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

Previous Issue
Volume 16, October
You seem to have javascript disabled. Please note that many of the page functionalities won't work as expected without javascript enabled.
 
 

Diversity, Volume 16, Issue 11 (November 2024) – 49 articles

  • Issues are regarded as officially published after their release is announced to the table of contents alert mailing list.
  • You may sign up for e-mail alerts to receive table of contents of newly released issues.
  • PDF is the official format for papers published in both, html and pdf forms. To view the papers in pdf format, click on the "PDF Full-text" link, and use the free Adobe Reader to open them.
Order results
Result details
Section
Select all
Export citation of selected articles as:
40 pages, 39717 KiB  
Article
Mites from the Suborder Uropodina (Acari: Mesostigmata) in Bory Tucholskie National Park—One of the Youngest National Parks in Poland
by Jerzy Błoszyk, Jacek Wendzonka, Karolina Lubińska, Marta Kulczak and Agnieszka Napierała
Diversity 2024, 16(11), 699; https://doi.org/10.3390/d16110699 - 14 Nov 2024
Viewed by 273
Abstract
The state of research into acarofauna in Polish national parks is very uneven. One of the least examined areas in this regard is Bory Tucholskie National Park (BTNP), established in 1996. The aim of the current research was to explore the species diversity [...] Read more.
The state of research into acarofauna in Polish national parks is very uneven. One of the least examined areas in this regard is Bory Tucholskie National Park (BTNP), established in 1996. The aim of the current research was to explore the species diversity and community structure of mites from the suborder Uropodina (Acari: Mesostigmata), inhabiting different forest, open, and unstable microhabitats in the area of BTNP. Based on the analysis of over 300 samples collected in BTNP between 2004 and 2024, 29 taxa of Uropodina were identified, with 3839 specimens found in the analyzed material. The highest species diversity has been observed in different types of pine forests (19 species), in transformed alder and alder forests (15 species, each), and in reeds (12 species), while the lowest diversity occurred in peat bog areas (8 species) and inland dunes (5 species). The spatial distribution analyses for Uropodina in the area of BTNP have been made and distribution maps for each species have been drawn. Moreover, the Maturity Index (MI) was also calculated to compare the species diversity of the Uropodina communities in BTNP with those in other Polish national parks. The Uropodina community in BTNP ranked eighth in terms of species richness among 13 national parks explored in Poland so far. Finally, the comparative analysis of the MI for the selected Polish national parks has revealed that BTNP could be ranked fourth in terms of the faunistic value for the discussed mite group. Full article
(This article belongs to the Special Issue Diversity and Ecology of the Acari)
10 pages, 1860 KiB  
Article
How to Count Parrots: Comparing the Performance of Point and Transect Counts for Surveying Tasman Parakeets (Cyanoramphus cookii)
by Michael John Adam Skirrow, Luis Ortiz-Catedral and Adam N. H. Smith
Diversity 2024, 16(11), 698; https://doi.org/10.3390/d16110698 - 14 Nov 2024
Viewed by 273
Abstract
Obtaining precise estimates of population size and trends through time is important for the effective management and conservation of threatened species. For parrots (Psittaciformes: Psittacidae), obtaining such estimates can be challenging, particularly for cryptic species that occur in low densities in complex and/or [...] Read more.
Obtaining precise estimates of population size and trends through time is important for the effective management and conservation of threatened species. For parrots (Psittaciformes: Psittacidae), obtaining such estimates can be challenging, particularly for cryptic species that occur in low densities in complex and/or fragmented habitats. We used a statistical resampling approach with the aim to compare the reliability and precision of counts for the critically endangered Tasman parakeet (Cyanoramphus cookii) that were taken using two methods on Norfolk Island (Pacific Ocean), namely, fixed-point counts and line transect counts. The detections obtained during fixed-point counts had better estimated precision (0.274) than line transect counts (0.476). The fixed-point method was also more efficient, yielding 1.338 parakeet detections per count compared to the 0.642 parakeet detections per count obtained by the line transect method. Although Tasman parakeets can be detected by either of these methods, our research demonstrates that the fixed-point method is more precise and reliable. These findings can help prioritise resources for the long-term monitoring of recovering populations of this species and similar island species. Full article
(This article belongs to the Special Issue Ecology and Conservation of Parrots)
Show Figures

Figure 1

Figure 1
<p>Maps of the Mount Pitt section of the Norfolk Island National Park where we conducted counts of the Tasman parakeet (<span class="html-italic">Cyanoramphus cookii</span>). (<b>A</b>) shows the distribution of survey points, and (<b>B</b>) shows the distribution of walking tracks and roads. Dark grey lines show the increase in elevation at 20 m intervals.</p>
Full article ">Figure 1 Cont.
<p>Maps of the Mount Pitt section of the Norfolk Island National Park where we conducted counts of the Tasman parakeet (<span class="html-italic">Cyanoramphus cookii</span>). (<b>A</b>) shows the distribution of survey points, and (<b>B</b>) shows the distribution of walking tracks and roads. Dark grey lines show the increase in elevation at 20 m intervals.</p>
Full article ">Figure 2
<p>Precision of estimates from the bootstrap analysis of data obtained by two methods used to monitor the Tasman Parakeet (<span class="html-italic">Cyanoramphus cookii</span>) population in the Mount Pitt section of the Norfolk Island National Park.</p>
Full article ">Figure 3
<p>Cumulative detections from the bootstrap analysis of data obtained by two methods used to monitor the Tasman Parakeet (<span class="html-italic">Cyanoramphus cookii</span>) population in the Mount Pitt section of the Norfolk Island National Park.</p>
Full article ">
9 pages, 4975 KiB  
Interesting Images
Finding a Pied-à-Terre: Harbour Infrastructure Facilitates the Settlement of Non-Native Corals (Tubastraea spp.) in the Southern Caribbean
by Bert W. Hoeksema, Roeland J. van der Schoot and Kaveh Samimi-Namin
Diversity 2024, 16(11), 697; https://doi.org/10.3390/d16110697 - 14 Nov 2024
Viewed by 211
Abstract
Semi-submersible platforms are used in the offshore oil and gas industry. They are specialised marine vessels that float on submersed drafts, which are composed of pontoons and columns and can serve as habitats for biofouling marine benthic communities. When these vessels sail from [...] Read more.
Semi-submersible platforms are used in the offshore oil and gas industry. They are specialised marine vessels that float on submersed drafts, which are composed of pontoons and columns and can serve as habitats for biofouling marine benthic communities. When these vessels sail from one place to another, either by using their own propellers or being towed, they can act as vectors for introducing non-native marine species. To establish themselves in new areas, these exotic species require suitable benthic habitats. Artificial substrates, such as harbour infrastructure where such vessels are moored, appear to be highly suitable for this purpose. In the present study, a mooring buoy and a harbour piling at Curaçao (southern Caribbean), frequently used by semi-submersible platforms, were found to be colonised by the sun corals Tubastraea coccinea and T. tagusensis at shallow depths. This report presents the first record of T. tagusensis as an introduced non-native species in the southern Caribbean, highlighting the potential role of harbour infrastructure in facilitating coral settlement at depths shallower than those typically observed. These findings underscore the ecological impact of artificial substrates in supporting invasive species and emphasise the need for monitoring programs and defouling facilities. Full article
(This article belongs to the Collection Marine Invasive Species)
Show Figures

Figure 1

Figure 1
<p>Map of Curaçao showing the position of long-term mooring locations for semi-submersible platforms: 1. Mooring buoy at Boca Sami Beach, St. Michiel Bay; 2. Caracas Bay mooring jetties and harbour pilings near Tugboat Beach.</p>
Full article ">Figure 2
<p>Aerial view of a semi-submersible platform attached to a mooring buoy (arrow) near Boca Sami Beach (St. Michiel Bay, Curaçao), on which two <span class="html-italic">Tubastraea</span> species were found. Source: Google Maps; Imagery, Airbus Maxar ©2024.</p>
Full article ">Figure 3
<p>Aerial view of two semi-submersible platforms moored at jetties in Caracas Bay, Curaçao. Arrow: the position of a harbour piling at Tugboat Beach where two <span class="html-italic">Tubastraea</span> species were found. Source: Google Maps; Imagery, Airbus Maxar ©2024.</p>
Full article ">Figure 4
<p>Mooring buoy at Boca Sami, St. Michiel Bay, Curaçao (March 2023). (<b>a</b>) Underside of the buoy showing a high density of <span class="html-italic">Tubastraea coccinea</span>, 0.5 m depth; (<b>b</b>) Part of the anchor chain underneath the buoy with 100% cover of fouling organisms, including <span class="html-italic">Tubastraea</span> spp., 3 m depth; and (<b>c</b>) Chain section, partially overgrown by native reef corals attached to a concrete block to hold the buoy in position, 5 m depth. (<b>d</b>) Chain links not covered by fouling animals (40 cm wide), 8 m depth. Photo credit: B.W.H.</p>
Full article ">Figure 5
<p>Non-native <span class="html-italic">Tubastraea</span> corals as fouling organisms on harbour infrastructure in Curaçao (March 2023). (<b>a</b>) <span class="html-italic">Tubastraea coccinea</span> (two colour morphs) attached to the underside of a mooring buoy at Boca Sami, 0.5 m depth; (<b>b</b>,<b>c</b>) <span class="html-italic">Tubastraea tagusensis</span> attached to the buoy’s anchor chain at Boca Sami, 7–8 m depth; and (<b>d</b>) Adjacent colonies of <span class="html-italic">Tubastraea coccinea</span> (top) and <span class="html-italic">T. tagusensis</span> (bottom) on a harbour piling at Tugboat Beach in Caracas Bay, 1 m depth. Scale bars: 1 cm. Photo credit: B.W.H.</p>
Full article ">
17 pages, 4874 KiB  
Article
Aggregate Size Mediated the Changes in Soil Microbial Communities After the Afforestation of a Former Dryland in Northwestern China
by Deming Zhang, Ling Bai, Wei Wang, Yanhe Wang, Tiankun Chen, Quan Yang, Haowen Chen, Shuning Kang, Yongan Zhu and Xiang Liu
Diversity 2024, 16(11), 696; https://doi.org/10.3390/d16110696 - 13 Nov 2024
Viewed by 254
Abstract
Although the afforestation of former arable lands is a common global land-use conversion, its impact on soil microbial communities at the aggregate scale has not been adequately addressed. In this study, soil samples were categorized into large macroaggregates (LM, >2 mm), small macroaggregates [...] Read more.
Although the afforestation of former arable lands is a common global land-use conversion, its impact on soil microbial communities at the aggregate scale has not been adequately addressed. In this study, soil samples were categorized into large macroaggregates (LM, >2 mm), small macroaggregates (SM, 2–0.25 mm), and microaggregates (MI, <0.25 mm) to assess the changes in microbial composition, diversity, network complexity, and network stability within soil aggregates after the afforestation of a former dryland in northwestern China. The results revealed that afforestation enhanced the relative abundance of Actinobacteriota, Chloroflexi, Ascomycota, and Mortierellomycota within the soil aggregates, suggesting that these phyla may have greater advantages in microbial communities post-afforestation. The Shannon–Wiener and Pielou indices for bacterial communities showed no significant differences between land-use types across all aggregate fractions. However, the alpha diversity of fungal communities within the LM and SM significantly increased after afforestation. Bray–Curtis dissimilarity indices showed that afforestation altered bacterial beta diversity within the LM and MI but had a minimal impact on fungal beta diversity across all three aggregate fractions. The topological features of cross-kingdom microbial co-occurrence networks within the soil aggregates generally exhibited a decreasing trend post-afforestation, indicating a simplification of microbial community structure. The reduced robustness of microbial networks within the LM and SM fractions implies that afforestation also destabilized the structure of microbial communities within the macroaggregates. The composition of the soil microbial communities correlated closely with soil carbon and nitrogen contents, especially within the two macroaggregate fractions. The linkages suggests that improved resource conditions could be a key driver behind the shifts in microbial communities within soil aggregates following afforestation. Our findings indicate that the impact of afforestation on soil microbial ecology can be better understood by soil aggregate fractionation. Full article
(This article belongs to the Special Issue Microbial Community Dynamics in Soil Ecosystems)
Show Figures

Figure 1

Figure 1
<p>The characteristics of microbial community compositions ((<b>a</b>,<b>c</b>): bacteria; (<b>b</b>,<b>d</b>): fungi) within the soil aggregates of different land-use types. Different uppercase letters indicate significant differences among aggregate fractions within a specific land-use type, whereas different lowercase letters denote significant differences between land-use types within a specific aggregate fraction. DLM, DSM, and DMI represent large macroaggregates, small macroaggregates, and microaggregates in the dryland, respectively, while PLM, PSM, and PMI represent large macroaggregates, small macroaggregates, and microaggregates in the <span class="html-italic">C. korshinskii</span> plantation, respectively.</p>
Full article ">Figure 2
<p>Shannon–Wiener and Pielou indices of bacterial (<b>a</b>,<b>b</b>) and fungal communities (<b>c</b>,<b>d</b>) within soil aggregates of different land-use types. Different uppercase letters denote significant differences among aggregate fractions, while asterisks indicate significant differences between land-use types. * <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; ns, non-significant.</p>
Full article ">Figure 3
<p>Beta diversity of bacterial (<b>a,c</b>) and fungal communities (<b>b,d</b>) within soil aggregates of different land-use types as revealed by principal coordinates analysis and Bray–Curtis dissimilarity. DLM, DSM, and DMI denote large macroaggregates, small macroaggregates, and microaggregates in dryland, respectively, while PLM, PSM, and PMI represent large macroaggregates, small macroaggregates, and microaggregates in <span class="html-italic">C. korshinskii</span> plantation, respectively. Different uppercase letters indicate significant differences between aggregate fractions, while asterisks indicate significant differences between land-use types. *** <span class="html-italic">p</span> &lt; 0.001; ns, non-significant.</p>
Full article ">Figure 4
<p>Cross-kingdom co-occurrence networks of microbial taxa within soil aggregates of different land-use types ((<b>a</b>–<b>c</b>): dryland; (<b>d</b>–<b>f</b>): <span class="html-italic">C. korshinskii</span> plantation). Different colors represent different modules, and modules with &lt;5 nodes are represented in gray. LM, SM, and MI represent large macroaggregates, small macroaggregates, and microaggregates, respectively.</p>
Full article ">Figure 5
<p>Weighted (<b>a</b>) and unweighted robustness (<b>b</b>) of microbial cross-kingdom co-occurrence networks within the soil aggregates of different land-use types. Error bars represent the standard deviations of the mean. * <span class="html-italic">p</span> &lt; 0.05; *** <span class="html-italic">p</span> &lt; 0.001; ns, non-significant.</p>
Full article ">Figure 6
<p>Factors that influenced the compositions of bacterial (<b>a</b>) and fungal communities (<b>b</b>) within the soil aggregates as revealed by the Mantel test. LM, large macroaggregates; SM, small macroaggregates; MI, microaggregates; MP, mass proportion; TOC, total organic carbon; TN, total nitrogen; TP, total phosphorus; C:N, carbon to nitrogen ratio; AN, alkali-hydrolyzable nitrogen; AP, available phosphorus. * <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 ">
14 pages, 12628 KiB  
Article
The Range of the Colonial Microcystis’ Biomass for Shift to Diatom Aggregates Under Aeration Mixing and Low Light
by Xiaodong Wang, Xuan Che, Xingguo Liu, Xinfeng Li, Xiaolong Chen, Yiming Li and Lin Zhu
Diversity 2024, 16(11), 695; https://doi.org/10.3390/d16110695 - 13 Nov 2024
Viewed by 212
Abstract
In order to investigate non-cyanobacteria dominance succession from Microcystis blooms, particularly to diatom dominance, an experiment using varying colonial Microcystis biomasses expressed as bulk concentrations of 2.0, 4.0, 6.0, 8.0, 10.0, 12.0, 14.0, 16.0, 18.0, 20.0, 22.0, and 24.0 mL L−1 was [...] Read more.
In order to investigate non-cyanobacteria dominance succession from Microcystis blooms, particularly to diatom dominance, an experiment using varying colonial Microcystis biomasses expressed as bulk concentrations of 2.0, 4.0, 6.0, 8.0, 10.0, 12.0, 14.0, 16.0, 18.0, 20.0, 22.0, and 24.0 mL L−1 was undertaken under continuous aeration mixing in a greenhouse during a hot summer where shading had reduced light level by 97%. The results showed that the algal shift process was affected by the initial biomass of the colonial Microcystis, and the algal community diversified. When the Microcystis bulk concentration was between 2.0 and 16.0 mL L−1, the bloom became dominated by diatom Nitzschia palea, which aggregated on the mucilage sheathes of the Microcystis colonies. The diatom density at bulk concentration biomass of 2.0 mL L−1 reached a maximum at 2.8 × 105 cells mL−1 on day 27. When the bulk concentration was at 18.0–24.0 mL L−1, no diatom dominance appeared. The shift from a Microcystis bloom to diatom dominance was affected by the initial Microcystis biomass, and the most suitable bulk concentration biomass for colonial Microcystis was at 2–12 mL L−1, in which the chlorophyll-a level was about from 285 to 1714 μg L−1. The mechanism underlying this algal shift may be that the low light and nutrient levels in the Microcystis bloom promoted diatom aggregation growth on the mucilage sheaths of Microcystis colonies under continuous aeration mixing. Full article
(This article belongs to the Special Issue Eutrophication, Aquaculture and Aquatic Ecosystem Restoration)
Show Figures

Figure 1

Figure 1
<p>Changes in water temperature at 9:00–10:00 and 14:00–15:00 h every day in treatment B4.</p>
Full article ">Figure 2
<p>Changes in dissolved oxygen (DO) and pH value at 14:00–15:00 h in each treatment.</p>
Full article ">Figure 3
<p>Changes in Chl-<span class="html-italic">a</span> in the 12 treatments. Day 0 was 17 July 2020, and day 41 was 27 August 2020.</p>
Full article ">Figure 4
<p>Changes in the phenomenon of all the jars in different treatments under the black polyolefin sun-shading net in the greenhouse. The 0 d, 4 d, 7 d, 10 d, 14 d, 17 d, 20 d, 24 d, 27 d, 32 d, and 36 d in the figure represent the picture on days 0, 4, 7, 10, 14, 17, 20, 24, 27, 32, and 36, respectively. The distribution of the jars in different treatments in the greenhouse is presented in 4a. The last subfigure Pla represents the placement diagram of the jars in different treatments under the black polyolefin sun-shading net in the greenhouse.</p>
Full article ">Figure 5
<p>The changes in <span class="html-italic">Nitzschia</span> cell density in each treatment from day 17 to day 41. The <span class="html-italic">Nitzschia</span> density was measured from day 17 on, when the water in treatments B2 and B4 turned brownish.</p>
Full article ">Figure 6
<p>Linear regression between the square root transformed data of the first three samples of Chl-<span class="html-italic">a</span> and the three highest densities of diatom <span class="html-italic">Nitzschia</span>. The red line represents linear regression, and the black squares are scatter plots.</p>
Full article ">Figure 7
<p>Representative process for the algal shift from <span class="html-italic">Microcystis</span> colonies to <span class="html-italic">Nitzschia</span> dominance with aggregation. The <span class="html-italic">Nitzschia</span> dominance declines in treatment B2 (2 mL L<sup>−1</sup> bulk concentration). The 14 d, 17 d, 26 d, and 34 d in the figure represent microscope photographs on days 14, 17, 26, and 34, respectively.</p>
Full article ">Figure 8
<p>Representative process showing the algal shift from <span class="html-italic">Microcystis</span> colonies to <span class="html-italic">Nitzschia</span> aggregations in treatment B10 (10 mL L<sup>−1</sup> bulk concentration). The 10 d, 13 d, 17 d, and 26 d in the figure represent microscope photographs on days 10, 13, 17, and 26, respectively.</p>
Full article ">Figure 9
<p>Representative process showing the algal shift from <span class="html-italic">Microcystis</span> colonies to bacteria dominance without <span class="html-italic">Nitzschia</span> aggregation in treatment B18 (18 mL L<sup>−1</sup> bulk concentration). The 3 d, 20 d, 26 d, and 34 d in the figure represent microscope photographs on days 3, 20, 26, and 34, respectively.</p>
Full article ">
12 pages, 1087 KiB  
Article
Traditional Use, Chemical Constituents, and Pharmacological Activity of Maytenus elaeodendroides Stem Bark
by Trina H. García, Iraida Spengler, Antonio Fernández, Idania Rodeiro, Ivones Hernández-Balmaseda, Ilianet Céspedes, Gabino Garrido, Lourdes Campaner dos Santos, Wagner Vilegas, Rita Celano and Maria D’Elia
Diversity 2024, 16(11), 694; https://doi.org/10.3390/d16110694 - 13 Nov 2024
Viewed by 258
Abstract
Plants belonging to the genus Maytenus are members of the Celastraceae family. They have been widely used by different peoples as treatment for curing many diseases. The aim of this study was to explore the anti-inflammatory and antioxidant properties of Maytenus elaeodendroides stem [...] Read more.
Plants belonging to the genus Maytenus are members of the Celastraceae family. They have been widely used by different peoples as treatment for curing many diseases. The aim of this study was to explore the anti-inflammatory and antioxidant properties of Maytenus elaeodendroides stem bark extracts, an endemic Cuban plant. The antioxidant activity of four extracts (EtOH, EtOAc, n-BuOH, and diethyl ether/petroleum ether 1:1) was determined using DPPH and FRAP methods. Meanwhile, anti-inflammatory effects by the edema method were induced by croton oil in the mouse ear. The investigated extracts showed radical reduction capacity and prevented ear inflammation at doses of 4 mg/ear. In addition, FIA/ESI/IT/MSn was used to determine the qualitative chemical composition of the EtOAc extract and allowed the identification of five flavan-3-ol monomers, four dimers, and other proanthocyanidin oligomers. From this extract three flavan-3-ol compounds (elaeocyanidin and 4′-O-methylgallocatechin), one of them new (2′-hydroxy-4′-methoxy-epigallocatechin), and a proanthocyanidin dimer (afzelechin-(4β8)-4′-O-methylepigallocatechin) were isolated and identified by the chromatographic method and spectroscopic techniques, mainly ESI-MS and NMR spectroscopic methods. Full article
Show Figures

Figure 1

Figure 1
<p>Compounds <b>1</b>–<b>4</b> isolated from <span class="html-italic">M. elaeodendroides</span> stem bark.</p>
Full article ">Figure 2
<p>Total ion mass spectra of EtOAc extract from stem bark of <span class="html-italic">Maytenus elaeodendroides</span>.</p>
Full article ">Figure 3
<p>Anti-inflammatory effects of <span class="html-italic">M. elaeodendroides</span> extracts. Each bar represents the means ± ESM of seven animals per group. Values of * <span class="html-italic">p</span> &lt; 0.05 represent statistically significant differences with respect to the control group. Positive control Indometacine (C+, 3 mg/ear).</p>
Full article ">
19 pages, 4217 KiB  
Article
Midge Paleo-Communities (Diptera Chironomidae) as Indicators of Flood Regime Variations in a High-Mountain Lake (Italian Western Alps): Implications for Global Change
by Marco Bertoli, Gianguido Salvi, Rachele Morsanuto, Elena Pavoni, Paolo Pastorino, Giuseppe Esposito, Damià Barceló, Marino Prearo and Elisabetta Pizzul
Diversity 2024, 16(11), 693; https://doi.org/10.3390/d16110693 - 12 Nov 2024
Viewed by 293
Abstract
Sediments of alpine lakes serve as crucial records that reveal the history of lacustrine basins, offering valuable insights into the effects of global changes. One significant effect is the variation in rainfall regimes, which can substantially influence nutrient loads and sedimentation rates in [...] Read more.
Sediments of alpine lakes serve as crucial records that reveal the history of lacustrine basins, offering valuable insights into the effects of global changes. One significant effect is the variation in rainfall regimes, which can substantially influence nutrient loads and sedimentation rates in lacustrine ecosystems, thereby playing a pivotal role in shaping biotic communities. In this study, we analyze subfossil chironomid assemblages within a sediment core from an alpine lake (western Italian Alps) to investigate the effects of rainfall and flood regime variations over the past 1200 years. Sediment characterization results highlight changes in sediment textures and C/N ratio values, indicating phases of major material influx from the surrounding landscape into the lake basin. These influxes are likely associated with intense flooding events linked to heavy rainfall periods over time. Flooding events are reflected in changes in chironomid assemblages, which in our samples are primarily related to variations in sediment texture and nutrient loads from the surrounding landscape. Increased abundances of certain taxa (i.e., Brillia, Chaetocladius, Cricotopus, Psectrocladius, Cricotopus/Orthocladius Parorthocladius) may be linked to higher organic matter and vegetation inputs from the surrounding landscape. Biodiversity decreased during certain periods along the core profile due to intense flood regimes and extreme events. These results contribute to our understanding of alpine lake system dynamics, particularly those associated with intense flooding events, which are still understudied. Full article
(This article belongs to the Section Biodiversity Loss & Dynamics)
Show Figures

Figure 1

Figure 1
<p>(<b>a</b>) Study area and (<b>b</b>) location of the sampling site in Upper Balma Lake.</p>
Full article ">Figure 2
<p>Stratigraphic diagram of the sedimentological and geochemical parameters measured in core sections sampled in the Upper Balma Lake. Results of the element analysis in light blue color are also reported.</p>
Full article ">Figure 3
<p>Bayesian age-depth model calculated for the Upper Balma Lake, based on 4000 interactions Markov Chain Monte Carlo. The dark gray areas represent the more precise dates, those in light gray the less precise dates; the red line indicates the best estimate of age for each level, and the black dashed lines the 95% confidence intervals.</p>
Full article ">Figure 4
<p>Relative abundances of the chironomid taxa observed in the Upper Balma Lake core sections and trends of the main community indices calculated along the core. Time periods with high flood regimes are indicated by the light blue bands superimposed on the graphs; identification of these periods is based on Giguet-Covex et al. [<a href="#B8-diversity-16-00693" class="html-bibr">8</a>] and Wilhelm et al. [<a href="#B55-diversity-16-00693" class="html-bibr">55</a>]. Group colors highlighted by cluster analysis and used in the RDA are reported (see <a href="#diversity-16-00693-f005" class="html-fig">Figure 5</a> and <a href="#diversity-16-00693-f006" class="html-fig">Figure 6</a>a).</p>
Full article ">Figure 5
<p>Cluster analysis defining stratigraphic zones (groups of sections) along the core based on the chironomid assemblages in the Upper Balma Lake (<b>a</b>) and broken sticks analysis defining the proper number of groups (<b>b</b>) (<span class="html-italic">n</span> = 6). Obtained stratigraphic zones are indicated with the same colors used for RDA analysis (see <a href="#diversity-16-00693-f006" class="html-fig">Figure 6</a>a).</p>
Full article ">Figure 6
<p>(<b>a</b>) Redundancy Analyses (RDA) illustrate the associations between chironomid taxa and the variables under consideration and (<b>b</b>) Venn diagrams depict the results of variance partitioning analysis (VPA) for the four variable groups: nutrients (TOC and C/N ratio), trace elements (Pb, Mo), sediment characteristics (first percentile Cμ and median diameter Mμ), and the presence of fish in relation to chironomid taxa. Variance that is unexplained or accounts for less than 1% is omitted. The group colors used in the RDA analysis correspond to those in the cluster analysis (refer to <a href="#diversity-16-00693-f005" class="html-fig">Figure 5</a>).</p>
Full article ">
17 pages, 1600 KiB  
Article
Comparison of Orchid Conservation Between China and Other Countries
by Shixing Li, Cuiyi Liang, Shuwen Deng, Chen Chen, Liangchen Yuan, Zhen Liu, Shasha Wu, Siren Lan, Ziang Tang, Zhongjian Liu and Junwen Zhai
Diversity 2024, 16(11), 692; https://doi.org/10.3390/d16110692 - 12 Nov 2024
Viewed by 227
Abstract
Global attention is highly focused on biodiversity conservation. Various countries are actively implementing relevant conservation measures. To advance these efforts in China, it is essential to understand global conservation actions. The orchid family, one of the most diverse groups of flowering plants, has [...] Read more.
Global attention is highly focused on biodiversity conservation. Various countries are actively implementing relevant conservation measures. To advance these efforts in China, it is essential to understand global conservation actions. The orchid family, one of the most diverse groups of flowering plants, has become a “flagship” group for plant conservation. In this study, we summarized 3418 policies and regulations related to orchid conservation in 45 countries. We found that orchid conservation actions in various countries have focused on in situ conservation, with 1469 policies and regulations issued for nature reserves, while ex situ conservation has been seriously neglected, with only seven relevant regulations. Most developing countries have experienced an increase in orchid conservation actions, while developed countries have plateaued. We amassed 370 non-governmental organizations (NGOs) for orchid conservation. At present, the total number of policies and regulations for orchid protection in China is approximately 84, with 67 issued since 2000. Two non-governmental organizations have been established for orchid conservation. Although the benefit of orchid conservation in China is significant, it still requires continuous improvement compared to many other countries. We recommend that the Chinese government draws on the experiences of the United States, Canada, and Australia in areas such as policy and regulation formulation, optimization of non-governmental organizations, and implementation of related conservation projects. Through learning and collaboration, challenges can be transformed into opportunities for development. Full article
(This article belongs to the Special Issue Plant Diversity Hotspots in the 2020s)
Show Figures

Figure 1

Figure 1
<p>Comparison of the number of policies and regulations in 45 (14 developed and 31 developing) countries. The larger the sector area, the greater the number of policies and regulations. FRA, France; CAN, Canada; ITA, Italy; DEU, Germany; GBR, Great Britain; AUS, Australia; BEL, Belgium; AUT, Austria; GRC, Greece; USA, United States of America; NZL, New Zealand; RUS, Russia; BRA, Brazil; ARG, Argentina; MEX, Mexico; CHN, China; BLR, Belarus; CRI, Costa Rica; ECU, Ecuador; CHL, Chile; GTM, Guatemala; IND, Indonesia; ZAF, South Africa; COL, Colombia; PAN, Panama; BOL, Bolivia; VNM, Vietnam; TUR, Turkey; MYS, Malaysia.</p>
Full article ">Figure 2
<p>(<b>a</b>) Trends in the number of policies and regulations in 31 developing countries over ten-year intervals (for simplification, only countries with significant changes are highlighted). The time was determined based on the dates of policy enactment. Detailed data are provided in the <a href="#app1-diversity-16-00692" class="html-app">Supplementary Materials</a>. RUS, Russia; MEX, Mexico; ARG, Argentina; BRA, Brazil; CHN, China; CRI, Costa Rica; CHL, Chile; ECU, Ecuador; BLR, Belarus. (<b>b</b>) Trends in the number of policies and regulations in 14 developed countries over ten-year intervals (for simplification, only countries with significant changes are highlighted). The time was determined based on the dates of policy enactment. Detailed data are provided in the <a href="#app1-diversity-16-00692" class="html-app">Supplementary Materials</a>. FRA, France; GBR, Great Britain; DEU, Germany; AUS, Australia; ITA, Italy; CAN, Canada; BEL, Belgium; AUT, Austria.</p>
Full article ">Figure 3
<p>The differences in the number of non-governmental organizations (NGOs) across countries using a logarithmic scale. The logarithmic scale allows for the simultaneous visualization of smaller and larger quantities in the same chart: (<b>a</b>) Comparison of the number of orchid conservation NGOs in 24 developing countries; LAO, PNG, KHM, TUR, VNM, BLR, and RUS are not displayed due to the absence of NGO data. (<b>b</b>) Comparison of the number of orchid conservation NGOs in 14 developing countries. SGP, Singapore; LVA, Latvia; GRC, Greece; JPN, Japan; AUT, Austria; BEL, Belgium; ITA, Italy; NZL, New Zealand; DEU, Germany; FRA, France; GBR, Great Britain; CAN, Canada; AUS, Australia; USA, United States of America; LAO, Laos; PNG, Papua New Guinea; KHM, Cambodia; TUR, Turkey; VNM, Vietnam; BLR, Belarus; RUS, Russia; ROU, Romania; BRB, Barbados; BLZ, Belize; KEN, Kenya; NPL, Nepal; IDN, India; HUN, Hungary; MYS, Malaysia; BOL, Bolivia; PAN, Panama; GTM, Guatemala; CHL, Chile; ECU, Ecuador; CRI, Costa Rica; MEX, Mexico; BTN, Bhutan; POL, Poland; BGR, Bulgaria; IND, Indonesia; CHN, China; ARG, Argentina; COL, Colombia; ZAF, South Africa; BRA, Brazil.</p>
Full article ">
13 pages, 1596 KiB  
Article
Assessing the Viability of Translocated Mongolian Dung Beetles (Gymnopleurus mopsus) for Ecological Restoration in Republic of Korea: An Analysis of Environmental Adaptability
by Hwang Kim, Doo-Hyung Lee, Sun-Hee Hong, Jong-Seok Park, Jung-Wook Kho and Young-Joong Kim
Diversity 2024, 16(11), 691; https://doi.org/10.3390/d16110691 - 12 Nov 2024
Viewed by 229
Abstract
This study investigates the reintroduction and ecological adaptation of the endangered dung beetle, Gymnopleurus mopsus, in South Korea, a region from which it has been absent since the 1970s. To facilitate this, we imported genetically identical populations of G. mopsus from Mongolia [...] Read more.
This study investigates the reintroduction and ecological adaptation of the endangered dung beetle, Gymnopleurus mopsus, in South Korea, a region from which it has been absent since the 1970s. To facilitate this, we imported genetically identical populations of G. mopsus from Mongolia and embarked on a comprehensive restoration research project. A key focus of this endeavor was to evaluate the adaptability of these beetles to the local environment, an essential aspect of successful reintegration of species from foreign ecosystems. Under meticulously controlled field cage conditions, we conducted an in-depth monitoring of the life history traits of G. mopsus. This monitoring revealed that the adult beetles, which entered hibernation in September 2019, began emerging in stages from late April to May 2020. Following hibernation, we observed that the adults engaged in reproductive activities from late-May until early-August, with the emergence of the first-generation (F1) adults occurring from late-July to mid-September. This led to a notable tripling in population size, increasing from 34 to 109 individuals. The successful survival and reproductive behaviors of these Mongolian dung beetles in the climatic conditions of Korea suggest a promising potential for their adaptation when reintroduced into native habitats. We are now directing our efforts towards long-term monitoring, focusing on the survival and reproductive efficacy of these new generations, to further validate the success of this ecological restoration project. Full article
(This article belongs to the Special Issue Biodiversity Conservation Planning and Assessment)
Show Figures

Figure 1

Figure 1
<p>Emergence rates of introduced <span class="html-italic">G. mopsus</span> adults and second-generation (F1) adults in semi-field conditions following hibernation. Pink circles indicate the introduced females, purple circles represent the introduced males, yellow circles denote the newly emerged females, and green circles represent the newly emerged males.</p>
Full article ">Figure 2
<p>Cumulative rates of hibernation onset for introduced <span class="html-italic">G. mopsus</span> adults and newly born second-generation adults (F1). Pink circles indicate introduced the females, purple circles represent the introduced males, yellow circles denote the newly emerged female adults, and green circles represent the newly emerged male adults.</p>
Full article ">Figure 3
<p>Comparison of body sizes between introduced <span class="html-italic">G. mopsus</span> adults from Mongolia and newly born second-generation adults ((<b>A</b>) Thorax width and Body length; (<b>B</b>) Body weight). Black bars represent introduced adults, while gray bars represent newly born second-generation adults (means ± SE). Student’s <span class="html-italic">t</span>-tests, * <span class="html-italic">p</span> &lt; 0.001.</p>
Full article ">Figure 4
<p>Mean weekly oviposition per female of <span class="html-italic">G. mopsus</span> when exposed to semi-filed cage conditions. Numbers in the plot represent N of each data point. Black vertical lines represent SE.</p>
Full article ">Figure 5
<p>Survival of <span class="html-italic">G. mopsus</span> when exposed to semi-field cage condition and Kaplan–Meier survival estimates with 95% confidence limits ((<b>A</b>) introduced adults; (<b>B</b>) second generation adults).</p>
Full article ">
37 pages, 35096 KiB  
Article
Seaweed-Associated Diatoms (Bacillariophyta) in Dokdo of South Korea: I. Subphyla Melosirophytina, Coscinodiscophytina, and Class Mediophyceae
by Joon Sang Park, Kyun-Woo Lee, Seung Won Jung, Han Jun Kim and Jin Hwan Lee
Diversity 2024, 16(11), 690; https://doi.org/10.3390/d16110690 - 12 Nov 2024
Viewed by 386
Abstract
Dokdo is an island located in the easternmost part of Korea, which has high levels of biodiversity of birds and fish, especially marine invertebrates. However, the biodiversity of microalgae, especially diatoms (Bacillariophyta), is relatively unknown, despite their ecological importance as primary producers of [...] Read more.
Dokdo is an island located in the easternmost part of Korea, which has high levels of biodiversity of birds and fish, especially marine invertebrates. However, the biodiversity of microalgae, especially diatoms (Bacillariophyta), is relatively unknown, despite their ecological importance as primary producers of the marine food web and bioindicators of environmental conditions associated with climate change. To understand the biodiversity of seaweed-associated diatoms from Dokdo, we collected macroalgae present at a depth 5–15 m by SCUBA diving on 17 October 2017. There were a large number of diatoms (over 130 species), even though it was a one-time survey. As it includes too many taxa to cover at once, voucher flora for other taxonomic groups will be provided through the continuous serial papers. This is the first series of seaweed-associated diatoms, with 26 species belonging to the subphyla Melosirophytina and Coscinodisophytina, and the class Mediophyceae. Among these, seven species including one new taxon were reported for the first time in Korea, which, along with the geopolitical characteristics of the survey area, proved that there is no domestic interest in seaweed-related diatoms. In particular, the appearance of species that have been reported in subtropical waters, such as the order Ardissoneales, requires continuous monitoring of marine seaweed-associated diatoms to confirm whether their colonization in Dokdo waters was due to climate change or species-specific water temperature tolerance. Full article
(This article belongs to the Section Marine Diversity)
Show Figures

Figure 1

Figure 1
<p>Sampling site of seaweed-associated diatoms in Dokdo. (<b>A</b>) Location of Dokdo in the East Sea (square), (<b>B</b>) Sampling site of seaweeds by SCUBA (black circle).</p>
Full article ">Figure 2
<p>Light microscope images of seven diatoms in Dokdo. (<b>A</b>–<b>C</b>) <span class="html-italic">Hyalodiscus scoticus</span>, (<b>D</b>–<b>G</b>) <span class="html-italic">Podosira hormoides</span>, (<b>H</b>) <span class="html-italic">Podosira montagnei</span>, (<b>I</b>) <span class="html-italic">Asteromphalus heptactis</span>, (<b>J</b>) <span class="html-italic">Coscinodiscus asteromphalus</span>, (<b>K</b>) <span class="html-italic">Coscinodiscus marginatus</span>, (<b>L</b>,<b>M</b>) <span class="html-italic">Coscinodiscus radiatus</span>. Scale bar = 10 µm.</p>
Full article ">Figure 3
<p>Light microscope images of four diatoms in Dokdo. (<b>A</b>–<b>E</b>) <span class="html-italic">Actinocyclus pruinosus</span>, (<b>F</b>–<b>L</b>) <span class="html-italic">Actinocyclus subtilis</span>, (<b>M</b>) <span class="html-italic">Pseudosolenia calcar-avis</span>. Scale bar = 10 µm.</p>
Full article ">Figure 4
<p>Light microscope images of four diatoms in Dokdo. (<b>A</b>–<b>C</b>) <span class="html-italic">Ardissonea formosa</span>, (<b>D</b>–<b>F</b>) <span class="html-italic">Ardissonea</span> aff. <span class="html-italic">formosa</span>, (<b>G</b>–<b>J</b>) <span class="html-italic">Ardissoneopsis dokdoensis</span>, (<b>G</b>) Isotype image, (<b>J</b>) Holotype image, (<b>K</b>,<b>L</b>) <span class="html-italic">Ardissoneopsis fulgicans</span>. Scale bar = 10 µm.</p>
Full article ">Figure 5
<p>Light microscope images of six diatoms in Dokdo. (<b>A</b>–<b>C</b>) <span class="html-italic">Climacosphenia moniligera</span>, (<b>D</b>,<b>E</b>) <span class="html-italic">Grunowago pacifica</span>, (<b>F</b>,<b>G</b>) <span class="html-italic">Synedrosphenia crystallina</span>, (<b>H</b>–<b>J</b>) <span class="html-italic">Synedrosphenia</span> cf. <span class="html-italic">gomphonema</span>, (<b>K</b>) <span class="html-italic">Toxarium henndeyanum</span>, (<b>L</b>–<b>N</b>) <span class="html-italic">Toxarium undulatum</span>. Scale bar = 10 µm.</p>
Full article ">Figure 6
<p>Light microscope images of five diatoms in Dokdo. (<b>A</b>–<b>C</b>) <span class="html-italic">Biddulphiopsis titiana</span>, (<b>D</b>,<b>E</b>) <span class="html-italic">Trigonoium</span> cf. <span class="html-italic">arcticum</span>, (<b>F</b>,<b>G</b>) <span class="html-italic">Lampriscus shadboltianus</span>, (<b>H</b>) <span class="html-italic">Pseudictyota reticulata</span>, (<b>I</b>–<b>K</b>) <span class="html-italic">Thalassiosira</span> sp. Scale bar = 10 µm.</p>
Full article ">Figure 7
<p>Scanning electron microscope images of four species in the family Hyalodiscaceae in Dokdo. (<b>A</b>–<b>C</b>) <span class="html-italic">Hyalodiscus ambiguus</span>, (<b>A</b>) Internal valve view, (<b>B</b>) Central hyaline area, (<b>C</b>) Internal structure of valve face rimoportula, (<b>D</b>–<b>I</b>) <span class="html-italic">Hyalodiscus scoticus</span>, (<b>D</b>,<b>E</b>) External valve view, (<b>F</b>) External opening of marginal rimoportula (arrowhead), (<b>G</b>,<b>H</b>) Internal valve view, (<b>I</b>) Internal structure of marginal rimoportula, (<b>J</b>–<b>M</b>) <span class="html-italic">Podosira hormoides</span>, (<b>J</b>) External valve view, (<b>K</b>) Fasciculate areolation and central area covered by mucilage pad, (<b>L</b>) Internal valve view, (<b>M</b>) Fasciculate areolation and central annulus (<b>N</b>,<b>O</b>) <span class="html-italic">Podosira montagnei</span>, (<b>N</b>) External valve view, (<b>O</b>) Decussate areolation and scattered valve face rimoportulae. Scale bar = 10 µm (<b>A</b>,<b>J</b>,<b>L</b>,<b>N</b>); 5 µm (<b>D</b>,<b>E</b>,<b>G</b>,<b>H</b>,<b>K</b>); 1 µm (<b>B</b>,<b>C</b>,<b>F</b>,<b>I</b>,<b>M</b>,<b>O</b>).</p>
Full article ">Figure 8
<p>Scanning electron microscope images of two <span class="html-italic">Actinocyclus</span> species in Dokdo. (<b>A</b>–<b>F</b>) <span class="html-italic">Actinocyclus pruinosus</span>, (<b>A</b>) External valve view, (<b>B</b>) Fasciculate areolation and covered by cribra, (<b>C</b>) external opening of pseudonodule and mantle rimoportula, (<b>D</b>) Internal valve view (<b>E</b>) Internal opening of areolae and fasciculate areolation, (<b>F</b>) Internal structure of rimoportula and pseudonodule, (<b>G</b>–<b>L</b>) <span class="html-italic">Actinocyclus subtilis</span>, (<b>G</b>) External valve view, (<b>H</b>) Spicate areolation (<b>I</b>) External opening of pseudonodule and mantle rimoportula, (<b>J</b>) Internal valve view, (<b>K</b>) spicate areolation, (<b>L</b>) Internal structure of rimoportula and pseudonodule. Scale bar = 10 µm (<b>A</b>,<b>D</b>,<b>G</b>,<b>J</b>); 5 µm (<b>B</b>,<b>E</b>,<b>H</b>,<b>K</b>); 1 μm (<b>C</b>,<b>F</b>,<b>I</b>,<b>L</b>).</p>
Full article ">Figure 9
<p>Scanning electron microscope images of <span class="html-italic">Ardissonea formosa</span> in Dokdo. (<b>A</b>–<b>C</b>) Gridle view, (<b>A</b>,<b>C</b>) Apical poles showing valve, valvocopula, and copula, (<b>B</b>) Valve middle showing valve, valvocopula, and copula, (<b>D</b>) External valve view, (<b>E</b>,<b>G</b>) External apical poles (<b>F</b>) External valve middle with bifacial annulus in the half between the center and margin, (<b>H</b>) Internal valve view, (<b>I</b>,<b>K</b>) Internal apical poles covering the valve chamber with three longitudinal rows of foramina and apical openings, (<b>J</b>) Internal chamber with three longitudinal rows of small circular foramina lines. Scale bars = 10 µm (<b>D</b>,<b>H</b>); 5 µm (<b>A</b>–<b>C</b>,<b>E</b>–<b>G</b>,<b>I</b>–<b>K</b>).</p>
Full article ">Figure 10
<p>Scanning electron microscope images of <span class="html-italic">Ardissonea</span> aff. <span class="html-italic">formosa</span> in Dokdo. (<b>A</b>) External valve view, (<b>B</b>,<b>D</b>) Apical ends in the external valve view, (<b>C</b>) Valve middle with grooved annulus (arrowhead), (<b>F</b>,<b>H</b>) Apical ends in the internal valve view, (<b>G</b>) Internal chamber with four longitudinal rows of small circular foramina lines. Scale bars = 10 µm (<b>A</b>,<b>E</b>); 5 µm (<b>B</b>–<b>D</b>,<b>F</b>–<b>H</b>).</p>
Full article ">Figure 11
<p>Scanning electron microscope images of <span class="html-italic">Ardissoneopsis dokdoensis</span> in Dokdo. (<b>A</b>) External valve view, (<b>B</b>,<b>C</b>) External apical poles with numerous marginal spines and radiate striae, (<b>D</b>,<b>E</b>) External basal poles with marginal spines and radiate striae, (<b>F</b>,<b>G</b>) Marginal spines along valve edges (arrowheads in subfigure (<b>G</b>)), (<b>H</b>–<b>J</b>) Valve middle with grooved annulus (arrowheads) lines. Scale bars = 10 µm (<b>A</b>); 1 µm (<b>B</b>–<b>J</b>).</p>
Full article ">Figure 12
<p>Scanning electron microscope images of <span class="html-italic">Ardissoneopsis dokdoensis</span> in Dokdo. (<b>A</b>) Internal valve view, (<b>B</b>,<b>C</b>) Internal apical poles with blurring transverse costae and absence of pseudosepta, (<b>D</b>,<b>E</b>) Internal basal poles without transverse costae and radiate striae, (<b>F</b>,<b>G</b>) Valve middle with grooved annulus (arrowhead in subfigure (<b>F</b>)). Scale bars = 10 µm (<b>A</b>); 1 µm (<b>B</b>–<b>G</b>).</p>
Full article ">Figure 13
<p>Scanning electron microscope images of <span class="html-italic">Ardissoneopsis fulgicans</span> in Dokdo. (<b>A</b>) External valve view, (<b>B</b>) External apical pole with radiate striae, (<b>C</b>) External valve middle showing marginal bifacial annulus with groove (arrowhead), (<b>D</b>) Internal valve view, (<b>E</b>) Internal apical poles with weakened costae and without pseudosepta (arrowhead), (<b>F</b>) Internal valve middle showing transverse costae and marginal bifacial annulus with groove (arrowhead), Scale bars = 10 µm (<b>A</b>,<b>D</b>); 1 µm (<b>B</b>,<b>C</b>,<b>E</b>,<b>F</b>).</p>
Full article ">Figure 14
<p>Scanning electron microscope images of <span class="html-italic">Climacosphenia moniligera</span> in Dokdo. (<b>A</b>) External valve view, (<b>B</b>) External apical pole with radiate striae and apical spines, (<b>C</b>) External bifacial annulus with groove (arrowhead), (<b>D</b>) External basal pole with sparse areolae, (<b>E</b>) Internal valve view, (<b>F</b>) Internal apical pole with craticular bar, (<b>G</b>) Internal transapical bifacial costae with groove (arrowhead), (<b>H</b>) Internal basal pole without costae. Scale bars = 10 µm (<b>A</b>,<b>E</b>); 1 µm (<b>B</b>–<b>D</b>,<b>F</b>–<b>H</b>).</p>
Full article ">Figure 15
<p>Scanning electron microscope images of <span class="html-italic">Grunowago pacifica</span> in Dokdo. (<b>A</b>) External valve view, (<b>B</b>,<b>D</b>) External apical poles with marginal spines (arrowheads), (<b>C</b>) External valve middle with distinct trace of central long costa, (<b>E</b>) Internal valve view, (<b>F</b>,<b>H</b>) Internal apical poles with distinct costae, (<b>G</b>) Internal valve middle with distinct central long costa and transverse costae. Scale bars = 10 µm (<b>A</b>,<b>E</b>); 1 µm (<b>B</b>–<b>D</b>,<b>F</b>–<b>H</b>).</p>
Full article ">Figure 16
<p>Scanning electron microscope images of <span class="html-italic">Synedrosphenia crystallina</span> in Dokdo. (<b>A</b>) External valve view, (<b>B</b>,<b>D</b>) External apical poles with radiate striae, (<b>C</b>) External valve middle showing bifacial annulus with groove (arrowhead), (<b>E</b>) Internal valve view, (<b>F</b>,<b>H</b>) Internal apical poles with weakened costae and pseudosepta (arrowhead). (<b>G</b>) Internal valve middle showing transapical costae with groove (arrowhead), Scale bars = 10 µm (<b>A</b>,<b>E</b>); 1 µm (<b>B</b>–<b>D</b>,<b>F</b>–<b>H</b>).</p>
Full article ">Figure 17
<p>Scanning electron microscope images of <span class="html-italic">Synedrosphyenia</span> cf. <span class="html-italic">gomphonema</span> in Dokdo. (<b>A</b>,<b>B</b>) External valve view and both individuals showing the different valve face undulation, (<b>C</b>) External apical pole with rostrate apex and slightly radial striae toward apex, (<b>D</b>) External valve middle showing the marginal bifacial annulus with groove (arrowhead), (<b>E</b>) External basal pole with rounded apex, (<b>F</b>) Internal valve view, (<b>G</b>) Internal apical pole with slightly radiate costae and pseudosepta (arrowhead), (<b>H</b>) Internal valve middle showing the marginal bifacial annulus with grooved annulus (arrowhead) and transverse costae, (<b>I</b>) Internal basal pole without costae and with pseudosepta (arrowhead). Scale bar = 10 µm (<b>A</b>,<b>B</b>,<b>F</b>); 1 µm (<b>C</b>–<b>E</b>,<b>G</b>–<b>I</b>).</p>
Full article ">Figure 18
<p>Scanning electron microscope images of two multipolar diatoms in Dokdo. (<b>A</b>–<b>I</b>) <span class="html-italic">Biddulphiopsis titiana</span>, (<b>B</b>) Girdle view, (<b>C</b>) Valvocopula, (<b>D</b>) Valve mantle and pseudocelli in the corner, (<b>E</b>) External valve view, (<b>E</b>) Central annulus in the external valve view, (<b>F</b>) Pseudoocelli in the external valve view, (<b>G</b>) Internal valve view, (<b>H</b>) Central annulus in the internal valve view, (<b>I</b>) Internal structure of two rimoportulae, (<b>J</b>–<b>O</b>) <span class="html-italic">Lampriscus shadboltianus</span>, (<b>J</b>) External valve view, (<b>K</b>) Central annulus in the external valve view, (<b>L</b>) Ocelli in the external valve view, (<b>M</b>) Internal valve view, (<b>N</b>) Central annulus in the internal valve view, (<b>O</b>) Internal structure of two rimoportulae. Scale bar = 10 µm (<b>A</b>,<b>B</b>,<b>D</b>,<b>G</b>,<b>J</b>,<b>M</b>); 1 µm (<b>C</b>,<b>E</b>,<b>F</b>,<b>H</b>,<b>I</b>,<b>K</b>,<b>L</b>,<b>N</b>,<b>O</b>).</p>
Full article ">
14 pages, 1877 KiB  
Article
Description of Oryzobacter telluris sp. nov., a New Species Isolated from Bank-Side Soil in Seomjin River, South Korea
by Ahyoung Choi, Sumin Jang and Jaeduk Goh
Diversity 2024, 16(11), 689; https://doi.org/10.3390/d16110689 - 12 Nov 2024
Viewed by 321
Abstract
A novel bacterial strain, designated 24SJ04S-52T, was isolated from bank-side soil in the Osucheon Stream of the Seomjin River, Republic of Korea. This strain is aerobic, Gram-stain-positive, and short-rod-shaped, with optimal growth observed at 30 °C, pH 7, and 0% salinity, [...] Read more.
A novel bacterial strain, designated 24SJ04S-52T, was isolated from bank-side soil in the Osucheon Stream of the Seomjin River, Republic of Korea. This strain is aerobic, Gram-stain-positive, and short-rod-shaped, with optimal growth observed at 30 °C, pH 7, and 0% salinity, and growth occurring across a temperature range of 15–37 °C, pH 5–9, and salinity of 0–4%. Phylogenetic analysis of the 16S rRNA gene showed that strain 24SJ04S-52T shares 98.3% sequence similarity with Oryzobacter terrae PSGM2-16T. However, the average nucleotide identity (ANI) and digital DNA–DNA hybridization (dDDH) values were 85.0% and 50.1%, respectively, which are well below the species delineation thresholds of 95–96% for ANI and 70% for dDDH, confirming the novelty of this species. Genomic analysis identified a genome size of 3.98 Mb with a G+C content of 72.9 mol%. Functional annotation revealed various genes involved in amino acid, carbohydrate, and protein metabolism, suggesting metabolic versatility that may support adaptation to nutrient-variable environments. Chemotaxonomic analyses revealed distinctive profiles, including major fatty acids such as C17:1 ω8c, iso-C16:0, and iso-C14:0, with MK-8(H4) as the predominant menaquinone. The polar lipids included diphosphatidylglycerol, phosphatidylethanolamine, and phosphatidylinositol, and the peptidoglycan was of type A4γ with meso-diaminopimelic acid as the diagnostic diamino acid. These comprehensive analyses support the classification of strain 24SJ04S-52T as a novel species within the genus Oryzobacter, for which the name Oryzobacter telluris sp. nov. is proposed. The type strain is 24SJ04S-52T (=KACC 23836T = FBCC-B16192T). Full article
Show Figures

Figure 1

Figure 1
<p>Neighbor-joining phylogenetic tree based on 16S rRNA gene sequences showing positions of strain 24SJ04S-52<sup>T</sup> and other closely related members of the family <span class="html-italic">Intrasporangiaceae</span>. Bootstrap values (expressed as percentages of 1000 replications) over 70% are shown at nodes for neighbor-joining, maximum-likelihood, and maximum-parsimony methods, respectively. Filled circles indicate that the corresponding nodes were recovered by all treeing methods. An open circle indicates that the corresponding node was recovered by the neighbor-joining and maximum-likelihood methods. The minus sign indicates bootstrap values below 70% in the maximum-parsimony method. The bold font represents the novel species identified in this study. <span class="html-italic">Acidimicrobium ferrooxidans</span> DSM 10331<sup>T</sup> (CP001631) was used as an out-group. Bar, 0.02 substitutions per nucleotide position.</p>
Full article ">Figure 2
<p>Phylogenetic tree of strain 24SJ04S-52<sup>T</sup> and closely related type strains, inferred from GBDP distances calculated from genome sequences. Branch lengths are scaled according to the GBDP distance formula d5. Numbers near the branches represent GBDP pseudo-bootstrap support values &gt;70% based on 100 replications. The tree is rooted at the midpoint [<a href="#B49-diversity-16-00689" class="html-bibr">49</a>]. Leaf labels indicate species and subspecies affiliations, genomic G+C content, δ values, overall genome sequence length, protein count, and strain type [<a href="#B42-diversity-16-00689" class="html-bibr">42</a>]. The colors of the leaf labels denote species and subspecies clusters, with the novel species identified in this study highlighted in bold.</p>
Full article ">Figure 3
<p>Circular map of the strain 24SJ04S-52<sup>T</sup> genome. From outside to the center: the colored bands in ring 1 represent contigs; ring 2 represents the annotated genes on the forward strand (color determined by COG category); ring 3 shows the annotated genes on the reverse strand (color determined by COG category); ring 4 displays the RNA genes (rRNAs are displayed in red and tRNAs are displayed in purple); ring 5 shows the GC skew (higher-than-average values are displayed in green, while lower-than-average values are displayed in red); and ring 6 shows the GC ratio (higher-than-average values in blue and lower-than-average values in yellow).</p>
Full article ">Figure 4
<p>Orthologous gene cluster analysis using OrthoVenn3, depicting shared orthologous clusters among two <span class="html-italic">Oryzobacter</span> species. Numbers adjacent to species indicate the total clusters in each list.</p>
Full article ">Figure 5
<p>Transmission electron micrograph of strain 24SJ04S-52<sup>T</sup> cells. Bar, 0.5 µm.</p>
Full article ">
21 pages, 4216 KiB  
Article
Linking Biodiversity and Functional Patterns of Estuarine Free-Living Nematodes with Sedimentary Organic Matter Lability in an Atlantic Coastal Lagoon (Uruguay, South America)
by Noelia Kandratavicius, Luis Giménez, Catalina Pastor de Ward, Natalia Venturini and Pablo Muniz
Diversity 2024, 16(11), 688; https://doi.org/10.3390/d16110688 - 12 Nov 2024
Viewed by 330
Abstract
We examined the taxonomical and functional traits of free-living nematodes, focusing on their density by genus, maturity index (MI), and trophic diversity index (ITD) to determine whether these indices are sensitive to changes in the organic content of the sediment. Samples were collected [...] Read more.
We examined the taxonomical and functional traits of free-living nematodes, focusing on their density by genus, maturity index (MI), and trophic diversity index (ITD) to determine whether these indices are sensitive to changes in the organic content of the sediment. Samples were collected in autumn and spring from 12 subtidal sampling stations in Rocha Lagoon, distributed between the outer (near the mouth) and the inner sector. We identified 26 genera, with higher abundance in the inner sector, likely due to increased organic matter and biopolymers. In spring, both sectors had sediments rich in fresh organic matter, dominated by deposit-feeding nematodes and showing low trophic diversity (high ITD values). In autumn, the inner sector maintained similar characteristics to spring sampling, while the outer one was dominated by older organic matter, predatory nematodes and higher trophic diversity. The MI showed low variation between sectors, suggesting a disturbed environment. Our findings support the use of ITD to assess other aspects of communities such as the response of trophic groups to the freshness of organic matter, while the MI seems less effective for assessing the ecological status of Rocha Lagoon. Understanding nematode biodiversity and functional traits is crucial for effective ecological quality assessments. Full article
(This article belongs to the Special Issue Biodiversity as Tools to Assess Impacts on Coastal Ecosystems)
Show Figures

Figure 1

Figure 1
<p>Map of the study area showing the sampling sites located in the inner (I1–I6) and outer (O1–O6) sectors of Rocha Lagoon.</p>
Full article ">Figure 2
<p>Grain size characteristics in sites of the inner (I) and outer (O) sectors of Rocha Lagoon, for both sampling times: (<b>a</b>) spring; (<b>b</b>) autumn.</p>
Full article ">Figure 3
<p>Organic matter and photosynthetic pigment content in sediments of sites located in the outer (O1–O6) and inner (I1–I6) sectors of Rocha Lagoon for both sampling times. Error bars are standard deviations. (<b>a</b>) Organic matter. (<b>b</b>) Phaeopigments. (<b>c</b>) Chlorophyll a.</p>
Full article ">Figure 4
<p>Concentrations of biopolymers in sediments of sites in the inner (I) and outer (O) sectors of Rocha Lagoon for the two sampling times. Error bars are standard deviations. (<b>a</b>) Spring. (<b>b</b>) Autumn.</p>
Full article ">Figure 5
<p>Principal component analysis of the inner (<span style="color:red">I</span>, <span style="color:red">I+</span>) and outer (<span style="color:red">O</span>) sectors of Rocha Lagoon for both sampling times based on environmental and biochemical variables. OM: organic matter, C sand: %coarse sand; M sand: %medium sand; F sand: %fine sand; BPC: biopolymeric carbon; PRT: total proteins, CHO: total carbohydrates. Oa: outer autumn, Ia: inner autumn; Os: outer spring; Is: inner spring; I+: inner sites in old sandbar.</p>
Full article ">Figure 6
<p>Proportion of the most abundant genera (%) discriminated by time and sector.</p>
Full article ">Figure 7
<p>Abundance (mean and SD) of the main genera in sediments from the inner (I) and outer (O) sites of Rocha Lagoon for both sampling times. Error bars are standard deviations.</p>
Full article ">Figure 8
<p>Mean and SD of total abundance of nematodes (<b>a</b>) and richness of genera (<b>b</b>) from the inner (I) and outer (O) sites of Rocha Lagoon for both sampling times. Error bars are standard deviations.</p>
Full article ">Figure 9
<p>nMDS diagram of the analysis performed on the nematode genus matrix (<b>left</b>) and on the trophic group matrix (<b>right</b>) for autumn sampling (<b>top</b>) and spring sampling (<b>bottom</b>): (<b>a</b>) genera/autumn; (<b>b</b>) genera/spring; (<b>c</b>) trophic groups/autumn; (<b>d</b>) trophic groups/spring.</p>
Full article ">Figure 10
<p>Mean and SD of trophic diversity index (ITD) and maturity index (MI) of nematode assemblages in sediments from the inner (I) and outer (O) sites of Rocha Lagoon for both sampling times. Error bars are standard deviations.</p>
Full article ">Figure 11
<p>Mean and SD abundance of feeding types (%) of nematode assemblages in sediments from the inner (I) and outer (O) sites of Rocha Lagoon for both sampling samplings. Error bars are standard deviations. Deposit feeder (<b>1A</b>), nonselective deposit feeder (<b>1B</b>), epigrowth feeder (<b>2A</b>), and omnivore/predator (<b>2B</b>).</p>
Full article ">
23 pages, 4168 KiB  
Article
The Taxonomic and Functional Diversity of Leaf-Litter Dwelling Ants in the Tropical Dry Forest of the Colombian Caribbean
by Jose Camargo-Vanegas, Sebastian de la Hoz-Pedraza, Hubert Sierra-Chamorro and Roberto J. Guerrero
Diversity 2024, 16(11), 687; https://doi.org/10.3390/d16110687 - 11 Nov 2024
Viewed by 485
Abstract
There have been few advances in understanding the organization and dynamics of ants in tropical dry forests. The latter are a seriously threatened ecosystem, and ants are important indicators of diversity, disturbance, and restoration in forest ecosystems. Using diversity data and morphofunctional traits, [...] Read more.
There have been few advances in understanding the organization and dynamics of ants in tropical dry forests. The latter are a seriously threatened ecosystem, and ants are important indicators of diversity, disturbance, and restoration in forest ecosystems. Using diversity data and morphofunctional traits, we evaluated the spatial and temporal variation of taxonomic and functional ant groups; in addition, we explored the variation in functional traits and diversity among communities. Ants were sampled during the dry and rainy seasons using mini-Winkler bags. A total of 9 subfamilies, 57 genera, and 146 species were collected. Ant species composition and richness varied both spatially (75 to 119 species) and temporally (121 and 127 species). The fragments from N2 and N3 showed higher diversity than those from N1. The dissimilarity among all areas was moderate (50–60%), mainly attributable to species turnover processes (77%). Twenty functional groups were identified. The N3 fragments had the highest functional diversity, with lower resistance to species loss, while the N1 and N2 fragments reduced functional diversity and increased similarity among species. Our results highlight the importance of integrating a functional analysis with the taxonomic assessment of ants as an important contribution to understanding the organization and dynamics of this community of insects that inhabit the tropical dry forest. Full article
(This article belongs to the Special Issue Biodiversity in Arid Ecosystems)
Show Figures

Figure 1

Figure 1
<p>A map showing sampling sites of tropical dry forest in the northwest of the Colombian Caribbean. N1a: Luriza Integrated Regional Management District; N1b: “Palmar del Titi” Integrated Regional Management District; N2a: “Los Colorados” Flora and Fauna Sanctuary; N2b: Brasilar Tropical Dry Forest Reserve; N3a: “CARACOLÍ” Civil Society Nature Reserve; N3b: Coraza and Montes de María Protective Reserve. The green polygons correspond to dry forest fragments distributed both in the study area and in adjacent regions.</p>
Full article ">Figure 2
<p>Number of ant species by subfamilies (horizontal axis) for TDF fragments in each area of the northwestern Colombian Caribbean.</p>
Full article ">Figure 3
<p>Diversity expressed as the effective number of ant species (<sup>q</sup>D) in TDF fragments in each area. (<b>a</b>) Richness (<sup>0</sup>D); (<b>b</b>) common species (<sup>1</sup>D); (<b>c</b>) dominant species (<sup>2</sup>D). The bars indicate the confidence intervals (CI) of each of the measurements.</p>
Full article ">Figure 3 Cont.
<p>Diversity expressed as the effective number of ant species (<sup>q</sup>D) in TDF fragments in each area. (<b>a</b>) Richness (<sup>0</sup>D); (<b>b</b>) common species (<sup>1</sup>D); (<b>c</b>) dominant species (<sup>2</sup>D). The bars indicate the confidence intervals (CI) of each of the measurements.</p>
Full article ">Figure 4
<p>Rank–abundance curves showing the distribution of capture frequencies (relative abundance) for the ant assemblage in the TDF fragments in each area. The names of the species with the highest capture frequency (≥50%) are shown.</p>
Full article ">Figure 5
<p>Ordination analysis using non-metric multidimensional scaling (nMDS) for the leaf-litter-associated ant community in the studied areas. Letters in numbers are defined in the study area section. D corresponds to the dry season, while R is the rainy season.</p>
Full article ">Figure 6
<p>Partition of beta diversity (βjac) into its turnover (βjtu) and nestedness (βjne) components between pairs of sites in each area based on the Jaccard index. (<b>a</b>) Rainy season; (<b>b</b>) dry season.</p>
Full article ">Figure 6 Cont.
<p>Partition of beta diversity (βjac) into its turnover (βjtu) and nestedness (βjne) components between pairs of sites in each area based on the Jaccard index. (<b>a</b>) Rainy season; (<b>b</b>) dry season.</p>
Full article ">Figure 7
<p>The spatial variation in the functional groups recorded in the TDF fragments in each area.</p>
Full article ">Figure 8
<p>The functional trait weighted mean (CWM) of the eight functional traits of the ant communities located in the TDF fragments in each area: (<b>a</b>) head length (HL); (<b>b</b>) head width (HW); (<b>c</b>) mandible length (ML); (<b>d</b>) eye length (EL); (<b>e</b>) interocular distance (DI); (<b>f</b>) scape length (SL); (<b>g</b>) femur length (FL); and (<b>h</b>) Weber length (WL). In each graph, the different letters associated with the TDF fragments represent significant differences in the CWM values evaluated by Tukey’s post hoc analysis. All morphometric traits are expressed in mm. The boxes represent the interquartile range between the first and third quartiles (25th and 75th percentile, respectively), and the horizontal line indicates the median. The whiskers represent the maximum and minimum values. The solid dots represent outliers. Orange boxes: N1; Green boxes: N2; Blue boxes: N3.</p>
Full article ">Figure 9
<p>Estimated functional diversity indices for ant communities located in the TDF fragments in each area. (<b>a</b>) Functional richness (FRic); (<b>b</b>) functional evenness (FEve); (<b>c</b>) functional redundancy (Fred); and (<b>d</b>) Rao’s quadratic entropy (Rao’s Q). In each graph, the different letters associated with the TDF fragments represent significant differences in the values of each index evaluated by Tukey’s post hoc analysis. The boxes represent the interquartile range between the first and third quartiles (25th and 75th percentile, respectively), and the horizontal line indicates the median. The whiskers represent the maximum and minimum values. The solid dots represent outliers. Orange boxes: N1; Green boxes: N2; Blue boxes: N3.</p>
Full article ">
27 pages, 3348 KiB  
Article
Spatial Distribution and Temporal Variation of Megafauna in Circalittoral and Bathyal Soft Bottoms of the Westernmost Biodiversity Hotspot of the Mediterranean Sea: The Alboran Ridge
by Cristina García-Ruiz, Manuel Hidalgo, Cristina Ciércoles, María González-Aguilar, Pedro Torres, Javier Urra and José L. Rueda
Diversity 2024, 16(11), 686; https://doi.org/10.3390/d16110686 - 10 Nov 2024
Viewed by 511
Abstract
The Alboran Sea is the westernmost sub-basin of the Mediterranean Sea, and it is connected to the Atlantic Ocean through the Strait of Gibraltar. The Alboran Ridge is located in the middle of the Alboran Sea and represents a hotspot of biodiversity in [...] Read more.
The Alboran Sea is the westernmost sub-basin of the Mediterranean Sea, and it is connected to the Atlantic Ocean through the Strait of Gibraltar. The Alboran Ridge is located in the middle of the Alboran Sea and represents a hotspot of biodiversity in the Mediterranean Sea. Besides their critical importance, there are few studies on the communities and changes in biodiversity, and they mostly concentrate on infralittoral and circalittoral bottoms. In this work, the composition, structure and bathymetric and temporal changes of megafauna of the Alboran Ridge were examined. Samples were collected from MEDITS surveys carried out between 2012 and 2022 at depths ranging from 100 to 800 m. Analyses were performed separately for each of the taxonomic groups: osteichthyes, chondrichthyes, crustaceans, molluscs, echinoderms and “other groups”. There was no common spatial organization for each of the faunistic groups studied, although most of them displayed differences between the shelf and the slope. The continental shelf was characterized by the highest values of community metrics such as abundance, biomass, species richness and mean weight of species for all groups except for chondrichthyes and crustaceans. Decreasing trends of some community metrics were detected in some of the faunistic groups throughout the study period. Full article
(This article belongs to the Section Marine Diversity)
Show Figures

Figure 1

Figure 1
<p>Map of the study area (Alboran Ridge) showing samples from the MEDITS survey from 2012 to 2022 (black points). Bottom trawling is not allowed at depths of less than 100 m depth and more than 1000 m depth (Ministerial Order, B.O.E. 233, ref. A-1998-22,628). No survey could be conducted in 2020 due to COVID-19 lockdown restrictions.</p>
Full article ">Figure 2
<p>Results of cluster and nMDS analyses for each faunistic group carried out with abundance data (log (x + 1) transformed) from MEDITS 2012–2022. CS: 100–200 m depth; US: 201–500 m; MS: 501–800 m.</p>
Full article ">Figure 3
<p>Mean values (± standard errors) of community metrics: (<b>a</b>) abundance (number of individuals/km<sup>2</sup>) (log (x + 1) transformed); (<b>b</b>) biomass (g/km<sup>2</sup>) (log (x + 1) transformed); (<b>c</b>) mean weight (g); (<b>d</b>) species richness (S); (<b>e</b>) Shannon–Wiener diversity index (H’). CS: 100–200 m depth; US: 201–500 m depth; MS: 501–800 m depth.</p>
Full article ">Figure 4
<p>Mean values (±standard errors) of community metrics of different faunistic groups on CS (left panel), US (middle panel) and MS (right panel). (<b>a</b>) Abundance (number of individuals/km<sup>2</sup>) (log (x + 1) transformed); (<b>b</b>) biomass (g/km<sup>2</sup>) (log (x + 1) transformed); (<b>c</b>) mean weight (g) (logarithmic scale); (<b>d</b>) species richness (S); (<b>e</b>) Shannon–Wiener diversity index (H’). CS: 100–200 m depth; US: 201–500 m depth; MS: 501–800 m depth.</p>
Full article ">
12 pages, 2854 KiB  
Article
Human Activity Changed the Genetic Pattern of the Orchid Phaius flavus Population
by Cuiyi Liang, Jun Li, Shixing Li, Huayuan Zhang, Jiahao Zheng, Jianglin Miao, Siyuan Hao, Shasha Wu, Zhongjian Liu and Junwen Zhai
Diversity 2024, 16(11), 685; https://doi.org/10.3390/d16110685 - 8 Nov 2024
Viewed by 334
Abstract
Human activity often has profound effects on plant growth and evolution. Orchids are the most diverse group of flowering plants and are threatened by habitat fragmentation, over-harvesting, and urbanization. A population of Phaius flavus from Beikengding Mount (BM) in the Fujian Province of [...] Read more.
Human activity often has profound effects on plant growth and evolution. Orchids are the most diverse group of flowering plants and are threatened by habitat fragmentation, over-harvesting, and urbanization. A population of Phaius flavus from Beikengding Mount (BM) in the Fujian Province of China was divided into two patches by road construction. This study evaluated its genetic characteristics using restriction site-associated DNA sequencing (RAD-seq) data, more than seven years post-road construction. The purpose of this study was to explore the impact of road construction on the evolution of isolated patches within a population. The analysis revealed that the genetic diversity of patch B was slightly higher than that of patch A in the BM population of P. flavus. Principal component and phylogenetic analyses, genetic structure and genetic differentiation analysis, and bottleneck detection indicated relatively independent genetic differentiation between the two patches. Thus, the construction of the Y013 village road may have influenced different patches of this population on a genetic level. This study provides a case for understanding the impact of specific human activities on plant populations, and then biodiversity conservation. It is conducive to formulating more effective biological protection strategies to mitigate the damage inflicted by human activities on biodiversity. Full article
Show Figures

Figure 1

Figure 1
<p>Location of the studied area: (<b>a</b>) BM population of the <span class="html-italic">P. flavus</span> with the location of patches A and B; (<b>b</b>) habitat of patch A; (<b>c</b>) habitat of patch B. Map from <a href="http://lbs.tianditu.gov.cn/home.html" target="_blank">http://lbs.tianditu.gov.cn/home.html</a> (accessed on 25 November 2023) and <a href="http://www.arcgis.com/index.html" target="_blank">http://www.arcgis.com/index.html</a> (accessed on 25 November 2023).</p>
Full article ">Figure 2
<p>Principal component analysis of patch A and B of the <span class="html-italic">P. flavus</span> populations. Patch A is clustered in the second quadrant (with a few individuals distributed across the other three quadrants), while patch B is distributed in the first and fourth quadrants.</p>
Full article ">Figure 3
<p>The phylogenetic analysis of <span class="html-italic">P. flavus</span> populations. Patch A and patch B can be clearly distinguished. The red color marks the branch of patch A, and the yellow color marks the branch of patch B. The numbers on the branches indicate bootstrap support values (in %).</p>
Full article ">Figure 4
<p>Population structure between patches A and B of the <span class="html-italic">P. flavus</span> population. It shows ancestry coefficients for K = 2. The <span class="html-italic">x</span>-axis shows the different individuals from patch A and patch B; the <span class="html-italic">y</span>-axis quantifies the proportion of an individual’s variation from inferred ancestral populations.</p>
Full article ">
15 pages, 3083 KiB  
Article
Bio-Cultural Diversity for Food Security: Traditional Wild Food Plants and Their Folk Cuisine in Lakki Marwat, Northwestern Pakistan
by Tehsin Ullah, Shujaul Mulk Khan, Abdullah Abdullah, Naji Sulaiman, Ateef Ullah, Muhammad Sirab Khan, Shakil Ahmad Zeb and Andrea Pieroni
Diversity 2024, 16(11), 684; https://doi.org/10.3390/d16110684 - 8 Nov 2024
Viewed by 448
Abstract
Ethnobotanical studies on foraging are essential for documenting neglected or previously unknown wild food plants, which may be crucial for promoting the diversification of food sources and contributing to food security and sovereignty. The Pashtuns of the Marwat tribe in NW Pakistan are [...] Read more.
Ethnobotanical studies on foraging are essential for documenting neglected or previously unknown wild food plants, which may be crucial for promoting the diversification of food sources and contributing to food security and sovereignty. The Pashtuns of the Marwat tribe in NW Pakistan are renowned for their traditional customs and food systems. Studying the wild food plants (WFPs) and their associated bio-cultural diversity is quintessential for fostering food security and sovereignty in the region. The research presented here investigated the area’s wild food plants traditionally gathered and consumed. The field survey was conducted in 2023 with 87 study participants. A total of 41 plant species belonging to 24 botanical families was documented. The findings include food uses for Atriplex tatarica, Amaranthus graecizans, and Beta vulgaris subsp. maritima that have rarely been recorded in Pakistan. Moreover, the use of Citrulus colocynthus fruits in jam and Zygophyllum indicum leaves and stems in beverages are novel contributions to the gastronomy of NW Pakistan. The comparison with other food ethnobotanical studies conducted in North Pakistan suggests some similarities between the Lakki Marwat traditional WFPs and those from other semi-arid areas in North Pakistan, both Pashtun and non-Pashtun. While the findings underline the significant role of WFPs in local cuisine, we observed that this local knowledge is also threatened: the rapid spread of fast and industrialized food, modernization, and cultural dilution has led to an alarming reduction in these practices among the younger generations. Therefore, suitable measures to safeguard traditional plants, food knowledge, practices, and the associated culture are urgently needed. The urgency of this situation cannot be overstated, and it is crucial that we act now. Furthermore, preserving wild food plant-related cultural heritage may be fundamental to promoting food security and public health. Full article
Show Figures

Figure 1

Figure 1
<p>Typical landscape of Lakki Marwat (photo credit: T.U.).</p>
Full article ">Figure 2
<p>Study area.</p>
Full article ">Figure 3
<p>Chord diagram showing WFP consumption categories in the study area.</p>
Full article ">Figure 4
<p>Some popular wild, leafy vegetables in the study area: (<b>A</b>): leaves of <span class="html-italic">Eruca vesicaria</span>; (<b>B</b>): <span class="html-italic">Beta vulgaris</span> subsp. <span class="html-italic">maritima</span>; (<b>C</b>): <span class="html-italic">Oxalis corniculata</span>; (<b>D</b>): <span class="html-italic">Medicago polymorpha</span>. (photo credit: T.U.).</p>
Full article ">
41 pages, 10663 KiB  
Article
Forty-Five Years of Caterpillar Rearing in Area de Conservación Guanacaste (ACG) Northwestern Costa Rica: DNA Barcodes, BINs, and a First Description of Plant–Caterpillar–Ichneumonoid Interactions Detected
by Donald L. J. Quicke, Daniel H. Janzen, Winnie Hallwachs, Mike J. Sharkey, Paul D. N. Hebert and Buntika A. Butcher
Diversity 2024, 16(11), 683; https://doi.org/10.3390/d16110683 - 7 Nov 2024
Viewed by 1008
Abstract
Foliage-feeding wild caterpillars have been collected and reared year-round by 1–30 rural resident parataxonomists in the Area de Conservación Guanacaste (ACG) in northwestern Costa Rica since 1978. The aim of the work was to describe the diversity and interactions of Lepidoptera and their [...] Read more.
Foliage-feeding wild caterpillars have been collected and reared year-round by 1–30 rural resident parataxonomists in the Area de Conservación Guanacaste (ACG) in northwestern Costa Rica since 1978. The aim of the work was to describe the diversity and interactions of Lepidoptera and their associations with larval food plants and parasitoids in a diverse tropical community. A total of 457,816 caterpillars developed into a moth or butterfly, and these were identified to the family and species/morphospecies, with 151,316 having been successfully barcoded and assigned a Barcode Index Number (BIN) and/or “scientific name”. The host food plant was usually identified to the species or morphospecies. In addition to adult moths and butterflies, rearings also yielded many hundreds of species of parasitic wasps and tachinid flies, many of which were also DNA-barcoded and assigned a name and/or BIN. Increasingly over recent years, these have been identified or described by expert taxonomists. Here, we provide a summary of the number of species of ichneumonoid (Ichneumonidae and Braconidae) parasitoids of the caterpillars, their hosts, the host food plants involved, the bi- and tritrophic interactions, and their relationships to the caterpillar sampling effort. The dataset includes 16,133 and 9453 independent rearings of Braconidae and Ichneumonidae, respectively, collectively representing 31 subfamilies, all with parasitoid barcodes and host and host food plant species-level identifications. Host caterpillars collectively represented 2456 species, which, in turn, were collectively eating 1352 species of food plants. Species accumulation curves over time for parasitoids, hosts, and plants show various asymptotic trends. However, no asymptotic trends were detected for numbers of unique parasitoid–host and host–plant bitrophic interactions, nor for tritrophic interactions, after 1983, because climate change then began to conspicuously reduce caterpillar densities. Parasitoid host ranges, the proportions of specialists at the host species and host genus levels, host family utilisation, and host guild sizes show some differences among taxa and are discussed in turn. Ichneumonidae are shown to preferentially parasitise caterpillars of larger-bodied hosts compared to Braconidae. Several of the host plant species from which caterpillars were collected have been introduced from outside of the Americas and their utilisation by endemic parasitoids is described. The obligately hyperparasitoid ichneumonid subfamily Mesochorinae is dealt with separately and its strong association with microgastrine braconid primary parasitoids is illustrated. We discuss the implications for studies of tropical insect community food web ecology and make suggestions for future work. The aim was to make available the data from this remarkable study and to provide an overview of what we think are some of the more interesting relationships that emerge—other scientists/readers are expected to have different questions that they will go on to explore the data to answer. Full article
(This article belongs to the Section Animal Diversity)
Show Figures

Figure 1

Figure 1
<p>Map showing the main habitat zones (zonas de vida or Life Zones), the centre of the population (pobaciones), and posts or biological stations (puestos o estaciones) in the ACG.</p>
Full article ">Figure 2
<p>Photographs from the late 1990s of rearing and voucher preparation at the ACG; the system has not changed dramatically since (but see below for current protocols). (<b>A</b>,<b>B</b>) Rearing bags containing caterpillars and food plants suspended in purpose-built airy rearing barns that keep them in the shade but under otherwise ambient conditions; (<b>C</b>) A parataxonomist refreshing the caterpillar food plant; (<b>D</b>) Setting up parasitoid cocoons for eclosion and record-keeping in the Santa Rosa rearing barn; (<b>E</b>) Rearing bottles inverted for the easier visibility of the parasitoids, with parasitoid cocoons waiting for eclosion (as eclosions occur, date labels are attached to the jar with details); (<b>F</b>) Earlier gas-powered drying cabinet for preserving/oven-drying larger eclosed insects after mounting; (<b>G</b>) Spread Lepidoptera with their associated data; (<b>H</b>) Pinned larger ichneumonids (mostly Ophioninae) with data labels (‘microhymenoptera’ are preserved in microvials of 95% ethanol); (<b>I</b>) Gelatine capsules, puparia and small cocoons, etc.</p>
Full article ">Figure 3
<p>Relationship between the numbers of braconid and ichneumonid species (BINs) recovered and the numbers reared by subfamily. Braconid subfamilies are indicated in red, and Ichneumonidae subfamilies in blue.</p>
Full article ">Figure 4
<p>Species accumulation curves for Braconidae (BINs), their hosts, and host food plants over time.</p>
Full article ">Figure 5
<p>Species accumulation curves for Ichneumonidae (BINs), their hosts, and host food plants over time.</p>
Full article ">Figure 6
<p>Unique bi- and tritrophic species interaction accumulation curves for all primary parasitoid Ichneumonoidea.</p>
Full article ">Figure 7
<p>Example tritrophic food web for the braconid subfamily Orgilinae (because it does not comprise too many species to illustrate). Columns from left to right are the parasitoid BINs, host species, host family, food plant species, and food plant family.</p>
Full article ">Figure 8
<p>Ranked relative proportions of braconid and ichneumonid species attacking those families from which more than 100 ichneumonoid wasp rearings were obtained.</p>
Full article ">Figure 9
<p>Histogram of the numbers of species of braconid and ichneumonid parasitoids in relation to the mean host forewing length. Colours have 50% transparency so the Ichneumonidae appear as blue and purple.</p>
Full article ">Figure 10
<p>Mosaic plot showing the numbers of rearings of each braconid subfamily from each host family. Parasitoid subfamilies with fewer than 15 rearings, and host families with fewer than 15 parasitisation records, are excluded. (Abbreviations: Agath = Agathidinae; Br = Braconinae; C = Cardiochilinae; Ch = Cheloninae; Eu = Euphorinae; H = Homolobinae; Ho = Hormiinae; Mac = Macrocentrinae; O = Orgilinae; R = Rhysipolinae; Ro = Rogadinae).</p>
Full article ">Figure 11
<p>Mosaic plot showing the numbers of rearings of each ichneumonid subfamily from each host family. Parasitoid subfamilies with fewer than 15 rearings, and host families with fewer than 15 parasitisation records, are excluded. (Abbreviations: An = Anomaloninae; Banch = Banchinae; Ich = Ichneumoninae; L = Lycorininae; Met = Metopiinae; Oph = Ophioninae; P = Pimplinae; T = Tryphoninae).</p>
Full article ">Figure 12
<p>Logistic regression plots of the numbers of host families, host species, host food plant families, and host food plant species parasitised by braconid wasps. Points represent each parasitoid species (BIN). The best-fit models were calculated using generalised linear models (glms) with binomial errors, and the fitted curves were produced using the predict (model) function.</p>
Full article ">Figure 13
<p>Logistic regression plots of numbers of host families, host species, host food plant families, and host food plant species parasitised by ichneumonid wasps. The fitted models were calculated using a glm with binomial errors (see the legend to <a href="#diversity-16-00683-f012" class="html-fig">Figure 12</a> for more details).</p>
Full article ">Figure 14
<p>Relationship between the logarithm of the number of rearings and the parasitoid assemblage size for each host species, plotted separately for Braconidae and Ichneumonidae. The top (presented as a bar chart rather than a histogram for clarity) shows the numbers of rearings for five bins of the log rearing number. The lower part is a scatterplot for each species with points jittered for clarity.</p>
Full article ">Figure 15
<p>Relationship between the mean parasitoid assemblage size for each host family and the logarithm of the number of rearings, plotted separately for Braconidae and Ichneumonidae. Both relationships are highly significant—Braconidae: F = 21.43<sub>(1,39)</sub>, <span class="html-italic">p</span> ≪ 0.001, and residual standard error = 0.111; Ichneumonidae: F = 8.909<sub>(1,28)</sub>, <span class="html-italic">p</span> ≪ 0.001, and residual standard error = 0.241.</p>
Full article ">Figure 16
<p>Preference of Mesochorinae for primary parasitoid host subfamilies (red) and numbers of primary hosts available, i.e., unparasitised (dark green), all scaled so the maxima of unparasitised and hyperparasitised Microgastrinae primaries are the same. The figure shows that mesochorines, although attacking a wide range of primary parasitoids, show a strong preference for microgastrines.</p>
Full article ">
17 pages, 1909 KiB  
Article
Diet Diversity of Two Sculpin Species (Cottidae) in Midwestern USA Trout Streams: Patterns Across Nine Years After Severe Summer Flooding
by Neal D. Mundahl
Diversity 2024, 16(11), 682; https://doi.org/10.3390/d16110682 - 7 Nov 2024
Viewed by 304
Abstract
The geographic ranges of slimy (Uranidea cognata) and mottled (Uranidea bairdii) sculpin overlap broadly across cool and coldwater streams and rivers in North America, where they can serve very important roles in fish community dynamics. The diet diversities of [...] Read more.
The geographic ranges of slimy (Uranidea cognata) and mottled (Uranidea bairdii) sculpin overlap broadly across cool and coldwater streams and rivers in North America, where they can serve very important roles in fish community dynamics. The diet diversities of slimy and mottled sculpin were examined in early March (late winter) during eight out of nine years after the August 2007 catastrophic flooding in four streams to assess potential diet shifts as benthic invertebrate prey communities recovered post-flood. In total, 10,823 prey items, representing 39 invertebrate taxa and three fish taxa were identified from the stomachs of 532 slimy sculpins (present in Garvin Brook, Gilmore Creek, and Trout Run) and 179 mottled sculpins (present in Middle Fork Whitewater River). Only four prey taxa were consumed by sculpin in all streams: midge larvae and pupae (Diptera: Chironomidae), blackfly larvae and pupae (Diptera: Simuliidae), Hydropsyche caddisfly larvae (Trichoptera: Hydropsychidae), and Baetis mayfly nymphs (Ephemeroptera: Baetidae). Midges dominated diets of both slimy (61% of prey by number) and mottled (76%) sculpin across all years. Consequently, Shannon diversities of diets were typically low across all years and streams for slimy sculpin (annual site range 0.07–0.83) and across years for mottled sculpin (annual range 0.11–0.46). Diversities and taxa richness of slimy sculpin diets increased in Garvin Brook and Trout Run across the study years (driven by significant declines in midge dominance) but remained relatively unchanged for slimy sculpin in Gilmore Creek and mottled sculpin in the Middle Fork. Individual slimy and mottled sculpin differed significantly both in the numbers of taxa consumed per fish (<2 versus 2.5 taxa/fish, respectively) and in the numbers of individual prey per fish (11 versus 26 prey, respectively). Slimy sculpin in two streams displayed modest shifts in diets as benthic prey communities recovered during the 9-year period post-flood, whereas slimy and mottled sculpin in other streams displayed little to no changes in diets. Differing flood severity among streams may have produced the different responses observed in sculpin diets. Full article
(This article belongs to the Section Freshwater Biodiversity)
Show Figures

Figure 1

Figure 1
<p>Map of the four study sites (red stars) in southeastern Minnesota, USA, where sculpin diets were examined. Inset at the lower left indicates the location of the study area (yellow star) in North America.</p>
Full article ">Figure 2
<p>Total length–wet mass relationships for sculpin from four streams in southeastern Minnesota, USA, March 2008–March 2016. Slimy sculpin were found in Garvin Brook, Gilmore Creek, and Trout Run, and mottled sculpin were collected in the Middle Fork Whitewater River.</p>
Full article ">Figure 3
<p>Percentages (100 × [number of midges/total number of prey organisms]) of sculpin diets consisted of midges (Chironomidae) in four streams in southeastern Minnesota, USA, March 2008–March 2016. Values represent totals for all fish combined. Slimy sculpin were found in Garvin Brook, Gilmore Creek, and Trout Run, and mottled sculpin were collected in the Middle Fork Whitewater River.</p>
Full article ">Figure 4
<p>Percentages of the diets of individual sculpin consisted of midges (Chironomidae) in four streams in southeastern Minnesota, USA, March 2008–March 2016. Slimy sculpin were found in Garvin Brook, Gilmore Creek, and Trout Run, and mottled sculpin were collected in the Middle Fork Whitewater River. Dotted lines represent best-fit simple least-squares linear regression lines. See text for regression details.</p>
Full article ">Figure 4 Cont.
<p>Percentages of the diets of individual sculpin consisted of midges (Chironomidae) in four streams in southeastern Minnesota, USA, March 2008–March 2016. Slimy sculpin were found in Garvin Brook, Gilmore Creek, and Trout Run, and mottled sculpin were collected in the Middle Fork Whitewater River. Dotted lines represent best-fit simple least-squares linear regression lines. See text for regression details.</p>
Full article ">Figure 5
<p>Shannon diversity and taxa richness of sculpin diets in four streams in southeastern Minnesota, USA, March 2008–March 2016. Slimy sculpin were found in Garvin Brook, Gilmore Creek, and Trout Run, and mottled sculpin were collected in the Middle Fork Whitewater River. Values are based on the combined diets of all fish on each site and date combination.</p>
Full article ">Figure 6
<p>Mean numbers of taxa and individual prey organisms in sculpin stomachs in four streams in southeastern Minnesota, USA, March 2008–March 2016. Slimy sculpin were found in Garvin Brook, Gilmore Creek, and Trout Run, and mottled sculpin were collected in the Middle Fork Whitewater River. Error bars excluded for clarity.</p>
Full article ">Figure 7
<p>Mean standardized prey mass in sculpin stomachs in four streams in southeastern Minnesota, USA, March 2008–March 2016. Slimy sculpin were found in Garvin Brook, Gilmore Creek, and Trout Run, and mottled sculpin were collected in the Middle Fork Whitewater River. Error bars excluded for clarity.</p>
Full article ">
21 pages, 6519 KiB  
Article
The Mosquitoes (Diptera: Culicidae) of Sonora: Distribution, Ecology, and the First Records of Aedes deserticola Zavortink and Toxorhynchites septentrionalis (Dyar and Knab) in México
by Aldo I. Ortega-Morales, Juan Manuel Quijano-Barraza, Mario A. Rodríguez-Pérez, Luis M. Hernández-Triana, Francisco Wong-Corral and Fabián Correa-Morales
Diversity 2024, 16(11), 681; https://doi.org/10.3390/d16110681 - 6 Nov 2024
Viewed by 1012
Abstract
The diversity and distribution of mosquitoes from Sonora, Mexico, was documented through entomologic surveys conducted in the four physiographic regions and sub-regions of Sonora: the Sonoran Plain, the Sierra Madre Occidental, the Northern Mountains and Plains, and the Pacific Coastal Plain. Immature stages [...] Read more.
The diversity and distribution of mosquitoes from Sonora, Mexico, was documented through entomologic surveys conducted in the four physiographic regions and sub-regions of Sonora: the Sonoran Plain, the Sierra Madre Occidental, the Northern Mountains and Plains, and the Pacific Coastal Plain. Immature stages were collected from aquatic habitats, while adult mosquitoes were collected using Shannon traps, resting in vegetation, and by human landing collections. Overall, 11,316 specimens, which comprised 493 larvae, 224 larval exuviae, 400 pupal exuviae, 33 pupae, 4552 females, 5607 males, and seven male genitalia, were identified. Two subfamilies: Anophelinae and Culicinae, seven tribes, 10 genera, 23 subgenera, and 56 species are reported below. Of these, one tribe, one genus, five subgenera, and 15 species were recorded for the first time in Sonora. Two species, Aedes deserticola Zavortink and Toxorhynchites septentrionalis (Dyar and Knab), are the first recordings of their kind nationwide. Toxorhynchites septentrionalis was also barcoded with the cytochrome oxidase subunit 1 (COI) gene. The presence of Psorophora columbiae (Dyar and Knab) is confirmed in Mexico. Taxonomic notes, new geographic distribution limits of mosquitoes in Sonora, and information regarding their importance as disease vectors are provided. By adding Ae. deserticola and Tx. septentrionalis and confirming Ps. columbiae in Mexico, there are currently 251 mosquito species in the Country. Full article
(This article belongs to the Special Issue Biodiversity in Arid Ecosystems)
Show Figures

Figure 1

Figure 1
<p>Physiography of Sonora state. Red box shows the study area. Map originally created for this article.</p>
Full article ">Figure 2
<p>Species accumulation curve for 45 mosquito species collected at 139 site collections in Sonora from 2021 through 2022.</p>
Full article ">Figure 3
<p>Bayesian phylogenetic tree of <span class="html-italic">COI</span> DNA barcoding for <span class="html-italic">Toxorhynchites</span> spp. Red box shows <span class="html-italic">Tx. septentrionalis</span> of this study.</p>
Full article ">Figure 4
<p><span class="html-italic">Aedes deserticola</span>. female. PSc: postcoxal area. SA: subspiracular area. Mem: metameron. Ppn: postpronotum.</p>
Full article ">Figure 5
<p><span class="html-italic">Toxorhynchites</span> septentrionalis, male.</p>
Full article ">
13 pages, 1702 KiB  
Article
Multiple Speciation and Extinction Rate Shifts Shaped the Macro-Evolutionary History of the Genus Lycium Towards a Rather Gradual Accumulation of Species Within the Genus
by Haikui Chen, Kowiyou Yessoufou, Xiu Zhang, Shouhe Lin and Ledile Mankga
Diversity 2024, 16(11), 680; https://doi.org/10.3390/d16110680 - 6 Nov 2024
Viewed by 467
Abstract
The Neotropics are the most species-rich region on Earth, and spectacular diversification rates in plants are reported in plants, mostly occurring in oceanic archipelagos, making Neotropical and island plant lineages a model for macro-evolutionary studies. The genus Lycium in the Solanaceae family, originating [...] Read more.
The Neotropics are the most species-rich region on Earth, and spectacular diversification rates in plants are reported in plants, mostly occurring in oceanic archipelagos, making Neotropical and island plant lineages a model for macro-evolutionary studies. The genus Lycium in the Solanaceae family, originating from the Neotropics and exhibiting a unique disjunct geography across several islands, is therefore expected to experience exceptional diversification events. In this study, we aimed to quantify the diversification trajectories of the genus Lycium to elucidate the diversification events within the genus. We compiled a DNA matrix of six markers on 75% of all the species in the genus to reconstruct a dated phylogeny. Based on this phylogeny, we first revisited the historical biogeography of the genus. Then, we fitted a Compound Poisson Process on Mass Extinction Time model to investigate the following key evolutionary events: speciation rate, extinction rate, as well as mass extinction events. Our analysis confirmed that South America is the origin of the genus, which may have undergone a suite of successive long-distance dispersals. Also, we found that most species arose as recently as 5 million years ago, and that the diversification rate found is among the slowest rates in the plant kingdom. This is likely shaped by the multiple speciation and extinction rate shifts that we also detected throughout the evolutionary history of the genus, including one mass extinction at the early stage of its evolutionary history. However, both speciation and extinction rates remain roughly constant over time, leading to a gradual species accumulation over time. Full article
(This article belongs to the Special Issue 2024 Feature Papers by Diversity’s Editorial Board Members)
Show Figures

Figure 1

Figure 1
<p>A dated phylogenetic tree of the genus <span class="html-italic">Lycium</span> from combined six genes based on Bayesian inference. The numbers above the branches represent a Bayesian posterior probability greater than 0.5 (PP &gt; 0.5), and the branches without PP values were PP &lt; 0.5.</p>
Full article ">Figure 2
<p>A graphical output from RASP showing the results of the ancestral reconstruction area from the BBM (Bayesian Binary Method) analysis. Pie charts at each node show the probabilities of alternative ancestral ranges. The green circles around the node represent vicariance events, and the blue circles represent dispersal events. The probability of the origin at these nodes are also indicated (%).</p>
Full article ">Figure 3
<p>Patterns of speciation within the genus <span class="html-italic">Lycium</span>. (<b>a</b>) Frequency of speciation events showing that the highest frequency of speciation occurred within the last five years (red) and the remaining speciation events in the last 20 years (gray); (<b>b</b>) actual gamma value (x) in comparison to simulated gamma values under a model of constant diversification.</p>
Full article ">Figure 4
<p>Summary of all evolutionary events reported in this study by fitting the CoMET model. Results reported are for the diversification hyperpriors specified a priori. (<b>a</b>) Speciation rate; (<b>b</b>) speciation shift times; (<b>c</b>) extinction rates; (<b>d</b>) extinction shift times; (<b>e</b>) mass extinction Bayes factors; (<b>f</b>) mass extinction times.</p>
Full article ">
18 pages, 326 KiB  
Review
Pork as a Source of Diverse Viral Foodborne Infections: An Escalating Issue
by Anna Szczotka-Bochniarz and Maciej Kochanowski
Diversity 2024, 16(11), 679; https://doi.org/10.3390/d16110679 - 6 Nov 2024
Viewed by 651
Abstract
This review synthesizes current knowledge on the risks posed by viral foodborne infections associated with pork, emphasizing their global prevalence and the complexity of managing such pathogens. It covers a range of significant viruses, including hepatitis A and E, norovirus, rotavirus, sapovirus, enterovirus, [...] Read more.
This review synthesizes current knowledge on the risks posed by viral foodborne infections associated with pork, emphasizing their global prevalence and the complexity of managing such pathogens. It covers a range of significant viruses, including hepatitis A and E, norovirus, rotavirus, sapovirus, enterovirus, astrovirus, and enteric adenovirus. The role of pigs as reservoirs for diverse pathogens with zoonotic potential further complicates safety challenges, extending risks to individuals involved in pork production and processing. Various factors influencing viral contamination throughout the meat production chain are explored, from farm-level practices to processing and handling procedures. Emphasis is placed on the critical importance of implementing effective control measures at each stage, including enhanced biosecurity, rigorous hygiene practices, and appropriate thermal processing techniques. Additionally, the need for improved surveillance and detection methods to effectively identify and monitor viral presence in meat products is highlighted. In conclusion, the necessity of adopting a One Health approach that integrates efforts in animal health, food safety, and public health to mitigate the risks of viral foodborne infections associated with meat consumption is underscored. This holistic strategy is essential for safeguarding consumer health and ensuring the safety of the global food supply. Full article
(This article belongs to the Special Issue Diversity, Occurrence and Distribution of Foodborne Pathogens)
27 pages, 11437 KiB  
Article
Species Diversity and Spatial Distribution of Some Oribatid Mites in Bory Tucholskie National Park (N Poland)
by Wojciech Niedbała, Agnieszka Napierała, Jacek Wendzonka, Karolina Lubińska, Marta Kulczak and Jerzy Błoszyk
Diversity 2024, 16(11), 678; https://doi.org/10.3390/d16110678 - 5 Nov 2024
Viewed by 625
Abstract
There are 23 national parks in Poland, and only a few of them have been studied thoroughly with regard to acarofauna so far. One of the least-examined areas in this regard is Bory Tucholskie National Park (BTNP), established in 1996. The aim of [...] Read more.
There are 23 national parks in Poland, and only a few of them have been studied thoroughly with regard to acarofauna so far. One of the least-examined areas in this regard is Bory Tucholskie National Park (BTNP), established in 1996. The aim of this research study was to explore the species diversity, community structure, and spatial distribution of mites from the order Oribatida: ptyctimous mites (Acari: Oribatida) and species from the families Nothridae and Camisiidae (Acari: Oribatida: Crotonioidea) inhabiting different forests open and unstable microhabitats in the area of Bory Tucholskie National Park (BTNP). In the case of ptyctimous mites, the communities were compared to those in other Polish national parks. Based on the analysis of 285 samples collected in BTNP between 2022 and 2024, 8 species of Crotonioidea with dominant Heminothrus peltifer (C. L. Koch, 1839) and 21 species of ptyctimous mites with the most numerous Atropacarus (Atropacarus) striculus (C. L. Koch, 1835) were identified in the analyzed material. The highest species diversity was observed in different types of pine forests (25 species) and in alder forests (24 species), while the lowest diversity occurred in areas with reeds (11 species). The comparison of the number of ptyctimous mites in Polish national parks revealed that BTNP can be ranked second in terms of species diversity among 12 national parks examined in Poland so far. Full article
(This article belongs to the Special Issue Diversity and Ecology of the Acari)
Show Figures

Figure 1

Figure 1
<p>Distribution of examined plots (black dots) in the area of Bory Tucholskie National Park and localization of the national park (green dot). Blue parts—water tanks.</p>
Full article ">Figure 2
<p>Spatial distribution in BTNP (black dots): (<b>A</b>) <span class="html-italic">Heminothrus peltifer</span>, (<b>B</b>) <span class="html-italic">Nothrus anauniensis</span>; (<b>C</b>) <span class="html-italic">Nothrus silvestris</span>, and (<b>D</b>) <span class="html-italic">Camisia segnis</span>, against all examined plots (white dots).</p>
Full article ">Figure 3
<p>Spatial distribution in BTNP (black dots): (<b>A</b>) <span class="html-italic">Camisia biurus</span>, (<b>B</b>) <span class="html-italic">Nothrus pratensis</span>, (<b>C</b>) <span class="html-italic">Heminothrus targionii</span>, and (<b>D</b>) <span class="html-italic">Camisia horrida</span>, against all examined plots (white dots).</p>
Full article ">Figure 4
<p>Spatial distribution in BTNP (black dots): (<b>A</b>) <span class="html-italic">Phthiracarus longulus</span>, (<b>B</b>) <span class="html-italic">Atropacarus</span> (<span class="html-italic">Atropacarus</span>) <span class="html-italic">striculus</span>, (<b>C</b>) <span class="html-italic">Acrotritia ardua</span>, and (<b>D</b>) <span class="html-italic">Steganacarus</span> (<span class="html-italic">Tropacarus</span>) <span class="html-italic">carinatus</span> against all examined plots (white dots).</p>
Full article ">Figure 5
<p>Spatial distribution in BTNP (black dots): (<b>A</b>) <span class="html-italic">Phthiracarus nitens</span>, (<b>B</b>) <span class="html-italic">Euphthiracarus cribrarius</span>, (<b>C</b>) <span class="html-italic">Acrotritia duplicata</span>, and (<b>D</b>) <span class="html-italic">Phthiracarus laevigatus</span> against all examined plots (white dots).</p>
Full article ">Figure 6
<p>Spatial distribution in BTNP (black dots): (<b>A</b>) <span class="html-italic">Microtritia minima</span>, (<b>B</b>) <span class="html-italic">Phthiracarus globosus</span>, (<b>C</b>) <span class="html-italic">Phthiracarus crinitus</span>, and (<b>D</b>) <span class="html-italic">Phthiracarus clavatus</span>, against all examined plots (white dots).</p>
Full article ">Figure 7
<p>Spatial distribution in BTNP (black dots): (<b>A</b>) <span class="html-italic">Phthiracarus bryobius</span>, (<b>B</b>) <span class="html-italic">Phthiracarus ferrugineus</span>, (<b>C</b>) <span class="html-italic">Phthiracarus boresetosus</span>, and (<b>D</b>) <span class="html-italic">Steganacarus</span> (<span class="html-italic">Steganacarus</span>) <span class="html-italic">magnus</span>, against all examined plots (white dots).</p>
Full article ">Figure 8
<p>Spatial distribution in BTNP (black dots): (<b>A</b>) <span class="html-italic">Euphthiracarus monodactylus</span>, (<b>B</b>) <span class="html-italic">Atropacarus</span> (<span class="html-italic">Atropacarus</span>) <span class="html-italic">csiszarae</span>, (<b>C</b>) <span class="html-italic">Mesoplophora</span> (<span class="html-italic">Parplophora</span>) <span class="html-italic">pulchra</span>, and (<b>D</b>) <span class="html-italic">Mesotritia nuda</span>, against all examined plots (white dots).</p>
Full article ">Figure 9
<p>Spatial distribution in BTNP (black dots): <span class="html-italic">Phthiracarus anonymus</span>, against all examined plots (white dots).</p>
Full article ">Figure 10
<p>Percentage of species assigned to discerned ecotypes: E—eurytopic, P—politopic, M—mesotopic, O—oligotopic, S—stenotopic.</p>
Full article ">Figure 11
<p>Similarities in species composition in different types of habitat: A—alder forest, B—pine forests, C—transformed alder forests, D—peatlands, E—meadows, F—inland dunes, G—reeds.</p>
Full article ">Figure 12
<p>Number of ptyctimous mites recorded in national parks (NP) in Poland: BIA—Białowieża NP, BTU—Bory Tucholskie NP, GOR—Gorczański NP, KAR—Karkonoski NP, OJC—Ojcowski NP, PIE—Pieniński PN, ROZ—Roztoczański NP, SŁO—Słowiński NP, ŚWI—Świętokrzyski NP, TAT—Tatrzański NP, WLK—Wielkopolski NP, WOL—Woliński NP.</p>
Full article ">Figure 13
<p>Similarities in species composition (S) in communities of ptyctimous mites in selected national parks (NP) in Poland: BTU—Bory Tucholskie NP, BIA—Białowieża NP, WOL—Woliński NP, ŚWI—Świętokrzyski NP, KAR—Karkonoski NP, SŁO—Słowiński NP, ROZ—Roztoczański NP, GOR—Gorczański NP, PIE—Pieniński NP, OJC—Ojcowski NP, WLK—Wielkopolski NP, TAT—Tatrzański NP.</p>
Full article ">
16 pages, 15953 KiB  
Article
New Material of Thylacocephala from the Early Ladinian (Middle Triassic) of Northern Grigna (Lecco, Lombardy, Northern Italy)
by Cheng Ji and Andrea Tintori
Diversity 2024, 16(11), 677; https://doi.org/10.3390/d16110677 - 4 Nov 2024
Viewed by 432
Abstract
Here we report and describe a new assemblage of Thylacocephala (Crustacea) from the Early Ladinian Buchenstein Fm. (Middle Triassic) of Grigna, Northern Italy. The assemblage consists of at least four species from three different genera: Ankitokazocaris lariensis sp. n., Ankitokazocaris sp., Austriocaris sp., [...] Read more.
Here we report and describe a new assemblage of Thylacocephala (Crustacea) from the Early Ladinian Buchenstein Fm. (Middle Triassic) of Grigna, Northern Italy. The assemblage consists of at least four species from three different genera: Ankitokazocaris lariensis sp. n., Ankitokazocaris sp., Austriocaris sp., Stoppanicaris grignaensis gen. et sp. n. This thylacocephalan assemblage is rather diverse compared to the others of the Triassic. The largest size and ornamentation type of thylacocephalan species is compared among different periods of the Triassic and indicates that taxa with ridges on the carapace are generally smaller than those with smooth carapaces. This may be related to their different modes of life, such as inside or above the sediment with low oxygen levels. Large and smooth taxa were possibly more adapted to a life above sandy bottoms in shallow waters, under a somewhat high wave energy, while small, ornamented taxa are better suited for deeper environments with muddy bottoms, inside which they could move freely. The EDS analysis of Austriocaris sp. reveals that the cuticle mainly consists of apatite, which is in accordance with previous interpretations. Full article
(This article belongs to the Special Issue Marine Biodiversity from the Triassic)
Show Figures

Figure 1

Figure 1
<p>Map of the section location yielding the thylacocephalans. The left star: Scudi Tremare in Northern Grigna, GPS: 45°56′25″ N, 9°23′42″ E. The right star: Rio Sacuz in Dolomites, GPS: 46°26′54″ N, 12°04′59″ E.</p>
Full article ">Figure 2
<p>Occurrences of the newly described thylacocephalan material from the “fish level” in the Buchenstein Formation of Northern Grigna. The red stars represent the beds yielding the thylacocephalans.</p>
Full article ">Figure 3
<p>Thylacocephalans from Northern Grigna. (<b>A</b>–<b>G</b>), <span class="html-italic">Ankitokazocaris lariensis</span> sp. nov. (<b>A</b>) MPUM 13464. (<b>B</b>) MPUM 13465. (<b>C</b>) MPUM 13468. (<b>D</b>) MPUM 13469. (<b>E</b>) MPUM 13466. (<b>F</b>) MPUM 13467. (<b>G</b>) MPUM 13470. (<b>H</b>) <span class="html-italic">Ankitokazocaris</span> cf. <span class="html-italic">A. lariensis</span>, MVC-1. The red arrows indicate the dorso-lateral carina. Scale bars equal 1 cm.</p>
Full article ">Figure 4
<p>Eye of <span class="html-italic">Ankitokazocaris lariensis</span> sp. nov. under microscope. (<b>A</b>) <span class="html-italic">Ankitokazocaris lariensis</span> sp. nov., MPUM 13465. (<b>B</b>) Enlargement of the rectangle area in (<b>A</b>). Scale bars equal 2 mm and 500 μm, respectively.</p>
Full article ">Figure 5
<p>Reconstruction of <span class="html-italic">Ankitokazocaris lariensis</span> sp. nov.</p>
Full article ">Figure 6
<p>(<b>A</b>) <span class="html-italic">Ankitokazocaris</span> sp. gigantism specimen, MPUM 13471. (<b>B</b>) <span class="html-italic">Austriocaris</span> sp., MPUM 13472. (<b>C</b>) ? <span class="html-italic">Diplacanthocaris</span>, MPUM 13473. (<b>D</b>) <span class="html-italic">Stoppanicaris grignaensis</span>, MPUM 13474. The posterior margin is encompassed beween the two blue dots. The red arrows indicate the indentations on the posterior margin. (<b>E</b>) Enlargement on the rostrum of (<b>D</b>). Scale bars equal 1 cm in (<b>A</b>–<b>D</b>) and 5 mm in (<b>E</b>).</p>
Full article ">Figure 7
<p>Size range change of thylacocephalans through Triassic.</p>
Full article ">Figure 8
<p>Element mapping results of the <span class="html-italic">Austriocaris</span> sp. MPUM 13472.</p>
Full article ">
13 pages, 2976 KiB  
Article
Natural Resources of Rhaponticum carthamoides in the Tarbagatai State National Nature Park
by Anar Myrzagaliyeva, Serik Irsaliyev, Shynar Tustubayeva, Talant Samarkhanov, Aidyn Orazov and Zhanylkan Alemseitova
Diversity 2024, 16(11), 676; https://doi.org/10.3390/d16110676 - 4 Nov 2024
Viewed by 411
Abstract
The study of medicinal plants and having a protected status is an urgent issue for the conservation of biodiversity in Kazakhstan. Rhaponticum carthamoides (Willd.) Ilijn is a medicinal plant, and its excessive harvesting and destruction of habitats, as well as its conservation status, [...] Read more.
The study of medicinal plants and having a protected status is an urgent issue for the conservation of biodiversity in Kazakhstan. Rhaponticum carthamoides (Willd.) Ilijn is a medicinal plant, and its excessive harvesting and destruction of habitats, as well as its conservation status, are of concern. We conducted a study to assess the ecological characteristics of the habitat of the species in the Tarbagatai National Natural Park, calculate the amount of medicinal raw materials, calculate both aboveground and underground phytomass, and calculate biological and operational reserves. A map has been developed to show the distribution of the species and potential harvest sites. In addition, a correlation analysis was performed to understand how population size affects productivity. The results highlight the need for continuous monitoring and protection of endangered species. The conservation of Rhaponticum carthamoides in the Tarbagatai State National Nature Park is currently ensured by its protected status. However, the study emphasizes the importance of developing a sustainable use regime to effectively manage plant resources and ensure their preservation for future generations. Full article
Show Figures

Figure 1

Figure 1
<p>Location of Tarbagatai State National Nature Park.</p>
Full article ">Figure 2
<p><span class="html-italic">Rhaponticum carthamoides</span> population in the surveyed area.</p>
Full article ">Figure 3
<p>Graph of the trend line of correlation between the density of underground raw materials and the number of species (black bullets—number of species, blue line—trend line, gray color around the blue line—confidence interval).</p>
Full article ">
17 pages, 3017 KiB  
Article
Effects of Polycyclic Aromatic Hydrocarbons on Soil Bacterial and Fungal Communities in Soils
by Chunyong Wang, Haitao Wu, Weinong Zhao, Bo Zhu and Jiali Yang
Diversity 2024, 16(11), 675; https://doi.org/10.3390/d16110675 - 3 Nov 2024
Viewed by 895
Abstract
Soil organic pollution (such as heavy metals, PAHs, etc.) has caused serious environmental problems, which have resulted in unexpected effects on contaminated soil ecosystems. However, knowledge of the interactions between environmental PAHs and bacterial and fungal communities is still limited. In this study, [...] Read more.
Soil organic pollution (such as heavy metals, PAHs, etc.) has caused serious environmental problems, which have resulted in unexpected effects on contaminated soil ecosystems. However, knowledge of the interactions between environmental PAHs and bacterial and fungal communities is still limited. In this study, soil samples from different PAH-contaminated areas including non-contaminated areas (NC), low-contaminated areas (LC), and high-contaminated areas (HC) were selected. Results of toxic equivalent quantity (TEQ) indicated that Benzo[a]pyrene (BaP) and Dibenzo[a,h]anthracene (DBahA) constituted the main TEQs of ∑16PAHs. Incremental lifetime cancer risk (ILCR) assessment revealed that the main pathway of exposure to soil PAHs was dermal contact in adults and children. Furthermore, adults faced a higher total cancer risk (including dermal contact, ingestion, and inhalation) from soil PAHs than children. The microbial community composition analysis demonstrated that soil PAHs could decrease the diversity of bacterial and fungal communities. The relative abundance of Acidobacteriota, Gemmatimonadota, Fimicutes, Bacteroidota, Ascomycota, and Basidiomycota exhibited varying degrees of changes under different concentrations of PAHs. Benzo[a]anthracene (BaA) and Chrysene (Chr) drove the bacterial community composition, while BaP and DBahA drove the fungal community compositions. Co-occurrence network analysis revealed the high contamination levels of PAHs that could change the relationships among different microorganisms and reduce the complexity and stability of fungal and bacterial networks. Overall, these findings provide comprehensive insight into the responses of bacterial and fungal communities to PAHs. Full article
(This article belongs to the Section Biodiversity Loss & Dynamics)
Show Figures

Figure 1

Figure 1
<p>Alpha diversity indices (Chao1 index and Shannon index) of bacterial and fungal communities in the soil samples from NC, LC, and HC areas. (<b>a</b>) Bacterial Chao1 index. (<b>b</b>) Bacterial Shannon index. (<b>c</b>) Fungal Chao1 index. (<b>d</b>) Fungal Shannon index.</p>
Full article ">Figure 2
<p>Circos diagram of bacterial and fungal communities at the phylum level in the soil samples from NC, LC, and HC areas. (<b>a</b>) Bacterial Circos diagram. (<b>b</b>) Fungal Circos diagram.</p>
Full article ">Figure 3
<p>Relative abundance of bacterial and fungal communities at the phylum level in the soil samples from NC, LC, and HC areas. (<b>a</b>) Bacteria. (<b>b</b>) Fungi.</p>
Full article ">Figure 3 Cont.
<p>Relative abundance of bacterial and fungal communities at the phylum level in the soil samples from NC, LC, and HC areas. (<b>a</b>) Bacteria. (<b>b</b>) Fungi.</p>
Full article ">Figure 4
<p>RDA of bacterial and fungal community composition among soil samples (NC, LC, and HC) and PAHs (BaA, Chr, BbF, BkF, BaP, IDP, DBahA, and BghiP). (<b>a</b>) Bacteria. (<b>b</b>) Fungi.</p>
Full article ">Figure 5
<p>Nodes are colored based on the dominant bacterial and fungal phyla. The connections represent strong (Spearman’s correlation coefficient (r) &gt; 0.6) and significant (<span class="html-italic">p</span> &lt; 0.05) correlations. Different node colors represent different phyla. The red lines represent positive correlations between two linked genera, whereas the black lines represent negative correlations. (<b>a</b>) Bacterial co-occurring network in the soil samples of NC areas. (<b>b</b>) Bacterial co-occurring network in the soil samples of LC areas. (<b>c</b>) Bacterial co-occurring network in the soil samples of HC areas. (<b>d</b>) Fungal co-occurring network in the soil samples of NC areas. (<b>e</b>) Fungal co-occurring network in the soil samples of LC areas. (<b>f</b>) Fungal co-occurring network in the soil samples of HC areas.</p>
Full article ">
15 pages, 3025 KiB  
Article
Integrated Genetic and Statolith Shape Analysis Reveals the Population Structure of Loliolus (Nipponololigo) uyii (Cephalopoda: Loliginidae) in the Coastal Waters of China
by Xiaorong Wang, Chi Zhang and Xiaodong Zheng
Diversity 2024, 16(11), 674; https://doi.org/10.3390/d16110674 - 2 Nov 2024
Viewed by 574
Abstract
Understanding population structure is a priority for evaluating population dynamics of commercially fished cephalopods under fishing pressure and environmental changes. This study employed a multidisciplinary approach to clarify the population structure of Loliolus (Nipponololigo) uyii, a common squid in inshore [...] Read more.
Understanding population structure is a priority for evaluating population dynamics of commercially fished cephalopods under fishing pressure and environmental changes. This study employed a multidisciplinary approach to clarify the population structure of Loliolus (Nipponololigo) uyii, a common squid in inshore fisheries. Sampling was conducted multiple times to cover the distribution range across the East China Sea and South China Sea. High haplotype diversity was revealed by three gene markers (COI, 16S and ODH). Two geographical clades with significant genetic differentiation were divided through phylogenetic trees and haplotype networks. The boundary between the two clades is delineated by the Dongshan population in the southern East China Sea. Furthermore, the neutrality tests and mismatch analysis suggested that L. (N.) uyii populations may have undergone population expansion. Correspondingly, statolith differences in lateral dome and posterior indentation, along with high classification success, further supported the genetic division. The overall difference in statolith shape also efficiently identified seasonal groups in the Beibu Gulf lacking genetic differentiation. This result offers new insights into the influence of genetic and environmental factors on statolith shape. The integrated results provide a comprehensive understanding of the population structure of L. (N.) uyii, laying the foundation for resource development and the conservation of the species. Full article
Show Figures

Figure 1

Figure 1
<p>The sampling map for <span class="html-italic">L.</span> (<span class="html-italic">N.</span>) <span class="html-italic">uyii</span> from different locations. Zhoushan (ZS), Ningbo (NB), Ningde (ND), Dongshan (DS), Maoming (MM) and Beihai (BH). The green color represents the northern groups (ZS, NB and ND), and the orange color represents the southern groups (DS, MM and BH). The black arrows indicate ocean currents.</p>
Full article ">Figure 2
<p>Maximum likelihood (ML) and Bayesian inference (BI) phylogenetic tree and haplotype network of <span class="html-italic">L.</span> (<span class="html-italic">N.</span>) <span class="html-italic">uyii</span> based on <span class="html-italic">COI</span> haplotypes. Branch numbers are bootstraps (<b>left</b>) and posterior probability (<b>right</b>). Black dots represent hypothetical missing intermediates.</p>
Full article ">Figure 3
<p>Maximum likelihood (ML) and Bayesian inference (BI) phylogenetic tree and haplotype network of <span class="html-italic">L.</span> (<span class="html-italic">N.</span>) <span class="html-italic">uyii</span> based on <span class="html-italic">16S</span> haplotypes. Branch numbers are bootstraps (<b>left</b>) and posterior probability (<b>right</b>). Black dots represent hypothetical missing intermediates.</p>
Full article ">Figure 4
<p>Maximum likelihood (ML) and Bayesian inference (BI) phylogenetic tree and haplotype network of <span class="html-italic">L.</span> (<span class="html-italic">N.</span>) <span class="html-italic">uyii</span> based on <span class="html-italic">ODH</span> haplotypes. Branch numbers are bootstraps (<b>left</b>) and posterior probability (<b>right</b>). Black dots represent hypothetical missing intermediates.</p>
Full article ">Figure 5
<p>Phylogenetic trees constructed using the concatenation of mitochondrial and nuclear genes. The numbers in each node represent bootstraps of maximum likelihood (ML) and posterior probabilities of Bayesian inference (BI) analyses, respectively. The purple color represents the Clade A and the green color represents the Clade B.</p>
Full article ">Figure 6
<p>MDS plot of statolith morphology for different regions (<b>a</b>) and different seasons in BBG (<b>b</b>).</p>
Full article ">Figure 7
<p>Reconstructed statolith outlines for different regions (<b>a</b>) and different seasons in BBG (<b>b</b>).</p>
Full article ">
14 pages, 1994 KiB  
Article
Plant-Parasitic and Free-Living Nematode Community Associated with Oak Tree of Magoebaskloof Mountains, Limpopo Province, South Africa
by Ebrahim Shokoohi and Peter Masoko
Diversity 2024, 16(11), 673; https://doi.org/10.3390/d16110673 - 1 Nov 2024
Viewed by 459
Abstract
A study was conducted in the mountains of Magoebaskloof, Limpopo Province, where oak trees grow along the banks of the Broederstroom River. This study revealed that 22 nematode genera were associated with oak trees (Quercus robur). The most frequently occurring nematodes [...] Read more.
A study was conducted in the mountains of Magoebaskloof, Limpopo Province, where oak trees grow along the banks of the Broederstroom River. This study revealed that 22 nematode genera were associated with oak trees (Quercus robur). The most frequently occurring nematodes were Aphelenchus sp. (100%) and Plectus sp. (100%), followed by Helicotylenchus sp. (90%). This study examined the relationship between nematodes and the physicochemical properties of the soil using Pearson correlation. It uncovered that the organic matter content (OMC) had a negative correlation with the number of Panagrolaimus sp. (r = −0.770) and Hemicycliophora sp. (r = −0.674). Conversely, the sand percentage positively correlated (r = 0.695) with the number of Hemicycliophora sp. The clay content of the soil showed a positive correlation (r = 0.617) with the number of Ditylenchus. Soil pH demonstrated a significant negative correlation with Acrobeloides sp. (r = −0.877). The canonical correspondence analysis (CCA) explained 63.3% of the relationship between nematodes and soil physicochemical properties. The CCA results indicated that Ditylenchus exhibited a positive correlation with OMC, while the Panagrolaimus and Hemicycliophora species showed a negative correlation with OMC. The results indicated that none of the soil sample sites were under stress. The soil food web analysis revealed that most soil samples were nutrient-enriched with a low C/N ratio. In conclusion, this study revealed that oak trees harbor a high diversity of plant-parasitic and free-living nematodes. The results suggest that soil nematodes, particularly free-living bacterivores, such as Panagrolaimus, can indicate organic matter content in the soil. Full article
Show Figures

Figure 1

Figure 1
<p>Nematodes mean population density (MPD), frequencies of occurrence (FO%), and prominence values (PVs) were studied in oak trees of Magoebaskloof, Limpopo Province, South Africa.</p>
Full article ">Figure 2
<p>Correlation between soil variables and nematodes of oak trees of Magoebaskloof in Limpopo Province, South Africa.</p>
Full article ">Figure 3
<p>CCA plot of the relationship of the soil variables and nematodes associated with oak trees in Magoebaskloof, Limpopo Province, South Africa.</p>
Full article ">Figure 4
<p>The c-p (colonizer–persister) triangle depicting soil status based on the c-p groups of nematodes found in oak trees in Magoebaskloof, Limpopo Province, South Africa.</p>
Full article ">Figure 5
<p>Food web analysis of various soil samples collected from oak trees in Magoebaskloof, Limpopo Province, South Africa, and their position as soil health bioindicators.</p>
Full article ">
14 pages, 5392 KiB  
Article
First Report of Middle Eocene Micromorphic Brachiopods from Northeastern Libya: Taxonomy and Paleobiogeography Implications
by Sayed M. Abd El-Aziz, Ibrahim M. Abd El-Gaied, Mansour H. Al-Hashim, Maria Aleksandra Bitner, Yasser F. Salama, Petra Heinz and Mostafa M. Sayed
Diversity 2024, 16(11), 672; https://doi.org/10.3390/d16110672 - 31 Oct 2024
Viewed by 421
Abstract
Two brachiopod species, Terebratulina tenuistriata (Leymerie, 1846) and Orthothyris pectinoides (von Koenen, 1984), have been recorded for the first time from the Middle Eocene (Late Lutetian) nummulitic limestone beds in the Darnah Formation at the Wadi Darnah area in Northeast Libya. These brachiopod [...] Read more.
Two brachiopod species, Terebratulina tenuistriata (Leymerie, 1846) and Orthothyris pectinoides (von Koenen, 1984), have been recorded for the first time from the Middle Eocene (Late Lutetian) nummulitic limestone beds in the Darnah Formation at the Wadi Darnah area in Northeast Libya. These brachiopod species are associated here with Nummulites discorbinus (Schlotheim), Nummulites praelyelli (Boukhary and Kamal), and Nummulites bullatus (Schaub) and are widely distributed on this Middle Eocene Nummulites carbonate platform. The two recorded species are common in the Eocene rocks of Europe and the Arabian Gulf. In northern Africa, the brachiopod species Terebratulina tenuistriata (Leymerie) was only recorded from the Middle Eocene (Bartonian) of Egypt, while Orthothyris pectinoides (von Koenen) is firstly recorded from the Middle Eocene of the southern Tethyan Province (NE Libya) in the present work. Full article
Show Figures

Figure 1

Figure 1
<p>(<b>a</b>) Location of the study area (after El Mehaghag and Ashahomi, [<a href="#B23-diversity-16-00672" class="html-bibr">23</a>]); (<b>b</b>) Geologic map of Gabal Al Akhdar (after: Klen, [<a href="#B24-diversity-16-00672" class="html-bibr">24</a>]; Rohlich, [<a href="#B16-diversity-16-00672" class="html-bibr">16</a>]; and Zert, [<a href="#B15-diversity-16-00672" class="html-bibr">15</a>]).</p>
Full article ">Figure 2
<p>Lithology of the studied succession at the Wadi Darnah area.</p>
Full article ">Figure 3
<p>(<b>A</b>) Rock units of the study section; (<b>B</b>,<b>C</b>) gray nummulitic limestone at the lower part; (<b>D</b>,<b>E</b>) yellowish gray nummulitic beds at the middle part; (<b>F</b>,<b>G</b>) light gray nummulitic limestone with large-sized bivalves at the upper part.</p>
Full article ">Figure 4
<p>Distribution chart for the Nummulite and Brachiopod species in the Wadi Darnah section.</p>
Full article ">Figure 5
<p>(<b>1</b>,<b>6</b>–<b>8</b>) Ventral views of the complete specimens; (<b>2</b>–<b>5</b>,<b>10</b>–<b>12</b>) dorsal views of the complete specimens; (<b>9</b>) lateral view of complete specimen.</p>
Full article ">Figure 6
<p>(<b>1</b>,<b>3</b>,<b>5</b>–<b>7</b>,<b>9</b>–<b>11</b>) Dorsal views of the complete specimens; (<b>2</b>,<b>4</b>,<b>8</b>,<b>12</b>) ventral views of the complete specimens.</p>
Full article ">Figure 7
<p>Paleobiogeographic distribution of <span class="html-italic">Terebratulina tenuistriata</span> and <span class="html-italic">Orthothyris pectinoides</span> during the Middle Eocene [<a href="#B62-diversity-16-00672" class="html-bibr">62</a>] (The map is modified from <a href="http://scotese.com" target="_blank">http://scotese.com</a>).</p>
Full article ">
15 pages, 7720 KiB  
Article
First Attempts at DNA Barcoding Lepidoptera in North Cyprus Reveal Unexpected Complexities in Taxonomic and Faunistic Issues
by Peter Huemer and Özge Özden
Diversity 2024, 16(11), 671; https://doi.org/10.3390/d16110671 - 31 Oct 2024
Viewed by 611
Abstract
The fauna of Lepidoptera in the Mediterranean is still inadequately documented. As a result, even remotely complete DNA barcode libraries (mt. COI (cytochrome c oxidase 1) gene) are lacking in most areas. This proposed gap is being analyzed for the first time for [...] Read more.
The fauna of Lepidoptera in the Mediterranean is still inadequately documented. As a result, even remotely complete DNA barcode libraries (mt. COI (cytochrome c oxidase 1) gene) are lacking in most areas. This proposed gap is being analyzed for the first time for the fauna of North Cyprus. In the initial phase, 248 morphospecies from 29 families (exclusive Heterocera) were sampled, sequenced and compared with existing DNA reference sequences in the global BOLD database (Barcode of Life Data Systems) via BINs (Barcode Index Numbers). A total of 194 species could be unequivocally assigned to a Linnaean taxon. Additionally, six species previously unidentified in BOLD, as well as fourteen species without reference barcodes, were identified at the species level. Twenty-four of these species were new records for Cyprus. In addition, 25 taxa with new BINs could not be assigned to a valid species due to potential cryptic diversity or the lack of relevant revisions. Furthermore, a few species could not be identified due to barcode sharing and/or potential misidentifications in BOLD. Overall, approximately 20% of the samples could not be identified using the existing DNA barcode libraries, a significant deficit for European standards, which should be addressed as a priority issue in future studies. Full article
(This article belongs to the Special Issue DNA Barcodes for Evolution and Biodiversity—2nd Edition)
Show Figures

Figure 1

Figure 1
<p>Sampling localities in North Cyprus (1 = Dipkarpaz; 2 = Bafra/Vokolida; 3 = Ilgaz; 4 = Selvili tepe; 5 = Hisarköy). Copyrights: OpenStreetMap contributors (<a href="http://www.openstreetmap.org/copyright" target="_blank">http://www.openstreetmap.org/copyright</a> (accessed on 20 September 2024); SRTM 30 m by NASA EOSDIS Land Processes Distributed Active Archive Center (LP DAAC, <a href="https://lpdaac.usgs.gov/" target="_blank">https://lpdaac.usgs.gov/</a> (accessed on 20 September 2024).</p>
Full article ">Figure 2
<p>Largely unspoiled sand dunes near Dipkarpaz.</p>
Full article ">Figure 3
<p>Identification success based on BIN system (numbers refer to species).</p>
Full article ">
17 pages, 4335 KiB  
Review
Bibliometric Analysis of the Status and Trends of Seamounts’ Research and Their Conservation
by Maria Luisa Pica, Francesco Rendina, Adele Cocozza di Montanara and Giovanni Fulvio Russo
Diversity 2024, 16(11), 670; https://doi.org/10.3390/d16110670 - 31 Oct 2024
Viewed by 1051
Abstract
Seamounts are prominent volcanic seafloor features whose morphology affects many ocean processes, sustaining deep-sea communities and providing many ecosystem functions and services. Their study contributes to the understanding of many geological, oceanographic, biological, and ecological processes. Despite their acknowledged vulnerability to human activities [...] Read more.
Seamounts are prominent volcanic seafloor features whose morphology affects many ocean processes, sustaining deep-sea communities and providing many ecosystem functions and services. Their study contributes to the understanding of many geological, oceanographic, biological, and ecological processes. Despite their acknowledged vulnerability to human activities and climate change, the recovery time and ecological implications need to be properly understood. Moreover, only recently conservation measures have been considered. In this study, a bibliometric analysis of the scientific literature related to seamounts and their conservation was conducted. The analysis allowed for the generation of network maps displaying the relationships among keywords and countries. A total of 8019 articles were found regarding seamounts, 332 of which were related to their conservation. The results show that the main research fields concerned with seamounts are geology, seismology, geochemistry, oceanography, and biodiversity, whereas those regarding their conservation are corals, marine protected areas, benthos, community structure, fisheries, and management measures. Scientific papers about seamounts were published by 191 authors across 50 countries, while 19 authors across 25 countries published about their conservation. This study highlights the necessity to expand scientific knowledge on seamounts, especially regarding their ecological processes, to provide useful data for the successful management and conservation of these still mostly unexplored habitats. Full article
Show Figures

Figure 1

Figure 1
<p>Temporal trend of publications on (<b>a</b>) seamount* or deep sea pinnacle* or guyot* (1921–2024); (<b>b</b>) seamount* or deep sea pinnacle* or guyot* and conservation (1979–2024). Data exported from the Scopus database.</p>
Full article ">Figure 2
<p>Network map of the co-occurrence of keywords for the string “seamount* or deep sea pinnacle* or guyot<sup>*</sup>”. (<b>a</b>) Cluster visualization; (<b>b</b>) overlay visualization. Node size is representative of the number of occurrences. Clusters are defined by distinct colours; overlay visualization colors change according to the average publication year.</p>
Full article ">Figure 3
<p>Network map of the co-occurrence of keywords for the string “seamount* or deep sea pinnacle* or guyot<sup>*</sup> and conservation”. (<b>a</b>) Cluster visualization; (<b>b</b>) overlay visualization. Node size is representative of the occurrence. Clusters are defined by distinct colours; overlay visualization colors change according to the average publication year.</p>
Full article ">Figure 4
<p>Network map of co-authorship across countries in cluster visualization (<b>a</b>) for the string “seamount* or deep sea pinnacle* or guyot<sup>*</sup>” and (<b>b</b>) for the string “seamount* or deep sea pinnacle* or guyot<sup>*</sup> and conservation”. Node size is representative of the occurrence; clusters are defined by distinct colours.</p>
Full article ">
Previous Issue
Back to TopTop