandem repeats forming a stable secondary structure is often associated with replication origin in mitochondria [12,58,59], it appears likely that the repeat regions found in the LNCR are embedded within the control region. Results of our phylogenetic analyses are partially consistent with the relationships obtained using 18S rDNA and 28S rDNA, in which subfamilies Dactylogyrinae, Ancylodiscoidinae, and Pseudodactylogyrinae were embedded within the subfamily Ancyrocephalinae [60–63], thus rendering it paraphyletic. However, this result supports the argument that, regardless of the exact status and interrelationships of these subfamilies, all of them originated within the family Dactylogyridae [64]. As nuclear and mitogenomic data reject the monophyly of Ancyrocephalinae, and even the morphological data do not provide full support for it [64], we can conclude with some confidence that a taxonomic revision shall be needed with regard to this taxon. However, the resolution of our approach is too limited (small number of samples) to officially propose it here, so future studies should address this issue using a larger number of samples and combined datasets (ideally all three data types: nuclear, mitochondrial, and morphological). Although only a small proportion of monogeneans are mesoparasitic (less than 5% in estimation), they have a wide host range (fishes, mammals, and amphibians), exhibit a wide range of parasitic lifestyles with respect to the exact location in the body where they parasitize (stomach, heart, esophagus, cloaca, urinary bladder, and inside the eyelid), and they form phylogenetically distant clades (different subclasses and families) [1,6]. This indicates multiple independent origins of the mesoparasitic lifestyle within the Monogenea. Due to limited availability of data, here, we only compared the selection pressures between ectoparasitic and stomach-inhabiting mesoparasitic monogeneans. Among the 15 species that we included in the selection analyses, 14 are ectoparasitic monogeneans found on the gills, skin, and fins of fish hosts, whereas E. malmbergi is a mesoparasitic monogenean found in the stomach of fish hosts. We hypothesized that starkly different environments of ectoparasitic and mesoparasitic monogeneans (especially the supposed reduced availability of oxygen in the stomach in comparison to the gills) would require metabolic adaptations, which might be reflected on the mitochondrial genome. Our analyses do show that dozens of codons of mitochondrial OXPHOS genes were subjected to strong positive selection pressures (Table2 and Supplementary Table S4) in E. malmbergi. Particularly, several positive selection sites in cox2 and cox3 were consistently validated by three different methods (Table2), which provides strong evidence that these genes may have sustained adaptive evolution in the mesoparasitic E. malmbergi. This was further supported by the evidence that many positively-selected codons were posited in, or close to, the predicted functional regions (helical transmembrane region and protein binding region) in the structure of mitochondrial OXPHOS proteins (Figure4 and Table2). cox2 and cox3, both of which were confirmed to have undergone positive selection in E. malmbergi by three methods, belong to the multi-subunit OXPHOS enzyme complex IV [9]. As this complex is believed to consume a majority of the O2 [65], this finding is indicative of metabolic adaptations to a comparatively (in relation to gills) hypoxic environment in the stomach in E. malmbergi. However, it remains unclear whether this adaptation is a reflection of an enhanced capacity for oxygen usage or a switch to a predominantly anaerobic metabolism. Some of the genes in other enzymatic complexes were also predicted to exhibit positive selection by one or two methods: Complex I, which is the largest OXPHOS enzymatic complex, complex III, which transfers electrons, and complex V, which mainly synthesizes the ATP [66]. Therefore, there is some evidence for positive selection in all four mitochondrial OXPHOS complexes, which suggests life-history innovation-driven adaptive evolution in the mitogenome of E. malmbergi. The interpretation of these findings is hampered by the scarcity of physiological studies on these parasitic species. Notably, the exact concentration of oxygen in the stomach of their fish hosts remains unknown, it is unclear whether mesoparasitic species employ aerobic or anaerobic metabolic pathways, and we do not know whether these parasites may be able to gain oxygen from the host’s Genes 2019, 10, 863 12 of 16 tissues and body fluids. However, as the majority of parasites do not use the oxygen available within the host [25], we can tentatively reject the last hypothesis. Furthermore, in mammals, the oxygen content in the small intestine is about 25% of that in the environment (air) [25], and in humans, the gas in the stomach contains approximately 15% oxygen (as opposed to 21% oxygen in the air) [67]. Therefore we can also safely assume that oxygen availability is significantly different between gills (which are constantly flushed by water) and the stomach environment, which is isolated from the surrounding water in freshwater fish most of the time (apart from when they are ingesting food) [68,69]. It should be noted that O. niloticus is a euryhaline fish that ingests water in high salinity environments, which may increase the availability of oxygen in the stomach. However, studies in a closely related species mossambica, which exhibits much higher salinity tolerance than O. niloticus [70], have shown that its ingestion of water is fairly low in a freshwater environment [71]. This implies that any stomach-dwelling parasite of freshwater populations would face limited accessibility of oxygen. Finally, as only one stomach-inhabiting mesoparasitic species was sequenced, we could not establish whether the amino-acid replacements observed in E. malmbergi are unique to this species or fixed in all species within this genus (all are mesoparasitic species).

5. Conclusions The existence of several unique architectural characteristics of the mitogenome of A. mogurndae is intriguing and indicates that the sequencing of further mitogenomes belonging to this and closely related genera may be interesting from the aspect of the evolution of mitogenomic architecture in the Monogenea. With regards to the status of the subfamily Ancyrocephalinae, the addition of two new mitogenomes (now seven available in total) did not manage to stabilize the topology, and the subfamily was paraphyletic. Although this is partially in agreement with previous molecular studies, support values for the topology were low, so we cannot draw any conclusions with confidence. More data is needed to resolve the status of this subfamily. Finally, in agreement with our hypothesis, we found evidence for adaptive evolution in the mitogenome of E. malmbergi. Although we tentatively attributed it to the adaptation to the evolution from the ectoparasitic to stomach-inhabiting mesoparasitic lifestyle, where the hypoxic environment of the stomach required adaptive changes in some mitochondrial OXPHOS genes, it remains unclear whether this adaptation is a reflection of an enhanced capacity for oxygen usage or a switch to a predominantly anaerobic metabolism. Also, due to a lack of physiological studies on these species, we cannot exclude other, less obvious explanations for these adaptations. In order to investigate the adaptive evolution of stomach-inhabiting mesoparasitic monogeneans comprehensively, we will have to sequence more data from other mesoparasitic species and simultaneously study their physiological and genetic adaptations to the mesoparasitic lifestyle.

Supplementary Materials: The following are available online at http://www.mdpi.com/2073-4425/10/11/863/s1, Figure S1: Hierarchical clustering of monogeneans based on their A + T content. The Family of each species is shown on the right. Table S1: Primers used to amplify and sequence mitochondrial genomes of Ancyrocephalus mogurndae and Enterogyrus malmbergi. Table S2: The Monogenea species used for comparative mitogenomic and phylogenetic analyses. Table S3: General statistics for the protein-coding and rRNA genes of the 30 selected monogenean species. Table S4: Codons under positive or negative selection inferred by two HYPHY methods: FEL and MEME. P: positive selection; N: negative selection. Table S5: Null and alternative formulations of Branch-site Model A and their parameter estimates for monogenean mitochondrial sequences. Author Contributions: G.T.W., W.X.L. and D.Z. designed the study. D.Z., H.Z., J.Z., S.G.W., M.L. and R.C. conducted the experiments. D.Z. conducted the data analysis. D.Z., W.X.L. and I.J. wrote the paper. All authors have revised the manuscript critically for important intellectual content and approved the final version of the manuscript. Funding: This work was funded by the National Natural Science Foundation of China (31872604), the Earmarked Fund for China Agriculture Research System (CARS-45-15). Acknowledgments: The authors would like to thank Ting B. Yang and Shuai Zhang for kindly providing the Enterogyrus malmbergi sample. Conflicts of Interest: The authors declare that they have no conflict of interest. Genes 2019, 10, 863 13 of 16

References

1. Euzet, L.; Combes, C. The selection of habitats among the Monogenea. Int. J. Parasitol. 1998, 28, 1645–1652. [CrossRef] 2. Paperna, I. Enterogyrus cichlidarum n gen. n. sp. a Monogenetic Trematode parasitic in the intestine of a fish. Bull. Res. Counc. Israel 1963, 11, 183–187. 3. Zhang, S.; Zhi, T.; Xu, X.; Zheng, Y.; Bilong Bilong, C.F.; Pariselle, A.; Yang, T. Monogenean fauna of alien (Cichlidae) in south China. Parasite 2019, 26, 4. [CrossRef] 4. Madanire-Moyo, G.; Avenant-Oldewage, A. A new locality and host record for Enterogyrus coronatus (Pariselle Lambert & Euzet (1991) in South Africa and a review of the morphology and distribution of Enterogyrus () species. Helminthologia 2014, 51, 13–22. 5. Madanire-Moyo, G.N.; Avenant-Oldewage, A. The histopathology of Enterogyrus coronatus Pariselle, Lambert & Euzet, 1991 (Monogenoidea) in the stomach of the southern mouthbrooder Pseudocrenilabrus philander (Weber, 1897) (Cichlidae). Afr. Zool. 2015, 50, 175–180. 6. Pariselle, A.; Boeger, W.A.; Snoeks, J.; Bilong Bilong, C.F.; Morand, S.; Vanhove, M.P. The monogenean parasite fauna of cichlids: A potential tool for host biogeography. Int. J. Evol. Biol. 2011, 2011, 471480. [CrossRef][PubMed] 7. Vanhove, M.P.M.; Hablutzel, P.I.; Pariselle, A.; Simkova, A.; Huyse, T.; Raeymaekers, J.A.M. Cichlids: A host of opportunities for evolutionary parasitology. Trends Parasitol. 2016, 32, 820–832. [CrossRef] 8. Bychowsky, B.E. Monogenetic Trematodes, Their Systematics and Phylogeny; Academy of Sciences: Moscow/Leningrad, Russia, 1957; p. 509. (In Russian) 9. Almeida, D.; Maldonado, E.; Vasconcelos, V.; Antunes, A. Adaptation of the mitochondrial genome in cephalopods: Enhancing proton translocation channels and the subunit interactions. PLoS ONE 2015, 10, e0135405. [CrossRef] 10. Shao, R.; Barker, S. Mitochondrial genomes of parasitic : Implications for studies of population genetics and evolution. Parasitology 2007, 134, 153–167. [CrossRef] 11. Perkins, E.M.; Donnellan, S.C.; Bertozzi, T.; Whittington, I.D. Closing the mitochondrial circle on of the Monogenea (Platyhelminthes) infers evolution in the diet of parasitic flatworms. Int. J. Parasitol. 2010, 40, 1237–1245. [CrossRef] 12. Park, J.K.; Kim, K.H.; Kang, S.; Kim, W.; Eom, K.S.; Littlewood, D.T. A common origin of complex life cycles in parasitic flatworms: Evidence from the complete mitochondrial genome of Microcotyle sebastis (Monogenea: Platyhelminthes). BMC Evol. Biol. 2007, 7, 11. [CrossRef][PubMed] 13. Huyse, T.; Buchmann, K.; Littlewood, D.T. The mitochondrial genome of derjavinoides (Platyhelminthes: Monogenea)—A mitogenomic approach for Gyrodactylus species and strain identification. Gene 2008, 417, 27–34. [CrossRef][PubMed] 14. Avise, J.C.; Ellis, D.; Clarke, B.C.; Robertson, A.; Jeffreys, A.J. Mitochondrial DNA and the evolutionary genetics of higher . Philos. Trans. R. Soc. B Biol. Sci. 1986, 312, 325–342. [CrossRef][PubMed] 15. Ballard, J.W.O.; Kreitman, M. Is mitochondrial DNA a strictly neutral marker? Trends Ecol. Evol. 1995, 10, 485–488. [CrossRef] 16. Ballard, J.W.O.; Pichaud, N.; Fox, C. Mitochondrial DNA: More than an evolutionary bystander. Funct. Ecol. 2014, 28, 218–231. [CrossRef] 17. Botero-Castro, F.; Tilak, M.K.; Justy,F.; Catzeflis, F.; Delsuc, F.; Douzery,E.J.P.In cold blood: Compositional bias and positive selection drive the high evolutionary rate of vampire bats mitochondrial genomes. Genome Biol. Evol. 2018, 10, 2218–2239. [CrossRef] 18. McNab, B.K. Energy constraints on carnivore diet. Nature 2000, 407, 584. [CrossRef] 19. Scott, G.R.; Schulte, P.M.; Egginton, S.; Scott, A.L.; Richards, J.G.; Milsom, W.K. Molecular evolution of cytochrome C oxidase underlies high-altitude adaptation in the bar-headed goose. Mol. Biol. Evol. 2011, 28, 351–363. [CrossRef] 20. Guo, H.; Yang, H.; Tao, Y.; Tang, D.; Wu, Q.; Wang, Z.; Tang, B. Mitochondrial OXPHOS genes provides insights into genetics basis of hypoxia adaptation in anchialine cave shrimps. Genes Genom. 2018, 40, 1169–1180. [CrossRef] Genes 2019, 10, 863 14 of 16

21. Yu, L.; Wang, X.; Ting, N.; Zhang, Y. Mitogenomic analysis of Chinese snub-nosed monkeys: Evidence of positive selection in NADH dehydrogenase genes in high-altitude adaptation. Mitochondrion 2011, 11, 497–503. [CrossRef] 22. Speakman, J.R. Body size, energy metabolism and lifespan. J. Exp. Biol. 2005, 208, 1717–1730. [CrossRef] [PubMed] 23. Siddiqi, A.H.; Nizami, W.A. Gas content of swim bladder of Wallago attu and oxygen consumption in Isoparorchis hypselobagri (Trematoda). Z. Parasitenkd. 1975, 47, 263–268. [CrossRef][PubMed] 24. Stairs, C.W.; Leger, M.M.; Roger, A.J. Diversity and origins of anaerobic metabolism in mitochondria and related organelles. Philos. Trans. R. Soc. Lond. B Biol. Sci. 2015, 370, 20140326. [CrossRef][PubMed] 25. Kita, K.; Hirawake, H.; Miyadera, H.; Amino, H.; Takeo, S. Role of complex II in anaerobic respiration of the parasite mitochondria from suum and Plasmodium falciparum. Biochim. Biophys. Acta Bioenerg. 2002, 1553, 123–139. [CrossRef] 26. Wu, B.h.; Lang, S.; Wang, W.J. Fauna Sinica: Platyhelminthes: Monogenea; Science Press: Beijing, China, 2000. 27. Hassouna, N.; Mithot, B.; Bachellerie, J.-P. The complete nucleotide sequence of mouse 28S rRNA gene. Implications for the process of size increase of the large subunit rRNA in higher eukaryotes. Nucleic Acids Res. 1984, 12, 3563–3583. [CrossRef] 28. Thompson, J.D.; Gibson, T.J.; Higgins, D.G. Multiple sequence alignment using ClustalW and ClustalX. Curr. Protoc. Bioinform. 2002.[CrossRef] 29. Lalitha, S. Primer Premier 5. Biot. Soft. Int. Rep. 2000, 1, 270–272. [CrossRef] 30. Altschul, S.F.; Gish, W.; Miller, W.; Myers, E.W.; Lipman, D.J. Basic local alignment search tool. J. Mol. Biol. 1990, 215, 403–410. [CrossRef] 31. Zhang, D.; Zou, H.; Wu, S.G.; Li, M.; Jakovli´c,I.; Zhang, J.; Chen, R.; Wang, G.T.; Li, W.X. Sequencing of the complete mitochondrial genome of a fish-parasitic flatworm Paratetraonchoides inermis (Platyhelminthes: Monogenea): tRNA gene arrangement reshuffling and implications for phylogeny. Parasit. Vectors 2017, 10, 462. [CrossRef] 32. Zhang, D.; Zou, H.; Wu, S.G.; Li, M.; Jakovlic, I.; Zhang, J.; Chen, R.; Wang, G.T.; Li, W.X. Sequencing, characterization and phylogenomics of the complete mitochondrial genome of Dactylogyrus lamellatus (Monogenea: Dactylogyridae). J. Helminthol. 2017, 92, 455–466. [CrossRef] 33. Li, W.X.; Zhang, D.; Boyce, K.; Xi, B.W.; Zou, H.; Wu, S.G.; Li, M.; Wang, G.T. The complete mitochondrial DNA of three monozoic tapeworms in the Caryophyllidea: A mitogenomic perspective on the phylogeny of eucestodes. Parasit. Vectors 2017, 10, 314. [CrossRef][PubMed] 34. Zou, H.; Jakovlic, I.; Chen, R.; Zhang, D.; Zhang, J.; Li, W.X.; Wang, G.T. The complete mitochondrial genome of parasitic Camallanus cotti: Extreme discontinuity in the rate of mitogenomic architecture evolution within the Chromadorea class. BMC Genom. 2017, 18, 840. [CrossRef][PubMed] 35. Burland, T.G. DNASTAR’s Lasergene sequence analysis software. Methods Mol. Biol. 2000, 132, 71–91. [PubMed] 36. Bernt, M.; Donath, A.; Juhling, F.; Externbrink, F.; Florentz, C.; Fritzsch, G.; Putz, J.; Middendorf, M.; Stadler, P.F. MITOS: Improved de novo metazoan mitochondrial genome annotation. Mol. Phylogenet. Evol. 2013, 69, 313–319. [CrossRef][PubMed] 37. Laslett, D.; Canback, B. ARWEN: A program to detect tRNA genes in metazoan mitochondrial nucleotide sequences. Bioinformatics 2008, 24, 172–175. [CrossRef][PubMed] 38. Zhang, D.; Gao, F.; Jakovli´c,I.; Zou, H.; Zhang, J.; Li, W.X.; Wang, G.T. PhyloSuite: An integrated and scalable desktop platform for streamlined molecular sequence data management and evolutionary phylogenetics studies. Mol. Ecol. Resour. 2019.[CrossRef] 39. Benson, G. Tandem repeats finder: A program to analyze DNA sequences. Nucleic Acids Res. 1999, 27, 573. [CrossRef] 40. Zuker, M. Mfold web server for nucleic acid folding and hybridization prediction. Nucleic Acids Res. 2003, 31, 3406–3415. [CrossRef] 41. Kalyaanamoorthy, S.; Minh, B.Q.; Wong, T.K.F.; von Haeseler, A.; Jermiin, L.S. ModelFinder: Fast model selection for accurate phylogenetic estimates. Nat. Methods 2017, 14, 587–589. [CrossRef] 42. Stamatakis, A. RAxML version 8: A tool for phylogenetic analysis and post-analysis of large phylogenies. Bioinformatics 2014, 30, 1312–1313. [CrossRef] Genes 2019, 10, 863 15 of 16

43. Lartillot, N.; Rodrigue, N.; Stubbs, D.; Richer, J. PhyloBayes MPI: Phylogenetic reconstruction with infinite mixtures of profiles in a parallel environment. Syst. Biol. 2013, 62, 611–615. [CrossRef][PubMed] 44. Letunic, I.; Bork, P. Interactive tree of life (iTOL) v3: An online tool for the display and annotation of phylogenetic and other trees. Nucleic Acids Res. 2016, 44, W242–W245. [CrossRef][PubMed] 45. Yang, Z. PAML 4: Phylogenetic analysis by maximum likelihood. Mol. Biol. Evol. 2007, 24, 1586–1591. [CrossRef][PubMed] 46. Pond, S.L.; Frost, S.D.; Muse, S.V. HyPhy: Hypothesis testing using phylogenies. Bioinformatics 2005, 21, 676–679. [CrossRef][PubMed] 47. Zhang, Y. I-TASSER server for protein 3D structure prediction. BMC Bioinform. 2008, 9, 40. [CrossRef] [PubMed] 48. DeLano, W.L. PyMOL: An open-source molecular graphics tool. Ccp4 Newslett. Protein Crystallogr. 2002, 40, 82–92. 49. Rost, B.; Yachdav, G.; Liu, J. The PredictProtein server. Nucleic Acids Res. 2004, 32, W321–W326. [CrossRef] 50. Wey-Fabrizius, A.R.; Podsiadlowski, L.; Herlyn, H.; Hankeln, T. Platyzoan mitochondrial genomes. Mol. Phylogenet. Evol. 2013, 69, 365–375. [CrossRef] 51. Morales, H.E.; Pavlova, A.; Joseph, L.; Sunnucks, P. Positive and purifying selection in mitochondrial genomes of a bird with mitonuclear discordance. Mol. Ecol. 2015, 24, 2820–2837. [CrossRef] 52. Pavlova, A.; Gan, H.M.; Lee, Y.P.; Austin, C.M.; Gilligan, D.M.; Lintermans, M.; Sunnucks, P. Purifying selection and genetic drift shaped Pleistocene evolution of the mitochondrial genome in an endangered Australian freshwater fish. Heredity 2017, 118, 466–476. [CrossRef] 53. Kang, S.; Kim, J.; Lee, J.; Kim, S.; Min, G.S.; Park, J.K. The complete mitochondrial genome of an ectoparasitic monopisthocotylean fluke Benedenia hoshinai (Monogenea: Platyhelminthes). Mitochondrial DNA 2012, 23, 176–178. [CrossRef][PubMed] 54. Zhang, D.; Li, W.X.; Zou, H.; Wu, S.G.; Li, M.; Jakovli´c,I.; Zhang, J.; Chen, R.; Wang, G.T. Homoplasy or plesiomorphy? Reconstruction of the evolutionary history of mitochondrial gene order rearrangements in the subphylum Neodermata. Int. J. Parasitol. 2019, 49, 819–829. [CrossRef][PubMed] 55. Boore, J.L.; Fuerstenberg, S.I. Beyond linear sequence comparisons: The use of genome-level characters for phylogenetic reconstruction. Philos. Trans. R. Soc. Lond. B Biol. Sci. 2008, 363, 1445–1451. [CrossRef] [PubMed] 56. Zhang, D.; Li, W.X.; Zou, H.; Wu, S.G.; Li, M.; Jakovlic, I.; Zhang, J.; Chen, R.; Wang, G.T. Mitochondrial genomes of two diplectanids (Platyhelminthes: Monogenea) expose paraphyly of the order and extensive tRNA gene rearrangements. Parasit. Vectors 2018, 11, 601. [CrossRef] 57. Zhang, J.; Wu, X.; Xie, M.; Li, A. The complete mitochondrial genome of Pseudochauhanea macrorchis (Monogenea: Chauhaneidae) revealed a highly repetitive region and a gene rearrangement hot spot in Polyopisthocotylea. Mol. Biol. Rep. 2012, 39, 8115–8125. [CrossRef] 58. Le, T.H.; Blair, D.; McManus, D.P. Mitochondrial genomes of parasitic flatworms. Trends Parasitol. 2002, 18, 206–213. [CrossRef] 59. Fumagalli, L.; Taberlet, P.; Favre, L.; Hausser, J. Origin and evolution of homologous repeated sequences in the mitochondrial DNA control region of shrews. Mol. Biol. Evol. 1996, 13, 31–46. [CrossRef] 60. Simkova, A.; Plaisance, L.; Matejusova, I.; Morand, S.; Verneau, O. Phylogenetic relationships of the Dactylogyridae Bychowsky, 1933 (Monogenea: Dactylogyridea): The need for the systematic revision of the Ancyrocephalinae Bychowsky, 1937. Syst. Parasitol. 2003, 54, 1–11. [CrossRef] 61. Plaisance, L.; Littlewood, D.T.J.; Olson, P.D.; Morand, S. Molecular phylogeny of gill monogeneans (Platyhelminthes, Monogenea, Dactylogyridae) and colonization of Indo-West Pacific butterflyfish hosts (Perciformes, Chaetodontidae). Zool. Scr. 2005, 34, 425–436. [CrossRef] 62. Blasco-Costa, I.; Miguez-Lozano, R.; Sarabeev, V.; Balbuena, J.A. Molecular phylogeny of species of Ligophorus (Monogenea: Dactylogyridae) and their affinities within the Dactylogyridae. Parasitol. Int. 2012, 61, 619–627. [CrossRef] 63. Mendoza-Palmero, C.A.; Blasco-Costa, I.; Scholz, T. Molecular phylogeny of Neotropical monogeneans (Platyhelminthes: Monogenea) from catfishes (Siluriformes). Parasit. Vectors 2015, 8, 164. [CrossRef] [PubMed] 64. Kritsky, D.C.; Boeger, W.A. The phylogenetic status of the Ancyrocephalidae Bychowsky, 1937 (Monogenea: Dactylogyroidea). J. Parasitol. 1989, 75, 207–211. [CrossRef][PubMed] Genes 2019, 10, 863 16 of 16

65. Ferguson-Miller, S.; Hiser, C.; Liu, J. Gating and regulation of the cytochrome c oxidase proton pump. Biochim. Biophys. Acta 2012, 1817, 489–494. [CrossRef][PubMed] 66. Dowling, D.K.; Friberg, U.; Lindell, J. Evolutionary implications of non-neutral mitochondrial genetic variation. Trends Ecol. Evol. 2008, 23, 546–554. [CrossRef] 67. The Editors of Encyclopaedia Britannica. Intestinal Gas. Available online: https://www.britannica.com/ science/intestinal-gas (accessed on 20 October 2019). 68. Smith, M.W. The in vitro absorption of water and solutes from the intestine of goldfish, Carassius auratus. J. Physiol. 1964, 175, 38–49. [CrossRef] 69. Ern, R.; Huong, D.T.T.; Cong, N.V.; Bayley, M.; Wang, T. Effect of salinity on oxygen consumption in fishes: A review. J. Fish Biol. 2014, 84, 1210–1220. [CrossRef] 70. Yamaguchi, Y.; Breves, J.P.; Haws, M.C.; Lerner, D.T.; Grau, E.G.; Seale, A.P. Acute salinity tolerance and the control of two prolactins and their receptors in the Nile tilapia (Oreochromis niloticus) and Mozambique tilapia (O. mossambicus): A comparative study. Gen. Comp. Endocrinol. 2018, 257, 168–176. [CrossRef] 71. Potts, W.; Foster, M.; Rudy, P.; Howells, G.P.Sodium and water balance in the cichlid teleost, Tilapia mossambica. J. Exp. Biol. 1967, 47, 461–470.

© 2019 by the authors. Licensee MDPI, Basel, Switzerland. This article is an open access article distributed under the terms and conditions of the Creative Commons Attribution (CC BY) license (http://creativecommons.org/licenses/by/4.0/).