Nothing Special   »   [go: up one dir, main page]

You seem to have javascript disabled. Please note that many of the page functionalities won't work as expected without javascript enabled.
 
 
Sign in to use this feature.

Years

Between: -

Subjects

remove_circle_outline
remove_circle_outline
remove_circle_outline
remove_circle_outline
remove_circle_outline
remove_circle_outline
remove_circle_outline
remove_circle_outline
remove_circle_outline

Journals

Article Types

Countries / Regions

Search Results (130)

Search Parameters:
Keywords = amphipod

Order results
Result details
Results per page
Select all
Export citation of selected articles as:
19 pages, 4812 KiB  
Article
UV Sensitivities of Two Littoral and Two Deep-Freshwater Amphipods (Amphipoda, Crustacea) Reflect Their Preferred Depths in the Ancient Lake Baikal
by Elizaveta Kondrateva, Anton Gurkov, Yaroslav Rzhechitskiy, Alexandra Saranchina, Anastasiia Diagileva, Polina Drozdova, Kseniya Vereshchagina, Zhanna Shatilina, Inna Sokolova and Maxim Timofeyev
Biology 2024, 13(12), 1004; https://doi.org/10.3390/biology13121004 - 2 Dec 2024
Viewed by 595
Abstract
Solar ultraviolet (UV) is among the most important ecological factors shaping the composition of biota on the planet’s surface, including the upper layers of waterbodies. Inhabitants of dark environments recently evolving from surface organisms provide natural opportunities to study the evolutionary losses of [...] Read more.
Solar ultraviolet (UV) is among the most important ecological factors shaping the composition of biota on the planet’s surface, including the upper layers of waterbodies. Inhabitants of dark environments recently evolving from surface organisms provide natural opportunities to study the evolutionary losses of UV adaptation mechanisms and better understand how those mechanisms function at the biochemical level. The ancient Lake Baikal is the only freshwater reservoir where deep-water fauna emerged, and its diverse endemic amphipods (Amphipoda, Crustacea) now inhabit the whole range from highly transparent littoral to dark depths of over 1600 m, which makes them a convenient model to study UV adaptation. With 10-day-long laboratory exposures, we show that adults of deep-water Baikal amphipods Ommatogammarus flavus and O. albinus indeed have high sensitivity to environmentally relevant UV levels in contrast to littoral species Eulimnogammarus cyaneus and E. verrucosus. The UV intolerance was more pronounced in deeper-dwelling O. albinus and was partially explainable by lower levels of carotenoids and carotenoid-binding proteins. Signs of oxidative stress were not found but UV-B specifically seemingly led to the accumulation of toxic compounds. Overall, the obtained results demonstrate that UV is an important factor limiting the distribution of deep-water amphipods into the littoral zone of Lake Baikal. Full article
(This article belongs to the Section Ecology)
Show Figures

Figure 1

Figure 1
<p>Photographs of adult endemic amphipods from Lake Baikal used in the study: littoral species <span class="html-italic">Eulimnogammarus cyaneus</span> (Dybowski, 1874) and <span class="html-italic">E. verrucosus</span> (Gerstfeldt, 1858) and deep-water species <span class="html-italic">Ommatogammarus flavus</span> (Dybowsky, 1874) and <span class="html-italic">O. albinus</span> (Dybowsky, 1874). Depth ranges for each species distribution are given according to [<a href="#B13-biology-13-01004" class="html-bibr">13</a>,<a href="#B20-biology-13-01004" class="html-bibr">20</a>,<a href="#B21-biology-13-01004" class="html-bibr">21</a>]. Photo credits: Elizaveta Kondrateva and Polina Drozdova.</p>
Full article ">Figure 2
<p>UV monitoring on the shoreline of Lake Baikal (Bolshie Koty village, 4–10 of July). (<b>a</b>) UV-A and (<b>b</b>) UV-B intensities corresponding to the 24 h time format. (<b>c</b>) Relation between visible solar illumination and UV-A intensity. (<b>d</b>) Relation between UV-A and UV-B intensities. Horizontal solid lines in (<b>a</b>,<b>b</b>) indicate medians. Dotted lines in (<b>c</b>,<b>d</b>) show linear regressions, and <span class="html-italic">R</span><sup>2</sup> indicates adjusted coefficient of determination for the models.</p>
Full article ">Figure 3
<p>Survival rates of endemic amphipods from Lake Baikal (littoral species <span class="html-italic">E. cyaneus</span> and <span class="html-italic">E. verrucosus</span> and deep-water species <span class="html-italic">O. flavus</span> and <span class="html-italic">O. albinus</span>) during laboratory exposures under (<b>a</b>) UV-A and (<b>b</b>) UV-B treatments. Solid lines indicate the Kaplan–Meier survival curves, colored bands show the 95% confidence intervals for the curves and vertical dashed lines indicate reaching LT<sub>50</sub>. <span class="html-italic">n</span> = 160 for <span class="html-italic">E. cyaneus</span>; <span class="html-italic">n</span> = 80 (UV-A) or 120 (UV-B) for <span class="html-italic">E. verrucosus</span>; <span class="html-italic">n</span> = 40 for each <span class="html-italic">Ommatogammarus</span> species (see <a href="#app1-biology-13-01004" class="html-app">Table S1</a> for raw data).</p>
Full article ">Figure 4
<p>Horizontal locomotor activity of endemic amphipods from Lake Baikal (littoral species <span class="html-italic">E. cyaneus</span> and <span class="html-italic">E. verrucosus</span> and deep-water species <span class="html-italic">O. flavus</span> and <span class="html-italic">O. albinus</span>) during laboratory exposures under (<b>a</b>) UV-A and (<b>b</b>) UV-B treatments. <span class="html-italic">n</span> ≥ 8 for <span class="html-italic">Eulimnogammarus</span> species; <span class="html-italic">n</span> ≥ 4 for <span class="html-italic">Ommatogammarus</span> species (see <a href="#app1-biology-13-01004" class="html-app">Table S1</a> for raw data). * <span class="html-italic">p</span> &lt; 0.05, ** <span class="html-italic">p</span> &lt; 0.01, *** <span class="html-italic">p</span> &lt; 0.001.</p>
Full article ">Figure 5
<p>Activities of antioxidant enzymes (<b>a</b>) peroxidase (POD), (<b>b</b>) catalase (CAT), and (<b>c</b>) glutathione-S-transferase (GST) in endemic amphipods from Lake Baikal after laboratory exposures under UV-A and UV-B treatments. Note the difference in chosen exposure times: 10 days for littoral species (<span class="html-italic">E. cyaneus</span> and <span class="html-italic">E. verrucosus</span>) and 3 days for deep-water species (<span class="html-italic">O. flavus</span> and <span class="html-italic">O. albinus</span>). Horizontal colored solid lines indicate medians and dots show biological replicates. <span class="html-italic">n</span> ≥ 3 for <span class="html-italic">E. cyaneus</span>; <span class="html-italic">n</span> ≥ 5 for <span class="html-italic">E. verrucosus</span>; <span class="html-italic">n</span> ≥ 6 for <span class="html-italic">Ommatogammarus</span> species (see <a href="#app1-biology-13-01004" class="html-app">Table S1</a> for raw data). ** <span class="html-italic">p</span> &lt; 0.01, *** <span class="html-italic">p</span> &lt; 0.001.</p>
Full article ">Figure 6
<p>DNA breaks (expressed in tail moment, TM) in hemocytes of endemic amphipods from Lake Baikal (littoral species <span class="html-italic">E. cyaneus</span> and <span class="html-italic">E. verrucosus</span> and deep-water species <span class="html-italic">O. flavus</span> and <span class="html-italic">O. albinus</span>) after laboratory exposure under UV-A and UV-B treatments during 3 days. Horizontal colored solid lines indicate medians and dots show biological replicates. <span class="html-italic">n</span> ≥ 5 for <span class="html-italic">E. verrucosus</span>; <span class="html-italic">n</span> ≥ 8 for other species (see <a href="#app1-biology-13-01004" class="html-app">Table S1</a> for raw data). *** <span class="html-italic">p</span> &lt; 0.001.</p>
Full article ">Figure 7
<p>Parameters related to UV protection in endemic amphipods from Lake Baikal (littoral <span class="html-italic">E. verrucosus</span> and deep-water species <span class="html-italic">O. flavus</span> and <span class="html-italic">O. albinus</span>). (<b>a</b>) Total carotenoid content in the body (expressed in parts per million—ppm). (<b>b</b>) The level of a carotenoid-binding protein (CBP) in hemolymph (expressed in arbitrary units—AU). Note that the animals were kept in laboratory conditions for a short time without stressful exposures, used for hemolymph extraction and, finally, frozen for measurement of carotenoid content. Horizontal colored solid lines indicate medians and dots show biological replicates. <span class="html-italic">n</span> ≥ 6 for all species (see <a href="#app1-biology-13-01004" class="html-app">Table S1</a> for raw data). ** <span class="html-italic">p</span> &lt; 0.01.</p>
Full article ">Figure 8
<p>Survival rates of deep-water <span class="html-italic">O. flavus</span> kept either with littoral <span class="html-italic">E. cyaneus</span> or littoral <span class="html-italic">E. verrucosus</span> in two aquaria with large surfaces and with stones. UV radiation was excluded, while visible illumination was dim and followed the natural diurnal cycle. Dots show the results of counting amphipods in these two aquaria. See <a href="#app1-biology-13-01004" class="html-app">Table S1</a> for raw data.</p>
Full article ">
15 pages, 3174 KiB  
Article
Extent of Benthic Habitat Disturbance by Offshore Infrastructure
by Robert M. Cerrato, Roger D. Flood, Justin Bopp and Henry J. Bokuniewicz
J. Mar. Sci. Eng. 2024, 12(12), 2142; https://doi.org/10.3390/jmse12122142 - 24 Nov 2024
Viewed by 408
Abstract
The effects of the interaction between sandy, mobile, low-relief (sorted) bedforms and two sewage outfalls were investigated along the south shore of Long Island, NY. Sand bedforms at scales from ripples to ridges are common on continental shelves. In dynamic environments, these features [...] Read more.
The effects of the interaction between sandy, mobile, low-relief (sorted) bedforms and two sewage outfalls were investigated along the south shore of Long Island, NY. Sand bedforms at scales from ripples to ridges are common on continental shelves. In dynamic environments, these features can migrate 10s to 100s of meters per year, especially during storms. Beyond engineering considerations, little is known of the interaction between these mobile features and anthropogenic structures. Modification of bedform topography and sediment grain-size distribution can be expected to alter the species composition, abundance, and diversity of the benthic community. At the study site, the interaction increased the scour of modern fine- to medium-grained sediments extending out to a kilometer and uncovered coarser-grained late Pleistocene sediments. This alteration of the seafloor in turn resulted in changes in composition, higher abundance, and lower diversity in the species assemblage found in the impacted area. The most advantaged species was Pseudunciola obliquua, a sightless, tube-building, surface deposit-feeding amphipod that is known to prefer a dynamic coarse sand habitat. Overall, the ecological effects of artificial structures on a wave-dominated seabed with sorted bedforms have not been adequately assessed. In particular, and of great importance, is the pending large-scale development of wind farms off the East Coast of the U.S. Full article
(This article belongs to the Special Issue Morphological Changes in the Coastal Ocean)
Show Figures

Figure 1

Figure 1
<p>Location map south of Long Island, NY, USA (<b>A</b>) and ocean outfalls study areas (<b>B</b>).</p>
Full article ">Figure 2
<p>Sonar backscatter maps covering an area of about 3.0 × 1.6 km around (<b>A</b>) the Cedar Creek outfall and (<b>B</b>) the SWSD outfall. The outfall diffuser is represented as a black line. Acoustic backscatter was taken as a proxy for bottom type or habitat. High backscatter (light grey) usually corresponds to coarser sediments and/or coarser-scale morphology relative to low backscatter areas (dark greys).</p>
Full article ">Figure 3
<p>Sand wave provinces around (<b>A</b>) the Cedar Creek outfall and (<b>B</b>) the SWSD outfall. The outfall diffuser is represented as a black line. The red line bisects the study area into north and south regions and is oriented in the same direction as the outfall. Colors delineate the extent of provinces. The colors were arbitrarily chosen and do not represent similar conditions between outfall areas.</p>
Full article ">Figure 4
<p>Sampling transects (A–J), sampling locations along a transect (1–6), and the split-plot factor design used in the RDA analysis to test the impact of the diffusers on benthic fauna community structure.</p>
Full article ">Figure 5
<p>RDA ordination triplot for the benthic fauna. Species codes are tabulated in the <a href="#app1-jmse-12-02142" class="html-app">Supplementary Materials</a>. Different sample symbols were assigned to each transect.</p>
Full article ">Figure 6
<p>A contour-based attribute plot of mean grain size in phi units overlaid on the RDA ordination triplot for the benthic fauna. The contour plot was generated from a locally weighted polynomial regression (loess) between mean grain size and sample scores. Species codes are tabulated in the <a href="#app1-jmse-12-02142" class="html-app">Supplementary Materials</a>.</p>
Full article ">Figure 7
<p>Contour-based attribute plots of abundance (<b>A</b>), Shannon diversity (<b>B</b>), evenness (<b>C</b>), and species richness (<b>D</b>) overlaid on the <a href="#jmse-12-02142-f005" class="html-fig">Figure 5</a> RDA ordination triplot for the benthic fauna. The contour plots were generated from a locally weighted polynomial regression (loess) between the metric and sample scores. Species codes are tabulated in the <a href="#app1-jmse-12-02142" class="html-app">Supplementary Materials</a>.</p>
Full article ">
15 pages, 3033 KiB  
Article
Congruent and Hierarchical Intra-Lake Subdivisions from Nuclear and Mitochondrial Data of a Lake Baikal Shoreline Amphipod
by Risto Väinölä, Tytti Kontula, Kazuo Mashiko and Ravil M. Kamaltynov
Diversity 2024, 16(11), 706; https://doi.org/10.3390/d16110706 - 20 Nov 2024
Viewed by 547
Abstract
A central goal of molecular studies on ancient lake faunas is to resolve the origin and phylogeny of their strikingly diverse endemic species flocks. Another equally intriguing goal is to understand the integrity of individual morphologically diagnosed species, which should help to perceive [...] Read more.
A central goal of molecular studies on ancient lake faunas is to resolve the origin and phylogeny of their strikingly diverse endemic species flocks. Another equally intriguing goal is to understand the integrity of individual morphologically diagnosed species, which should help to perceive the nature and speed of the speciation process, and the true biological species diversity. In the uniquely diverse Lake Baikal amphipod crustaceans, molecular data from shallow-water species have often disclosed their cryptic subdivision into geographically segregated genetic lineages, but the evidence so far is mainly based on mitochondrial DNA. We now present a lake-wide parallel survey of both mitochondrial and multilocus nuclear genetic structuring in the common shoreline amphipod Eulimnogammarus verrucosus, known to comprise three deep, parapatric mtDNA lineages. Allele frequencies of seven nuclear allozyme loci divide the data into three main groups whose distributions exactly match the distributions of the main mitochondrial lineages S, W, and E and involve a further division of the W cluster into two subgroups. The inter-group differences involve one to four diagnostic loci and additional group-specific alleles. The transition zones are either abrupt (1 km), occur over a long segment of uninhabitable shoreline, or may be gradual with non-coincident clinal change at different loci. Mitochondrial variation is hierarchically structured, each main lineage further subdivided into 2–4 parapatric sublineages or phylogroups, and patterns of further local segregation are seen in some of them. Despite the recurring observations of cryptic diversity in Baikalian amphipods, the geographical subdivisions and clade depths do not match in different taxa, defying a common explanation for the diversification in environmental history. Full article
(This article belongs to the Special Issue Diversity and Evolution within the Amphipoda)
Show Figures

Figure 1

Figure 1
<p>Index map of sampling localities. Black dots are sites for the allozyme + mtDNA data. Plain numbers represent 1993–1995 samples, with site codes from Mashiko et al. [<a href="#B14-diversity-16-00706" class="html-bibr">14</a>]; codes with a letter are adjacent sites from the same expeditions. Numbers with an asterisk are sites from 1999. Open circles are sites with mtDNA data only. A full list of the localities with sample information is presented in <a href="#app1-diversity-16-00706" class="html-app">Table S1</a>. Open squares are additional sites of Gurkov et al.’s mtDNA data [<a href="#B19-diversity-16-00706" class="html-bibr">19</a>]. The photographs display <span class="html-italic">E. verrucosus</span> E from the Chivyrkui Bay, near site 22.</p>
Full article ">Figure 2
<p>Principal components plot (PC1 vs. PC2) of the <span class="html-italic">Eulimnogammarus verrucosus</span> complex allozyme frequency data. Locality codes are those in <a href="#diversity-16-00706-f001" class="html-fig">Figure 1</a> and <a href="#app1-diversity-16-00706" class="html-app">Table S1</a>.</p>
Full article ">Figure 3
<p>NJ trees (<b>A</b>) from uncorrected <span class="html-italic">p</span>-distances from the original <span class="html-italic">Eulimnogammarus verrucosus</span> COI sequence data in this study, (<b>B</b>) from allozyme data (7-locus Euclidean distances among populations, data as in <a href="#diversity-16-00706-f002" class="html-fig">Figure 2</a>). (<b>C</b>) Distribution of the three main genetic groups S, W, and E and the subgroups of W along the shores of Baikal, congruently in the two datasets. The grey reference sequence is of <span class="html-italic">E. oligacanthus</span>.</p>
Full article ">Figure 4
<p>Frequencies of major alleles at each allozyme locus plotted against geographical distance along the shoreline around Baikal, clockwise from site 1 in <a href="#diversity-16-00706-f001" class="html-fig">Figure 1</a>. Samples 28, 19*, and 27 from the Olkhon and Ogoi islands in the W(a)/W(b) borderline are encircled, those from the continental strand are interconnected by the lines. Original data are given in <a href="#app1-diversity-16-00706" class="html-app">Table S2</a>.</p>
Full article ">Figure 5
<p>Examples of allele frequency variation at three nuclear allozyme loci. The two most common alleles and the pooled frequency of remaining minor alleles at each locus are shown in the pie diagrams; data from <a href="#app1-diversity-16-00706" class="html-app">Table S2</a>.</p>
Full article ">Figure 6
<p><span class="html-italic">COI</span> haplotype trees for each of the three main mitochondrial lineages of <span class="html-italic">Eulimnogammarus verrucosus</span>, color-coded for geographically demarcated sublineages or phylogroups. The topologies are examples from larger sets of equally parsimonious MP trees, and detailed local relationships are not significant. Haplogroups with restricted distribution within a phylogroup range are surrounded by a dashed line, reciprocally in the tree and on the map. Individuals clustering in a clade typical of another region are marked with the color of their own region. Apart from original sequences, the trees include data from Gurkov et al. [<a href="#B19-diversity-16-00706" class="html-bibr">19</a>] and Saranchina et al. [<a href="#B20-diversity-16-00706" class="html-bibr">20</a>] from sites indicated by open squares.</p>
Full article ">
19 pages, 2564 KiB  
Article
Genome Structure, Evolution, and Host Shift of Nosema
by Xiao Xiong, Christopher J. Geden, Yongjun Tan, Ying Zhang, Dapeng Zhang, John H. Werren and Xu Wang
Biology 2024, 13(11), 952; https://doi.org/10.3390/biology13110952 - 19 Nov 2024
Viewed by 658
Abstract
Nosema is a diverse fungal genus of unicellular, obligate symbionts infecting various arthropods. We performed comparative genomic analyses of seven Nosema species that infect bees, wasps, moths, butterflies, and amphipods. As intracellular parasites, these species exhibit significant genome reduction, retaining only about half [...] Read more.
Nosema is a diverse fungal genus of unicellular, obligate symbionts infecting various arthropods. We performed comparative genomic analyses of seven Nosema species that infect bees, wasps, moths, butterflies, and amphipods. As intracellular parasites, these species exhibit significant genome reduction, retaining only about half of the genes found in free-living yeast genomes. Notably, genes related to oxidative phosphorylation are entirely absent (p < 0.001), and those associated with endocytosis are significantly diminished compared to other pathways (p < 0.05). All seven Nosema genomes display significantly lower G-C content compared to their microsporidian outgroup. Species-specific 5~12 bp motifs were identified immediately upstream of start codons for coding genes in all species (p ≤ 1.6 × 10−72). Our RNA-seq data from Nosema muscidifuracis showed that this motif is enriched in highly expressed genes but depleted in lowly expressed ones (p < 0.05), suggesting it functions as a cis-regulatory element in gene expression. We also discovered diverse telomeric repeats within the genus. Phylogenomic analyses revealed two major Nosema clades and incongruency between the Nosema species tree and their hosts’ phylogeny, indicating potential host switch events (100% bootstrap values). This study advances the understanding of genomic architecture, gene regulation, and evolution of Nosema, offering valuable insights for developing strategies to control these microbial pathogens. Full article
(This article belongs to the Special Issue Advances in Evolutionary Ecology of Host–Parasite Interactions)
Show Figures

Figure 1

Figure 1
<p>A novel type of telomere in the <span class="html-italic">Nosema Muscidifuracis</span> genome. (<b>A</b>) Presence of telomeric sequences at the termini of 28 <span class="html-italic">N. Muscidifuracis</span> genome contigs. (<b>B</b>) Plot of GC content along contig14 showing the high GC content at telomeric regions. (<b>C</b>) Sequence alignment at the telomere-subtelomere boundaries, showing the novel composite 4 bp and 5 bp telomeric repeat motifs. (<b>D</b>) Total length and relative abundance of telomeric repeat motifs (TAGG, TTAGG, and TAGGG) in telomeric regions. (<b>E</b>) Phylogenetic tree of 27 subtelomeric sequences from different genomic contigs in <span class="html-italic">N. muscidifuracis</span>. (Yellow shading, subtelomeric region. Red color, positions that are not identical across all contigs. Purple shading: TTAGG repeats in telomeric region. Green shading: TAGG repeats in telomeric region).</p>
Full article ">Figure 2
<p><b>Functional pathway specific genome reduction in <span class="html-italic">Nosema muscidifuracis</span>.</b> (<b>A</b>) Gene number in 23 pathways in <span class="html-italic">Nosema muscidifuracis</span> and <span class="html-italic">Saccharomyces cerevisiae</span> (Chi-squared test, *, <span class="html-italic">p</span> &lt; 0.05; ***, <span class="html-italic">p</span> &lt; 0.001). (<b>B</b>) KEGG pathway analysis of <span class="html-italic">Nosema muscidifuracis</span> mitochondrial proteins suggested that the entire electron transport chain and eukaryotic F-type ATPase were completely missing in the mitochondrial oxidative phosphorylation metabolic pathway. The enzymes/proteins that are present in the <span class="html-italic">N. muscidifuracis</span> genome are shaded in red.</p>
Full article ">Figure 3
<p>A motif associated with translation start sites and gene expression levels in <span class="html-italic">Nosema muscidifuracis</span>. (<b>A</b>) A sequence motif enriched upstream of <span class="html-italic">N. muscidifuracis</span> genes, containing a homopolymer of seven thymine (T) nucleotides, followed by an adenine (<b>A</b>) and three consecutive cytosine (C) nucleotides. (<b>B</b>) Distribution of the motif upstream of the gene regions. The <span class="html-italic">x</span>-axis measures the distance from the first nucleotide of the motif to the start codon in bases, and the <span class="html-italic">y</span>-axis indicates the number of detected motifs. (<b>C</b>) Average RNA-seq coverage across protein-coding gene regions in <span class="html-italic">N. muscidifuracis</span>. (<b>D</b>) The percentage of genes with the motif in gene groups with different expression levels.</p>
Full article ">Figure 4
<p>Phylogenomic analysis revealed a host switch event and conserved sequence motifs in <span class="html-italic">Nosema</span>. A maximum-likelihood tree of <span class="html-italic">N. muscidifuracis</span> isolated in parasitoid wasps <span class="html-italic">Muscidifurax zaraptor</span> (NosMusMzar) and <span class="html-italic">M. raptor</span> (NosMusMrap) with other <span class="html-italic">Nosema</span> was constructed based on 449 shared proteins. The <span class="html-italic">Nosema</span> species/strains included are <span class="html-italic">N. apis</span> strain BRL01 (NosApis), <span class="html-italic">N. ceranae</span> strain PA08 1199 (NcerPA08), <span class="html-italic">N. ceranae</span> strain BRL (NcerBRL), <span class="html-italic">N. ceranae</span> strain BRL01 (NcerBRL01), the tussar moth <span class="html-italic">Antheraea pernyi Nosema strain YNPr</span> (NosYNPr), <span class="html-italic">N. antheraeae</span> strain YY (NosYY), <span class="html-italic">N. bombycis</span> strain CQ1 (NosBomCQ1), and <span class="html-italic">N. granulosis</span> strain Ou3-Ou53 (NosGranOu53). The <span class="html-italic">Encephalitozoon cuniculi</span> GB-M1 strain (Ecuniculi) was included as the outgroup. The bootstrap value is indicated by dots, with red representing a support level of 100/100. The length of each branch is indicated beneath the branches. The sequence logos displayed the conserved motifs located upstream of the start codons, as predicted by MEME using 449 shared orthologous genes and other gene models in the seven <span class="html-italic">Nosema</span> species and <span class="html-italic">E. cuniculi</span>. Genome-wide average G-C content for each species is displayed beneath their respective logos. The inferred ancestral G-C content, along with the standard deviation, is labeled near the nodes and shaded in orange.</p>
Full article ">Figure 5
<p>Codon bias and evolution toward AT-rich genomes in <span class="html-italic">Nosema</span>. (<b>A</b>) Boxplot of GC3 (G-C content at the 3rd codon position) over G-C content at all condo positions, rank ordered by the genome average G-C content in <span class="html-italic">Encephalitozoon cuniculi</span> (Ecuni), <span class="html-italic">Nosema granulosis</span> (NgOu53), <span class="html-italic">Nosema bombycis</span> (NosBom), <span class="html-italic">Nosema antheraeae</span> (NosYY), <span class="html-italic">Nosema ceranae</span> (Ncer), <span class="html-italic">Nosema</span> sp. <span class="html-italic">YNPr</span> (NosYNPr), <span class="html-italic">Nosema muscidifuracis</span> (Nmus), and <span class="html-italic">Nosema apis</span> (Napis). (<b>B</b>) The correlation between coding region G-C content (<span class="html-italic">x</span>-axis) and GC3/GC (<span class="html-italic">y</span>-axis). (<b>C</b>) Proportion of codon usage for glutamic acid and tyrosine in eight microsporidian genomes. The proportion of arginine codon usage across eight microsporidian genomes. (<b>D</b>) Proportion of codon usage for alanine and threonine in eight microsporidian genomes (Chi-squared test, ***, <span class="html-italic">p</span> &lt; 0.001). (<b>E</b>) Proportion of codon usage for arginine in eight microsporidian genomes. (<b>F</b>) Proportion of STOP codon usage in eight microsporidian genomes.</p>
Full article ">
16 pages, 835 KiB  
Article
Diversity and Distribution of Australian Stygobiont and Other Groundwater-Associated Amphipods (Crustacea: Malacostraca: Peracarida)
by Rachael A. King, Steven J. B. Cooper, Benjamin Schwartz, Remko Leijs, Danielle N. Stringer, William F. Humphreys, Jake Thornhill and Michelle T. Guzik
Diversity 2024, 16(10), 650; https://doi.org/10.3390/d16100650 - 21 Oct 2024
Viewed by 1006
Abstract
Numerous and diverse groundwater habitats suitable for sustaining aquatic invertebrate communities exist across Australia. These habitats include enclosed subterranean aquifer systems, fractured rock, alluvial aquifers, perched aquifers, artesian springs, and spring-fed seeps and marshes. Crustaceans are a dominant member of these groundwater-associated invertebrate [...] Read more.
Numerous and diverse groundwater habitats suitable for sustaining aquatic invertebrate communities exist across Australia. These habitats include enclosed subterranean aquifer systems, fractured rock, alluvial aquifers, perched aquifers, artesian springs, and spring-fed seeps and marshes. Crustaceans are a dominant member of these groundwater-associated invertebrate communities, and amphipods, both stygobiont and associated epigean species, are particularly diverse yet are still relatively poorly known. We review both the diversity and distributions of Australian amphipods associated with groundwater habitats, describing hotspots of diversity, providing notes on the unique Australian habitats, and examining the extraordinary species diversity and endemism of the Australian species. Our review highlights the significance of Australian groundwater ecosystems, their associated biodiversity, and the importance in considering these ecosystems in groundwater conservation management plans. Full article
(This article belongs to the Special Issue Diversity and Evolution within the Amphipoda)
Show Figures

Figure 1

Figure 1
<p>The distribution of described species of Australian groundwater-associated amphipods (colored by family) (the numbers are linked to species details in <a href="#diversity-16-00650-t001" class="html-table">Table 1</a>). The figure was created with QGIS [<a href="#B31-diversity-16-00650" class="html-bibr">31</a>] using a States and Territories digital boundary file from Australian Bureau of Statistics: Australian Statistical Geography Standard (ASGS) Edition 3.</p>
Full article ">
22 pages, 968 KiB  
Article
Diet of Three Cryptobenthic Clingfish Species and the Factors Influencing It
by Domen Trkov, Danijel Ivajnšič, Marcelo Kovačić and Lovrenc Lipej
Animals 2024, 14(19), 2835; https://doi.org/10.3390/ani14192835 - 1 Oct 2024
Viewed by 820
Abstract
Cryptobenthic fish are small benthic fish species that normally live in various hiding places. Due to their large numbers, they are very important for energy transfer to higher trophic levels. However, due to their small size and hidden lifestyle, knowledge about them and [...] Read more.
Cryptobenthic fish are small benthic fish species that normally live in various hiding places. Due to their large numbers, they are very important for energy transfer to higher trophic levels. However, due to their small size and hidden lifestyle, knowledge about them and their ecology, including their diet, is still limited. Using a non-destructive method based on faecal pellets, we investigated the diet of three clingfish species, Lepadogaster lepadogaster, L. candolii, and Apletodon incognitus, in the shallow northern Adriatic Sea. To better understand the results, we studied the fauna of potential prey in the habitats of the fish studied and also took fish specimens to observe their behaviour in the laboratory. The three species feed predominantly on crustaceans, particularly amphipods, copepods, and decapods. The proportion of the different taxa in the diet depends on the species of clingfish, the size of the specimens, and the size of the prey. In addition, the behaviour of the fish, the home range of the specimens, and the availability of food played an important role. The presence of certain crustacean groups in the environment also determines the occurrence of clingfish of different species and sizes. Full article
Show Figures

Figure 1

Figure 1
<p>Sampling sites (black dots) along the Slovenian coast where clingfish were searched for.</p>
Full article ">Figure 2
<p>Index of the relative importance (%) of different prey groups for three clingfish species.</p>
Full article ">
13 pages, 5475 KiB  
Article
Taxonomic Exploration of Rare Amphipods: A New Genus and Two New Species (Amphipoda, Iphimedioidea, Laphystiopsidae) Described from Seamounts in the Western Pacific
by Yanrong Wang, Zhongli Sha and Xianqiu Ren
Diversity 2024, 16(9), 564; https://doi.org/10.3390/d16090564 - 10 Sep 2024
Viewed by 663
Abstract
During two expeditions to the seamounts in the Yap-Caroline area of the Western Pacific, a new genus, Phoxirostus gen. nov., in the family Laphystiopsidae Stebbing, 1899, is erected for two new species, P. longicarpus sp. nov. (type species) and P. yapensis sp. nov. [...] Read more.
During two expeditions to the seamounts in the Yap-Caroline area of the Western Pacific, a new genus, Phoxirostus gen. nov., in the family Laphystiopsidae Stebbing, 1899, is erected for two new species, P. longicarpus sp. nov. (type species) and P. yapensis sp. nov. The new genus can be distinguished from the other three laphystiopsid genera by the acute rostrum not overreaching the distal end of the first peduncular article of antenna 1, the outer plate of maxilla 1 bearing 10–11 spines, and the elongated carpus of pereopods 3–7 being distinctly longer than half the length of the propodus. Phoxirostus longicarpus sp. nov. differs from P. yapensis sp. nov. by the shape of the eyes and coxa 4, the presence of posterodistal protrusions on pleonite 1, and the number of posterodistal protrusions on pleonite 2. Generic analysis of one mitochondrial (COI) and one nuclear (H3) gene using maximum likelihood and Bayesian inference clarified the phylogenetic position of the Laphystiopsidae within the superfamily Iphimedioidea Boeck, 1871. Full article
(This article belongs to the Special Issue Diversity and Evolution within the Amphipoda)
Show Figures

Figure 1

Figure 1
<p><span class="html-italic">Phoxirostus longicarpus</span> <b>sp. nov.</b>, MBM 286818, holotype, female (6.0 mm): showing that it is associated with the sponge and photographed after being fixed in 95% ethanol.</p>
Full article ">Figure 2
<p><span class="html-italic">Phoxirostus longicarpus</span> <b>sp. nov.</b>, MBM 286818, holotype, female (6.0 mm): A1, antenna 1; A2, antenna 2; G1 L, left gnathopod 1; G2 L, left gnathopod 2; H, head, U1 R, right uropod1; U2 L, left uropod 2; U3 R, right uropod 3, and the arrow points to the ventral view of inner ramus; T, telson.</p>
Full article ">Figure 3
<p><span class="html-italic">Phoxirostus longicarpus</span> <b>sp. nov.</b>, MBM 286818, holotype, female (6.0 mm): UL, upper lip; LL, lower lip; Md L, left mandible; Mx1, maxilla 1; Mx2, maxilla 2; Mxp, maxilliped.</p>
Full article ">Figure 4
<p><span class="html-italic">Phoxirostus longicarpus</span> <b>sp. nov.</b>, MBM 286818, holotype, female (6.0 mm): P3 L, left pereopod 3; P4 L, left pereopod 4; P5 L, left pereopod 5; P6 R, right pereopod 6; P7 R, right pereopod 7.</p>
Full article ">Figure 5
<p><span class="html-italic">Phoxirostus yapensis</span> <b>sp. nov.</b>, MBM 286617, holotype, female (7.1 mm): photographed immediately after being collected by Wei Jiang.</p>
Full article ">Figure 6
<p><span class="html-italic">Phoxirostus yapensis</span> <b>sp. nov.</b>, MBM 286617, holotype, female (7.1 mm): UL, upper lip; LL, lower lip; Md L, left mandible, and the arrow points to details of two distal articles of palp; Md R, only shows the incisor and accessory spines; Mx1 R, right maxilla 1; Mx2, maxilla 2; Mxp, maxilliped; A1, antenna 1; A2, antenna 2.</p>
Full article ">Figure 7
<p><span class="html-italic">Phoxirostus yapensis</span> <b>sp. nov.</b>, MBM 286617, holotype, female (7.1 mm): G1 R, right gnathopod 1; G2 R, right gnathopod 2; P4 R, right pereopod 4; P5 R, right pereopod 5; P6 R, right pereopod 6; P7 R, right pereopod 7; H, head, arrow points acute rostrum; T, telson.</p>
Full article ">Figure 8
<p><span class="html-italic">Phoxirostus yapensis</span> <b>sp. nov.</b>, MBM 286617, paratype, male (5.3 mm): G1 R, right gnathopod 1; G2 R, right gnathopod 2; P3 L, left pereopod 3; P4 L, left pereopod 4; P5 L, left pereopod 5; P6 L, left pereopod 6; P7 L, left pereopod 7; E1–3, epimeron plates 1–3.</p>
Full article ">Figure 9
<p>Phylogenetic tree of the superfamily Iphimedioidea Boeck, 1871, taxa resolved based on the combined dataset of four genes (COI and H3): (<b>A</b>) Bayesian inference (BI) tree; (<b>B</b>) maximum likelihood tree.</p>
Full article ">Figure 10
<p>The distribution of laphystiopsid species: the location of the sampling site of two new species (red and yellow rhombus); distribution of <span class="html-italic">Laphystiopsis</span> species (green, pink, red, yellow, and blue square); distribution of <span class="html-italic">Prolaphystius</span> species (pink triangle) and Prolaphystiopsis species (red and yellow circle).</p>
Full article ">
9 pages, 3799 KiB  
Communication
The Assessment of Methyl Methanesulfonate Absorption by Amphipods from the Environment Using Lux-Biosensors
by Uliana S. Novoyatlova, Anna A. Kudryavtseva, Sergey V. Bazhenov, Anna A. Utkina, Vadim V. Fomin, Shamil A. Nevmyanov, Bagila S. Zhoshibekova, Maria A. Fedyaeva, Mikhail Y. Kolobov and Ilya V. Manukhov
Biosensors 2024, 14(9), 427; https://doi.org/10.3390/bios14090427 - 5 Sep 2024
Viewed by 1223
Abstract
The ability of aquatic mesofauna representatives involved in trophic chains to sorb and accumulate toxicants is important for understanding the functioning of aquatic ecosystems and for fishing industry. This study investigated the capacity of marine amphipod Gammarus oceanicus and freshwater amphipods Eulimnogammarus vittatus [...] Read more.
The ability of aquatic mesofauna representatives involved in trophic chains to sorb and accumulate toxicants is important for understanding the functioning of aquatic ecosystems and for fishing industry. This study investigated the capacity of marine amphipod Gammarus oceanicus and freshwater amphipods Eulimnogammarus vittatus and Gammarus lacustris to absorb the DNA-alkylating agent methyl methanesulfonate (MMS). The presence of alkylating agents in the environment and in the tissues of the amphipods was determined using whole-cell lux-biosensor Escherichia coli MG1655 pAlkA-lux, in which the luxCDABE genes from Photorhabdus luminescens, enabling the luminescence of the cell culture, are controlled by the PalkA promoter of DNA glycosylase. It was shown that within one day of incubation in water containing MMS at a concentration above 10 μM, the amphipods absorbed the toxicant and their tissues produce more alkylation damage to biosensor cells than the surrounding water. Concentrations of MMS above 1 mM in the environment caused the death of the amphipods before the toxicant could be significantly concentrated in their tissues. The sensitivity and the capacity to absorb MMS were found to be approximately the same for the marine amphipod G. oceanicus and the freshwater amphipods E. vittatus and G. lacustris. Full article
(This article belongs to the Section Environmental Biosensors and Biosensing)
Show Figures

Figure 1

Figure 1
<p>Scheme of the experiment. Amphipods are divided into groups and incubated in water with various concentrations of MMS or without it. After 24 h, the incubation alkylation ability of amphipod’s tissues and water from the flasks/jars was tested using the whole-cell lux-biosensors.</p>
Full article ">Figure 2
<p>(<b>A</b>–<b>C</b>)—Luminescent signal of <span class="html-italic">E. coli</span> MG1655 pAlkA-lux biosensor upon the addition (1/10 V) of water and the liquid fraction of homogenized amphipods incubated in this water. Curve “water 100”—MMS-supplemented (100 µM) water sampled after incubation of amphipods in it. “gam 100”—amphipods after incubation in MMS-supplemented (100 µM) water; “gam 0”—amphipods after incubation in MMS-free water, “K-”—MMS-free water sampled after incubation of amphipods (negative control). As test objects, the following amphipods were used: (<b>A</b>)—<span class="html-italic">G. lacustris</span>, (<b>B</b>)—<span class="html-italic">E. vittatus</span> (<b>C</b>)—<span class="html-italic">G. oceanicus</span>, <span class="html-italic">s</span>. (<b>D</b>)—Luminescent signal of <span class="html-italic">E. coli</span> MG1655 pAlkA-lux biosensor upon the addition (1/10 V) of water with MMS added in various concentrations (calibration samples).</p>
Full article ">Figure 3
<p>Comparison of the induction coefficient of <span class="html-italic">E. coli</span> MG1655 pAlkA-lux after 5 h of incubation following the addition of the liquid fraction of homogenized amphipods, incubated in MMS-containing water, or the water, in which amphipods were incubated. Abscissa values correspond to MMS concentration in water in the beginning of the experiment. All the values from three replicates of experiments from three gammarid species are combined. The dashed line show the value of 1 for induction coefficient (the value indicating the absence of biosensor induction).</p>
Full article ">
16 pages, 3279 KiB  
Article
Short-Term Chronic Toxicity of Copper to Hyalella azteca: Contrast in Terms of Equilibrating Diet, Diet Type, and Organic Matter Source
by Nafis Fuad, Rebecca Williams and Timothy M. Vadas
Toxics 2024, 12(8), 608; https://doi.org/10.3390/toxics12080608 - 20 Aug 2024
Viewed by 1074
Abstract
The most up-to-date regulatory guidelines for establishing acute and chronic numeric limits for copper in freshwaters are based on a biotic ligand model for various species, but the model for Cu lacks data on dietary uptake. In addition, some common macroinvertebrate toxicity assay [...] Read more.
The most up-to-date regulatory guidelines for establishing acute and chronic numeric limits for copper in freshwaters are based on a biotic ligand model for various species, but the model for Cu lacks data on dietary uptake. In addition, some common macroinvertebrate toxicity assay parameters are less representative of the ecosystem. We investigated the effects of diet and its type in the experimental setup and as an exposure pathway to an established amphipod (crustacean) Hyalella azteca (H. azteca) for Cu toxicity assays. We also investigated another overlooked aspect, the organic matter (OM) source. Our experiments compared the toxicity of pre-equilibrated and unequilibrated natural diets and a laboratory-favored diet in effluent and stormwater sources of organic matter adjusted to standard water characteristics. The experiments indicated a more toxic effect of the pre-equilibrated diet and natural dietary sources, and less toxic effects in the presence of effluent OM compared with stormwater OM, shifting LC50 or EC20 values by as much as 67% compared with the controls. The use of a pre-equilibrated natural diet in toxicity assays provides the advantage of producing toxicity data more representative of field conditions. Considering organic matter type, especially in dietary exposures, will better predict toxicity, accounting for copper complexation with OM from different sources and partitioning to the food supply. Adapting these ecologically relevant parameters in whole effluent toxicity testing or other assays will also provide safer regulatory oversite of discharges to surface waters. Full article
(This article belongs to the Section Ecotoxicology)
Show Figures

Figure 1

Figure 1
<p>Measured vs. nominal copper after the 14-day exposure experiment. Squares, diamonds, crosses, triangles, pluses, filled circles, and hollow circles represent 2.5 ppm stormwater, 4 ppm, stormwater, 4 ppm effluent, 2.5 ppm effluent periphyton pre-equilibrated, 2.5 ppm effluent periphyton unequilibrated, 2.5 ppm effluent TetraMin<sup>®</sup> pre-equilibrated, and 2.5 ppm effluent TetraMin<sup>®</sup> unequilibrated, respectively. The brown line represents the 1:1 line.</p>
Full article ">Figure 2
<p>Relationships between measured copper concentration (μg L<sup>−1</sup>) and % survival following 14-day exposures: (<b>a</b>) effluent TetraMin<sup>®</sup> unequilibrated; (<b>b</b>) effluent TetraMin<sup>®</sup> pre-equilibrated; (<b>c</b>) effluent periphyton unequilibrated; and (<b>d</b>) effluent periphyton pre-equilibrated. The shaded region represents the confidence band. All exposures contained 2.5 ppm organic carbon.</p>
Full article ">Figure 3
<p>Relationships between measured copper concentration (μg L<sup>−1</sup>) and length (mm) following 14-day exposures: (<b>a</b>) effluent TetraMin<sup>®</sup> unequilibrated; (<b>b</b>) effluent TetraMin<sup>®</sup> pre-equilibrated; (<b>c</b>) effluent periphyton unequilibrated; and (<b>d</b>) effluent periphyton pre-equilibrated. The shaded region represents the confidence band. All exposures contained 2.5 ppm organic carbon.</p>
Full article ">Figure 4
<p>Relationships between measured copper concentration (μg L<sup>−1</sup>) and % survival following 14-day exposures: (<b>a</b>) stormwater periphyton pre-equilibrated; (<b>b</b>) effluent periphyton pre-equilibrated; and (<b>c</b>) stormwater periphyton pre-equilibrated. The shaded region represents the confidence band. (<b>a</b>) contained 2.5 ppm organic carbon and (<b>b</b>,<b>c</b>) contained 4 ppm organic carbon.</p>
Full article ">Figure 5
<p>Relationships between measured copper concentration (μg L<sup>−1</sup>) and length (mm) following 14-day exposures: (<b>a</b>) stormwater periphyton pre-equilibrated; (<b>b</b>) effluent periphyton pre-equilibrated; and (<b>c</b>) stormwater periphyton pre-equilibrated. The shaded region represents the confidence band. (<b>a</b>) contained 2.5 ppm organic carbon and (<b>b</b>,<b>c</b>) contained 4 ppm organic carbon.</p>
Full article ">
12 pages, 1225 KiB  
Article
Macroinvertebrates Associated with Macroalgae within Integrated Multi-Trophic Aquaculture (IMTA) in Earthen Ponds: Potential for Accessory Production
by Rafael Vieira, Miguel Ângelo Mateus, Carlos Manuel Lourenço Afonso, Florbela Soares, Pedro Pousão-Ferreira and Sofia Gamito
J. Mar. Sci. Eng. 2024, 12(8), 1369; https://doi.org/10.3390/jmse12081369 - 11 Aug 2024
Viewed by 1151
Abstract
The present work aims to evaluate the macroinvertebrate community associated with macroalgae in earthen pond systems to better understand their potential in detritus recycling and as an accessory production. Sampling took place on the settling pond of an aquaculture research station, where macroalgae [...] Read more.
The present work aims to evaluate the macroinvertebrate community associated with macroalgae in earthen pond systems to better understand their potential in detritus recycling and as an accessory production. Sampling took place on the settling pond of an aquaculture research station, where macroalgae permanently occurred at high densities. The results suggest differentiation between seasons but not between sites within the settling pond. Seasonal variation was observable in terms of macroinvertebrate density, biomass, and diversity. Two non-indigenous species of invertebrates were found, the crustaceans Grandidierella japonica and Paracerceis sculpta Amphipods were the most abundant group, and their high nutritional value can be exploited. Detritus and the epiphyte layer are the main food items for the invertebrates, reinforcing the advantages of these organisms being present to enhance the recycling of excess detritus and to transfer organic matter to upper trophic levels. These species, naturally present in aquaculture facilities, can improve the water quality and increase the variability of food nutrients for reared species. Full article
(This article belongs to the Topic Aquatic Environment Research for Sustainable Development)
Show Figures

Figure 1

Figure 1
<p>Seasonal variation plots with standard error for (<b>a</b>) dried algae biomass by cubic metre of water; (<b>b</b>) invertebrates’ biomass by cubic metre of water; (<b>c</b>) invertebrates’ density by gram of algae dry weight; (<b>d</b>) invertebrates’ density by cubic metre of water.</p>
Full article ">Figure 2
<p>Scatter plots of taxa richness (<b>A</b>) and number of individuals (<b>B</b>) per algae dry weight (DW) based on season. A: Autumn (Y = 0.1608x + 5.030); Winter (Y = 0.3252x + 4.921); Spring (Y = 0.2446x + 12.96). B: Autumn (Y = 6.681x + 196.0); Winter (Y = 0.7874x + 582.9); Spring (Y = 166.2x − 565.5).</p>
Full article ">Figure 3
<p>Variation, per season, in the six most abundant species (<span class="html-italic">Monocorophium insidiosum</span>, <span class="html-italic">Hydrobia glyca</span>, <span class="html-italic">Paracerceis sculpta</span>, <span class="html-italic">Grandidierella japonica</span>, <span class="html-italic">Cymadusa filose</span> and <span class="html-italic">Microdeutopus gryllotalpa</span>). (<b>a</b>) Density per gram of dry algae and (<b>b</b>) density per cubic metre.</p>
Full article ">Figure 4
<p>Multidimensional scaling (MDS) carried out with macroinvertebrates associated with a macroalgae abundance dataset using Bray–Curtis similarity index and root transformation. Four samples were collected for each season: Autumn (Aut1–4), Winter (Win1–4), and Spring (Spr1–4).</p>
Full article ">Figure 5
<p>(<b>a</b>)Variation per season in the main taxonomic groups; (<b>b</b>) density of the most abundant benthic macroinvertebrates in Autumn sampling (<span class="html-italic">Capitella</span> sp., <span class="html-italic">Neanthes acuminata</span>, <span class="html-italic">Monocorophium insidiosum</span>, <span class="html-italic">Gammarella fucicola</span>, <span class="html-italic">Grandidierella japonica</span>, <span class="html-italic">Hydrobia glyca,</span> and <span class="html-italic">Peringia ulvae</span>).</p>
Full article ">
13 pages, 1396 KiB  
Communication
Detection of Pharmaceutical Contamination in Amphipods of Lake Baikal by the HPLC-MS Method
by Tamara Y. Telnova, Maria M. Morgunova, Sophie S. Shashkina, Anfisa A. Vlasova, Maria E. Dmitrieva, Victoria N. Shelkovnikova, Ekaterina V. Malygina, Natalia A. Imidoeva, Alexander Y. Belyshenko, Alexander S. Konovalov, Evgenia A. Misharina and Denis V. Axenov-Gribanov
Antibiotics 2024, 13(8), 738; https://doi.org/10.3390/antibiotics13080738 - 6 Aug 2024
Cited by 1 | Viewed by 1099
Abstract
Pollution by active ingredients is one of the most significant and widespread forms of pollution on Earth. Medicines can have a negative impact on ecosystems, and contamination can have unpredictable consequences. An urgent and unexplored task is to study the Lake Baikal ecosystem [...] Read more.
Pollution by active ingredients is one of the most significant and widespread forms of pollution on Earth. Medicines can have a negative impact on ecosystems, and contamination can have unpredictable consequences. An urgent and unexplored task is to study the Lake Baikal ecosystem and its organisms for the presence of trace concentrations of active pharmaceutical ingredients. Our study aimed to conduct a qualitative analysis of active pharmaceutical ingredients, and quantitative analysis of ibuprofen in endemic amphipods of Lake Baikal, using methods of high-performance liquid chromatography and mass spectrometry (HPLC-MS). Acetylsalicylic acid (aspirin), ibuprofen, acetaminophen, azithromycin, dimetridazole, metronidazole, amikacin, spiramycin, and some tetracycline antibiotics were detected in the studied littoral amphipods. We also detected different annual loads of active pharmaceutical ingredients on amphipods. Using the multiple reaction monitoring (MRM) mode mentioned in GOST International Technical Standards, we detected molecules, fragmented as amikacin, chlortetracycline, doxycycline, oxytetracycline, dimetridazole, metronidazole and spiramycin. Thus, we first revealed that invertebrates of Lake Baikal can uptake pharmaceutical contaminants in the environment. Full article
Show Figures

Graphical abstract

Graphical abstract
Full article ">Figure 1
<p>Chromatogram of metronidazole detected in samples of amphipod <span class="html-italic">E. verrucosus</span> (2020).</p>
Full article ">Figure 2
<p>Chromatogram of ibuprofen detected in samples of amphipod <span class="html-italic">E. verrucosus</span>.</p>
Full article ">Figure 3
<p>Correlation between concentration of ibuprofen and the wet weight of species <span class="html-italic">E. verrucosus</span> amphipods, ng/g (white marker—2020; black marker—2022).</p>
Full article ">
15 pages, 7602 KiB  
Article
Development of Single-Nucleotide Polymorphism Markers and Population Genetic Analysis of the Hadal Amphipod Alicella gigantea across the Mariana and New Britain Trenches
by Lei Chen, Shouwen Jiang, Binbin Pan and Qianghua Xu
J. Mar. Sci. Eng. 2024, 12(7), 1117; https://doi.org/10.3390/jmse12071117 - 3 Jul 2024
Viewed by 976
Abstract
Alicella gigantea, the largest amphipod scavengers found to date, play key roles in the food web of the hadal ecosystem. However, the genetic structure of A. gigantea populations among different trenches has not been reported yet. In this study, SNP (single-nucleotide polymorphism) [...] Read more.
Alicella gigantea, the largest amphipod scavengers found to date, play key roles in the food web of the hadal ecosystem. However, the genetic structure of A. gigantea populations among different trenches has not been reported yet. In this study, SNP (single-nucleotide polymorphism) markers were developed for three A. gigantea geographic populations collected from the southern Mariana Trench (SMT), the central New Britain Trench (CNBT), and the eastern New Britain Trench (ENBT), based on the SLAF-seq (specific locus amplified fragment sequencing) technology. A total of 570,168 filtered SNPs were screened out for subsequent population genetic analysis. Results showed that the inbreeding levels across the three geographic populations were relatively low, and the genomic inbreeding coefficients of the three populations were similar in magnitude. Based on the results of phylogenetic analysis, population structure analysis, and principal component analysis, it is believed that the three A. gigantea geographic populations belong to the same population, and the kinship relationship between the ENBT and CNBT populations is close. Moreover, the differential candidate adaptive sites on the SNPs suggest that there may be variations in metabolic rates among the three geographic populations, possibly linked to differences in food availability and sources in different trenches, ultimately resulting in different survival strategies in A. gigantea populations within distinct trenches. Compared with the Mariana Trench, the New Britain Trench has a richer organic matter input, and it is speculated that the A. gigantea Mariana Trench population may adopt a lower metabolic rate to cope with the harsher environment of nutrient deficiency. Full article
Show Figures

Figure 1

Figure 1
<p>Sampling locations of the three <span class="html-italic">A. gigantea</span> geographic populations.</p>
Full article ">Figure 2
<p>Genomic inbreeding coefficients and phylogenetic analysis regarding the 30 individuals of <span class="html-italic">A. gigantea</span>. (<b>a</b>) Genomic inbreeding coefficients of the 30 sequenced individuals and average genomic inbreeding coefficients of three populations. (<b>b</b>) Neighbor-joining phylogenetic tree for 30 sequenced individuals based on 570,168 SNPs.</p>
Full article ">Figure 3
<p>Population structure analysis and principal component analysis. (<b>a</b>) Plots of the individual ancestry inference for K = 1 (upper), K = 2 (middle), and K = 3 (lower). (<b>b</b>) Cross-validation error of <span class="html-italic">A. gigantea</span> K = 1–5. (<b>c</b>) Principal component analysis of <span class="html-italic">A. gigantea</span>.</p>
Full article ">Figure 4
<p>Bayesian factor numerical diagram of SNP sites. A blue dashed line represents log<sub>10</sub>(BF) = 0.5 and a red dashed line represents log<sub>10</sub>(BF) = 1.</p>
Full article ">Figure 5
<p>Functional annotation of environmental adaptation loci. (<b>a</b>) KOG function classification of selected genes. (<b>b</b>) KEGG function classification of environmental adaptation loci. (<b>c</b>) GO function annotation of environmental adaptation loci.</p>
Full article ">Figure 6
<p>Two SLAF tags annotated by all three databases. (<b>a</b>) Venn diagram of three-database-annotated tags. (<b>b</b>) Sequence logo diagram of these two genes. The red box in the figure indicates the environmental adaptation site. One gene’s sequence number is Marker63205, and the selected site is at the 164th base (<b>upper</b>), and the other gene’s sequence number is Marker137036, and the selected site is at the 103rd base (<b>down</b>).</p>
Full article ">
16 pages, 1934 KiB  
Review
Phenotype Variation in Niphargus (Amphipoda: Niphargidae): Possible Explanations and Open Challenges
by Cene Fišer and Ester Premate
Diversity 2024, 16(7), 375; https://doi.org/10.3390/d16070375 - 28 Jun 2024
Viewed by 1008
Abstract
Understanding phenotype variation is among the central topics in biology. We revise and reanalyze studies of the amphipod genus Niphargus to confront two potential mechanisms driving its phenotype variation, namely, cladogenesis and adaptive evolution. We found evidence for both mechanisms. Reanalysis of a [...] Read more.
Understanding phenotype variation is among the central topics in biology. We revise and reanalyze studies of the amphipod genus Niphargus to confront two potential mechanisms driving its phenotype variation, namely, cladogenesis and adaptive evolution. We found evidence for both mechanisms. Reanalysis of a subset of traits using molecular phylogeny showed moderate phylogenetic signal, consistent with the hypothesis that overall phylogenetic variation increases with phylogeny. The phylogenetic signal in Niphargus traits seems to be stronger at the tips of the phylogeny than at basal splits. Indirect evidence suggests that much of the phenotype variation can be attributed to adaptive evolution. Both lines of evidence are consistent with the hypothesis that Niphargus evolved in several adaptive radiations, where theory predicts that most of the phenotype variation evolves early, when ecological niches are vacant. As the niches fill up, the rate of phenotype variation slows down and becomes associated with cladogenetic events. This hypothesis can explain the high level of trait-convergence and unresolved taxonomy above the species level. The main caveats to these hypotheses comprise lack of experimental evidence for trait function and nonquantified heritable component of trait variation. Promising venues towards better understanding of phenotypic variation include studies of ontogenetic variation, functional interactions between traits, and genome–phenotype associations. Full article
(This article belongs to the Special Issue Diversity and Evolution within the Amphipoda)
Show Figures

Figure 1

Figure 1
<p>Phenotypic variability of <span class="html-italic">Niphargus</span> in the Western Balkans. (<b>a</b>) <span class="html-italic">N. pachytelson</span>, (<b>b</b>) <span class="html-italic">N. croaticus</span>, (<b>c</b>) <span class="html-italic">N.</span> sp. n., (<b>d</b>) <span class="html-italic">N. stygius</span>, (<b>e</b>) <span class="html-italic">N. castellanus</span>, (<b>f</b>) <span class="html-italic">N. subtypicus</span>, (<b>g</b>) <span class="html-italic">N.</span> sp. n. (<b>a</b>,<b>b</b>) and (<b>f</b>) are species living in cave lakes, (<b>d</b>,<b>e</b>) are species living in cave streams, and (<b>c</b>,<b>g</b>) are species from river interstitial. The scale is approximate. Photos by Teo Delić, SubBio Lab.</p>
Full article ">Figure 2
<p>Latest <span class="html-italic">Niphargus</span> molecular phylogeny comprising 562 molecular operational taxonomic units (MOTUs) and reconstructed using Bayesian inference [<a href="#B69-diversity-16-00375" class="html-bibr">69</a>]. The phylogeny was used in calculations of phylogenetic signal (<a href="#diversity-16-00375-t001" class="html-table">Table 1</a>). Tips colored according to MOTU ecology (Herrera-Alsina et al., submitted).</p>
Full article ">Figure 3
<p>(<b>a</b>) <span class="html-italic">Niphargus</span> functional model (left, modified from [<a href="#B72-diversity-16-00375" class="html-bibr">72</a>]) and detailed schematic representation of the ventral channel determining overall body shape (right, modified from [<a href="#B72-diversity-16-00375" class="html-bibr">72</a>]). The colored dots correspond to functional traits listed in the table below (<b>b</b>). (<b>b</b>) An overview of functional traits according to their recognized inter- and intraspecific variability and underlying causes. (<b>c</b>) A summary of future possible directions in the studies of <span class="html-italic">Niphargus</span> inter- and intraspecific variability.</p>
Full article ">
17 pages, 4812 KiB  
Article
Rediscovering the Evasive Amphipod Idunella spinifera (Dauvin and Gentil, 1983) in the Northwest Coast of the Iberian Peninsula
by Juan Moreira, Puri Veiga and Marcos Rubal
J. Mar. Sci. Eng. 2024, 12(7), 1043; https://doi.org/10.3390/jmse12071043 - 21 Jun 2024
Viewed by 664
Abstract
Idunella spinifera (Dauvin and Gentil, 1983) (Crustacea: Amphipoda: Liljeborgiidae) is reported for the first time after the original description that was based on one immature female. Specimens were collected in the Ría de Muros (NW Iberian Peninsula) during the course of a study [...] Read more.
Idunella spinifera (Dauvin and Gentil, 1983) (Crustacea: Amphipoda: Liljeborgiidae) is reported for the first time after the original description that was based on one immature female. Specimens were collected in the Ría de Muros (NW Iberian Peninsula) during the course of a study on macrofauna diversity in shallow sublittoral biogenic sands along a one-year period. The male and the ovigerous female are fully described, and the data on ecology and temporal variation of abundance is provided as well. Full article
(This article belongs to the Special Issue Taxonomy, Biodiversity, and Distribution of Marine Invertebrates)
Show Figures

Figure 1

Figure 1
<p>Mean abundance (+SE; n = 13) of <span class="html-italic">I. spinifera</span> (<b>A</b>) per month; % of males, females, and ovigerous females at each month (<b>B</b>); size distribution (total body length, mm) of males, females, and ovigerous females (<b>C</b>).</p>
Full article ">Figure 2
<p><span class="html-italic">Idunella spinifera</span> (Dauvin and Gentil, 1983). Male, Ría de Muros (NW Spain), lateral view.</p>
Full article ">Figure 3
<p><span class="html-italic">Idunella spinifera</span> (Dauvin and Gentil, 1983). Male, Ría de Muros (NW Spain). (<b>A</b>) head and anterior thoracic segments, lateral view. (<b>B</b>) antenna 1. (<b>C</b>) antenna 2. (<b>D</b>) left mandible (palp upturned) and detail of rows of D3 (inferior) and B3-setae (superior; only some clusters of setae illustrated). (<b>E</b>) right mandible, distomedial part. (<b>F</b>) right maxilla 1. (<b>G</b>) left maxilla 2. (<b>B</b>,<b>C</b>,<b>F</b>,<b>G</b>) same scale.</p>
Full article ">Figure 4
<p><span class="html-italic">Idunella spinifera</span> (Dauvin and Gentil, 1983). Male, Ría de Muros (NW Spain). (<b>A</b>) right maxilliped (palp upturned). (<b>B</b>) right maxilliped, proximal part (opposite face). (<b>C</b>) left maxilliped (palp upturned; opposite face). (<b>D</b>) gnathopod 1. (<b>E</b>) gnathopod 1 propodus, palm inner face. (<b>F</b>) gnathopod 2. (<b>A</b>–<b>C</b>) same scale. (<b>C</b>) palp setae not illustrated.</p>
Full article ">Figure 5
<p><span class="html-italic">Idunella spinifera</span> (Dauvin and Gentil, 1983). Male, Ría de Muros (NW Spain). Pereopods and coxae. (<b>A</b>) right pereopod 3, inner face. (<b>B</b>) right pereopod 4, inner face. (<b>C</b>) left pereopod 5, outer face. (<b>D</b>) right pereopod 5, proximal part, inner face. (<b>E</b>) right pereopod 6, outer face. (<b>F</b>) right pereopod 7, outer face. (<b>A</b>–<b>F</b>) same scale bar. (<b>E</b>,<b>F</b>) broken at the carpus-merus level.</p>
Full article ">Figure 6
<p><span class="html-italic">Idunella spinifera</span> (Dauvin and Gentil, 1983). Male, Ría de Muros (NW Spain). (<b>A</b>) pleonites, right side, lateral view. (<b>B</b>) urosomites, uropods, and telson, right side, lateral view. (<b>C</b>) right uropod 1, laterodorsal view. (<b>D</b>) right uropod 2, laterodorsal view. (<b>E</b>) right uropod 3, lateral view. (<b>F</b>) telson, dorsal view. (<b>C</b>–<b>E</b>) same scale. (<b>B</b>) only some setae illustrated.</p>
Full article ">Figure 7
<p><span class="html-italic">Idunella spinifera</span> (Dauvin and Gentil, 1983). Ovigerous female, Ría de Muros (NW Spain). (<b>A</b>) antenna 1. (<b>B</b>) antenna 2. (<b>C</b>) right mandible. (<b>D</b>) right mandible, distomedial part, inner face. (<b>E</b>) right maxilla 1. (<b>F</b>) right maxilla 2. (<b>E</b>,<b>F</b>) same scale. (<b>A</b>,<b>B</b>,<b>E</b>,<b>F</b>) only some setae illustrated.</p>
Full article ">Figure 8
<p><span class="html-italic">Idunella spinifera</span> (Dauvin and Gentil, 1983). Ovigerous female, Ría de Muros (NW Spain). (<b>A</b>) left maxilliped. (<b>B</b>) right maxilliped. (<b>C</b>) gnathopod 1. (<b>D</b>) gnathopod 2 coxa, gill, and oostegite, inner face. (<b>E</b>) gnathopod 2. (<b>B</b>) palp setae, not illustrated. (<b>E</b>) only some setae illustrated. (<b>A</b>,<b>B</b>,<b>C</b>,<b>E</b>) same scale.</p>
Full article ">Figure 9
<p><span class="html-italic">Idunella spinifera</span> (Dauvin and Gentil, 1983). Ovigerous female, Ría de Muros (NW Spain). Pereopods and coxae. (<b>A</b>) right pereopod 3, inner face. (<b>B</b>) right pereopod 4, inner face. (<b>C</b>) left pereopod 5, inner face. (<b>D</b>) right pereopod 6, outer face. (<b>E</b>) right pereopod 7, outer face. (<b>A</b>–<b>E</b>) same scale bar. (<b>D</b>,<b>E</b>) broken at the carpus-merus level.</p>
Full article ">Figure 10
<p><span class="html-italic">Idunella spinifera</span> (Dauvin and Gentil, 1983). Ovigerous female (<b>A</b>–<b>E</b>), non-ovigerous females (<b>F</b>,<b>G</b>), Ría de Muros (NW Spain). (<b>A</b>) pleonites, left side, lateral view. (<b>B</b>) right uropod 1, laterodorsal view. (<b>C</b>) right uropod 2, laterodorsal view. (<b>D</b>) right uropod 3, lateral view. (<b>E</b>–<b>G</b>) telson, dorsal view. (<b>B</b>–<b>D</b>,<b>F</b>,<b>G</b>) same scale. (<b>F</b>,<b>G</b>) dorsal and terminal setae, with setules not illustrated.</p>
Full article ">
29 pages, 3927 KiB  
Article
Baseline Inventory of Benthic Macrofauna in German Marine Protected Areas (2020–2022) before Closure for Bottom-Contact Fishing
by Mayya Gogina, Sarah Joy Hahn, Ramona Ohde, Angelika Brandt, Stefan Forster, Ingrid Kröncke, Martin Powilleit, Katharina Romoth, Moritz Sonnewald and Michael L. Zettler
Biology 2024, 13(6), 389; https://doi.org/10.3390/biology13060389 - 28 May 2024
Viewed by 1663
Abstract
The response of benthic habitats and organisms to bottom-contact fishing intensity is investigated in marine protected areas (MPAs) of the German EEZ in the North and Baltic Seas. We examined the current state of macrofauna biodiversity in 2020–2022. Comparative analysis for macrofauna (in- [...] Read more.
The response of benthic habitats and organisms to bottom-contact fishing intensity is investigated in marine protected areas (MPAs) of the German EEZ in the North and Baltic Seas. We examined the current state of macrofauna biodiversity in 2020–2022. Comparative analysis for macrofauna (in- and epifauna) inhabiting nine Natura 2000 MPAs constitutes a baseline to assess the effects of bottom-contact fishing exclusion in the future. Aspects of spatial and temporal variability are briefly summarized and discussed. We provide a species list for each region, including 481 taxa, of which 79 were found in both regions, 183 only in the North Sea, and 219 only in the Baltic Sea. The Baltic Sea dataset surprisingly included higher numbers of taxa and revealed more Red List species. The share of major taxonomic groups (polychaetes, bivalves and amphipods) in species richness showed peculiar commonalities between the two regions. In the North Sea, multivariate analysis of community structure revealed significantly higher within-similarity and stronger separation between the considered MPAs compared to the Baltic MPAs. Salinity, temperature and sediment fractions of sand were responsible for over 60% of the variation in the North Sea macrofauna occurrence data. Salinity, mud fraction and bottom-contact fishing were the most important drivers in the Baltic Sea and, together with other considered environmental drivers, were responsible for 53% of the variation. This study identifies aspects of macrofauna occurrence that may be used to assess (causes of) future changes. Full article
Show Figures

Figure 1

Figure 1
<p>Maps of Natura 2000 sites (green polygons) and the MGF focus areas (thick red line boxes) in (<b>A</b>) the North Sea and (<b>B</b>) the Baltic Sea. The small grayscale inlet (inserted in (<b>A</b>)) shows a general view of the North and Baltic Seas. The thin red line marks the boundaries of the German Exclusive Economic Zone (EEZ). Black dots show the sampled stations. Focus areas in the Baltic Sea are zoomed in on the three small inlet maps. Dots inside the focus areas are stations sampled within the MGF Baltic Sea project, whereas other stations were mostly visited within the LEGRA and ATLAS projects. The half-transparent red line outlines the initial focus area in Rønne Bank, later shifted due to proximity to wind farms that inhibited later sampling. Intense green background outlines the future OB closure area if it will only take place in part of the MPA.</p>
Full article ">Figure 2
<p>Overview of the mobile bottom-contact fishing intensity, (<b>A</b>) subsurface swept area ratio (&gt;2 cm, subsurSwAR) in 2020 in the North Sea EEZ, based on ICES [<a href="#B62-biology-13-00389" class="html-bibr">62</a>,<a href="#B63-biology-13-00389" class="html-bibr">63</a>], and (<b>B</b>) subsurSwAR in 2020 in each 0.05° × 0.05°-degree c-square from ICES [<a href="#B64-biology-13-00389" class="html-bibr">64</a>] data in the Baltic Sea EEZ; red = high mobile bottom-contact fishing intensity; blue = non or low mobile bottom-contact fishing intensity.</p>
Full article ">Figure 3
<p>Number and (after semicolon) percentage of taxa found per group in the MPAs in the North Sea (<b>upper pane</b>) and the Baltic Sea (<b>lower pane</b>). The groups used here in order to facilitate the summary should be rather considered as functional, i.e., not strictly taxonomic, as they vary in rank ranging from Phylum to Order level. In the North Sea MPAs (upper pane), the category “other” includes Isopoda (4), Cirripedia (3 taxa), Nemertea (2), Sipuncula (2) and single taxa of Ascidiacea, Leptocardii, Oligochaeta, Phoronida, Platyhelminthes, Priapulida, Pycnogonida and Tanaidacea. In the Baltic Sea MPAs (lower pane), the category “other” includes Oligochaeta (6), Isopoda (5), Mysida (5), Nemertea (5), Ascidiacea (4 taxa), Cirripedia (4), Priapulida (2), Pycnogonida (2), Tanaidacea (2) and single taxa of Arachnida, Entoprocta, Hirudinea, Leptocardii, Phoronida, Platyhelminthes and Polyplacophora.</p>
Full article ">Figure 4
<p>Multidimensional scaling (MDS) plots for the North Sea and the Baltic Sea areas based on presence/absence transformed data. The North Sea plot only includes stations where both data sets for in- and epifauna were available. Labeling is according to the MPAs and the sampling years.</p>
Full article ">Figure 5
<p>Most common species in the North Sea MPAs. (<b>A</b>) <span class="html-italic">Echinocardium cordatum</span> (Pennant, 1777), (<b>B</b>) <span class="html-italic">Spiophanes bombyx</span> (Claparède, 1870), (<b>C</b>) <span class="html-italic">Liocarcinus holsatus</span> (Fabricius, 1798), (<b>D</b>) <span class="html-italic">Aonides paucibranchiata</span> Southern, 1915, (<b>E</b>) <span class="html-italic">Asterias rubens</span> Linnaeus, 1758, (<b>F</b>) <span class="html-italic">Bathyporeia elegans</span> Watkin, 1940, (<b>G</b>) <span class="html-italic">Abra alba</span> (W. Wood, 1802), (<b>H</b>) <span class="html-italic">Ophiura ophiura</span> (Linnaeus, 1758), (<b>I</b>) <span class="html-italic">Spisula solida</span> (Linnaeus, 1758). Indicated sizes are approximate total lengths (of longest dimension) for all species, with two exceptions: for <span class="html-italic">L. holsatus</span> (<b>C</b>), the value corresponds to carapace length, and for <span class="html-italic">O. ophiura</span>, the disc diameter is specified. These sizes were measured with calipers and are provided only for visualization and to show scale differences between species; they are not relevant for any other reported results.</p>
Full article ">Figure 6
<p>Most common species in the Baltic Sea MPAs. Size measures for each species are given in mm. (<b>A</b>) <span class="html-italic">Mytilus edulis</span> Linnaeus, 1758, (<b>B</b>) <span class="html-italic">Mya arenaria</span> Linnaeus, 1758, (<b>C</b>) <span class="html-italic">Cerastoderma glaucum</span> (Bruguière, 1789), (<b>D</b>) <span class="html-italic">Peringia ulvae</span> (Pennant, 1777), (<b>E</b>) <span class="html-italic">Macoma balthica</span> (Linnaeus, 1758), (<b>F</b>) <span class="html-italic">Abra alba</span> (W. Wood, 1802), (<b>G</b>) <span class="html-italic">Diastylis rathkei</span> (Krøyer, 1841), (<b>H</b>) <span class="html-italic">Scoloplos armiger</span> (Müller, 1776), (<b>I</b>) <span class="html-italic">Carcinus maenas</span> (Linnaeus, 1758), (<b>J</b>) <span class="html-italic">Pygospio elegans</span> Claparède, 1863, (<b>K</b>) <span class="html-italic">Crangon crangon</span> (Linnaeus, 1758). Indicated sizes are approximate total lengths (of longest dimension) for all species but I (for <span class="html-italic">C. maenas</span>, the value corresponds to carapace length). These sizes were measured with calipers and are provided only for visualization and to show scale differences between species; they are not relevant for any other reported results.</p>
Full article ">Figure 7
<p>dbRDA ordination of stations in the North Sea MPAs along environmental (depth (m), sediment parameters (shell fraction &gt; 2 mm, sand fraction &lt; 2 mm to &gt;0.063 mm, mud fraction &lt; 0.063 mm, and gravel fraction), temperature (°C) and salinity (psu)) and anthropogenic (bottom-contact fishing expressed as subsusSwAR) drivers. Labeling according to the MPAs.</p>
Full article ">Figure 8
<p>dbRDA ordination of stations in the Baltic Sea MPAs along environmental and anthropogenic (bottom-contact fishing expressed as subsurSwAR) drivers. Labeling according to the MPAs.</p>
Full article ">
Back to TopTop