Skip to main content

Comparative mitogenomic analysis of subterranean and surface amphipods (Crustacea, Amphipoda) with special reference to the family Crangonyctidae

Abstract

Mitochondrial genomes play important roles in studying genome evolution, phylogenetic analyses, and species identification. Amphipods (Class Malacostraca, Order Amphipoda) are one of the most ecologically diverse crustacean groups occurring in a diverse array of aquatic and terrestrial environments globally, from freshwater streams and lakes to groundwater aquifers and the deep sea, but we have a limited understanding of how habitat influences the molecular evolution of mitochondrial energy metabolism. Subterranean amphipods likely experience different evolutionary pressures on energy management compared to surface-dwelling taxa that generally encounter higher levels of predation and energy resources and live in more variable environments. In this study, we compared the mitogenomes, including the 13 protein-coding genes involved in the oxidative phosphorylation (OXPHOS) pathway, of surface and subterranean amphipods to uncover potentially different molecular signals of energy metabolism between surface and subterranean environments in this diverse crustacean group. We compared base composition, codon usage, gene order rearrangement, conducted comparative mitogenomic and phylogenomic analyses, and examined evolutionary signals of 35 amphipod mitogenomes representing 13 families, with an emphasis on Crangonyctidae. Mitogenome size, AT content, GC-skew, gene order, uncommon start codons, location of putative control region (CR), length of rrnL and intergenic spacers differed between surface and subterranean amphipods. Among crangonyctid amphipods, the spring-dwelling Crangonyx forbesi exhibited a unique gene order, a long nad5 locus, longer rrnL and rrnS loci, and unconventional start codons. Evidence of directional selection was detected in several protein-encoding genes of the OXPHOS pathway in the mitogenomes of surface amphipods, while a signal of purifying selection was more prominent in subterranean species, which is consistent with the hypothesis that the mitogenome of surface-adapted species has evolved in response to a more energy demanding environment compared to subterranean amphipods. Overall, gene order, locations of non-coding regions, and base-substitution rates points to habitat as an important factor influencing the evolution of amphipod mitogenomes.

Peer Review reports

Introduction

Caves and other subterranean habitats, such as groundwater aquifers and superficial subterranean habitats (SSHs; [1]), represent some of the most challenging environments that exist on Earth. The primary characteristic of all subterranean habitats is the lack of light and associated photosynthesis [1, 2]. Though some subterranean ecosystems are supported by chemoautotrophic production by microbial communities [3, 4], chemoautotrophy rarely provides enough energy to support several trophic levels in most subterranean ecosystems [1, 5]. The primary source of energy input for many cave systems is the organic matter transferred from the surface hydrologically or by animals that frequently enter and exit caves [1, 6], which drive the structure and dynamics of subterranean communities [7,8,9]. Although most subterranean ecosystems are largely thought to be energy-limited [10], food availability can be highly variable both among and within cave systems [11, 12]. Previous studies have shown that many subterranean organisms living in such energy-limited habitats have undergone several physiological and metabolic adaptations to sustain themselves during extended food shortages [13, 14]. Among these troglomorphic traits, low metabolic rate is a key adaptation that occurs in both terrestrial and aquatic fauna of subterranean communities [15, 16].

Mitochondria are the primary sites of energy production in cells, generating ~ 95% of the adenosine triphosphate (ATP) required for everyday activities of life through oxidative phosphorylation [17,18,19]. The mitochondrial genome—mitogenome—encodes 13 essential proteins including two ATP synthases (atp6 and atp8), three cytochrome oxidases (cox1, cox2, and cox3), seven NADPH reductases (nad1, nad2, nad3, nad4, nad4l, nad5, and nad6), and cytochrome b (cytb) subunits. All mitochondrial protein-coding genes (PCGs) play a vital role in the electron transport chain [20,21,22]. Due to the unique characteristics of mitochondria, including maternal inheritance, small genomic size, absence of introns, and their surplus availability in cells, the use of mitochondrial DNA (mtDNA) loci and mitogenomes has a long history in population genetics, phylogenetics, and molecular evolution studies [23,24,25]. Previous studies have demonstrated a close association between mitochondrial loci and energy metabolism [18, 26, 26, 27]. Although considered to largely evolve under purifying selection, there is growing evidence that mitogenomes may undergo episodes of directional selection in response to shifts in physiological or environmental pressures [28, 29] leading to improved metabolic performance under new environmental conditions [26, 30, 31]. For example, previous studies that investigated varying selective pressures acting on mitochondrial PCGs of insects and mammals have revealed significant positive selective constraints at several loci that have comparatively increased energy demands [18, 19, 32]. Similarly, other studies have shown the various adaptive mitochondrial responses of organisms surviving in extreme environments including the deep sea and Tibetan Plateau [29, 32, 33]. However, these adaptations can occur at different metabolic levels, not just mitochondrial metabolism [34, 35]. Thus, variation in mitogenomes of species inhabiting different environments may reflect only a small portion of these adaptive metabolic changes. Despite this limitation, previous studies have detected signals of directional selection in the mitogenomes of organisms dwelling in contrasting habitats with varying energy demands [36,37,38].

Amphipods (Class Malacostraca: Order Amphipoda) are one of the most ecologically diverse crustacean groups including over 10,000 species [39, 40], occurring in a diverse array of aquatic and even terrestrial environments globally, from aphotic groundwater aquifers and hadal depths to freshwater streams and lakes in temperate and tropical forests, among other habitats [41, 42]. Several studies have demonstrated the genetic basis of subterranean adaptation in several taxa, including dytiscid diving beetles [43], cave dwelling-cyprinid fishes [44, 45], anchialine cave shrimps [46], and cave isopods [47]. However, we still have a limited understanding of the mechanisms of subterranean adaptations in amphipods. Although physiological adaptations to challenging environments like cave and groundwater ecosystems have been well-studied in amphipods [13, 16], no studies to date have addressed the selective pressures and the molecular evolution mechanisms of mitochondrial energy metabolism loci in amphipods occupying caves and other subterranean habitats. Subterranean amphipods likely experience different evolutionary pressures on energy management due to lower levels of predation, lower food resources, and more stable environments compared to surface-dwelling taxa that generally experience higher levels of predation and energy resources [48, 49].

In this study, we compared the mitogenomes of surface and subterranean amphipods, including the 13 mitochondrial PCGs involved in the OXPHOS pathway to understand the potential molecular mechanisms of energy metabolism in this diverse crustacean group. Our aims were to test whether the mitochondrial PCGs showed evidence of adaptive evolution in subterranean environments in amphipods. We tested the hypothesis that the mitogenome of surface-adapted amphipods will be imprinted by mitogenomic adaptations to the energy demanding environment with greater signal of directional selection when compared to their subterranean counterparts. We compared base composition, codon usage, gene order rearrangement, conducted comparative mitogenomic and phylogenomic analyses, and examined evolutionary signals using publicly available amphipod mitogenomes. In particular, we focused on the amphipod family Crangonyctidae, a diverse family that comprises species inhabiting a variety of surface and subterranean habitats and for which several mitogenomes have been sequenced and annotated recently [50, 51].

Materials and methods

We generated new mitogenomes recently for the following crangonyctid species: Stygobromus pizzinii, S. tenuis potomacus, Bactrurus brachycaudus, Stygobromus allegheniensis, and Crangonyx forbesi [51].

DNA extraction, library preparation, and sequencing

Whole genomic DNA from five crangonyctid species was isolated using the Qiagen DNA Easy Blood and Tissue kit and libraries were prepared using the Illumina TruSeq DNA Library Prep Kit (Illumina Inc., California). Libraries were then paired-end sequenced (2 × 150 bp) on an Illumina HiSeq 4000 platform at the Vincent J. Coates Genomics Sequencing Laboratory at the University of California, Berkeley. We assessed the quality of the raw reads using FastQC v0.11.5 [52], and the reads were trimmed and filtered using Trimmomatic v0.33 [53]. De-novo assembly was carried out using NOVOPlasty v2.6.3 assembler [54]. We then annotated the protein-coding genes, transfer RNAs (tRNAs), and ribosomal RNAs (rRNAs) for each of the five mitogenomes using the mitochondrial genome annotation web server MITOS [55]. The secondary structures of tRNAs were inferred using MITFI [56], a built-in module in MITOS. The locations of start and stop codons of protein coding genes were confirmed using NCBI ORFfinder [57] and by visual comparison to other published amphipod mitogenomes. The location of the control region was confirmed by the presence of a large intergenic spacer region with a string of thymines found immediately after rrnS and before trnl. We then downloaded from GenBank the annotated mitogenomes of 30 additional amphipod taxa that occupy aquatic habitats, including groundwater and springs, and three isopods that served as outgroups for comparative analyses.

Mitogenome analyses

Nucleotide composition, amino acid frequencies, and codon usage were calculated in PhyloSuite v1.1.15 [58, 59]. The web-based program CREx (http://pacosy.informatik.uni-leipzig.de/crex, [60]) was used to perform pair-wise comparison of the gene orders in the mitogenome to determine rearrangement events. CREx comparisons were based on common intervals, and it considers common rearrangement scenarios including transpositions, reversals, reverse transpositions, and tandem-duplication-random-losses (TDRLs). In addition, phylograms and gene orders were visualized in iTOL (https://itol.embl.de/, [61]) using files exported from PhyloSuite. AT and GC skew of entire mitogenomes and individual loci were calculated in PhyloSuite using the formulae: AT-skew = (A – T)/(A + T) and GC-skew = (G – C)/(G + C). Welch two sample t-tests were performed between the surface and subterranean amphipods for different mitogenomic features, including mitogenome length, AT content, AT and GC skew, and rRNA length using R [62]. Visualization of AT-skew, GC-skew, AT-content, and amino acid frequencies were generated in R.

Phylogenetic inference

The amino acid sequences of 13 PCGs of the five new mitogenomes [51], 30 previously published amphipod mitogenomes, and three isopod mitogenomes were aligned using MAFFT version 7 [63]. A total of 38 sequences with 350 positions in the alignment file were trimmed using Gblocks version 0.91b [64] to yield 255 positions (72%) in 6 selected blocks (parameters used: Supplementary Table S1). The alignment was partitioned by gene and then the best-fit partitioning strategy and evolutionary models for each partition were inferred using PartitionFinder v2.1.1 [65]; Supplementary Table S2). Phylogenetic relationships of the 35 amphipod mitogenomes and three isopod mitogenomes using the concatenated 13 PCG alignment were determined using Bayesian inference in MrBayes v3.2 [66]. The analyses contained two parallel runs with four chains each and were conducted for 5,000,000 generations (sampling every 100 generations). After dropping the first 25% “burn in” trees to ensure stationarity after examination of log-likelihood values for each Bayesian run using Tracer v1.7 [67], the remaining 37,500 sampled trees were used to estimate the consensus tree and the associated Bayesian posterior probabilities. All analyses were conducted within PhyloSuite.

Positive selection and selection pressure analyses of mitochondrial PCGs

We performed base-substitution analyses on entire mitogenomes as well as for each of the 13 PCGs individually to compare surface versus subterranean amphipod taxa. The non-synonymous to synonymous rate ratio (dN/dS or ω) for each PCG was estimated using the free-ratio model using the CodeML program implemented in PAML v4.8a [68]. The ω values were estimated for surface and subterranean species separately and visualized in R for comparison. To estimate the probability of positively selected sites in each PCG across all amphipod species, we implemented site models (M1 and M2, M8a and M8), where ω was allowed to vary among sites [69]. To further investigate if some lineages and sites in particular lineages have undergone positive selection, we conducted maximum likelihood analyses on all PCG using the branch model and branch-site model in EasyCodeML v1.21, a visual tool for analysis of selection using CodeML [70].

To determine if all 13 PCG are free of functional constraints in subterranean lineages, we compared alternative branch selection models on each PCG tree. First, we tested a model (M0) where a single ω was estimated for all branches. This model was compared to a model (M1) with two ratios, a background ω for surface branches and a separate foreground ω for subterranean branches. In addition, we included a two-ratio model where ω was fixed at 1.0 in subterranean branches (M1fixed) to determine if estimates of ω differed from rates of neutral evolution and a model similar to the M1 model but where each subterranean lineage (B. brachycaudus, B. jaraguensis, P. daviui, and the clade containing Stygobromus and Metacrangonyx) was allowed to have a separate ω (M1a). We also examined a saturated model (M2) where each branch had its own ω. Akaike’s information criterion (AIC) was used to compare models.

For both the branch and branch-site models, a likelihood ratio test (LRT) was conducted for each PCG to test whether ω was homogeneous across all branches. In the branch model, the null hypothesis assumes that the average ω values of branches of interest (ωF) is equal to that of other branches (ωB), whereas the alternative hypothesis assumes the opposite ωF ≠ ωB. If the alternative hypothesis is supported and ω > 1, the foreground lineage is under positive selection. The branch-site model allows ω to differ among codon sites in a foreground lineage when compared to background lineages. We implemented the branch-site model to identify sites on specific lineages regulated by positive selection. Selected sites were considered positively selected only if they passed a Bayes Empirical Bayes (BEB) analysis with a posterior probability of > 95%.

We performed selection pressure analyses on the concatenated 13 PCGs dataset aligned using the codon mode as well as on each PCG with the Bayesian topology (see Fig. 4) as the guidance tree using several approaches available from the Datamonkey Webserver [71]. First, we implemented aBSREL (Adaptive Branch-Site Random Effects Likelihood), an improved version of the commonly used “branch-site” models, to test if positive selection has occurred on a proportion of branches [72]. We implemented BUSTED (Branch-site Unrestricted Statistical Test for Episodic Diversification) to test for gene-wide (not site-specific) positive selection by querying if a gene has experienced positive selection in at least one site on at least one branch [73]. Finally, we implemented RELAX [74] to test whether the strength of selection has been relaxed or intensified along a specified set of test branches.

Results and discussion

We compared the mitogenomes for 35 surface and groundwater amphipods, including mitogenomes of one spring-dwelling and six groundwater species in the family Crangonyctidae by Aunins et al. [50] and Benito et al. [51], to determine whether subterranean species show evidence of adaptive evolution in subterranean habitats. Our study examined whether features of mitogenomes (e.g. base composition, codon usage, gene order) differed in relation to dominant habitat (surface vs. subterranean) and inferred the evolutionary forces potentially shaping mitogenome evolution in amphipods, with an emphasis on crangonyctid species.

Mitogenome length and content

Mitogenome sizes ranged from 14,113 to 18,424 bp for all amphipods and 14,661 to 15,469 bp for crangonyctid amphipods (Table 1). Mean mitogenome size of surface amphipods (15,770 ± 1206 bp; mean ± 1 standard deviation) was higher than that of the subterranean amphipods (14,716 ± 297 bp) (phylogenetic paired t-test: t = 0.586, df = 33, p-value = 0.562; Supplementary Figure S1). All crangonyctid amphipod mitogenomes possessed 13 PCGs, two rRNA genes, 22 tRNA genes, a control region, and intergenic spacers of varying number and lengths (Supplementary Figure S2, annotations of the genomes are presented in Supplementary Table S3) like other arthropods [75]. The length of the crangonyctid mitogenomes was similar to lengths reported for other amphipod families including Allocrangonyctidae, Caprellidae, Eulimnogammaridae, Gammaridae, Hadziidae, Lysianassidae, Metacrangonyctidae, Micruropodidae, Pallaseidae, Pontogeneiidae, Talitridae. Variation in mitogenome length within Crangonyctidae appears to be related to length variation in the nad5, rrnL, and rrnS loci, which were all notably longer in the C. forbesi mitogenome.

Table 1 Summary of mitogenomic characteristics, location, and habitat of subterranean and surface amphipods included for comparative mitogenome analyses

Base composition and AT- and GC-skews

Mitogenome AT% in all amphipods ranged from 62.2 to 76.9% (Table 1). Mean AT% of the subterranean amphipods (71.8 ± 3.6%) was higher than that of the surface amphipods (67.6 ± 3.4%) (phylogenetic paired t-test: t = 0.926, df = 33, p-value = 0.361). Mean AT% of all 13 PCG of the subterranean amphipods was significantly higher than that of the surface amphipods (Supplementary Figure S3a). Variation in AT% across crangonyctid amphipod taxa ranged 63.9–69.3%, with a mean of 67.9 ± 1.93% (Table 1). Across loci, AT% ranged from a minimum of 60.0% at the cox1 locus and a maximum of 75.5% at the nad4l locus (Fig. 1A). Variation in AT% across all PCGs combined ranged from 61.9% (B. brachycaudus) to 69.0% (S. indentatus). Genes encoded on the negative strand had a slightly higher AT-content values than those on the positive strand. The nad6 locus showed the greatest variation in AT-content across species. Bactrurus brachycaudus displayed the outlier lower AT% values for most of the PCG (Table 2; Fig. 1A). Similarly, Bactrurus brachycaudus had the lowest AT content (63.9%) among the crangonyctid mitogenomes, while all other mitogenomes had comparatively typical AT content reported for other arthropods [86, 87]. This could indicate that the evolution of the B. brachycaudus mitogenome has occurred under different evolutionary pressures (adaptive or non-adaptive) than other subterranean crangonyctids.

Fig. 1
figure 1

Crangonyctidae mitochondrial nucleotide composition. Box plots showing values of nucleotide composition (A + T percentage) (a), AT-skew (b), and GC-skew (c) across mitogenomes, protein coding genes (PCG), and ribosomal (rRNA) and transfer ribosomal (tRNA) RNA. The same features are shown for each protein-coding gene and pooled by codon position and coding strand. Genes coded on the (-) strand are represented by a “-“ sign and genes coded on the (+) strand are represented by “+” sign at the end of the gene label

Table 2 Comparison of mitogenomic characteristics of 35 amphipods discussed in this study

Mitogenome AT-skew in all amphipods ranged from − 0.062 to -0.037. Mean AT-skew of the surface amphipods (0.001 ± 0.02) was positive and slightly higher than that of the subterranean amphipods (-0.004 ± 0.02) (phylogenetic paired t-test: t = 0.045, df = 33, p-value = 0.965). Mean AT-skew of four PCG (cox1, cox2, nad2, nad3) of surface amphipods was significantly higher than that of the subterranean amphipods, whereas the mean AT-skew of nad4 of the subterranean amphipods was significantly higher than that of the surface amphipods (Supplementary Figure S3b). Among crangonyctid amphipods, mean AT-skew was 0.022 ± 0.02 (range 0.004 to 0.061), with all mitogenomes displaying positive skew. Mitogenome GC-skew ranged from − 0.431 − 0.120. Mean GC-skew of the subterranean amphipods (-0.129 ± 0.15) was negative and higher than that of the surface amphipods (-0.236 ± 0.05) (phylogenetic paired t-test: t = 0.349, df = 33, p-value = 0.729). Mean GC-skew of seven PCG (atp6, atp8, cox1, cox2, cox3, nad2, nad3) of subterranean amphipods was significantly higher than that of the surface amphipods, whereas the mean GC-skew of nad4 of the surface amphipods was significantly higher than that of the subterranean amphipods (Supplementary Figure S3c). Among crangonyctid amphipods, mean GC-skew was − 0.264 ± 0.01 (range − 0.275 to -0.248) with all mitogenomes displaying negative skew (Table 1). Strand asymmetry is commonly observed in mitogenomes [88, 89], however, at times it can hinder phylogenetic reconstruction and yield false phylogenetic artefacts especially when unrelated taxa display inverted skews [90, 91]. Bactrurus brachycaudus exhibited the lowest AT skew among the crangonyctid mitogenomes (0.004), while S. tenuis had the lowest GC skew (− 0.275). Crangonyctid amphipod mitogenomes exhibited positive GC-skew values for genes encoded on the (-) strand and negative GC-skew for genes encoded on the (+) strand (Fig. 1C), whereas all PCGs exhibited negative AT-skew values (Fig. 1B). Except the six loci (nad1, nad4, nad4L, nad5, rrnL, and rrnS) which were encoded on the (-) strand, most PCG had negative GC skews. Such strand bias is typical for most mitochondrial genomes in metazoan [81, 83]. This is consistent with the hypothesis that strand asymmetry is caused by spontaneous deamination of bases in the leading strand during replication [88]. All other mitogenomes had comparatively typical AT and GC skew values like other amphipod species [76, 92]. The only outlier to this pattern was the positive GC skew value of tRNAs encoded on the (+) strand of B. brachycaudus (0.012). In general, crangonyctid mitogenomes exhibited relatively consistent skews.

Rearrangements of mitochondrial genome

Comparisons of crangonyctid mitogenomes revealed at least six conserved gene blocks (Fig. 2B). The gene orders in subterranean species (genera Stygobromus and Bactrurus) are identical except for the transposition of tRNA-G,W. However, a few unique gene order arrangements were observed in the spring-dwelling C. forbesi. The gene order of C. forbesi differs from the four subterranean species in the locations of the conserved gene blocks (tRNA-H-nad4-nad4l and nad6-cytb-tRNA-S2 and tRNA-L1-rrnL and rrnS-tRNA-I and tRNA-Y,Q), seven tRNAs (P,T,M,V,G,C, and W), and two protein-coding loci: nad1 and nad2. Compared to the conserved mitogenome gene orders of other crangonyctid mitogenomes, another unique feature in the rearranged C. forbesi mitogenome was the presence of at least two long (~ 50 and 70 bp) non-coding regions (Supplementary Table S3). The locations of rRNA genes in all crangonyctid mitogenomes are mostly similar compared to the pancrustacean ground pattern except for C. forbesi where the rRNA genes had altered positions (Fig. 2A and B). Rearrangements in the mitogenome is common especially when it involves only tRNA-coding genes [93]. In case of ribosomal RNA genes or PCGs, rearrangements occur much less frequently, and they are commonly referred to as major rearrangements, as they might potentially affect the differential regulation of replication and transcription of mitogenomes [94].

Fig. 2
figure 2

Mitochondrial phylogenomics and gene orders: (a) Bayesian phylogram inferred using amino acid sequences of all mitochondrial PCGs (left) and gene orders (right). Three isopod outgroups are not shown. GenBank accession numbers are included as suffix next to the species names; (b) gene orders of mitochondrial genomes in three genera of crangonyctid amphipods, including Stygobromus, Bactrurus, and Crangonyx. Conserved gene clusters are indicated by different colors and gene rearrangements are highlighted by red border lines

CREx analysis indicated that transpositions and TDRL may have been responsible for the evolution of mitogenomes in crangonyctid amphipods. Two transpositions of tRNA-R,N,S1,E and two steps of TDRL from the ancestral pan-crustacean pattern were needed to generate the gene order observed in Stygobromus species. In addition to the same two transpositions, one TDRL, and a transposition within a second TDRL from the ancestral pattern were required to generate the gene order in Bactrurus. However, four different transpositions (tRNA-N,S1, tRNA-T,P, tRNA-W,C and gene block tRNA-H-nad4-nad4L-tRNA-P,T-nad6-cytb-tRNA-S2) and three steps of TDRL from the ancestral pattern were needed to generate the gene order observed in C. forbesi (Supplementary Figure S4).

Similar to C. forbesi, other surface amphipods including Gmelinoides fasciatus (Micruropodidae) and Onisimus nanseni (Lysianassidae) exhibited a highly rearranged gene order. Other surface amphipods that exhibited a moderate to highly rearranged gene order include Gondogeneia antarctica (Pontogeneiidae), Platorchestia parapacifica and P. japonica (Talitridae), Pallaseopsis kessleri (Pallaseidae), and the two basal amphipods Caprella scaura and C. mutica (Caprellidae) (Fig. 2A). Interestingly, a subterranean amphipod Pseudoniphargus daviui (Allocrangonyctidae) also exhibited a moderate rearranged gene order. The stark contrast between the highly conserved gene order in most subterranean amphipods and the highly volatile gene order in many of the surface amphipods may support the hypothesis that evolution of mitogenomic architecture could be highly discontinuous. A long period of inactivity in gene order and content could have been interspersed by a rearrangement event, this destabilized mitogenome is much more likely to undergo subsequent accelerated rate of mitogenomic rearrangements [95]. Thus, it is appealing to examine mitogenomes of surface amphipod families represented by just a single taxon in our dataset.

Codon usage and amino acid frequencies

In addition to the regular start codons (ATA and ATG) and uncommon start codons (ATT, ATC, TTG, and GTG), surface amphipods, particularly Caprella scaura, possessed one rare start codon CTG, whereas subterranean amphipods possessed three rare start codons including CTG, TTT, and AAT to initiate the mitochondrial PCGs (Supplementary Table S4). Codon usage analysis of the five crangonyctid amphipods mitogenomes identified the existence of all codon types typical for any invertebrate mitogenome. In addition to the regular start codons (ATA and ATG), uncommon start codons (ATT, ATC, TTG, and GTG) were also present to initiate the mitochondrial PCG. Such unusual start codons have been reported previously in other arthropods [96, 97]. A few PCG in the crangonyctid mitogenomes possessed truncated or incomplete stop codons (TA- and T–) that have been described in other crustaceans (Supplementary Table S3). These are presumably completed after a post-transcriptional polyadenylation [98,99,100]. Among the crangonyctid mitogenomes, the most frequently used codons are TTA (Leu2; 5.64–8.49%) and TTT (Phe; 5.94–6.78%). Other frequently used codons include ATT (Ile; 4.92–6.85%) and ATA (Met; 4.13–5.34%) (Supplementary Table S5). These four codons are also among the most abundant in non-crangonyctid amphipods included in this study. This bias towards the AT-rich codons is quite typical for arthropods [101]. Among crangonyctid amphipod mitogenomes, relative synonymous codon usage (RSCU) values, which is the measure of the extent that synonymous codons depart from random usage, showed a high prevalence of A or T nucleotides at third codon positions (Fig. 3). This trend was also observed in other subterranean and surface amphipods. This positive correlation between RSCU and AT content at third codon positions has been reported in mitochondrial genomes of the abalone and oyster [102,103,104].

Fig. 3
figure 3

The relative synonymous codon usage (RSCU) of the mitogenome of all crangonyctid amphipods. The RSCU value are provided on the Y-axis and the codon families are provided on the X-axis

In PCGs, the second copy of leucine (8.86–10.01%) and cysteine (0.95–1.17%) are the most and the least used amino acids, respectively. Amino acid frequency analysis of both surface and subterranean amphipods indicated that five amino acids (leucine, phenylalanine, isoleucine, methonine, and valine) account for more than half of the total amino acid composition and exhibited greater variation among species (Supplementary Figure S5; Supplementary Table S6).

Transfer RNA genes

All 22 tRNA genes were identified in the mitogenomes of crangonyctid amphipods. However, the locations of tRNA genes were highly variable among these mitogenomes, and they also displayed altered positions relative to the pancrustacean ground pattern (Fig. 2; Supplementary Figure S3). The secondary structures of all mitogenome-encoded tRNAs belonging to crangonyctid amphipods were predicted and ranged in length from 50 to 66 bp. Most of the tRNAs displayed the regular clover-leaf structures, however, a few displayed aberrant structures. The tRNA-Ser1 (UCU) lacked the DHU arm in all crangonyctid species. Similarly, the tRNA-Ser2 (UGA) lacked the DHU arm in all crangonyctid species except S. allegheniensis where tRNA-Ser2 (UGA) possessed the DHU arm. The DHU arm was also missing in the tRNA-Cys and tRNA-Arg of B. brachycaudus and tRNA-Arg of C. forbesi. The tRNA-Gln lacked the TψC arm in all crangonyctid species except C. forbesi where tRNA-Gln possessed the TψC arm. In addition to lacking the TψC arm, tRNA-Gln of B. brachycaudus lacked the acceptor stem as well (Supplementary Figure S6). The presence of aberrant structures in tRNAs have been observed in several other crustaceans and invertebrates [79, 105,106,107], which may be the result of replication slippage [108] or selection towards minimization of the mitogenome [109].

Ribosomal RNA genes

The length of rrnL genes in all amphipods ranged 976–1,137 bp and that of rrnS genes ranged 618–1,631 bp. rrnL length of the subterranean amphipods (1,055 ± 26 bp) was higher than that of the surface amphipods (1005 ± 46 bp) (phylogenetic paired t-test: t = 0.921, df = 33, p-value = 0.364). rrnS length of the surface amphipods (738 ± 258 bp) was higher than that of the subterranean amphipods (684 ± 16 bp) (phylogenetic paired t-test: t = -0.558, df = 33, p-value = 0.581). The length of rrnL genes in crangonyctid amphipods ranged 1,034–1,090 bp and that of rrnS genes ranged 671–695 bp. The length of rRNA genes in crangonyctid amphipods was similar to that of other amphipod mitogenomes except C. forbesi, which had long overhangs (~ 50 bp and ~ 25 bp) on the 5’ end of the rrnL and rrnS genes, respectively. AT content ranged 67.8–72.8% in the rrnL genes and 71.5–77.2% in the rrnS genes of crangonyctid species, respectively. GC-skew values for rRNA genes were positive (0.259 to 0.426) and comparable to that of PCGs encoded on the (-) strand (Supplementary Table S7).

Control region and intergenic spacers

In the mitogenome of S. pizzinii the putative control region (CR) was identified as a 1,021 bp sequence between the rrnS gene and the trnl-trnM-trnC-trnY-trnQ-nad2 gene cluster. Similarly, CR was observed in the other crangonyctid mitogenomes, including S. tenuis (556 bp), S. allegheniensis (991 bp), B. brachycaudus (531 bp), S. indentatus (535 bp), and S. tenuis potomacus (773 bp). The CR was similarly located between the rrnS and nad2 genes in some of the other mitogenomes of non-crangonyctid amphipods, including G. duebeni [81], O. nanseni [83], G. antarctica [84], P. daviui [77], and for the pancrustacean ground pattern. However, the adjacent tRNA genes were often different. In G. fasciatus, the CR region was located between the rrnS and nad5 genes [76]. In contrast, a control region 843 bp was observed in C. forbesi which is located between the nad1 and trnM-trnV-nad2 gene cluster and separated by few intergenic spacers was identified as the CR (Supplementary Figure S2; Supplementary Table S3). The only other surface amphipod that had a similar CR location to C. forbesi was P. kessleri with the CR located between nad1 and nad2 genes, although the adjacent tRNA genes were different [76]. Thus, the variable location of the CR in C. forbesi was in concordance with a few surface amphipods, while the subterranean amphipods mostly followed the universal pattern between rrnS and nad2 genes.

The non-coding regions or intergenic spacers identified in the crangonyctid mitogenomes varied in number and length. The number of intergenic spacers ranged from 7 to 17 and their lengths ranged from 1 to 99 bp (mean 13.0 bp ± 18.6). Two crangonyctid mitogenomes (S. allegheniensis and C. forbesi) possessed the largest intergenic spacers (a total of 220 and 249 bp, respectively; Supplementary Table S3). Among the non-crangonyctid amphipods, G. fasciatus and G. antarctica possessed relatively large non-coding intergenic spacers (a total of 3,863 bp and 4,354 bp, respectively; [76, 84].

Phylogenetic inference

The phylogenetic analyses of the 13 concatenated PCG from 35 amphipod species using Bayesian Inference (BI) resulted in a well-supported phylogeny, with the crangonyctid species forming a well-supported monophyletic group (Fig. 4). Within Crangonyctidae, Stygobromus species formed a monophyletic group sister to Bactrurus + Crangonyx; however, few crangonyctid taxa were included in our analysis. A previous study based on the cox1 gene found that Stygobromus was not monophyletic, but several relationships had low support [110]. Likewise, Stygobromus was recovered as polyphyletic in a multilocus concatenated phylogenetic analysis of the Crangonyctidae by Copilaş-Ciocianu et al. [111]. In addition, several well-supported clades were recovered within Crangonyctidae but relationships among these clades had low support. Other past studies have not supported monophyly of the widespread genera (i.e., Crangonyx, Stygobromus, and Synurella) in the family based on either morphological [112] or molecular data [113, 114]. A comprehensive phylogenomic study with robust taxonomic sampling is greatly needed to better elucidate evolutionary relationships and test biogeographic and ecological hypotheses regarding the origin and diversification of this diverse family of subterranean and surface-dwelling amphipods.

Fig. 4
figure 4

Bayesian phylogeny of aligned protein-coding loci (3,607 amino acids) for five new amphipod mitogenomes (Stygobromus allegheniensis, S. pizzinii, S. tenuis potomacus, Bactrurus brachycaudus, and Crangonyx forbesi) in addition to 30 additional amphipod mitogenomes available on Genbank. The three isopods Ligia oceanica, Limnoria quadripunctata, and Eophreatoicus sp.14 FK-2009 are included as an outgroup to root the phylogeny. New mitogenomes generated in this study are highlighted. GenBank accession numbers are included as suffix next to the species names. Values at nodes represent posterior probabilities

Selection in PCGs

Most of the energy required for active movement to escape predation and meet energy demands is supplied by the mitochondrial electron transport chain [99, 100. Mitochondrial genes encode for all of the protein complexes related to oxidative phosphorylation except for succinate dehydrogenase (complex II) [115,116,117]. Because of their importance in oxidative phosphorylation during cellular respiration, it is unsurprising that many studies have shown evidence of purifying (negative) selection in mitochondrial PCG [29, 118, 119]. While we found strong evidence for purifying selection in amphipod mitochondrial PCGs in our selection analyses, we also found signatures of positive selection in a few of the mitochondrial PCGs in the surface amphipods.

Using a free-ratio model (M2; [27]), we calculated the ω values for the 13 PCGs for the terminal branches to estimate the strength of selection between different primary habitats (i.e., subterranean vs. surface). The cox2 locus significantly differed in ω values between the amphipods of the two habitat types (p = 0.020), with higher ω values for the surface amphipods. Similarly, cox1 and cox3 genes also exhibited a similar trend (p = 0.095 and p = 0.057, respectively) (Fig. 5). This could be because the rate at which slightly deleterious mutations (ω) responsible for the mitochondrial gene evolution accumulates comparatively faster in cox gene family of the surface lineages. However, this result is quite contradictory to previous studies showing higher functional constraint and conserved pattern in the genes coding for cox proteins than in other mitochondrial genes [119, 120].

Fig. 5
figure 5

Ratio of non-synonymous to synonymous substitutions (ω) in the 13 PCGs of subterranean (coral color) and surface (cyan color) amphipods based on the free-ratio model. Boxes include 50% of values; ω is not significantly different between subterranean and surface amphipods for any gene except cox2* (P value = 0.02)

To test if the 13 PCGs in subterranean lineages evolve at different relative rates compared to surface lineages, we compared a series of ML branch-based selection models (Table 3). For all PCG loci except atp8, nad3, and nad4l the saturated model (M2) where each branch had its own ω was favored. For atp8, nad3, and nad4l the best models were the M0 (single ω for all branches) and M1 (two ω model with one for surface and one for subterranean linages). In addition, the M1a model (six ω model with one for surface and one for each subterranean linage) was included in the set of best models for the atp8 and nad4l loci. To further test if specific branches have undergone variable selective pressures, especially those amphipod branches adapted to surface habitats, we employed the two-ratio branch model. When the ω values for each PCG were compared between each amphipod terminal branch and the other 34 amphipod taxa, several loci in surface amphipod mitogenomes were found to be undergoing positive selection (ω1 > ω0; Fig. 6; Supplementary Table S8). This suggests that many surface amphipods have experienced directional selection in their mitochondrial loci perhaps due to high energy demands and was in accordance to previous studies in other arthropods [19, 32, 121, 122]. In contrast, several loci in subterranean amphipod mitogenomes have undergone purifying selection (ω1 < ω0). Surprisingly, a few loci in subterranean taxa displayed positive selection (ω1 > ω0; Fig. 6; Supplementary Table S8). To test if individual gene codons were subject to positive selection, we implemented two pairs of site models (M1a vs. M2a and M8a vs. M8). The M8 model identified one positively selected site on the atp8 locus (37 N; p = 0) and one positively selected site on the nad5 locus (482 Q; p = 0). Similarly, The M2a model identified two positively selected sites (37 N & 31 S; p = 0.0194) on the atp8 locus (Table 4).

Table 3 AIC scores and ω estimates for various branch-based selection models for the 13 PCGs (one ω for all branches (M0), two-ratio model with background (surface) ω and single ω for subterranean branches (M1), two-ratio model with background (surface) ω and single ω for subterranean branches fixed at neutral evolution (ω = 1) (M1fixed), six-ratio model with background (surface) ω and a single ω for each subterranean lineage, B. brachycaudus (C1), Stygobromus clade (C2), B. jaraguensis (C3), Metacrangonyx clade (C4), P. daviui (C5) (M1a), and ω for each branch (M2)). The best-fit model(s) for each PCG is highlighted in red color
Table 4 Evidence of positive selection on the mitochondrial PCGs of subterranean and surface-dwelling amphipods based on site model
Fig. 6
figure 6

Results of selective pressure analysis of mitochondrial PCGs with LRT P-value < 0.05 in subterranean and surface-dwelling lineages of amphipods based on branch 2 vs. 0 model. Different colored shapes represent different mitochondrial genes. Squares represent purifying selection and circles represent positive selection. Surface amphipod branches are colored blue and subterranean amphipod branches are colored red

Similar to flying grasshoppers that have evolved to adapt to increased energy demands to maintain the flight capacity [32], the mitochondrial loci of surface amphipods may have evolved mechanisms to meet increased energy demands due to predation, dispersal, and other factors. Although surface amphipods appear to be evolving under selective pressures different from those of the subterranean taxa and their mitochondrial loci have accumulated more nonsynonymous than synonymous mutations compared to subterranean taxa, the branch model tests did not clearly support positive selection on these branches, and we cannot rule out the influence of relaxed selection. Previous studies have demonstrated that positive selection will act on only a few sites for a short period of evolutionary time, and a signal of positive selection often is overwhelmed by continuous negative selection that sweeps across most sites in a gene sequence [123].

In contrast to branch models where ω varies only among branches, branch-site models allow selection to vary both among amino acid sites and lineages. Thus, branch-site models are considered quite useful in distinguishing positive selection from relaxed or purifying selection [123]. Using the more stringent branch-site model, we detected positive selection in 14 branches and 12 loci with a total of 308 amino acid sites under positive selection. Among them, 80 amino acid sites in seven loci (atp6, atp8, cox3, nad2, nad3, nad4, and nad5) were identified on the subterranean terminal branches, whereas 228 amino acid sites in 10 loci (atp6, atp8, cox1, cox2, cytb, nad1, nad2, nad3, nad5, and nad6) were identified on the surface terminal branches. Nearly three times as many positively selected amino acid sites were detected on surface branches compared to subterranean branches. Most of the positively selected loci on surface branches were found in C. forbesi with 114 sites (Fig. 7; Supplementary Table S9). In total, eight positive selected loci (atp6, atp8, cox1, cox2, cytb, nad1, nad4, and nad5) were identified by the branch-site model and by at least one other model on the surface branches, whereas only four positive selected genes (atp6, atp8, cox3, and nad5) were identified on the subterranean branches.

Fig. 7
figure 7

Evidence of positive selection on the mitochondrial PCGs (LRT P<0.05) and positively selected site (BEB: P≥95%) in subterranean and surface-dwelling lineages of amphipods based on branch-site models. Different colored circles represent different mitochondrial loci. The number within each circle represents the number of positive selection sites detected for the locus. Surface amphipod branches are colored blue and subterranean amphipod branches are colored red

The identification of many positively selected amino acid sites suggests that episodic positive selection has acted on mitochondrial PCGs of surface amphipods. In addition, we also identified a few positively selected sites on subterranean branches primarily in B. brachycaudus with 39 sites and P. daviui with 25 sites (Supplementary Table S9). Bactrurus brachycaudus is usually associated with springs and caves [124], while P. daviui is associated with groundwater wells [77].

Direction and magnitude of selection pressures

Given the crucial role played by the mitochondrial genome in metabolic energy production [125], we hypothesized that the mitogenome of surface amphipods may show evidence of adaptation (directional selection) to life in surface habitats where energy demand is higher relative to subterranean habitats. We found support for directional selection in surface lineages based on three different selection analyses (RELAX, aBSREL, and BUSTED). In summary, all tests confirmed the existence of a moderate signal of positive or diversifying selection, as well as signal for significant relaxed purifying selection in the mitogenome of surface amphipods. This supports a previous study by Carlini and Fong [126] who reported evidence for relaxation of functional evolutionary constraints (positive or diversifying selection) in the transcriptome of a subterranean amphipod Gammarus minus.

We implemented aBSREL on the concatenated 13 PCG dataset comprising all 35 species as test branches and detected episodic diversifying selection in seven species: P. daviui (p = 0), O. nanseni (p = 0.0008), G. fasciatus (p = 0.0298), G. fossarum (p = 0.045), B. jaraguensis (p = 0.0016), C. forbesi (p = 0), and B. brachycaudus (p = 0.0001). We then used aBSREL to conduct independent tests for the crangonyctid species as the test branch and the remaining species as reference branches. We detected evidence of episodic diversifying selection in C. forbesi (p = 0) and B. brachycaudus (p = 0.0001) (Table 5). Using BUSTED, which provides a gene-wide test for positive selection, we detected evidence of episodic diversifying selection in three of the surface species: C. forbesi (p = 0.011), G. fasciatus (p = 0.033), G. antarctica (p = 0.009), whereas evidence of gene-wide episodic diversifying selection was found in just one of the subterranean species, P. daviui (p = 0.020) (Table 5). Using RELAX, which tests whether the strength of selection has been relaxed or intensified along a specified set of test branches, we detected selection evidence of relaxed selection in C. forbesi (p = 0) and other surface species, including O. nanseni, G. fasciatus, G. fossarum, G. antarctica, and P. kessleri. Contrastingly, evidence of intensification of selection was detected in subterranean species, including S. tenuis (p = 0), S. allegheniensis (p = 0.0025), S. indentatus (p = 0), and S. pizzinii (p = 0). Surprisingly, a few of the surface species including C. mutica (p = 0.015), E. cyaneus (p = 0), and P. japonica (p = 0) exhibited intensification of selection and subterranean species including P. daviui (p = 0) and M. dominicanus (p = 0.015) exhibited relaxation of selection (Table 5).

Table 5 Selection signals in the mitogenomes of amphipods inferred using aBSREL, BUSTED, and RELAX algorithms. The dataset comprising all 13 concatenated protein-coding genes with 3,607 amino acid sites in the alignment. K column: a statistically significant K > 1 indicates that selection strength has been intensified, and K < 1 indicates that selection strength has been relaxed. LR is likelihood ratio and D indicates the direction of selection pressure change: intensified (I) or relaxed (R), where * highlights a statistically significant (p < 0.05) result. Mitogenomes with significant LRT P -value < 0.05 are highlighted in red color

In addition to the concatenated 13 PCG dataset, we also conducted selection analyses for each PCG to determine which genes might be evolving under unique selection pressures. We found evidence of directional selection in atp8 of C. forbesi (p = 0.026) and nad3 of S. pizzinii (p = 0.041) using aBSREL and cox3 of B. brachycaudus (p = 0.029) using BUSTED. Atp8 of the surface amphipod C. forbesi exhibited strong evidence of directional selection, which was quite surprising as atp8 is a small locus sometimes missing from metazoan mitogenomes and normally evolves under highly relaxed selection pressures [127]. RELAX analyses uncovered five loci (cox1, cox3, cytb, nad1, and nad3) that exhibited relaxed selection and one gene (atp6) that exhibited intensification of selection in C. forbesi. Similarly, three loci (cox3, nad5, and nad6) in B. brachycaudus showed evidence of relaxed selection. Several loci in other subterranean species, including S. tenuis, S. allegheniensis, and S. pizzinii, exhibited varying levels of intensification of selection, whereas none exhibited relaxed selection (Table 6). Some of these outliers were expected, as nad5 and nad6 are known to evolve faster among the mitochondrial loci [128]. Also, evidence for relaxation of functional evolutionary constraints (positive or diversifying selection) has been reported in the nad family of subterranean Gammarus species adapted to the subterranean environment [126]. Although this may explain outliers in the subterranean B. brachycaudus mitogenome, it remains unclear why cox3 exhibited signatures of relaxed selection. This gene is generally one of the most conserved mitochondrial loci in animals [92, 129, 130], and high levels of purifying selection has been observed in the cox family in other amphipod species [29]. In C. forbesi, atp6 showed signatures of positive selection, which contrasted most other PCGs in its mitogenome that exhibited relaxed selection. Overall, in accordance with the results obtained using the concatenated dataset, individual mitochondrial loci of subterranean amphipods mostly exhibited varying levels of purifying selection, whereas surface amphipods predominantly exhibited more relaxed selection.

Table 6 Selection signals in the mitochondrial PCGs of crangonyctid amphipods sequenced in this study inferred using aBSREL, BUSTED, and RELAX algorithms. K column: a statistically significant K > 1 indicates that selection strength has been intensified, and K < 1 indicates that selection strength has been relaxed. LR is likelihood ratio and D indicates the direction of selection pressure change: intensified (I) or relaxed (R), where * highlights a statistically significant (p < 0.05) result. PCGs with significant LRT P -value < 0.05 are highlighted in red color

To provide further evidence of positive selection, we implemented the RELAX, aBSREL, and BUSTED algorithms on the branch, branch-site, and site models. Eight loci (atp8, cox1, cox2, cytb, nad1, nad4, nad5, and nad6) all involved in the OXPHOS pathway were under positive selection in surface branches by at least two methods. The loci nad1, nad4, nad5, and nad6 encode the subunits of NADH dehydrogenase, also called Complex I, that initiates the oxidative phosphorylation process. Complex I is the largest and most complicated proton pump of the respiratory chain and is involved in electron transfer from NADH to ubiquinone to supply the proton motive force used for ATP synthesis [131], Complex I plays a key role in cellular energy metabolism by pumping gradient of protons across the mitochondrial membrane producing more than one-third of mitochondrial energy [132]. Genes cox1 & cox2 encode the catalytic core of Cytochrome c oxidase also called Complex IV. Complex IV is directly involved in electron transfer and proton translocation [133]. Gene atp8 encodes a part of ATP synthase, also called Complex V, and plays a major role in the final assembly of ATPase [133]. In summary, our selection analyses revealed signals of positive selection in several mitochondrial genes of surface amphipods, which may be associated with increased energy demands in surface environments. In contrast, subterranean amphipods showed signatures of purifying selection, which may be related to maintaining efficient energy metabolism in subterranean habitats.

Conclusion

In this study, we compared mitogenome features including AT/GC-skew, codon usage, gene order, phylogenetic relationships, and selection pressures acting upon amphipods inhabiting surface and subterranean habitats. We described a novel mitochondrial gene order for C. forbesi. We identified a signal of directional selection in the protein-encoding genes of the OXPHOS pathway in the mitogenomes of surface amphipods and a signal of purifying selection in subterranean species, which is consistent with the hypothesis that the mitogenome of surface-adapted amphipods has evolved in response to a more energy demanding environment compared to subterranean species. Our comparative analyses of gene order, locations of non-coding regions, and base-substitution rates points to habitat as an important factor influencing the evolution of amphipod mitogenomes. However, the generation and study of mitogenomes from additional amphipod taxa, including other crangonyctid species, are needed to better elucidate phylogenetic relationships and the evolution of mitogenomes of amphipods, as mitogenomes are available for just a tiny fraction of the more than 10,000 described amphipods. In addition, more evidence is needed to further validate our inferences, such as studying the effects of amino acid changes on three-dimensional protein structure and function. Nevertheless, our study provides a necessary foundation for the study of mitogenome evolution in amphipods and other crustaceans.

Availability of data and materials

The data that support the findings of this study are openly available in NCBI GenBank at https://www.ncbi.nlm.nih.gov/nuccore/MN175619.1 (Bactrurus brachycaudus), https://www.ncbi.nlm.nih.gov/nuccore/MN175620.1 (Stygobromus pizzinii), https://www.ncbi.nlm.nih.gov/nuccore/MN175621.1 (Stygobromus tenuis potomacus), https://www.ncbi.nlm.nih.gov/nuccore/MN175622.1 (Stygobromus allegheniensis), and https://www.ncbi.nlm.nih.gov/nuccore/MN175623.1 (Crangonyx forbesi).

The sample voucher numbers, related meta-data, and raw sequencing data are openly available in NCBI SRA RunSelector at https://www.ncbi.nlm.nih.gov/Traces/study/?acc=PRJNA657640.

References

  1. Culver DC, Pipan T. The biology of caves and other subterranean habitats. Oxford: Oxford University Press; 2009.

    Google Scholar 

  2. Soares D, Niemiller ML. Extreme adaptation in caves. Anat Rec. 2020;303(1):15–23.

    Article  Google Scholar 

  3. Dröse S, Krack S, Sokolova L, Zwicker K, Barth HD, Morgner N, Brandt U. Functional dissection of the proton pumping modules of mitochondrial complex I. PLoS Biol. 2011;9(8):e1001128.

    Article  PubMed  PubMed Central  Google Scholar 

  4. Pons J, Bauzà-Ribot MM, Jaume D, Juan C. Next-generation sequencing, phylogenetic signal and comparative mitogenomic analyses in Metacrangonyctidae (Amphipoda: Crustacea). BMC Genomics. 2014;15(1):1–16.

    Article  Google Scholar 

  5. Porter ML, Engel AS, Kane TC, Kinkle BK. Productivity-diversity relationships from chemolithoautotrophically based sulfidic karst systems. Int J Speleol. 2009;38(1):4.

    Article  Google Scholar 

  6. Simon KS, Benfield EF. Leaf and wood breakdown in cave streams. J North Am Benthological Soc. 2001;20(4):550–63.

    Article  Google Scholar 

  7. Gissi C, Iannelli F, Pesole G. Evolution of the mitochondrial genome of Metazoa as exemplified by comparison of congeneric species. Heredity. 2008;101(4):301–20.

    Article  CAS  PubMed  Google Scholar 

  8. Horton T, Thurston MH, Vlierboom R, Gutteridge Z, Pebody CA, Gates AR, Bett BJ. Are abyssal scavenging amphipod assemblages linked to climate cycles? Prog Oceanogr. 2020;184:102318.

    Article  Google Scholar 

  9. Huntsman BM, Venarsky MP, Benstead JP, Huryn AD. Effects of organic matter availability on the life history and production of a top vertebrate predator (Plethodontidae: Gyrinophilus palleucus) in two cave streams. Freshw Biol. 2011;56(9):1746–60.

    Article  Google Scholar 

  10. Venarsky MP, Huntsman BM, Huryn AD, Benstead JP, Kuhajda BR. Quantitative food web analysis supports the energy-limitation hypothesis in cave stream ecosystems. Oecologia. 2014;176(3):859–69.

    Article  PubMed  Google Scholar 

  11. Crozier RH, Crozier YC. The mitochondrial genome of the honeybee Apis mellifera: complete sequence and genome organization. Genetics. 1993;133(1):97–117.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  12. Ito A, Aoki MN, Yokobori SI, Wada H. The complete mitochondrial genome of Caprella scaura (Crustacea, Amphipoda, Caprellidea), with emphasis on the unique gene order pattern and duplicated control region. Mitochondrial DNA. 2010;21(5):183–90.

    Article  CAS  PubMed  Google Scholar 

  13. Hassanin A, Ropiquet A, Couloux A, Cruaud C. Evolution of the mitochondrial genome in mammals living at high altitude: new insights from a study of the tribe Caprini (Bovidae, Antilopinae). J Mol Evol. 2009;68(4):293–310.

    Article  CAS  PubMed  Google Scholar 

  14. Hyde J, Cooper SJ, Munguia P, Humphreys WF, Austin AD. The first complete mitochondrial genomes of subterranean dytiscid diving beetles (Limbodessus and Paroster) from calcrete aquifers of Western Australia. Aust J Zool. 2018;65(5):283–91.

    Article  Google Scholar 

  15. Bishop R, Humphreys WF, Longley G. Epigean and Hypogean Palaemonetes sp. (Decapoda, Palaemonidae) from Edwards Aquifer: an examination of trophic structure and metabolism. Subterr Biol. 2014;14:79–102.

    Article  Google Scholar 

  16. Murrell B, Weaver S, Smith MD, Wertheim JO, Murrell S, Aylward A, Kosakovsky Pond SL. Gene-wide identification of episodic selection. Mol Biol Evol. 2015;32(5):1365–71.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  17. Da Fonseca RR, Johnson WE, O’Brien SJ, Ramos MJ, Antunes A. The adaptive evolution of the mammalian mitochondrial genome. BMC Genomics. 2008;9(1):1–22.

    Article  Google Scholar 

  18. Shen YY, Shi P, Sun YB, Zhang YP. Relaxation of selective constraints on avian mitochondrial DNA following the degeneration of flight ability. Genome Res. 2009;19(10):1760–5.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  19. Yang Y, Xu S, Xu J, Guo Y, Yang G. Adaptive evolution of mitochondrial energy metabolism genes associated with increased energy demand in flying insects. PLoS One. 2014;9(6):e99120.

    Article  PubMed  PubMed Central  Google Scholar 

  20. Boore JL. Animal mitochondrial genomes. Nucleic Acids Res. 1999;27(8):1767–80.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  21. Burger G, Gray MW, Lang BF. Mitochondrial genomes: anything goes. Trends Genet. 2003;19(12):709–16.

    Article  CAS  PubMed  Google Scholar 

  22. Xu S, Luosang J, Hua S, He J, Ciren A, Wang W, Zheng X. High altitude adaptation and phylogenetic analysis of tibetan horse based on the mitochondrial genome. J Genet Genomics. 2007;34(8):720–9.

    Article  CAS  PubMed  Google Scholar 

  23. Ballard JWO, Pichaud N. Mitochondrial DNA: more than an evolutionary bystander. Funct Ecol. 2014;28(1):218–31.

    Article  Google Scholar 

  24. Bourguignon T, Tang Q, Ho SY, Juna F, Wang Z, Arab DA, Lo N. Transoceanic dispersal and plate tectonics shaped global cockroach distributions: evidence from mitochondrial phylogenomics. Mol Biol Evol. 2018;35(4):970–83.

    Article  CAS  PubMed  Google Scholar 

  25. Zou H, Jakovlić I, Zhang D, Chen R, Mahboob S, Al-Ghanim KA, Wang GT. The complete mitochondrial genome of Cymothoa indica has a highly rearranged gene order and clusters at the very base of the Isopoda clade. PLoS One. 2018;13(9):e0203089.

    Article  PubMed  PubMed Central  Google Scholar 

  26. Culver DC, Kane TC, Fong DW. Adaptation and natural selection in caves: the evolution of Gammarus minus. Cambridge: Harvard University Press; 1995. p. 223.

    Book  Google Scholar 

  27. Shen YY, Liang L, Zhu ZH, Zhou WP, Irwin DM, Zhang YP. Adaptive evolution of energy metabolism genes and the origin of flight in bats. Proc Natl Acad Sci. 2010;107(19):8666–71.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  28. Botero-Castro F, Tilak MK, Justy F, Catzeflis F, Delsuc F, Douzery EJ. In cold blood: compositional bias and positive selection drive the high evolutionary rate of vampire bats mitochondrial genomes. Genome Biol Evol. 2018;10(9):2218–39.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  29. Sun S, Wu Y, Ge X, Jakovlić I, Zhu J, Mahboob S, Fu H. Disentangling the interplay of positive and negative selection forces that shaped mitochondrial genomes of Gammarus pisinnus andGammarus lacustris. Royal Soc Open Sci. 2020;7(1):190669.

    Article  CAS  Google Scholar 

  30. Gao F, Chen C, Arab DA, Du Z, He Y, Ho SY. EasyCodeML: a visual tool for analysis of selection using CodeML. Ecol Evol. 2019;9(7):3891–8.

    Article  PubMed  PubMed Central  Google Scholar 

  31. Welch AJ, Bedoya-Reina OC, Carretero-Paulet L, Miller W, Rode KD, Lindqvist C. Polar bears exhibit genome-wide signatures of bioenergetic adaptation to life in the arctic environment. Genome Biol Evol. 2014;6(2):433–50.

    Article  PubMed  PubMed Central  Google Scholar 

  32. Li M, Jin L, Ma J, Tian S, Li R, Li X. Detecting mitochondrial signatures of selection in wild tibetan pigs and domesticated pigs. Mitochondrial DNA Part A. 2016;27(1):747–52.

    Article  CAS  Google Scholar 

  33. Mitterboeck TF, Liu S, Adamowicz SJ, Fu J, Zhang R, Song W, Zhou X. Positive and relaxed selection associated with flight evolution and loss in insect transcriptomes. GigaScience. 2017;6(10):gix073.

    Article  Google Scholar 

  34. Beall CM. Two routes to functional adaptation: Tibetan and Andean high-altitude natives. Proc Natl Acad Sci. 2007;104(suppl 1):8655–60.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  35. Scott GR, Schulte PM, Egginton S, Scott AL, Richards JG, Milsom WK. Molecular evolution of cytochrome c oxidase underlies high-altitude adaptation in the bar-headed goose. Mol Biol Evol. 2011;28(1):351–63.

    Article  CAS  PubMed  Google Scholar 

  36. Oliveira DC, Raychoudhury R, Lavrov DV, Werren JH. Rapidly evolving mitochondrial genome and directional selection in mitochondrial genes in the parasitic wasp Nasonia (Hymenoptera: Pteromalidae). Mol Biol Evol. 2008;25(10):2167–80.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  37. Yang Z. PAML 4: phylogenetic analysis by maximum likelihood. Mol Biol Evol. 2007;24(8):1586–91.

    Article  CAS  PubMed  Google Scholar 

  38. Letunic I, Bork P. Interactive tree of life (iTOL) v5: an online tool for phylogenetic tree display and annotation. Nucleic Acids Res. 2021;49(W1):W293–6.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  39. Arfianti T, Wilson S, Costello MJ. Progress in the discovery of amphipod crustaceans. PeerJ. 2018;6:e5187.

    Article  PubMed  PubMed Central  Google Scholar 

  40. Hervant F, Mathieu J, Barré H, Simon K, Pinon C. Comparative study on the behavioral, ventilatory, and respiratory responses of hypogean and epigean crustaceans to long-term starvation and subsequent feeding. Comp Biochem Physiol A. 1997;118(4):1277–83.

    Article  Google Scholar 

  41. Barnard JL, Karaman GS. The families and genera of marine gammaridean Amphipoda (except marine gammaroids). Records Aust Museum Suppl. 1991;13(1/2):1–866.

    Google Scholar 

  42. Bousfield EL. An updated phyletic classification and palaeohistory of the Amphipoda. Crustacean Issues. 1983;1:257–77.

    Google Scholar 

  43. Huppop K. How do cave animals cope with the food scarcity in caves? In: Wilkens H, Culver DC, Humphreys WF, editors. Ecosystems of the world. Subterranean ecosystems. Vol. 30. Amsterdam: Elsevier; 2000. p. 159–188.

  44. Dotson EM, Beard C. Sequence and organization of the mitochondrial genome of the Chagas disease vector, Triatoma dimidiata. Insect Mol Biol. 2001;10(3):205–15.

    Article  CAS  PubMed  Google Scholar 

  45. Wu X, Wang L, Chen S, Zan R, Xiao H, Zhang YP. The complete mitochondrial genomes of two species from Sinocyclocheilus (Cypriniformes: Cyprinidae) and a phylogenetic analysis within Cyprininae. Mol Biol Rep. 2010;37(5):2163–71.

    Article  CAS  PubMed  Google Scholar 

  46. Graening GO, Brown AV. Ecosystem dynamics and pollution effects in an Ozark cave stream 1. JAWRA J Am Water Resour Assoc. 2003;39(6):1497–507.

    Article  CAS  Google Scholar 

  47. Poulson TL, White WB. The Cave Environment: limestone caves provide unique natural laboratories for studying biological and geological processes. Science. 1969;165(3897):971–81.

    Article  CAS  PubMed  Google Scholar 

  48. Protas ME, Trontelj P, Patel NH. Genetic basis of eye and pigment loss in the cave crustacean, Asellus aquaticus. Proc Natl Acad Scie. 2011;108(14):5702–7.

    Article  CAS  Google Scholar 

  49. Qiu Q, Zhang G, Ma T, Qian W, Wang J, Ye Z, Liu J. The yak genome and adaptation to life at high altitude. Nat Genet. 2012;44(8):946–9.

    Article  CAS  PubMed  Google Scholar 

  50. Aunins AW, Nelms DL, Hobson CS, King TL. Comparative mitogenomic analyses of three north American stygobiont amphipods of the genus Stygobromus (Crustacea: Amphipoda). Mitochondrial DNA Part B. 2016;1(1):560–3.

    Article  PubMed  PubMed Central  Google Scholar 

  51. Benito JB, Porter ML, Niemiller ML. The mitochondrial genomes of five spring and groundwater amphipods of the family Crangonyctidae (Crustacea: Amphipoda) from eastern North America. Mitochondrial DNA B. 2021;6(6):1662–7.

    Article  Google Scholar 

  52. Andrews S. FastQC: a quality control tool for high throughput sequence data. 2010. Available online at: http://www.bioinformatics.babraham.ac.uk/projects/fastqc.

  53. Bolger AM, Lohse M, Usadel B. Trimmomatic: a flexible trimmer for Illumina sequence data. Bioinf (Oxford England). 2014;30(15):2114–20.

    CAS  Google Scholar 

  54. Das J. The role of mitochondrial respiration in physiological and evolutionary adaptation. BioEssays. 2006;28(9):890–901.

    Article  CAS  PubMed  Google Scholar 

  55. Bernt M, Donath A, Jühling F, Externbrink F, Florentz C, Fritzsch G, Pütz M, Middendorf M, Stadler PF. MITOS: improved de novo metazoan mitochondrial genome annotation. Mol Phylogenet Evol. 2013;69:313–9.

    Article  PubMed  Google Scholar 

  56. Juan C, Guzik MT, Jaume D, Cooper SJ. Evolution in caves: Darwin’s ‘wrecks of ancient life’in the molecular era. Mol Ecol. 2010;19(18):3865–80.

    Article  PubMed  Google Scholar 

  57. Wheeler DL, Church DM, Federhen S, Lash AE, Madden TL, Pontius JU, Schuler GD, Schriml LM, Sequeira E, Tatusova TA, Wagner L. Database resources of the National Center for Biotechnology. Nucleic Acids Res. 2003;31(1):28–33.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  58. Zhang D, Li WX, Zou H, Wu SG, Li M, Jakovlić I, Wang GT. Mitochondrial genomes of two diplectanids (Platyhelminthes: Monogenea) expose paraphyly of the order Dactylogyridea and extensive tRNA gene rearrangements. Parasites Vectors. 2018;11(1):1–13.

    Article  Google Scholar 

  59. Zhang D, Gao F, Jakovlić I, Zou H, Zhang J, Li WX, Wang GT. PhyloSuite: an integrated and scalable desktop platform for streamlined molecular sequence data management and evolutionary phylogenetics studies. Mol Ecol Resour. 2020;20(1):348–55.

    Article  PubMed  Google Scholar 

  60. Bernt M, Merkle D, Ramsch K, Fritzsch G, Perseke M, Bernhard D, Middendorf M. CREx: inferring genomic rearrangements based on common intervals. Bioinformatics. 2007;23(21):2957–8.

    Article  CAS  PubMed  Google Scholar 

  61. Lessinger AC, Junqueira M, Lemos AC, Kemper TA, Silva ELD, Vettore FR, Arruda ALP, Azeredo-Espin LAM. The mitochondrial genome of the primary screwworm fly Cochliomyia hominivorax (Diptera: Calliphoridae). Insect Mol Biol. 2000;9(5):521–9.

    Article  CAS  PubMed  Google Scholar 

  62. Qu Y, Zhao H, Han N, Zhou G, Song G, Gao B, Lei F. Ground tit genome reveals avian adaptation to living at high altitudes in the tibetan plateau. Nat Commun. 2013;4(1):1–9.

    Article  Google Scholar 

  63. Jühling F, Pütz J, Bernt M, Donath A, Middendorf M, Florentz C, Stadler PF. Improved systematic tRNA gene annotation allows new insights into the evolution of mitochondrial tRNA structures and into the mechanisms of mitochondrial genome rearrangements. Nucleic Acids Res. 2012;40(7):2833–45.

    Article  PubMed  Google Scholar 

  64. Castresana J. Selection of conserved blocks from multiple alignments for their use in phylogenetic analysis. Mol Biol Evol. 2000;17(4):540–52.

    Article  CAS  PubMed  Google Scholar 

  65. Krebes L, Bastrop R. The mitogenome of Gammarus duebeni (Crustacea Amphipoda): a new gene order and non-neutral sequence evolution of tandem repeats in the control region. Comp Biochem Physiol D Genomics Proteomics. 2012;7(2):201–11.

    Article  CAS  PubMed  Google Scholar 

  66. Ronquist F, Teslenko M, van der Mark P, Ayres DL, Darling A, Höhna S, Larget B, Liu L, Suchard MA, Huelsenbeck JP. MrBayes 3.2: efficient bayesian phylogenetic inference and model choice across a large model space. Syst Biol. 2012;61:539–42.

    Article  PubMed  PubMed Central  Google Scholar 

  67. Rambaut A, Drummond AJ, Xie D, Baele G, Suchard MA. Posterior summarization in bayesian phylogenetics using Tracer 1.7. Syst Biol. 2018;67(5):901.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  68. Xu B, Yang Z. PAMLX: a graphical user interface for PAML. Mol Biol Evol. 2013;30(12):2723–4.

    Article  CAS  PubMed  Google Scholar 

  69. Yang L, Wang Y, Zhang Z, He S. Comprehensive transcriptome analysis reveals accelerated genic evolution in a Tibet fish, Gymnodiptychus pachycheilus. Genome Biol Evol. 2015;7(1):251–61.

    Article  Google Scholar 

  70. Engel AS, Porter ML, Stern LA, Quinlan S, Bennett PC. Bacterial diversity and ecosystem function of filamentous microbial mats from aphotic (cave) sulfidic springs dominated by chemolithoautotrophic Epsilonproteobacteria. FEMS Microbiol Ecol. 2004;51(1):31–53.

    Article  CAS  PubMed  Google Scholar 

  71. Weaver S, Shank SD, Spielman SJ, Li M, Muse SV, Pond K, S. L. Datamonkey 2.0: a modern web application for characterizing selective and other evolutionary processes. Mol Biol Evol. 2018;35(3):773–7.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  72. Smith MD, Wertheim JO, Weaver S, Murrell B, Scheffler K, Pond K, S. L. Less is more: an adaptive branch-site random effects model for efficient detection of episodic diversifying selection. Mol Biol Evol. 2015;32(5):1342–53.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  73. Mu W, Liu J, Zhang H. Complete mitochondrial genome of Benthodytes marianensis (Holothuroidea: Elasipodida: Psychropotidae): Insight into deep sea adaptation in the sea cucumber. PLoS One. 2018;13(11):e0208051.

    Article  PubMed  PubMed Central  Google Scholar 

  74. Wertheim JO, Murrell B, Smith MD, Pond K, S. L., Scheffler K. RELAX: detecting relaxed selection in a phylogenetic framework. Mol Biol Evol. 2015;32(3):820–32.

    Article  CAS  PubMed  Google Scholar 

  75. Clary DO, Wolstenholme DR. The mitochondrial DNA molecule of Drosophila yakuba: nucleotide sequence, gene organization, and genetic code. J Mol Evol. 1985;22(3):252–71.

    Article  CAS  PubMed  Google Scholar 

  76. Romanova EV, Aleoshin VV, Kamaltynov RM, Mikhailov KV, Logacheva MD, Sirotinina EA, Sherbakov DY. Evolution of mitochondrial genomes in Baikalian amphipods. BMC Genomics. 2016;17(14):291–306.

    Google Scholar 

  77. Bauzà-Ribot MM, Jaume D, Juan C, Pons J. The complete mitochondrial genome of the subterranean crustacean Metacrangonyx longipes (Amphipoda): a unique gene order and extremely short control region: full-length Research Paper. Mitochondrial DNA. 2009;20(4):88–99.

    Article  PubMed  Google Scholar 

  78. Ki JS, Hop H, Kim SJ, Kim IC, Park HG, Lee JS. Complete mitochondrial genome sequence of the Arctic gammarid, Onisimus nanseni (Crustacea; Amphipoda): novel gene structures and unusual control region features. Comp Biochem Physiol D Genomics Proteomics. 2010;5(2):105–15.

    Article  PubMed  Google Scholar 

  79. Issartel J, Voituron Y, Guillaume O, Clobert J, Hervant F. Selection of physiological and metabolic adaptations to food deprivation in the Pyrenean newt Calotriton asper during cave colonisation. Comp Biochem Physiol A Mol Integr Physiol. 2010;155(1):77–83.

    Article  PubMed  Google Scholar 

  80. Rivarola-Duarte L, Otto C, Jühling F, Schreiber S, Bedulina D, Jakob L, Stadler PF. A first glimpse at the genome of the Baikalian amphipod Eulimnogammarus verrucosus. J Exp Zool Mol Dev Evol. 2014;322(3):177–89.

    Article  CAS  Google Scholar 

  81. Kornobis E, Pálsson SN, Svavarsson J. Classification of Crangonyx Islandicus (Amphipoda, Crangonyctidae) based on morphological characters and comparison with molecular phylogenies. Zootaxa. 2012;3233(1):52–66.

    Article  Google Scholar 

  82. Macey JR, Larson A, Ananjeva NB, Papenfuss TJ. Replication slippage may cause parallel evolution in the secondary structures of mitochondrial transfer RNAs. Mol Biol Evol. 1997;14(1):30–9.

    Article  CAS  PubMed  Google Scholar 

  83. Katoh K, Standley DM. MAFFT multiple sequence alignment software version 7: improvements in performance and usability. Mol Biol Evol. 2013;30(4):772–80.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  84. Shin SC, Cho J, Lee JK, Ahn DH, Lee H, Park H. Complete mitochondrial genome of the Antarctic Amphipod Gondogeneia Antarctica (Crustacea, amphipod). Mitochondrial DNA. 2012;23(1):25–7.

    Article  CAS  PubMed  Google Scholar 

  85. Yang HM, Song JH, Kim MS, Min GS. The complete mitochondrial genomes of two talitrid amphipods, Platorchestia Japonica and P. Parapacifica (Crustacea, Amphipoda). Mitochondrial DNA B. 2017;2(2):757–8.

    Article  Google Scholar 

  86. R Core Team.. R: A language and environment for statistical computing. R Foundation for Statistical Computing, Vienna, Austria. 2021. https://www.R-project.org/.

  87. Dierckxsens N, Mardulyn P, Smits G. NOVOPlasty: de novo assembly of organelle genomes from whole genome data. Nucleic Acids Res. 2017;45(4):e18–8.

    PubMed  Google Scholar 

  88. Reyes A, Gissi C, Pesole G, Saccone C. Asymmetrical directional mutation pressure in the mitochondrial genome of mammals. Mol Biol Evol. 1998;15(8):957–66.

    Article  CAS  PubMed  Google Scholar 

  89. Wei SJ, Shi M, Chen XX, Sharkey MJ, van Achterberg C, Ye GY, He JH. New views on strand asymmetry in insect mitochondrial genomes. PLoS One. 2010;5(9):e12708.

    Article  PubMed  PubMed Central  Google Scholar 

  90. Hao YJ, Zou YL, Ding YR, Xu WY, Yan ZT, Li XD, Chen B. Complete mitochondrial genomes of Anopheles stephensi and an. Dirus and comparative evolutionary mitochondriomics of 50 mosquitoes. Sci Rep. 2017;7(1):1–13.

    Article  Google Scholar 

  91. Zhang J, Nielsen R, Yang Z. Evaluation of an improved branch-site likelihood method for detecting positive selection at the molecular level. Mol Biol Evol. 2005;22(12):2472–9.

    Article  CAS  PubMed  Google Scholar 

  92. Peng Y, Yang Z, Zhang H, Cui C, Qi X, Luo X, Su B. Genetic variations in tibetan populations and high-altitude adaptation at the Himalayas. Mol Biol Evol. 2011;28(2):1075–81.

    Article  CAS  PubMed  Google Scholar 

  93. Macher JN, Leese F, Weigand AM, Rozenberg A. The complete mitochondrial genome of a cryptic amphipod species from the Gammarus fossarum complex. Mitochondrial DNA Part B. 2017;2(1):17–8.

    Article  PubMed  PubMed Central  Google Scholar 

  94. Rawlings TA, Collins TM, Bieler R. A major mitochondrial gene rearrangement among closely related species. Mol Biol Evol. 2001;18(8):1604–9.

    Article  CAS  PubMed  Google Scholar 

  95. Zou H, Jakovlić I, Chen R, Zhang D, Zhang J, Li WX, Wang GT. The complete mitochondrial genome of parasitic nematode Camallanus cotti: extreme discontinuity in the rate of mitogenomic architecture evolution within the Chromadorea class. BMC Genomics. 2017;18(1):1–17.

    Article  CAS  Google Scholar 

  96. Boore JL, Macey JR, Medina M. Sequencing and comparing whole mitochondrial genomes of animals. Methods Enzymol. 2005;395:311–48.

    Article  CAS  PubMed  Google Scholar 

  97. Lanfear R, Calcott B, Ho SY, Guindon S. PartitionFinder: combined selection of partitioning schemes and substitution models for phylogenetic analyses. Mol Biol Evol. 2012;29(6):1695–701.

    Article  CAS  PubMed  Google Scholar 

  98. Castresana J, Feldmaier-Fuchs G, Pääbo S. Codon reassignment and amino acid composition in hemichordate mitochondria. Proc Natl Acad Sci. 1998;95(7):3703–7.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  99. Nair P, Huertas M, Nowlin WH. Metabolic responses to long-term food deprivation in subterranean and surface amphipods. Subterr Biol. 2020;33:1–15.

    Article  Google Scholar 

  100. Niemiller ML, Porter ML, Keany J, Gilbert H, Fong DW, Culver DC, Taylor SJ. Evaluation of eDNA for groundwater invertebrate detection and monitoring: a case study with endangered Stygobromus (Amphipoda: Crangonyctidae). Conserv Genet Resour. 2018;10(2):247–57.

    Article  Google Scholar 

  101. Wilson K, Cahill V, Ballment E, Benzie J. The complete sequence of the mitochondrial genome of the crustacean Penaeus monodon: are malacostracan crustaceans more closely related to insects than to branchiopods? Mol Biol Evol. 2000;17(6):863–74.

    Article  CAS  PubMed  Google Scholar 

  102. Ren J, Shen X, Jiang F, Liu B. The mitochondrial genomes of two scallops, Argopecten irradians and Chlamys farreri (Mollusca: Bivalvia): the most highly rearranged gene order in the family Pectinidae. J Mol Evol. 2010;70(1):57–68.

    Article  CAS  PubMed  Google Scholar 

  103. Ren J, Liu X, Jiang F, Guo X, Liu B. Unusual conservation of mitochondrial gene order in Crassostrea oysters: evidence for recent speciation in Asia. BMC Evol Biol. 2010a;10(1):1–14.

    Article  Google Scholar 

  104. Xin Y, Ren J, Liu X. Mitogenome of the small abalone Haliotis diversicolor Reeve and phylogenetic analysis within Gastropoda. Mar Genom. 2011;4(4):253–62.

    Article  Google Scholar 

  105. Bauzà-Ribot MM, Juan C, Nardi F, Oromí P, Pons J, Jaume D. Mitogenomic phylogenetic analysis supports continental-scale vicariance in subterranean thalassoid crustaceans. Curr Biol. 2012;22(21):2069–74.

    Article  PubMed  Google Scholar 

  106. Li XD, Jiang GF, Yan LY, Li R, Mu Y, Deng WA. Positive selection drove the adaptation of mitochondrial genes to the demands of flight and high-altitude environments in grasshoppers. Front Genet. 2018;9:605.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  107. Ojala D, Montoya J, Attardi G. tRNA punctuation model of RNA processing in human mitochondria. Nature. 1981;290(5806):470–4.

    Article  CAS  PubMed  Google Scholar 

  108. Lin FJ, Liu Y, Sha Z, Tsang LM, Chu KH, Chan TY, Cui Z. Evolution and phylogeny of the mud shrimps (Crustacea: Decapoda) revealed from complete mitochondrial genomes. BMC Genomics. 2012;13(1):1–12.

    Article  Google Scholar 

  109. Yamazaki N, Ueshima R, Terrett JA, Yokobori SI, Kaifu M, Segawa R, Thomas RH. Evolution of pulmonate gastropod mitochondrial genomes: comparisons of gene organizations of Euhadra, Cepaeaand Albinaria and implications of unusual tRNA secondary structures. Genetics. 1997;145(3):749–58.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  110. Nardi F, Carapelli A, Fanciulli PP, Dallai R, Frati F. The complete mitochondrial DNA sequence of the basal hexapod Tetrodontophora bielanensis: evidence for heteroplasmy and tRNA translocations. Mol Biol Evol. 2001;18(7):1293–304.

    Article  CAS  PubMed  Google Scholar 

  111. Copilaş-Ciocianu D, Sidorov D, Gontcharov A. Adrift across tectonic plates: molecular phylogenetics supports the ancient laurasian origin of old limnic crangonyctid amphipods. Org Divers Evol. 2019;19:191–207.

    Article  Google Scholar 

  112. Kilpert F, Podsiadlowski L. The mitochondrial genome of the Japanese skeleton shrimp Caprella mutica (Amphipoda: Caprellidea) reveals a unique gene order and shared apomorphic translocations with Gammaridea. Mitochondrial DNA. 2010;21(3–4):77–86.

    Article  CAS  PubMed  Google Scholar 

  113. Koenemann S, Holsinger JR. Systematics of the north American subterranean amphipod genus Bactrurus. (Crangonyctidae) Beaufortia. 2001;51(1):1–56.

    Google Scholar 

  114. Kornobis E, Pálsson S, Sidorov DA, Holsinger JR, Kristjánsson BK. Molecular taxonomy and phylogenetic affinities of two groundwater amphipods, Crangonyx islandicus and Crymostygius Thingvallensis, endemic to Iceland. Mol Phylogenet Evol. 2011;58(3):527–39.

    Article  PubMed  Google Scholar 

  115. Carroll J, Fearnley IM, Wang Q, Walker JE. Measurement of the molecular masses of hydrophilic and hydrophobic subunits of ATP synthase and complex I in a single experiment. Anal Biochem. 2009;395(2):249–55.

    Article  CAS  PubMed  Google Scholar 

  116. Matsumoto Y, Yanase T, Tsuda T, Noda H. Species–specific mitochondrial gene rearrangements in biting midges and vector species identification. Med Vet Entomol. 2009;23(1):47–55.

    Article  CAS  PubMed  Google Scholar 

  117. Scheffler IE. Molecular genetics of succinate: quinone oxidoreductase in eukaryotes. Prog Nucl Acid Res Mol Biol. 1998;60:267–315.

    Article  CAS  Google Scholar 

  118. 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 Genomics. 2018;40(11):1169–80.

    Article  CAS  PubMed  Google Scholar 

  119. McKenzie M, Lazarou M, Ryan MT. Analysis of respiratory chain complex assembly with radiolabeled nuclear- and mitochondrial-encoded subunits. Methods Enzymol. 2009;456:321–39.

    Article  CAS  PubMed  Google Scholar 

  120. Zapelloni F, Jurado-Rivera JA, Jaume D, Juan C, Pons J. Comparative Mitogenomics in Hyalella (Amphipoda: Crustacea). Genes. 2021;12(2):292.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  121. Meiklejohn CD, Montooth KL, Rand DM. Positive and negative selection on the mitochondrial genome. Trends Genet. 2007;23(6):259–63.

    Article  CAS  PubMed  Google Scholar 

  122. Mitterboeck TF, Adamowicz SJ. Flight loss linked to faster molecular evolution in insects. Proc Royal Soc Biol Sci. 2013;280(1767):20131128.

    Google Scholar 

  123. Zhang H, Luo Q, Sun J, Liu F, Wu G, Yu J, Wang W. Mitochondrial genome sequences of Artemia tibetiana and Artemia urmiana: assessing molecular changes for high plateau adaptation. Sci China Life Sci. 2013a;56(5):440–52.

    Article  CAS  PubMed  Google Scholar 

  124. Taylor SJ, Niemiller ML. Biogeography and conservation assessment of Bactrurus groundwater amphipods (Crangonyctidae) in the central and eastern United States. Subterr Biol. 2016;17:1–29.

    Article  Google Scholar 

  125. Hassanin A, Leger NELLY, Deutsch J. Evidence for multiple reversals of asymmetric mutational constraints during the evolution of the mitochondrial genome of Metazoa, and consequences for phylogenetic inferences. Syst Biol. 2005;54(2):277–98.

    Article  PubMed  Google Scholar 

  126. Carlini DB, Fong DW. The transcriptomes of cave and surface populations of Gammarus minus (Crustacea: Amphipoda) provide evidence for positive selection on cave downregulated transcripts. PLoS One. 2017;12(10):e0186173.

    Article  PubMed  PubMed Central  Google Scholar 

  127. Garvin MR, Bielawski JP, Gharrett AJ. Positive Darwinian selection in the piston that powers proton pumps in complex I of the mitochondria of Pacific salmon. PLoS One. 2011;6(9):e24127.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  128. Zhang D, Zou H, Hua CJ, Li WX, Mahboob S, Al-Ghanim KA, Wang GT. Mitochondrial architecture rearrangements produce asymmetrical nonadaptive mutational pressures that subvert the phylogenetic reconstruction in Isopoda. Genome Biol Evol. 2019;11(7):1797–812.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  129. Okimoto R, Wolstenholme DR. A set of tRNAs that lack either the T psi C arm or the dihydrouridine arm: towards a minimal tRNA adaptor. EMBO J. 1990;9(10):3405–11.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  130. Xiao JH, Jia JG, Murphy RW, Huang DW. Rapid evolution of the mitochondrial genome in chalcidoid wasps (Hymenoptera: Chalcidoidea) driven by parasitic lifestyles. PLoS One. 2011;6(11):e26645.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  131. Wirth C, Brandt U, Hunte C, Zickermann V. Structure and function of mitochondrial complex I. Biochim et Biophys Acta (BBA)-Bioenergetics. 2016;1857(7):902–14.

    Article  CAS  Google Scholar 

  132. Dowling TE, Martasian DP, Jeffery WR. Evidence for multiple genetic forms with similar eyeless phenotypes in the blind cavefish, Astyanax mexicanus. Mol Biol Evol. 2002;19(4):446–55.

    Article  CAS  PubMed  Google Scholar 

  133. Zhang ZY, Chen B, Zhao DJ, Kang L. Functional modulation of mitochondrial cytochrome c oxidase underlies adaptation to high-altitude hypoxia in a Tibetan migratory locust. Proc Royal Soc B Biol Sci. 2013;280(1756):20122758.

    Article  Google Scholar 

Download references

Acknowledgements

The authors would like to thank David C. Culver, Daniel W. Fong, Steven J. Taylor, and Kayla Wilson for collecting specimens. We also thank David Young from Alabama Supercomputer Center for providing assistance with troubleshooting and program installation on the high-performance computing cluster.

Funding

This study was supported by a grant from the Friends of the Capital Crescent Trail, the town of Chevy Chase, Illinois Natural History Survey, and The University of Alabama in Huntsville.

Author information

Authors and Affiliations

Authors

Contributions

Authors Joseph B. Benito and Matthew L. Niemiller contributed to conceptualization, data collection and curation, and manuscript writing and reviewing. Author Megan L. Porter contributed to manuscript writing and reviewing.

Corresponding author

Correspondence to Matthew L. Niemiller.

Ethics declarations

Ethics approval and consent to participate

Not applicable.

Consent for publication

Not applicable.

Competing interests

The authors declare no competing interests.

Additional information

Publisher’s Note

Springer Nature remains neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Supplementary Information

Rights and permissions

Open Access This article is licensed under a Creative Commons Attribution 4.0 International License, which permits use, sharing, adaptation, distribution and reproduction in any medium or format, as long as you give appropriate credit to the original author(s) and the source, provide a link to the Creative Commons licence, and indicate if changes were made. The images or other third party material in this article are included in the article's Creative Commons licence, unless indicated otherwise in a credit line to the material. If material is not included in the article's Creative Commons licence and your intended use is not permitted by statutory regulation or exceeds the permitted use, you will need to obtain permission directly from the copyright holder. To view a copy of this licence, visit http://creativecommons.org/licenses/by/4.0/. The Creative Commons Public Domain Dedication waiver (http://creativecommons.org/publicdomain/zero/1.0/) applies to the data made available in this article, unless otherwise stated in a credit line to the data.

Reprints and permissions

About this article

Check for updates. Verify currency and authenticity via CrossMark

Cite this article

Benito, J.B., Porter, M.L. & Niemiller, M.L. Comparative mitogenomic analysis of subterranean and surface amphipods (Crustacea, Amphipoda) with special reference to the family Crangonyctidae. BMC Genomics 25, 298 (2024). https://doi.org/10.1186/s12864-024-10111-w

Download citation

  • Received:

  • Accepted:

  • Published:

  • DOI: https://doi.org/10.1186/s12864-024-10111-w

Keywords