[go: up one dir, main page]

 
 
Sign in to use this feature.

Years

Between: -

Subjects

remove_circle_outline
remove_circle_outline
remove_circle_outline
remove_circle_outline
remove_circle_outline
remove_circle_outline
remove_circle_outline
remove_circle_outline
remove_circle_outline

Journals

remove_circle_outline
remove_circle_outline
remove_circle_outline
remove_circle_outline
remove_circle_outline
remove_circle_outline

Article Types

Countries / Regions

remove_circle_outline
remove_circle_outline
remove_circle_outline
remove_circle_outline
remove_circle_outline
remove_circle_outline
remove_circle_outline

Search Results (1,038)

Search Parameters:
Keywords = new taxa

Order results
Result details
Results per page
Select all
Export citation of selected articles as:
15 pages, 5113 KiB  
Article
Stratocorticium sinensis gen. et sp. nov. and Cericium gloeocystidiatum sp. nov. (Cyphellaceae, Agaricales) from East Asia
by Yu-Peng Zhang, Yue Li, Karen K. Nakasone and Shuang-Hui He
J. Fungi 2024, 10(10), 722; https://doi.org/10.3390/jof10100722 - 17 Oct 2024
Viewed by 225
Abstract
Cyphellaceae, a small and under-studied family of Agaricales, includes mostly saprophytic taxa with varied basidiomes. In this study, we focus on wood-decay species with corticioid or stereoid basidiomes. Phylogenetic analyses of concatenated ITS-nrLSU sequences uncovered seven generic lineages of [...] Read more.
Cyphellaceae, a small and under-studied family of Agaricales, includes mostly saprophytic taxa with varied basidiomes. In this study, we focus on wood-decay species with corticioid or stereoid basidiomes. Phylogenetic analyses of concatenated ITS-nrLSU sequences uncovered seven generic lineages of corticioid or stereoid fungi—Acanthocorticium, Cericium, Chondrostereum, Cunninghammyces, Gloeostereum, Granulobasidium, and Stratocorticium gen. nov. The genus Cericium is shown to be in the Cyphellaceae family, and two new species, Cericium gloeocystidiatum and Stratocorticium sinensis, are described from East Asia. Morphologically, Ce. gloeocystidiatum is characterized by resupinate basidiomes with smooth hymenophores, a dimitic hyphal system with clamped generative hyphae and micro-binding hyphae, cystidia with resinous-like or golden yellow contents, and ellipsoid basidiospores. Stratocorticium is monotypic, differing from Cericium by a trimitic hyphal system of clamped generative, micro-binding, and brown, thick-walled skeletal-like hyphae, clavate to cylindrical cystidia with homogenous, colorless contents, and hyphidia. Descriptions and illustrations are provided for the new taxa and Cericium luteoincrustatum, and a key to corticioid or stereoid genera in Cyphellaceae is included. Full article
(This article belongs to the Special Issue Diversity, Phylogeny and Ecology of Forest Fungi)
Show Figures

Figure 1

Figure 1
<p>Phylogeny of <span class="html-italic">Cyphellaceae</span> generated by ML analyses based on combined <span class="html-italic">ITS</span>-<span class="html-italic">nrLSU</span> sequences. Branches are labeled with parsimony bootstrap values (≥50%, first), likelihood bootstrap values (≥50%, second), and Bayesian posterior probabilities (≥0.95, third). New taxa are highlighted and set in bold.</p>
Full article ">Figure 2
<p>Basidiomes of <span class="html-italic">Cericium luteoincrustatum</span> (from the holotype, F-450396). (<b>a</b>,<b>b</b>) Hymenophore.</p>
Full article ">Figure 3
<p><span class="html-italic">Cericium luteoincrustatum</span> (from the holotype F-450396). Scale bars: (<b>a</b>) = 50 µm; (<b>b</b>–<b>d</b>) = 20 µm. (<b>a</b>) Part of a vertical section through basidiome; (<b>b</b>) Close-up of a vertical section through basidiome; (<b>c</b>) hymenium with subfusiform hymenial cystidia and clavate cystidia with yellow, resinous-like contents; (<b>d</b>) partially collapsed, capitate cystidia with stalk from subhymenium.</p>
Full article ">Figure 4
<p>Basidiomes of <span class="html-italic">Cericium gloeocystidiatum</span>. Scale bars: (<b>a</b>,<b>b</b>) = 1 cm. (<b>a</b>) He 4332 (BJFC 023774, holotype); (<b>b</b>) He 4725 (BJFC 024244).</p>
Full article ">Figure 5
<p>Microscopic structures of <span class="html-italic">Cericium gloeocystidiatum</span> (from the holotype He 4332). Scale bars: (<b>a</b>–<b>f</b>) =10 µm. (<b>a</b>) Basidiospores; (<b>b</b>) Basidia; (<b>c</b>) Basidioles; (<b>d</b>) subfusiform cystidia from hymenium; (<b>e</b>) Cystidia with dark yellow, resinous-like contents; (<b>f</b>) Generative and micro-binding hyphae from subiculum.</p>
Full article ">Figure 6
<p>Basidiomata of <span class="html-italic">Stratocorticium sinensis</span>. Scale bars: (<b>a</b>–<b>d</b>) = 1 cm. (<b>a</b>) He 2264 (BJFC 020718); (<b>b</b>) He 5349 (BJFC 024867); (<b>c</b>) He 6875 (BJFC 033824); (<b>d</b>) He 6895 (BJFC 033844).</p>
Full article ">Figure 7
<p>Microscopic structures of <span class="html-italic">Stratocorticium sinensis.</span> (from the holotype He 3289). Scale bars: (<b>a</b>–<b>f</b>) = 10 µm. (<b>a</b>) Basidiospores; (<b>b</b>) Basidia and a basidiole; (<b>c</b>) Embedded cystidia; (<b>d</b>) Capitate cystidia, (<b>e</b>) Hyphidia; (<b>f</b>) Generative, micro-binding, and skeletal-like hyphae from subhymenium.</p>
Full article ">
19 pages, 7301 KiB  
Article
A Treasure Trove of Urban Microbial Diversity: Community and Diversity Characteristics of Urban Ancient Ginkgo biloba Rhizosphere Microorganisms in Shanghai
by Jieying Mao, Qiong Wang, Yaying Yang, Feng Pan, Ziwei Zou, Xiaona Su, Yi Wang, Wei Liu and Yaohua Tang
J. Fungi 2024, 10(10), 720; https://doi.org/10.3390/jof10100720 - 16 Oct 2024
Viewed by 285
Abstract
Rapid urbanization has exerted immense pressure on urban environments, severely constraining the growth of ancient trees. The growth of ancient trees is closely linked to the microbial communities in their rhizospheres, and studying their community characteristics may provide new insights into promoting the [...] Read more.
Rapid urbanization has exerted immense pressure on urban environments, severely constraining the growth of ancient trees. The growth of ancient trees is closely linked to the microbial communities in their rhizospheres, and studying their community characteristics may provide new insights into promoting the growth and rejuvenation of ancient trees. In this study, the rhizosphere soil and root systems of ancient Ginkgo biloba trees (approximately 200 years old) and adult G. biloba trees (approximately 50 years old) in Shanghai were selected as research subjects. Phospholipid fatty acid (PLFA) analysis and high-throughput sequencing were employed to investigate the diversity of microbial communities in the G. biloba rhizosphere. The results indicated that the 19 PLFA species selected to characterize the soil microbial community structure and biomass were present in the rhizosphere soil of both ancient and adult G. biloba trees. However, the total microbial biomass and the microbial biomass in the rhizosphere soil of ancient G. biloba were lower than the microbial biomass in the rhizosphere soil of adult G. biloba. The biomasses of Gram-negative bacteria (G), arbuscular mycorrhizal fungi (AMF), and protozoans (P) were significantly different. Total phosphorus, organic matter, and pH may be the key factors influencing the soil microbial community in the rhizosphere zone of ancient G. biloba. An in-depth study of AMF showed that the roots and rhizosphere soil of G. biloba contained abundant AMF resources, which were assigned to 224 virtual taxa using the MaarjAM reference database, belonging to four orders, ten families, and nineteen genera. The first and second most dominant genera were Glomus and Paraglomus, respectively. Archaeospora and Ambispora were more dominant in the rhizosphere than the roots. Furthermore, the abundance of live AMF was significantly higher in ancient G. biloba than in adult G. biloba. Therefore, future research should focus on the improvement of soil environmental characteristics and the identification and cultivation of indigenous dominant AMF in the rhizosphere of ancient G. biloba, aiming for their effective application in the rejuvenation of ancient trees. Full article
(This article belongs to the Special Issue Fungal Communities in Various Environments)
Show Figures

Figure 1

Figure 1
<p>Sampling positions and process. (<b>a</b>) Map of study sites; (<b>b</b>) sampling sites; (<b>c</b>) sampling process; and (<b>d</b>) sampling diagram.</p>
Full article ">Figure 2
<p>Microbial biomass in the rhizosphere soil of <span class="html-italic">Ginkgo biloba</span>. ns indicates <span class="html-italic">p</span> &gt; 0.05, * indicates <span class="html-italic">p</span> &lt; 0.05, ** indicates <span class="html-italic">p</span> &lt; 0.01. The following is the same.</p>
Full article ">Figure 3
<p>Analysis of soil environmental traits and microbial biomass. * indicates <span class="html-italic">p</span> &lt; 0.05, ** indicates <span class="html-italic">p</span> &lt; 0.01.</p>
Full article ">Figure 4
<p>Arbuscular mycorrhizal fungi (AMF) community composition of the root and rhizosphere soil of <span class="html-italic">Ginkgo biloba</span> at the genus level. XYXG represents the roots of adult <span class="html-italic">Ginkgo biloba</span>, XYXT represents the rhizosphere soil of adult <span class="html-italic">Ginkgo biloba</span>, GYXG represents the roots of ancient <span class="html-italic">Ginkgo biloba</span>, and GYXT represents the rhizosphere soil of ancient <span class="html-italic">Ginkgo biloba</span>.</p>
Full article ">Figure 5
<p>Venn diagram of arbuscular mycorrhizal fungi (AMF) community in the root and rhizosphere soil of <span class="html-italic">Ginkgo biloba</span>.</p>
Full article ">Figure 6
<p>Diversity analysis of the arbuscular mycorrhizal fungi (AMF) community in the root and rhizosphere soil of <span class="html-italic">Ginkgo biloba</span>. (<b>a</b>) the chart of Chao1 data analysis; (<b>b</b>) the chart of Shannon analysis; (<b>c</b>) the chart of Simpson analysis; (<b>d</b>) the chart of Pielou’s evenness analysis. Each point in the graph represents the data for each sample.</p>
Full article ">Figure 7
<p>Principal coordinate analysis (PCoA) of arbuscular mycorrhizal fungi (AMF) community diversity in the root and rhizosphere soil of <span class="html-italic">Ginkgo biloba</span>.</p>
Full article ">Figure 8
<p>Species clustering heat map of the arbuscular mycorrhizal fungi (AMF) community in the root and rhizosphere soil of <span class="html-italic">Ginkgo biloba</span>.</p>
Full article ">Figure 9
<p>Analysis of soil environmental traits and the arbuscular mycorrhizal fungi (AMF) community. * indicates <span class="html-italic">p</span> &lt; 0.05, ** indicates <span class="html-italic">p</span> &lt; 0.01.</p>
Full article ">
22 pages, 2208 KiB  
Article
The Final Pliocene and Early Pleistocene Faunal Dispersals from East to Europe and Correlation of the Villafranchian Biochronology between Eastern and Western Europe
by Nikolai Spassov
Quaternary 2024, 7(4), 43; https://doi.org/10.3390/quat7040043 - 11 Oct 2024
Viewed by 499
Abstract
The Villafranchian stage in the mammal fauna evolution in Eurasia (ca. 3.6/3.4 Ma—ca. 1.2 Ma) is associated with the beginning of the formation of the modern appearance of the mammal megafauna of today’s Palaearctic. The cooling and the aridification starting with the beginning [...] Read more.
The Villafranchian stage in the mammal fauna evolution in Eurasia (ca. 3.6/3.4 Ma—ca. 1.2 Ma) is associated with the beginning of the formation of the modern appearance of the mammal megafauna of today’s Palaearctic. The cooling and the aridification starting with the beginning of the Early Pleistocene gradually eliminated the quasi-tropical appearance of the Late Neogene landscapes and fauna of Europe. The time from the Mid-Piacenzian (ca. 3.3–3.0 Ma) to the end of the Early Pleistocene was a time of particularly intense dispersal of species, of faunal exchange between Eurasia and Africa, and of the entry of new mammals into Europe from the East. That is why the correlation of the biochronology of the Villafranchian fauna between Eastern and Western Europe is of particular interest. Accumulated data make possible a more precise correlation of these faunas today. A correlation of selected Eastern European localities with established faunal units and MNQ zones is made in the present work. Usually, the dispersal from Asia or from E. Europe to W. Europe is instantaneous from a geological point of view, but in a number of cases, reaching W. Europe happens later, or some species known to be from Eastern Europe do not reach Western Europe. The main driving forces of the faunal dispersals, which are the key bioevents in the faunal formation, are climate changes, which in turn, affect the environment. We can summarize the following more significant Villafranchian bioevents in Europe: the End Pliocene (Early Villafranchian: MNQ16) turnover related to the first appearance of a number of taxa, for example, felids, canids, proboscideans, and ungulates; the Quaternary beginning turnover. Correlated with this are the beginning of the Middle Villafranchian, which should be placed at about 2.6 Ma; the Coste San Giacomo faunal unit turnover (Senèze and Slivnitsa localities should be included here, and the FU itself, at the very beginning of the late Villafranchian (=MNQ18a)); the Pachycrocuta event at the very beginning of the Olivola FU; and the events related to the Late Villafranchian/Epivillafranchian bounfary. Full article
Show Figures

Figure 1

Figure 1
<p>Canids discussed in paragraph 1.1. (<b>a</b>): m1 of “<span class="html-italic">Canis</span>” from Vialette in occlusal view (cast of mandible 2003-5-401-VIA, Crozatier Museum of Le Puy-en-Velay). The arrow shows the hypoconid and the metaconid (though heavily worn) are not fused at their bases. (<b>b</b>–<b>d</b>): <span class="html-italic">Canis neschersensis</span> (MNHN.F.PET2010) stored in the coll. of the Laboratory. of Paleontology of the National Museum of Natural History, Paris, in occlusal, labial, and lingual views.</p>
Full article ">Figure 2
<p>Correlation of the Villafranchian biochronology between Eastern and Western Europe (geological age and biochronological position of selected Eastern European Villafranchian localities). The biochronology table and the correlation between the faunal units and the MNQ zones is based on Nomade et al. [<a href="#B18-quaternary-07-00043" class="html-bibr">18</a>] with some original modifications. The additional column on the far right presents the position of the Eastern European localities, discussed in <a href="#sec2dot3-quaternary-07-00043" class="html-sec">Section 2.3</a>. Abbreviations of polarity subchrons: Reu.—Réunion; Mamm.—Mammoth. Localities abbreviations: CER—Cernãteşti; TUL—Tulucesti; RSK—Ripa Skortselskaya; BOS—Bossilkovtsi; TSO—Tsotylio; DFN—Dafnero; SESK—Sesklo; VAR – Varshets; VLKS—Volakas; SLIV—Slivnitsa; LaPI—La Pietris; GER—Gerakarou; KARN—Karnezeika; VGRA—Vale Graunceanului; KRIM—Krimni; TRLI—Trlica; APOL—Apollonia.</p>
Full article ">
31 pages, 75566 KiB  
Article
Phylogenetic and Morphological Perspectives on Crepidotus subg. Dochmiopus: Exploratively Unveiling Hidden Diversity in China
by Menghui Han, Qin Na, Renxiu Wei, Hui Zeng, Yaping Hu, Libo Zhang, Jinhong Du, Li Zou, Weimin Tang, Xianhao Cheng and Yupeng Ge
J. Fungi 2024, 10(10), 710; https://doi.org/10.3390/jof10100710 - 11 Oct 2024
Viewed by 330
Abstract
Crepidotus subg. Dochmiopus contributes to more than half of Crepidotus species and exhibits highly hidden diversity. However, C. subg. Dochmiopus is challenging to study because the basidiomata of C. subg. Dochmiopus species are usually small and white, inconspicuous interspecific distinctions, and [...] Read more.
Crepidotus subg. Dochmiopus contributes to more than half of Crepidotus species and exhibits highly hidden diversity. However, C. subg. Dochmiopus is challenging to study because the basidiomata of C. subg. Dochmiopus species are usually small and white, inconspicuous interspecific distinctions, and possess a familiar complex. In this study, we utilized a variety of characteristics for species identification, including habitat, presence or absence of a stipe in mature specimens, pileipellis and cheilocystidia patterns, whether the lamellae edges are fimbriated, and other characteristics. Above all, cheilocystidia and pileipellis patterns will be important in C. subg. Dochmiopus research. Based on the present specimens, we constructed a multigene phylogenetic tree (ITS + LSU) and recognized four new species: C. lamellomaculatus sp. nov., C. capitatocystidiatus sp. nov., C. succineus sp. nov., C. clavocystidiatustustus sp. nov. Detailed morphological descriptions, photographs, line drawings and comparisons with closely related taxa for the new species are provided. The current phylogenetic analysis does not support the previously classifications, indicating that the classification of Crepidotus requires re-evaluation. But the existing molecular datasets and species’ descriptions are insufficient to fully resolve the classification. Further integration of new gene segments and a comprehensive review of morphological characteristics will reveal a natural classification for Crepidotus. Full article
(This article belongs to the Special Issue Taxonomy, Systematics and Evolution of Forestry Fungi, 2nd Edition)
Show Figures

Figure 1

Figure 1
<p>Phylogenetic tree inferred from the Bayesian Inference (BI) analysis based on a concatenated ITS and LSU dataset; bootstrap (BS) values over 75% and Bayesian posterior probabilities (BPP) over 0.90 are indicated. The new species are marked in red. In the top left corner, the figure caption indicates that different sections and series are marked with different colors.</p>
Full article ">Figure 1 Cont.
<p>Phylogenetic tree inferred from the Bayesian Inference (BI) analysis based on a concatenated ITS and LSU dataset; bootstrap (BS) values over 75% and Bayesian posterior probabilities (BPP) over 0.90 are indicated. The new species are marked in red. In the top left corner, the figure caption indicates that different sections and series are marked with different colors.</p>
Full article ">Figure 2
<p>Fresh basidiomata of <span class="html-italic">Crepidotus lamellomaculatus</span> sp. nov. (<b>a</b>,<b>i</b>) <span class="html-italic">FFAAS1307</span>; (<b>b</b>,<b>g</b>) <span class="html-italic">FFAAS1305</span> (holotype); (<b>c</b>–<b>f</b>) <span class="html-italic">FFAAS1306</span>; (<b>h</b>,<b>k</b>) <span class="html-italic">FFAAS1309</span>; (<b>j</b>,<b>l</b>,<b>m</b>) <span class="html-italic">FFAAS1308</span>; (<b>a</b>) lamellae edge when young and matured. Bars: (<b>a</b>,<b>b</b>,<b>d</b>–<b>f</b>,<b>h</b>–<b>k</b>) = 5 mm; (<b>c</b>,<b>g</b>) = 2 mm; (<b>l</b>,<b>m</b>) = 10 mm. Photos by Yupeng Ge and Menghui Han.</p>
Full article ">Figure 3
<p>Microscopic features of <span class="html-italic">Crepidotus lamellomaculatus</span> (<span class="html-italic">FFAAS1305,</span> holotype). (<b>a</b>–<b>e</b>) Lateral view of basidiospores; (<b>f</b>) frontal view of basidiospores; (<b>g</b>,<b>h</b>) basidia; (<b>i</b>–<b>r</b>) cheilocystidia; (<b>j</b>) clamp connection; (<b>s</b>) lamellae trama; (<b>t</b>) pileipellis, encrusted hyphae. Bars: (<b>a</b>–<b>f</b>) = 5 μm; (<b>g</b>–<b>r</b>) = 10 μm; (<b>s</b>,<b>t</b>) = 30 μm. Structures (<b>a</b>–<b>f</b>) were rehydrated in 5% KOH aqueous solution and (<b>g</b>–<b>t</b>) were stained in 1% Congo red aqueous solution.</p>
Full article ">Figure 4
<p>Morphological features of <span class="html-italic">Crepidotus lamellomaculatus</span> (<span class="html-italic">FFAAS1305,</span> holotype). (<b>a</b>) Basidiomata; (<b>b</b>) basidia; (<b>c</b>) basidiospores; (<b>d</b>) cheilocystidia; (<b>e</b>) pileipellis. Bars: (<b>a</b>) = 3 mm; (<b>b</b>) = 10 μm; (<b>c</b>) = 5 μm; (<b>d</b>,<b>e</b>) = 20 μm. Drawing by Menghui Han.</p>
Full article ">Figure 5
<p>Fresh basidiomata of <span class="html-italic">Crepidotus capitatocystidiatus</span> sp. nov. (<b>a</b>,<b>c</b>,<b>d</b>,<b>f</b>) <span class="html-italic">FFAAS1311</span>; (<b>b</b>) <span class="html-italic">FFAAS1312</span>; (<b>e</b>,<b>g</b>,<b>h</b>) <span class="html-italic">FFAAS1310</span> (holotype); (<b>a</b>) lamellae edge fimbriated when matured; (<b>c</b>) clear villosity near the point of attachment; (<b>d</b>,<b>h</b>) pileus margin tomentose. Bars: (<b>a</b>–<b>d</b>,<b>h</b>) = 5 mm; (<b>e</b>–<b>g</b>) = 1 mm. Photos by Yupeng Ge, Junqing Yan and Menghui Han.</p>
Full article ">Figure 6
<p>Microscopic features of <span class="html-italic">Crepidotus capitatocystidiatus</span> (<span class="html-italic">FFAAS1310</span>, holotype). (<b>a</b>–<b>e</b>) Lateral view of basidiospores; (<b>f</b>) frontal view of basidiospores; (<b>g</b>,<b>h</b>) basidia; (<b>i</b>–<b>r</b>) cheilocystidia; (<b>s</b>) lamellae trama; (<b>t</b>) pileipellis, clamp connection of pileipellis cell and oleiferous hyphae. Bars: (<b>a</b>–<b>f</b>) = 5 μm; (<b>g</b>–<b>r</b>) = 10 μm; (<b>s</b>,<b>t</b>) = 20 μm. Structures (<b>a</b>–<b>f</b>) were rehydrated in 5% KOH aqueous solution and (<b>g</b>–<b>t</b>) were stained in 1% Congo red aqueous solution.</p>
Full article ">Figure 7
<p>Morphological features of <span class="html-italic">Crepidotus capitatocystidiatus</span> (<span class="html-italic">FFAAS</span>1310, holotype). (<b>a</b>) Basidiomata; (<b>b</b>) basidia; (<b>c</b>) cheilocystidia; (<b>d</b>) basidiospores; (<b>e</b>) pileipellis. Bars: (<b>a</b>) = 5 mm; (<b>b</b>–<b>c</b>) = 10 μm; (<b>d</b>) = 5 μm; (<b>e</b>) = 20 μm. Drawing by Menghui Han.</p>
Full article ">Figure 8
<p>Fresh basidiomata of <span class="html-italic">Crepidotus succineus</span> sp. nov. (<b>a</b>,<b>c</b>,<b>d</b>,<b>f</b>,<b>h</b>) <span class="html-italic">FFAAS1313</span> (holotype); (<b>b</b>,<b>e</b>,<b>g</b>) <span class="html-italic">FFAAS1315</span>; (<b>i</b>,<b>j</b>) <span class="html-italic">FFAAS1314</span>; (<b>d</b>) short tomentums when immature in pileus; (<b>e</b>) lamellae in side view; (<b>h</b>) tomentums near the point of attachment; (<b>i</b>) pubescence when matured on pileus; (<b>j</b>) lamellae edge smooth when matured. Bars: (<b>a</b>,<b>c</b>–<b>f</b>,<b>h</b>–<b>j</b>) = 5 mm; (<b>b</b>,<b>g</b>) = 10 mm. Photos by Yupeng Ge and Menghui Han.</p>
Full article ">Figure 9
<p>Microscopic features of <span class="html-italic">Crepidotus succineus</span> (<span class="html-italic">FFAAS</span>1313, holotype). (<b>a</b>–<b>e</b>) Lateral view of basidiospores; (<b>f</b>) frontal view of basidiospores; (<b>g</b>,<b>h</b>) basidia; (<b>i</b>–<b>l</b>) pileocystidia; (<b>m</b>–<b>r</b>) cheilocystidia; (<b>s</b>) lamellae trama; (<b>t</b>) pileipellis. Bars: (<b>a</b>–<b>f</b>) = 5 μm; (<b>g</b>,<b>h</b>) = 10 μm; (<b>i</b>–<b>t</b>) = 20 μm. Structures (<b>a</b>–<b>f</b>) were rehydrated in 5% KOH aqueous solution and (<b>g</b>–<b>t</b>) were stained in 1% Congo red aqueous solution.</p>
Full article ">Figure 10
<p>Morphological features of <span class="html-italic">Crepidotus succineus</span> (<span class="html-italic">FFAAS</span>1313, holotype). (<b>a</b>) Basidiomata; (<b>b</b>) basidia; (<b>c</b>) basidiospores; (<b>d</b>) pileocystidia; (<b>e</b>) cheilocystidia; (<b>f</b>) pileipellis. Bars: (<b>a</b>) = 10 mm; (<b>b</b>) = 10 μm; (<b>c</b>) = 5 μm; (<b>d</b>–<b>f</b>) = 20 μm. Drawing by Menghui Han.</p>
Full article ">Figure 11
<p>Fresh basidiomata of <span class="html-italic">Crepidotus clavocystidiatus</span> sp. nov. (<b>a</b>,<b>b</b>,<b>d</b>,<b>e</b>,<b>g</b>–<b>j</b>) <span class="html-italic">FFAAS1319</span>; (<b>c</b>,<b>f</b>) <span class="html-italic">FFAAS1316</span> (holotype); (<b>a</b>) lamellae edge fimbriated when matured; (<b>c</b>) villosity and tomentums on pileus surface; (<b>i</b>) stipe pruinose. Bars: (<b>a</b>–<b>d</b>) = 5 mm; (<b>e</b>–<b>i</b>) = 3 mm; (<b>j</b>) = 1 mm. Photos by Yupeng Ge and Menghui Han.</p>
Full article ">Figure 12
<p>Microscopic features of <span class="html-italic">Crepidotus clavocystidiatus</span> (<span class="html-italic">FFAAS</span>1316, holotype). (<b>a</b>–<b>e</b>) Lateral view of basidiospores; (<b>f</b>) frontal view of basidiospores; (<b>g</b>,<b>h</b>) basidia; (<b>i</b>–<b>r</b>) cheilocystidia; (<b>s</b>) lamellae trama; (<b>t</b>) pileipellis, pigment and encrusted hyphae. Bars: (<b>a</b>–<b>f</b>) = 5 μm; (<b>g</b>–<b>r</b>) = 10 μm; (<b>s</b>,<b>t</b>) = 30 μm. Structures (<b>a</b>–<b>f</b>) were rehydrated in 5% KOH aqueous solution and (<b>g</b>–<b>t</b>) were stained in 1% Congo red aqueous solution.</p>
Full article ">Figure 13
<p>Morphological features of <span class="html-italic">Crepidotus clavocystidiatus</span> (<span class="html-italic">FFAAS</span>1316, holotype). (<b>a</b>) Basidiomata; (<b>b</b>) basidia; (<b>c</b>) basidiospores; (<b>d</b>) cheilocystidia; (<b>e</b>) pileipellis. Bars: (<b>a</b>) = 5 mm; (<b>b</b>) = 10 μm; (<b>c</b>) = 5 μm; (<b>d</b>,<b>e</b>) = 20 μm. Drawing by Menghui Han.</p>
Full article ">
18 pages, 4127 KiB  
Article
Karyotype’s Rearrangement in Some Hybrids of the Orchidinae Subtribe
by Alessio Turco, Robert Philipp Wagensommer, Antonella Albano, Pietro Medagli and Saverio D’Emerico
Plants 2024, 13(20), 2838; https://doi.org/10.3390/plants13202838 - 10 Oct 2024
Viewed by 404
Abstract
Based on our karyological findings in the Anacamptis Rich., Ophrys L., and Serapias L. genera, we have identified chromosomal markers within some hybrids and elucidated their interrelationships. Mitotic chromosomes of fifteen taxa were analyzed using the conventional Feulgen staining method. Only for Anacamptis [...] Read more.
Based on our karyological findings in the Anacamptis Rich., Ophrys L., and Serapias L. genera, we have identified chromosomal markers within some hybrids and elucidated their interrelationships. Mitotic chromosomes of fifteen taxa were analyzed using the conventional Feulgen staining method. Only for Anacamptis ×gennarii (Rchb. f.) H.Kretzschmar, Eccarius & Dietr. [A. morio (L.) R.M.Bateman, Pridgeon & M.W.Chase × A. papilionacea (L.) R.M.Bateman, Pridgeon & M.W.Chase] and its parental species were some data obtained and reported with the banding method with Giemsa, Hoechst 33258 fluorochrome, and the FISH techniques. Our research involved new chromosomal measurements of fifteen taxa, including six hybrids, along with schematic representations. Morphometric parameters, i.e., MCA and CVCL, were used to evaluate karyotype asymmetry. Of meaning were the analyses performed on chromosomal complements of selected hybrids, which distinctly revealed marker chromosomes present in one or both putative parental species. Among the parents identified in some hybrids, Ophrys tenthredinifera Willd. has shown some interest due to the presence in its karyotype of a pair of chromosomes (n.1) showing a notable secondary constriction on the long arm. Indeed, one of the homologs is clearly distinguishable in the analyzed hybrids, where it clearly emerges as one of the putative parents. Given the challenges in detecting certain karyomorphological features within the Orchidinae subtribe using alternative methods, such as Giemsa C-banding or fluorescence banding, the Feulgen method remains valuable for cytogenetic characterization. It helps us to understand the genomes of hybrids and parental species, thus contributing to a deeper understanding of their genetic composition. Full article
(This article belongs to the Special Issue Biosystematics and Phylogenetic Taxonomy of Plants)
Show Figures

Figure 1

Figure 1
<p>Metaphase chromosomes of (<b>A</b>) <span class="html-italic">Ophrys apulica</span>, 2n = 36; (<b>B</b>) <span class="html-italic">O. tenthredinifera</span>, 2n = 36; (<b>C</b>) <span class="html-italic">O. incubacea</span>, 2n = 36; (<b>D</b>) <span class="html-italic">O. bombyliflora</span>, 2n = 36; (<b>E</b>) <span class="html-italic">Anacamptis morio</span>, 2n = 36; (<b>F</b>) <span class="html-italic">A. collina</span>, 2n = 36; (<b>G</b>) <span class="html-italic">A. papilionacea</span>, 2n = 32; (<b>H</b>) <span class="html-italic">Serapias parviflora</span>, 2n = 36; (<b>I</b>) <span class="html-italic">Ophrys ×salentina</span>, 2n = 36; (<b>J</b>) <span class="html-italic">Ophrys ×franciniae</span>, 2n = 36; (<b>K</b>) <span class="html-italic">Ophrys ×sommieri</span>, 2n = 36; (<b>L</b>) <span class="html-italic">Anacamptis ×gennarii</span>, 2n = 34; (<b>M</b>) <span class="html-italic">Anacamptis ×semisaccata</span> nothosubsp. <span class="html-italic">murgiana</span>, 2n = 36; (<b>N</b>) <span class="html-italic">×Serapicamptis nelsoniana</span>, 2n = 36; (<b>O</b>) <span class="html-italic">Ophrys tardans</span>, 2n = 36. Scale bar = 5 µm.</p>
Full article ">Figure 2
<p>Karyotypes of (<b>A</b>) <span class="html-italic">Ophrys apulica</span>; (<b>B</b>,<b>B<sup>1</sup></b>) <span class="html-italic">O. ×salentina</span> (two karyotypes of different specimens); (<b>C</b>) <span class="html-italic">O. tenthredinifera</span>; (<b>D</b>) <span class="html-italic">O. apulica</span>; (<b>E</b>) <span class="html-italic">O. ×franciniae</span>; (<b>F</b>) <span class="html-italic">O. incubacea</span>; (<b>G</b>) <span class="html-italic">O. bombyliflora</span>; (<b>H</b>) <span class="html-italic">O. ×sommieri</span>; (<b>I</b>) <span class="html-italic">O. tenthredinifera</span>. Asterisks and plus signs indicate marker chromosomes observed in the analyzed taxa. Scale bar = 5 µm.</p>
Full article ">Figure 3
<p>Karyotypes of (<b>A</b>) <span class="html-italic">Anacamptis morio</span>; (<b>B</b>) <span class="html-italic">A. ×semisaccata nothosubsp. murgiana</span>; (<b>C</b>) <span class="html-italic">A. collina</span>; (<b>D</b>) <span class="html-italic">A. morio</span>; (<b>E</b>) <span class="html-italic">A.</span> ×<span class="html-italic">gennarii</span>; (<b>F</b>) <span class="html-italic">A. papilionacea</span>. Asterisks and plus signs indicate marker chromosomes observed in the analyzed taxa. Scale bar = 5 µm.</p>
Full article ">Figure 4
<p>Somatic metaphases of <span class="html-italic">Anacamptis</span> ×<span class="html-italic">gennarii</span> with possible separation of the haploid complement of <span class="html-italic">A. morio</span> (<b>A</b>–<b>D</b>) and haploid complement of <span class="html-italic">A. papilionacea</span> (<b>A1</b>–<b>D1</b>). It is possible to notice in the two kits a notable variation of the chromosomes following notable rearrangements during meiosis. (<b>D</b>,<b>D1</b>) Somatic metaphases of <span class="html-italic">Anacamptis</span> ×<span class="html-italic">gennarii</span> staining with Giemsa C-band (asterisks indicate telomeric bands). (<b>E</b>–<b>G</b>) Staining with Giemsa C-band (arrows indicate telomeric bands): <span class="html-italic">A. morio</span> (<b>E</b>), <span class="html-italic">A. papilionacea</span> (<b>F</b>), <span class="html-italic">A. ×gennarii</span> (<b>G</b>). (<b>H</b>–<b>J</b>) Somatic metaphases of <span class="html-italic">A. morio</span> (<b>H</b>), <span class="html-italic">A. papilionacea</span> (<b>I</b>), <span class="html-italic">A.</span> ×<span class="html-italic">gennarii</span> (<b>J</b>) treated with the fluorochrome Hoechst 33258. In <span class="html-italic">A. papilionacea,</span> we can observe four chromosomes with telomeric bands; in <span class="html-italic">A.</span> ×<span class="html-italic">gennarii</span>, we observe only two chromosomes with telomeric bands belonging to <span class="html-italic">A. papilionacea</span> (arrows). Differently, <span class="html-italic">A. morio</span> does not show any important banding. Scale bar = 5 µm.</p>
Full article ">Figure 5
<p>In situ hybridization applied to the chromosomes of <span class="html-italic">Anacamptis morio</span>, <span class="html-italic">A.</span> ×<span class="html-italic">gennarii</span>, and <span class="html-italic">A. papilionacea</span>. Blue DAPI staining shows chromosomal DNA, respectively, in <span class="html-italic">A. morio</span> (<b>A</b>), <span class="html-italic">A.</span> ×<span class="html-italic">gennarii</span> (<b>D</b>), and <span class="html-italic">A. papilionacea</span> (<b>G</b>). Red and green signals show sites of hybridization of 18S-25S rDNA and 5S rDNA: in <span class="html-italic">A. morio</span> (<b>C</b>), four 18S-25S rDNA sites and two 5S rDNA sites; in <span class="html-italic">A.</span> ×<span class="html-italic">gennarii</span> (<b>F</b>), three 18S-25S rDNA sites and three 5S rDNA sites; in <span class="html-italic">A. papilionacea</span> (<b>I</b>), two 18S-25S rDNA sites and four 5S rDNA sites. Red signals show sites of hybridization of 18S-25S rDNA (<b>B</b>,<b>E</b>,<b>H</b>) in the three taxa. Scale bar = 5 µm.</p>
Full article ">Figure 6
<p>Karyotypes of (<b>A</b>) <span class="html-italic">Anacamptis collina</span>; (<b>B</b>) ×<span class="html-italic">Serapicamptis nelsoniana</span>; (<b>C</b>) <span class="html-italic">Serapias parviflora</span>; (<b>D</b>) <span class="html-italic">Ophrys tardans</span> (Three karyotypes in different specimens); (<b>E</b>) <span class="html-italic">Ophrys tenthredinifera</span>. Asterisks and plus signs indicate marker chromosomes observed in the analyzed taxa. Scale bar = 5 µm.</p>
Full article ">Figure 7
<p>Diagram of the M<sub>CA</sub> and CV<sub>CL</sub> values of the karyotypes of the taxa examined in <span class="html-italic">Ophrys</span> species and hybrids. In the diagram, it is interesting to note the parameter values of the hybrids intermediate to the supposed parents. Codes: see <a href="#plants-13-02838-t001" class="html-table">Table 1</a>.</p>
Full article ">Figure 8
<p>Diagram of the M<sub>CA</sub> and CV<sub>CL</sub> values of the karyotypes of the taxa examined in <span class="html-italic">Anacamptis</span>, <span class="html-italic">Serapias</span> species, and hybrids. In the diagram, it is interesting to note the parameter values of the hybrids intermediate to the supposed parents. Codes: see <a href="#plants-13-02838-t001" class="html-table">Table 1</a>.</p>
Full article ">
19 pages, 7114 KiB  
Article
The Diversity of Arachnid Assemblages on the Endemic Tree Zelkova abelicea (Ulmaceae): An Evaluation of Fragmentation and Connectivity in Crete (Greece)
by Dariusz J. Gwiazdowicz, Laurence Fazan, Giulio Gardini, Dany Ghosn, Sławomir Kaczmarek, Alireza Nemati, Ilektra Remoundou, Tomasz Rutkowski, Piotr Skubała, Bogna Zawieja and Gregor Kozlowski
Insects 2024, 15(10), 788; https://doi.org/10.3390/insects15100788 - 10 Oct 2024
Viewed by 449
Abstract
Zelkova abelicea is an endemic tree growing only on eight mountain stands on the Greek island of Crete. The aim of this study was to determine the structure of the assemblages and analyze the diversity of the arachnid assemblages living on Zelkova abelicea [...] Read more.
Zelkova abelicea is an endemic tree growing only on eight mountain stands on the Greek island of Crete. The aim of this study was to determine the structure of the assemblages and analyze the diversity of the arachnid assemblages living on Zelkova abelicea, an endemic tree species in Crete. Material for the analyses was collected from tree trunks, oftentimes covered by bryophytes or lichens. In the examined material, 85 taxa were recorded. The most numerous groups represented in the analyzed material were Acari, including representatives of the orders Mesostigmata (78 ind. of 18 spp.) and Oribatida (1056 ind. of 51 spp.). In the order Mesostigmata the species represented by the highest numbers of specimens were Onchodellus karawaiewi (15 individuals) and Hypoaspisella sp. (13), which is probably a species new to science. In turn, representatives of the order Oribatida were much more numerous, with Zygoribatula exilis (284) and Eremaeus tuberosus (210) being identified in the largest numbers. Among the eight sampled localities, Gerakari (646 ind. and 50 spp.) and Omalos (409 ind. and 43 spp., respectively) had by far the richest assemblages. Statistical analyses confirmed the highly diverse character of the arachnid assemblages at the individual sites, which is a consequence not only of the varied numbers of arachnids found, but also of the presence of very rare species, such as Androlaelaps shealsi, Cosmolaelaps lutegiensis or Hoploseius oblongus. These results highlight the high species diversity of the arachnids found on Z. abelicea but also suggest the lack of connectivity between the isolated and fragmented forest stands on Crete. Full article
(This article belongs to the Section Insect Ecology, Diversity and Conservation)
Show Figures

Figure 1

Figure 1
<p>Sampled localities (red dots) in Crete (Greece) with <span class="html-italic">Zelkova abelicea</span> trees.</p>
Full article ">Figure 2
<p><span class="html-italic">Zelkova abelicea</span> trees with microhabitats for invertebrates. (<b>A</b>) Large trees (Omalos). (<b>B</b>) Dwarfed trees heavily browsed by goats (Omalos). (<b>C</b>) Bark of large tree covered by species of lichens from genus <span class="html-italic">Parmelina</span> (Gerakari). (<b>D</b>) Bark of large trees (Gerakari). (<b>E</b>) View of <span class="html-italic">Z. abelicea</span> population in Gerakari (Photos: G. Kozlowski).</p>
Full article ">Figure 3
<p>The range of arachnida specimens (<b>A</b>) and species (<b>B</b>) per study site. The width of the violin plots represents the number of individuals (<b>A</b>) or species (<b>B</b>).</p>
Full article ">Figure 3 Cont.
<p>The range of arachnida specimens (<b>A</b>) and species (<b>B</b>) per study site. The width of the violin plots represents the number of individuals (<b>A</b>) or species (<b>B</b>).</p>
Full article ">Figure 4
<p>The centroids determined from the NMDS analysis showing the numerical diversity among the samples for each study site for Acari. The center of each centroid is indicated by a lettered square representing the study site (i.e., O: Omalos, N: Niato, I: Impros, G: Gerakari, R: Rouvas, V: Viannou, K: Katharo and T: Thripti).</p>
Full article ">Figure 5
<p>A cluster analysis showing the similarity of study sites depending on their Acari communities.</p>
Full article ">Figure 6
<p>Heatmap showing species frequency.</p>
Full article ">
29 pages, 10333 KiB  
Article
How to Recognize Mosses from Extant Groups among Paleozoic and Mesozoic Fossils
by Michael S. Ignatov, Tatyana V. Voronkova, Ulyana N. Spirina and Svetlana V. Polevova
Diversity 2024, 16(10), 622; https://doi.org/10.3390/d16100622 - 8 Oct 2024
Viewed by 629
Abstract
This paper describes a range of Paleozoic and Mesozoic mosses and assesses how far they can be referred to extant taxa at the family, ordinal, or class levels. The present study provides new data on Paleozoic mosses of the order Protosphagnales, re-evaluating affinities [...] Read more.
This paper describes a range of Paleozoic and Mesozoic mosses and assesses how far they can be referred to extant taxa at the family, ordinal, or class levels. The present study provides new data on Paleozoic mosses of the order Protosphagnales, re-evaluating affinities of some groups previously thought to be unrelated. The leaf areolation pattern, combined with the leaf costa anatomy, results in the subdivision of Protosphagnales into five separate families: Protosphagnaceae (at least six genera), Polyssaieviaceae (at least three genera), and three monogeneric families: Rhizonigeritaceae, Palaeosphagnaceae, and Servicktiaceae. We urge caution in referring Paleozoic and Early Mesozoic fossil mosses as members of Dicranidae and Bryidae, as they may belong to the extinct moss order Protosphagnales. Additional evidence supports the relation of the Permian genus Arvildia to extant Andreaeopsida. We segregate Late Palaeozoic and Early Mesozoic mosses that are superficially similar to extant members of either Dicranales or Polytrichales, into the artificial informal group of Archaeodicranids, distinguishing them from ecostate Paleozoic and Mesozoic mosses, which are combined here into another artificial informal group, Bryokhutuliinids. The latter includes the genus Bryokhutuliinia, widespread in contemporary Asia, from the Middle Jurassic to the Lower Cretaceous, as well as other superficially similar ecostate plants from different regions worldwide, ranging from the Upper Palaeozoic to the Lower Cretaceous. A list of Paleozoic, Mesozoic, and Eocene moss fossils suitable for age calibration in phylogenetic trees is provided. Full article
(This article belongs to the Section Plant Diversity)
Show Figures

Figure 1

Figure 1
<p>Protosphagnalean mosses showing typical dimorphic areolation pattern for <span class="html-italic">Protosphagnum</span> (<b>H</b>,<b>I</b>), mostly monomorphic for <span class="html-italic">Intia</span> (<b>B</b>,<b>D</b>,<b>F</b>,<b>J</b>) and <span class="html-italic">Kosjunia</span> (<b>G</b>), and combining mono and dimorphic areolation types in different parts of leaves (<b>A</b>,<b>C</b>,<b>E</b>,<b>H</b>). (<b>A</b>) Stem with leaves, (<b>B</b>) young leaves crowded at stem tip, and (<b>C</b>–<b>J</b>) leaf fragments. Aristovo, Permian (Lopingian): (<b>A</b>) 126A-3A, (<b>B</b>) 49B-5, (<b>C</b>) 16B-6, (<b>D</b>) 39B-7, (<b>E</b>) 100A-9, (<b>F</b>) 44A-1, (<b>G</b>) 47B-9, (<b>H</b>) 31A-3, (<b>I</b>) 38A-6, and (<b>J</b>) 43B-9.</p>
Full article ">Figure 2
<p><span class="html-italic">Protosphagnum nervatum</span>: (<b>A</b>) fully developed apical areolation; (<b>B</b>,<b>D</b>) developing apical areolation; (<b>C</b>,<b>E</b>) young leaves with costa branches confluent with rows of laminal areolation; (<b>F</b>–<b>K</b>) branch primordia with surrounding leaves, red arrow in H points one of primordia magnified in G, LM (<b>F</b>–<b>J</b>) and SEM (<b>K</b>) images. Aristovo, Permian (Lopingian): (<b>A</b>) 5A-3, (<b>B</b>) 11B-2, (<b>C</b>) 19A-2, (<b>D</b>) 47A-1, (<b>E</b>) 19A-3, (<b>F</b>) 126B-1, (<b>G</b>) 126B-2, (<b>H</b>) 126B-2, (<b>I</b>) 126B-4, (<b>J</b>) 126B-1, and (<b>K</b>) CUT_SEM_4.</p>
Full article ">Figure 3
<p><span class="html-italic">Protosphagnum nervatum</span>. (<b>A</b>–<b>C</b>) Araldite-embedded leaf transverse sections, 2 μm thick, under LM (<b>A</b>) and 60 nm thick under TEM (<b>B</b>–<b>C</b>), showing the homogeneous structure of the fossil; (<b>D</b>,<b>E</b>) costa surface; (<b>D</b>–<b>J</b>,<b>L</b>,<b>M</b>) SEM images of leaves, showing dimorphic cells with very thin walls of the hyalocysts (<b>F</b>–<b>J</b>,<b>L</b>,<b>M</b>), partly broken (<b>G</b>,<b>H</b>,<b>J</b>) or mostly retained (<b>I</b>,<b>L</b>,<b>M</b>); (<b>K</b>) juvenile leaf (shown in whole in inset) areolation, LM image, showing the delicate nature of the hyalocyst cell walls. Unistratose part of costa is seen in (<b>I</b>) (arrowed). Aristovo, Permian (Lopingian): (<b>A</b>) CUT_S5_3_37, (<b>B</b>–<b>C</b>) CUT_TEM_5, (<b>D</b>–<b>J</b>,<b>L</b>,<b>M</b>) CUT_SEM_1, and (<b>K</b>) 100B-2.</p>
Full article ">Figure 4
<p><span class="html-italic">Protosphagnum nervatum</span>. (<b>A</b>,<b>B</b>) Transverse costa sections, SEM images, of leaf shown in (<b>D</b>,<b>E</b>), LM images; (<b>C</b>,<b>F</b>–<b>J</b>) transverse costa sections, SEM images, of leaf shown in (<b>H</b>,<b>I</b>), LM images. Note the unistratose costa, in which cells look to be filled with spongy material (<b>C</b>,<b>F</b>), likely an effect of fossilization. Aristovo, Permian (Lopingian): (<b>A</b>,<b>B</b>,<b>D</b>,<b>E</b>) CUT_SEM_2, and (<b>C</b>,<b>F</b>–<b>J</b>) CUT_SEM_3.</p>
Full article ">Figure 5
<p><span class="html-italic">Rhizinigerites neuburgae</span>. (<b>A</b>, with close up of rhizoidophore in <b>G</b>) habit, showing the branched stem with leaves and rhizoidophore; the A-inset shows the branch bud (where the arrow points) that has no foliate structure on the stem around it; the G-inset shows the end of the rhizoidophore with rhizoid clusters. (<b>B</b>) A leaf fragment showing areolation with some of the cells missing and no marginal border; the inset highlights the protosphagnalean areolation pattern. (<b>C</b>,<b>D</b>) The leaf apical parts and areolation in different parts of the leaves. (<b>E</b>,<b>F</b>) The laminal cells, showing the areolation variation. (<b>H</b>) A part of the rhizoidophore separated by places with abundant rhizoids. (<b>I</b>) A rhizoid cluster on the rhizoidophore. (<b>J</b>,<b>K</b>) Lower leaf parts showing the unistratose veins of long cells, diverging from the main costa (red arrows). Viled, Permian (Lopingian). See details of the locality in [<a href="#B15-diversity-16-00622" class="html-bibr">15</a>,<a href="#B16-diversity-16-00622" class="html-bibr">16</a>]. (<b>A</b>–<b>C</b>,<b>E</b>,<b>F</b>,<b>H</b>–<b>K</b>) GIN 3774/3B-10-1, (<b>D</b>) GIN 3774/3B-10-2, and (<b>G</b>) GIN 3774/3B-5-9.</p>
Full article ">Figure 6
<p><span class="html-italic">Palaeosphagnum meyenii</span>. (<b>A</b>–<b>D</b>) Leaf fragments and details of areolation. (<b>E</b>–<b>G</b>) Leaf fragment used for sectioning, shown in (<b>H</b>–<b>K</b>). (<b>H</b>–<b>K</b>) Transverse sections of leaf fragment F, which is 60 nm thick, TEM, showing a unistratose lamina with partly inflated cells (<b>H</b>,<b>J</b>). Multistratose costa (<b>I</b>), and bistratose area flanking the costa (<b>K</b>). Aristovo, Permian (Lopingian), (<b>A</b>) 124B-13, (<b>B</b>) 105B-8, (<b>C</b>) 125A-1, (<b>D</b>) 105B-8, and (<b>E</b>–<b>K</b>) CUT_TEM_P9.</p>
Full article ">Figure 7
<p><span class="html-italic">Servicktia undulata</span>. (<b>A</b>,<b>B</b>) Leaf fragment and the cells of its border. (<b>C</b>–<b>E</b>) Its transverse sections, showing a rough cell surface due to irregular papillae and probably prorate cell ends (<b>C</b>), unistratose lamina (<b>D</b>), and multistratose costa (<b>E</b>). Aristovo, Permian (Wushiapingian), (<b>A</b>–<b>E</b>) CUT_TEM_P11.</p>
Full article ">Figure 8
<p><span class="html-italic">Servicktia vorcutannularioides</span>. (<b>A</b>–<b>D</b>) Leaf fragments and details of dimorphic cell areolation. (<b>E</b>–<b>G</b>) Transverse sections under TEM (<b>E</b>,<b>F</b>) and LM (<b>G</b>), showing multistratose costa (<b>E</b>,<b>G</b>), unistratose lamina (<b>F</b>), and inflated cells on ventral side of costa (E, G). Aristovo, Permian (Lopingian), (<b>A</b>) 107A-2, (<b>B</b>) 106A-5, (<b>C</b>) 106A-5, (<b>D</b>) 106A-10, and (<b>E</b>–<b>G</b>) CUT_TEM_P6.</p>
Full article ">Figure 9
<p><span class="html-italic">Servicktia tatyanae</span>. Holotype: (<b>A</b>–<b>C</b>), whole leaf fragment and close ups of upper and lower leaf parts. Aristovo, Permian (Lopingian), (<b>A</b>–<b>C</b>) 100B-1.</p>
Full article ">Figure 10
<p><span class="html-italic">Polyssaievia spinulifolia</span> (<b>A</b>–<b>H</b>), <span class="html-italic">Polyssaievia deflexa</span> (<b>I</b>), and young leaf of <span class="html-italic">Protosphagnum nervatum</span> (<b>J</b>), showing variation in areolation in different parts of leaves of <span class="html-italic">Polyssaievia</span> and similarity in areolation to juvenile leaf of <span class="html-italic">P. nervatum</span>. (<b>A</b>) shoot, (<b>B</b>–<b>I</b>) leaf fragments, showing ‘net venation’ in proximal part of leaves and prosenchymatous cells in their distal part (<b>B</b>,<b>E</b>,<b>F</b>). (<b>J</b>) Juvenile leaf. Permian (Lopingian) specimens from localities described for A–I in [<a href="#B11-diversity-16-00622" class="html-bibr">11</a>] and for J in [<a href="#B23-diversity-16-00622" class="html-bibr">23</a>]: (<b>A</b>–<b>D</b>) Tunguska coal basin GIN 3087/1019-3, (<b>E</b>,<b>F</b>) Tunguska coal basin GIN 3087/1018-4, (<b>G</b>,<b>H</b>) Kuznetsk coal basin GIN 3026/95A, (<b>I</b>) Pechora coal basin GIN 3041_151c, and (<b>J</b>) Pechora coal basin MHA: Adzva_32M_20_4_A_3.</p>
Full article ">Figure 11
<p><span class="html-italic">Arvildia elenae</span> (<b>A</b>,<b>B</b>,<b>E</b>–<b>K</b>), compared with extant <span class="html-italic">Andreaea rothii</span> F. Weber &amp; D. Mohr (<b>C</b>) and <span class="html-italic">Andreaeobryum macrosporum</span> Steere &amp; B.M. Murray (<b>D</b>,<b>L</b>). (<b>A</b>,<b>B</b>) Leaf apical part showing uniseriate and then, from the 5th cell from the apex, abruptly a biseriate cell arrangement, typical of the Andreaeales and Andreaeobryales. (<b>C</b>,<b>D</b>) Leaf apices with uniseriate and then biseriate cells. (<b>E</b>,<b>F</b>) Apical parts of two leaves, apparently from the same shoot, and their transverse sections showing the most distal unicellular part of the leaf on the left, as shown in the 2 μm section under LM. (<b>G</b>,<b>H</b>) A densely foliate shoot and sublongitudinal section of the apical leaf part under TEM. (<b>I</b>–<b>K</b>) A leaf and its cross-section under TEM, showing only moderately differentiated costal cells, comparable to the <span class="html-italic">Andreaeobryum</span> sections, shown in L. (<b>L</b>) Transverse sections of the distal leaf parts above the stem apex of <span class="html-italic">Andreaeobryum</span>, showing a moderately differentiated costa, a 2 μm section, with fluorescent microscopy. <span class="html-italic">Arvildia</span> collections are from Aristovo, Permian (Lopingian): (<b>A</b>,<b>B</b>) 113-1, (<b>E</b>,<b>F</b>) CUT_TEM_P15, (<b>G</b>,<b>H</b>) CUT_TEM_P16, and (I–K) CUT_TEM_P12. Extant specimens are from: (<b>C</b>) Norway, MW9000070, and (<b>D</b>,<b>L</b>) Russia, MHA9022375.</p>
Full article ">Figure 12
<p><span class="html-italic">Kulindobryum taylorioides</span> (<b>A</b>,<b>B</b>,<b>D</b>,<b>E</b>) compared with extant <span class="html-italic">Tayloria splachnoides</span> (Schleich. ex Schwär.) Hook. (<b>C</b>). ((<b>A</b>), and its close up in (<b>B</b>)): capsule with long neck, partly broken at mouth, where pendent peristome teeth occur, (<b>C</b>) open capsule with 32 peristome teeth, (<b>D</b>) obliquely compressed capsule showing deoperculate mouth with peristome teeth fragments, and (<b>E</b>) still operculate capsule covered by calyptra. Peristome teeth or their fragments arrowed. Transbaikalia, Kulinda, Middle Jurassic (for locality information see references in [<a href="#B18-diversity-16-00622" class="html-bibr">18</a>]): ((<b>A</b>),(<b>B</b>): PIN 5648/2, (<b>D</b>): PIN 5648/3, (<b>E</b>): PIN 5648/1) and (<b>C</b>) Russia, Urals, MHA9020838.</p>
Full article ">
26 pages, 2268 KiB  
Article
Exploring Floristic Diversity, Propagation Patterns, and Plant Functions in Domestic Gardens across Urban Planning Gradient in Lubumbashi, DR Congo
by Yannick Useni Sikuzani, Bernard Kisangani Kalonda, Médard Mpanda Mukenza, Jonas Yona Mleci, Alex Mpibwe Kalenga, François Malaisse and Jan Bogaert
Ecologies 2024, 5(4), 512-537; https://doi.org/10.3390/ecologies5040032 - 2 Oct 2024
Viewed by 442
Abstract
Urbanization degrades natural habitats and creates new urban ecosystems like domestic gardens. The plant composition of these gardens varies with socio-economic factors and urban planning levels. However, the diversity and impact of introduced species are often poorly assessed, causing potential ecological imbalances (disruptions [...] Read more.
Urbanization degrades natural habitats and creates new urban ecosystems like domestic gardens. The plant composition of these gardens varies with socio-economic factors and urban planning levels. However, the diversity and impact of introduced species are often poorly assessed, causing potential ecological imbalances (disruptions in the natural functioning and stability of ecosystems), particularly in Lubumbashi (DR Congo). The objective was to analyze the spatial structure, plant diversity, propagation strategies, and ecological functions of domestic gardens. Three distinct neighborhoods were selected: a planned, unplanned, and residential neighborhood. Twenty avenues (with five plots per avenue) were chosen to represent the diversity within each neighborhood, and stratified random sampling of plots was conducted to analyze gardening practices. Gardens were classified into types, and their vegetation was evaluated based on species origin and ecological impact. The analysis of domestic gardens in Lubumbashi reveals significant variations across different neighborhood types. Residential neighborhoods exhibit larger average garden sizes (315.1 m2), higher species richness (22 species), and larger plot sizes (1032 m2) compared to unplanned and planned neighborhoods, where garden areas and species richness are notably lower. Rectangular gardens dominate in unplanned areas, while planned neighborhoods feature more intentional landscaping elements, such as flowerbeds and hedges. The use of gardens for food production is prominent in planned areas (40.7%), whereas residential neighborhoods prioritize ornamentation (51.4%). The study identified 232 taxa across 68 families, with a predominance of exotic species (80%) in all neighborhoods, particularly in unplanned areas (82.25%). The data revealed that Mangifera indica and Persea americana are abundant in all neighborhoods, illustrating their adaptability to different urban contexts. Herbaceous species are most common, followed by woody plants, with vines being sparse. Species dispersal is primarily driven by human activities (anthropochory), accounting for over 85% in all neighborhoods. These findings highlight the strong human influence on the composition and structure of domestic gardens in Lubumbashi, emphasizing the dominance of exotic species and the importance of anthropogenic factors in shaping urban green spaces. Urban policies should incorporate strategies to minimize the negative impacts of exotic species on native flora. Full article
Show Figures

Figure 1

Figure 1
<p>Geographical map of the Lubumbashi city in southeastern DRC. The map also shows the study sites: Gambela III, Bel-AirI, and Mampala neighborhoods. The red dot on the map locates the study area in the province of Haut-Katanga in the DRC.</p>
Full article ">Figure 2
<p>Origin status of plant species inventoried in domestic gardens across neighborhood types.</p>
Full article ">
15 pages, 7364 KiB  
Article
The Rediscovery of Noblella peruviana after More than 115 Years Helps Resolve the Molecular Phylogeny and Taxonomy of Noblella (Amphibia, Anura, Strabomantidae)
by Rudolf von May, M. Isabel Diaz, Alex Ttito, Roy Santa-Cruz and Alessandro Catenazzi
Diversity 2024, 16(10), 613; https://doi.org/10.3390/d16100613 - 1 Oct 2024
Viewed by 1410
Abstract
We revise the taxonomy of the frog genus Noblella on the basis of a molecular phylogeny. Previous studies recognized that Noblella is non-monophyletic, with one clade distributed from southeastern Peru to northeastern Bolivia and adjacent areas in Brazil and another clade distributed from [...] Read more.
We revise the taxonomy of the frog genus Noblella on the basis of a molecular phylogeny. Previous studies recognized that Noblella is non-monophyletic, with one clade distributed from southeastern Peru to northeastern Bolivia and adjacent areas in Brazil and another clade distributed from northern Peru to Ecuador and southeastern Colombia. The lack of sequences from the type species Noblella peruviana prevented the investigation of its phylogenetic position and the status of related taxa. Our rediscovery after more than 115 years allowed for the inclusion of DNA sequences of Noblella peruviana obtained from specimens collected at the type locality in southeastern Peru. We inferred a phylogeny based on a concatenated dataset (three mitochondrial and two nuclear loci) using Bayesian and maximum likelihood methods. Our phylogeny corroborated the non-monophyly of Noblella and helped resolve the status of related taxa, including Psychrophrynella bagrecito, the type species of the genus Psychrophrynella (rediscovered after 42 years). We identified a clade containing N. peruviana, P. bagrecito, and other species of Noblella and Psychrophrynella distributed in southern Peru. Given that the name Noblella predates Psychrophrynella, we propose that Psychrophrynella should be considered a junior synonym of Noblella. The second clade contains species of Noblella distributed in Ecuador and northern Peru, including N. myrmecoides, which used to be the type species of the genus Phyllonastes. Consequently, we propose to reinstate the genus Phyllonastes to accommodate all species of Noblella distributed in Ecuador, northern Peru, southeastern Colombia, and adjacent areas in Brazil. We present an updated taxonomy including new combinations for 12 species and reinstatements for three species. Full article
Show Figures

Figure 1

Figure 1
<p>(<b>Top</b>): Consensus Bayesian phylogeny based on a 2546-bp concatenated dataset (fragments of genes 16S, 12S, COI, RAG1, and Tyr) analyzed in MrBayes. Posterior probability values are indicated at each node. The area in gray indicates a clade with species distributed in northern Peru and Ecuador (“northern clade”), and the area in light blue indicates a clade with species distributed in southern Peru (“southern clade” containing the type species <span class="html-italic">Noblella peruviana</span>). Consensus Bayesian phylogeny based on a 2546-bp concatenated dataset (fragments of genes 16S, 12S, COI, RAG1, and Tyr) analyzed in MrBayes. Posterior probability values are indicated at each node. (<b>Bottom</b>): Part of the consensus Bayesian phylogeny showing closely related clades <span class="html-italic">Qosqophryne</span> and <span class="html-italic">Microkayla</span>, and more distantly related clades including <span class="html-italic">Lynchius</span>, <span class="html-italic">Oreobates</span>, and <span class="html-italic">Phrynopus</span>.</p>
Full article ">Figure 1 Cont.
<p>(<b>Top</b>): Consensus Bayesian phylogeny based on a 2546-bp concatenated dataset (fragments of genes 16S, 12S, COI, RAG1, and Tyr) analyzed in MrBayes. Posterior probability values are indicated at each node. The area in gray indicates a clade with species distributed in northern Peru and Ecuador (“northern clade”), and the area in light blue indicates a clade with species distributed in southern Peru (“southern clade” containing the type species <span class="html-italic">Noblella peruviana</span>). Consensus Bayesian phylogeny based on a 2546-bp concatenated dataset (fragments of genes 16S, 12S, COI, RAG1, and Tyr) analyzed in MrBayes. Posterior probability values are indicated at each node. (<b>Bottom</b>): Part of the consensus Bayesian phylogeny showing closely related clades <span class="html-italic">Qosqophryne</span> and <span class="html-italic">Microkayla</span>, and more distantly related clades including <span class="html-italic">Lynchius</span>, <span class="html-italic">Oreobates</span>, and <span class="html-italic">Phrynopus</span>.</p>
Full article ">Figure 2
<p><span class="html-italic">Noblella bagrecito</span>, adult specimens collected at the type locality (<b>A</b>–<b>F</b>) and at the nearby locality of Limacpunko (<b>G</b>,<b>H</b>). Snout-vent length (SVL) is given in mm. (<b>A</b>,<b>B</b>) male (CORBIDI 17482, SVL = 17.5 mm); (<b>C</b>,<b>D</b>) male (CORBIDI 17483, SVL = 12.5 mm); (<b>E</b>,<b>F</b>) female (CORBIDI 15727, SVL = 21.5 mm); (<b>G</b>,<b>H</b>) male (CORBIDI 15728, SVL = 12.5 mm).</p>
Full article ">Figure 2 Cont.
<p><span class="html-italic">Noblella bagrecito</span>, adult specimens collected at the type locality (<b>A</b>–<b>F</b>) and at the nearby locality of Limacpunko (<b>G</b>,<b>H</b>). Snout-vent length (SVL) is given in mm. (<b>A</b>,<b>B</b>) male (CORBIDI 17482, SVL = 17.5 mm); (<b>C</b>,<b>D</b>) male (CORBIDI 17483, SVL = 12.5 mm); (<b>E</b>,<b>F</b>) female (CORBIDI 15727, SVL = 21.5 mm); (<b>G</b>,<b>H</b>) male (CORBIDI 15728, SVL = 12.5 mm).</p>
Full article ">Figure 3
<p><span class="html-italic">Noblella peruviana</span>, adult specimens collected at the type locality (<b>A</b>–<b>F</b>) and at Oconeque, Aquele, Limbani, Puno (<b>G</b>,<b>H</b>). Snout-vent length (SVL) is given in mm. (<b>A</b>,<b>B</b>) female (CORBIDI 17510, SVL = 17.8 mm); (<b>C</b>,<b>D</b>) male (MUBI 19037, SVL = 13.3 mm); (<b>E</b>,<b>F</b>) female (CORBIDI 18700, SVL = 16.7 mm); (<b>G</b>,<b>H</b>) male (CORBIDI 18734, SVL = 11.0 mm).</p>
Full article ">Figure 3 Cont.
<p><span class="html-italic">Noblella peruviana</span>, adult specimens collected at the type locality (<b>A</b>–<b>F</b>) and at Oconeque, Aquele, Limbani, Puno (<b>G</b>,<b>H</b>). Snout-vent length (SVL) is given in mm. (<b>A</b>,<b>B</b>) female (CORBIDI 17510, SVL = 17.8 mm); (<b>C</b>,<b>D</b>) male (MUBI 19037, SVL = 13.3 mm); (<b>E</b>,<b>F</b>) female (CORBIDI 18700, SVL = 16.7 mm); (<b>G</b>,<b>H</b>) male (CORBIDI 18734, SVL = 11.0 mm).</p>
Full article ">Figure 4
<p>Map of northwestern South America showing the location of the type localities of species in the genera <span class="html-italic">Phyllonastes</span> (yellow dots) and <span class="html-italic">Noblella</span> (blue dots). (1 = <span class="html-italic">Phyllonastes worleyae</span>, 2 = <span class="html-italic">Phyllonastes mindo</span>, 3 = <span class="html-italic">Phyllonastes coloma</span>, 4 = <span class="html-italic">Phyllonastes naturetrekii</span>, 5 = <span class="html-italic">Phyllonastes personina</span>, 6 = <span class="html-italic">Phyllonastes lochites</span>, 7 = <span class="html-italic">Phyllonastes myrmecoides</span>, 8 = <span class="html-italic">Phyllonastes heyeri</span>, 9 = <span class="html-italic">Phyllonastes lynchi</span>, 10 = <span class="html-italic">Phyllonastes duellmani</span>, 11 = <span class="html-italic">Noblella vilcabambensis</span>, 12 = <span class="html-italic">Noblella madreselva</span>, 13 = <span class="html-italic">Noblella losamigos</span>, 14 = <span class="html-italic">Noblella pygmaea</span> and <span class="html-italic">Noblella usurpator</span>, 15 = <span class="html-italic">Noblella chirihampatu</span>, 16 = <span class="html-italic">Noblella bagrecito</span>, 17 = <span class="html-italic">Noblella glauca</span> and <span class="html-italic">Noblella thiuni</span>, 18 = <span class="html-italic">Noblella peruviana</span>, 19 = <span class="html-italic">Noblella ritarasquinae</span>, 20 = <span class="html-italic">Noblella carrascoicola</span>.</p>
Full article ">
16 pages, 3068 KiB  
Article
Differential Effects of Sulfur Fertilization on Soil Microbial Communities and Maize Yield Enhancement
by Siqi Dong, Bing Zhang, Wenfeng Hou, Xue Zhou and Qiang Gao
Agronomy 2024, 14(10), 2251; https://doi.org/10.3390/agronomy14102251 - 29 Sep 2024
Viewed by 382
Abstract
Sulfur (S) is an essential nutrient for plant growth, influencing not only crop yields but also the composition and function of soil microbial communities. However, the differential effects of S fertilization on abundant and rare taxa in agricultural soils remain poorly understood. This [...] Read more.
Sulfur (S) is an essential nutrient for plant growth, influencing not only crop yields but also the composition and function of soil microbial communities. However, the differential effects of S fertilization on abundant and rare taxa in agricultural soils remain poorly understood. This study investigates the impact of different S fertilizer types on maize yield and the structure and stability of soil microbial communities, with a particular focus on abundant and rare taxa. S fertilization led to significant increases maize yield on two typical soils (black soil and sandy soil) (5.3–24.3%) and altered soil properties, including reducing pH (0.04–0.20) and increasing the available sulfur (AS) content (3.8–8.0 mg kg−1), with ammonium sulfate having a more pronounced effect than elemental sulfur. Microbial analysis revealed distinct impacts on the diversity and community structure of both abundant and rare taxa. Elemental sulfur reduced the alpha diversity of abundant taxa more than ammonium sulfate, while NMDS indicated significant shifts in community structures, particularly among abundant taxa. Network analysis showed that S fertilization decreased the complexity of microbial interactions among rare taxa, with ammonium sulfate leading to simpler networks and elemental sulfur resulting in higher modularity. SEM highlighted that the diversity of rare taxa played a crucial role in influencing maize yield, alongside direct effects from soil properties such as AS and SAR (aryl sulfatase). Functional predictions demonstrated that amino acid metabolism and xenobiotic biodegradation and metabolism pathways were enriched in rare taxa, suggesting significant implications for soil health and crop productivity. This study provides new insights into the roles of abundant and rare bacterial taxa under S fertilization, emphasizing their importance in optimizing fertilization strategies for enhanced crop yield in specific soil types. Full article
Show Figures

Figure 1

Figure 1
<p>The histogram illustrates the response of maize yield to different S fertilizers (elemental sulfur and ammonium sulfate) in two soil types. Different letters indicate significant differences (<span class="html-italic">p</span> &lt; 0.05) between treatment groups within each soil type. For the analysis of variance results.</p>
Full article ">Figure 2
<p>Alpha diversity of abundant (<b>A</b>,<b>B</b>) and rare taxa (<b>C</b>,<b>D</b>) with different S application in black and sandy soil. Different letters indicate significant differences in treatments within each soil type (<span class="html-italic">p</span> &lt; 0.05). NMDS of abundant (<b>E</b>,<b>F</b>) and rare (<b>G</b>,<b>H</b>) taxa with different S application in black and sandy soil. ANOSIM = analysis of similarities. ANOSIM was used to test the difference in soil bacterial communities between different S application treatments within the same soil type. In the results of the analysis of variance, ** <span class="html-italic">p</span> = 0.01.</p>
Full article ">Figure 3
<p>Co-occurrence network of different S application. (<b>A</b>–<b>F</b>): Co-occurrence network colored by phylum under different S application treatments. (<b>G</b>–<b>I</b>): Properties of the co-occurrence network colored for abundant or rare species and composition of each co-occurrence network. The external and internal (black numbers) connections among each subcommunity are shown on the bottom right. In the results of the analysis of variance. The two numbers separated by a comma represent the number of edges (E) and nodes (N), respectively. *, **, and *** indicate significant differences between the treatments at the levels of <span class="html-italic">p</span> = 0.05, <span class="html-italic">p</span> = 0.01, and <span class="html-italic">p</span> = 0.001.</p>
Full article ">Figure 4
<p>VPA (<b>A</b>–<b>D</b>) for the contributions of soil properties, S variables to abundant and rare taxa communities’ variations in black and sandy soils.</p>
Full article ">Figure 5
<p>SEM between yield, soil properties, and rare and rich bacterial communities. S types are represented as different S types. The red line represents positive correlation, the blue line represents negative correlation, and the gray line indicates irrelevance. “*” indicates significance between root morphology or crops <span class="html-italic">p</span> index and maize yield (*, <span class="html-italic">p</span> &lt; 0.05; ***, <span class="html-italic">p</span> &lt; 0.001).</p>
Full article ">Figure 6
<p>Linear discriminant analysis effect size (LEfSe) analysis of functions of abundant and rare taxa in different S types. The potential function of bacteria was based on PICRUSt2. LDA score determined the differentiation size of abundant and rare bacterial sub-communities in S90 and s90 systems with a threshold value of 2.0.</p>
Full article ">
17 pages, 9466 KiB  
Article
Characterization of New Tropicoporus Species (Basidiomycota, Hymenochaetales, Hymenochaetaceae) Discovered in Tamil Nadu, India
by Elangovan Arumugam, Ramesh Murugadoss, Sugantha Gunaseelan, Samantha C. Karunarathna, Abdallah M. Elgorban, Pabulo Henrique Rampelotto and Malarvizhi Kaliyaperumal
Biology 2024, 13(10), 770; https://doi.org/10.3390/biology13100770 - 27 Sep 2024
Viewed by 480
Abstract
This study aimed to investigate the morphological characteristics and phylogenetic relationships of three new species of Tropicoporus from the southern parts of India. The analyses of the ITS and nLSU regions revealed the novelty of these species, which have been named T. pannaensis [...] Read more.
This study aimed to investigate the morphological characteristics and phylogenetic relationships of three new species of Tropicoporus from the southern parts of India. The analyses of the ITS and nLSU regions revealed the novelty of these species, which have been named T. pannaensis, T. subindicus, and T. xerophyticus. All three species possess pileate basidiomes, a monomitic hyphal system in the context, and the presence of cystidioles and setae. However, they differ significantly in their phylogenetic placements and other morpho-taxonomic features. Tropicoporus pannaensis is characterized by a meagrely ungulate basidiome, indistinct zones, and an obtuse margin. Tropicoporus subindicus has a triquetrous basidiome and a radially cracked, crusted pileal surface with an acute margin, while T. xerophyticus is distinguished by an imbricate, perennial basidiome with an abundantly warted pileal surface. A phylogenetic tree is provided to show the placement of the three new species, along with detailed descriptions and illustrations. Additionally, a key for the identification of the Asian species of Tropicoporus is presented. Full article
(This article belongs to the Collection Feature Papers in Microbial Biology)
Show Figures

Figure 1

Figure 1
<p>Molecular phylogeny of the three new <span class="html-italic">Tropicoporus</span> species from India inferred through a Bayesian analysis of the combined ITS and LSU sequence data. The phylogenetic tree presented shows the placement of the novel taxa in relation to other known <span class="html-italic">Tropicoporus</span> species. The numbers indicated at the nodes represent the Bayesian posterior probabilities and bootstrap support values, with only those equal to or above 0.8 and 60%, respectively, being displayed. The type specimens are shown in bold, while the new <span class="html-italic">Tropicoporus</span> species are highlighted in colour and presented in bold text.</p>
Full article ">Figure 2
<p><span class="html-italic">Tropicoporus pannaensis</span> (MUBL1094 holotype). (<b>A</b>) Holotype basidiomes. (<b>B</b>) Pore surface with enlarged pores. (<b>C</b>) Cross-section of a basidiome; yellow arrow indicates duplex context with blackline and white arrow indicates stratified tubes. (<b>D</b>) Hymenial setae. (<b>E</b>) Basidiospores in water. (<b>F</b>) Basidiospores in KOH. (<b>G</b>) Basidiospore in Melzer’s reagent. (<b>H</b>) Basidiospores in cotton blue. Scale bars: 5 µm (<b>D</b>–<b>H</b>).</p>
Full article ">Figure 3
<p>Microscopic structures of <span class="html-italic">Tropicoporus pannaensis</span> (from the Holotype). (<b>A</b>) Hyphae from context. (<b>B</b>) Hyphae from trama. (<b>C</b>) Hymenial setae. (<b>D</b>) Cystidioles. (<b>E</b>) Basidioles. (<b>F</b>) Basidia. (<b>G</b>) Basidiospores. Scale bars: 5 µm.</p>
Full article ">Figure 4
<p><span class="html-italic">Tropicoporus subindicus</span> (MUBL1093 Holotype) (<b>A</b>) Basidiome (Holotype). (<b>B</b>) Pore surface with enlarged pores. (<b>C</b>) Cross-section of a basidiome; yellow arrow represents stratified tubes. (<b>D</b>) Hymenial setae. (<b>E</b>) Basidiospores in water. (<b>F</b>) Basidiospores in KOH. (<b>G</b>) Basidiospores in cotton blue. (<b>H</b>) Basidiospores in Melzer’s reagent. Scale bars: (<b>D</b>–<b>H</b>) = 5 µm.</p>
Full article ">Figure 5
<p>Microscopic structures of <span class="html-italic">Tropicoporus subindicus</span> (from the Holotype). (<b>A</b>) Hyphae from context. (<b>B</b>) Hyphae from trama. (<b>C</b>) Hymenial setae. (<b>D</b>) Cystidioles. (<b>E</b>) Basidioles. (<b>F</b>) Basidia. (<b>G</b>) Basidiospores. Scale bars: 5 µm.</p>
Full article ">Figure 6
<p><span class="html-italic">Tropicoporus xerophyticus</span> (MUBL1091 holotype). (<b>A</b>) Basidiomes (Holotype). (<b>B</b>) Pore surface with enlarged pores. (<b>C</b>) Cross-section of a basidiome; yellow arrow indicates stratified tubes. (<b>D</b>,<b>E</b>) Hymenial setae. (<b>F</b>–<b>I</b>) Basidiospores: (<b>F</b>) Basidiospores in water. (<b>G</b>) Basidiopores in KOH. (<b>H</b>). Basidiopores in cotton blue. (<b>I</b>) Basidiopores in Melzer’s reagent. Scale bars: (<b>D</b>–<b>I</b>) = 5 µm.</p>
Full article ">Figure 7
<p><span class="html-italic">Tropicoporus xerophyticus</span> (MUBL1091 from the Holotype). (<b>A</b>) Contextual hyphae. (<b>B</b>) Tramal hyphae. (<b>C</b>) Hymenial setae. (<b>D</b>) Cystidioles. (<b>E</b>) Basidioles. (<b>F</b>) Basidia. (<b>G</b>) Basidiospores. Scale bars = 5 µm.</p>
Full article ">
27 pages, 6135 KiB  
Article
Hot Spots of Site-Specific Integration into the Sinorhizobium meliloti Chromosome
by Maria E. Vladimirova, Marina L. Roumiantseva, Alla S. Saksaganskaia, Victoria S. Muntyan, Sergey P. Gaponov and Alessio Mengoni
Int. J. Mol. Sci. 2024, 25(19), 10421; https://doi.org/10.3390/ijms251910421 - 27 Sep 2024
Viewed by 422
Abstract
The diversity of phage-related sequences (PRSs) and their site-specific integration into the genomes of nonpathogenic, agriculturally valuable, nitrogen-fixing root nodule bacteria, such as Sinorhizobium meliloti, were evaluated in this study. A total of 314 PRSs, ranging in size from 3.24 kb to [...] Read more.
The diversity of phage-related sequences (PRSs) and their site-specific integration into the genomes of nonpathogenic, agriculturally valuable, nitrogen-fixing root nodule bacteria, such as Sinorhizobium meliloti, were evaluated in this study. A total of 314 PRSs, ranging in size from 3.24 kb to 88.98 kb, were identified in the genomes of 27 S. meliloti strains. The amount of genetic information foreign to S. meliloti accumulated in all identified PRSs was 6.30 Mb. However, more than 53% of this information was contained in prophages (Phs) and genomic islands (GIs) integrated into genes encoding tRNAs (tRNA genes) located on the chromosomes of the rhizobial strains studied. It was found that phiLM21-like Phs were predominantly abundant in the genomes of S. meliloti strains of distant geographical origin, whereas RR1-A- and 16-3-like Phs were much less common. In addition, GIs predominantly contained fragments of phages infecting bacteria of distant taxa, while rhizobiophage-like sequences were unique. A site-specific integration analysis revealed that not all tRNA genes in S. meliloti are integration sites, but among those in which integration occurred, there were “hot spots” of integration into which either Phs or GIs were predominantly inserted. For the first time, it is shown that at these integration “hot spots”, not only is the homology of attP and attB strictly preserved, but integrases in PRSs similar to those of phages infecting the Proteobacteria genera Azospirillum or Pseudomonas are also present. The data presented greatly expand the understanding of the fate of phage-related sequences in host bacterial genomes and also raise new questions about the role of phages in bacterial–phage coevolution. Full article
(This article belongs to the Special Issue Bacteriophage: Molecular Ecology and Pharmacology, 2nd Edition)
Show Figures

Figure 1

Figure 1
<p>General scheme of PRS integration into bacteria genome (<b>a</b>) adopted from [<a href="#B5-ijms-25-10421" class="html-bibr">5</a>]. PRS predecessor—the PRS before integration into a bacterial chromosome. <span class="html-italic">att</span> site—attachment site for PRS integration. Scheme of integration of a phage (<b>b</b>) and a genomic island (<b>c</b>) into the tRNA gene, designed according to the data obtained in this work for <span class="html-italic">S. meliloti</span> strains and adopted from [<a href="#B5-ijms-25-10421" class="html-bibr">5</a>,<a href="#B48-ijms-25-10421" class="html-bibr">48</a>,<a href="#B49-ijms-25-10421" class="html-bibr">49</a>]. <span class="html-italic">attP</span>, <span class="html-italic">attB</span>—sites for phage integration (see text); <span class="html-italic">attL</span>, <span class="html-italic">attR</span>—recombinant sites (see text); DR-1 and -2—short direct repeat sequences flanking the GI (see text).</p>
Full article ">Figure 2
<p>Localization of essential tRNA genes on <span class="html-italic">S. meliloti</span> 1021 chromosome. Designations: 1-52—ordinal numbers of tRNA genes (in the text 001-052); <span class="html-italic">oriC</span> and <span class="html-italic">terC</span>—origin and terminus of replication, respectively; I–IV—quarters of the chromosome, defined relative to <span class="html-italic">oriC</span> and <span class="html-italic">terC</span>; in bold—ordinal number of tRNA genes with integrated PRSs; *—“hot spots” of PRS integration (genes 030 and 039) and spot of preferential integration (gene 010) (see text).</p>
Full article ">Figure 3
<p>Frequencies of PRS integration in essential tRNA genes of <span class="html-italic">S. meliloti</span>. Designations: along the abscissa axis—the signature combines the ordinal number of tRNA genes, with a corresponding amino acid and anticodon following the dash (see <a href="#ijms-25-10421-t003" class="html-table">Table 3</a>); along the abscissa axis—frequency of occurrence of PRSs integrated into different tRNA genes; *—tRNA genes at “hot spots” of integration (genes 030 and 039) and spot of preferential integration (gene 010); I–IV—chromosome quarters defined relative to <span class="html-italic">oriC</span> and <span class="html-italic">terC</span> (see <a href="#ijms-25-10421-f002" class="html-fig">Figure 2</a>).</p>
Full article ">Figure 4
<p>Phylogenetic analysis of genes encoding essential tRNA<sup>Met</sup>(CAT) and tRNA<sup>fMet</sup>(CAT) of <span class="html-italic">S. meliloti</span> strains. Designations: the signatures of tRNA genes are in the form X_Y/N, where X—the name of the strain, Y—amino acid Met or fMet, N—ordinal number of the tRNA gene in which the PRS with an additional tRNA gene is integrated; see also the designations given in <a href="#ijms-25-10421-t003" class="html-table">Table 3</a>; *—additional tRNA<sup>fMet</sup>(CAT) located in PRS that is not integrated into the tRNA gene.</p>
Full article ">Figure 5
<p>The synteny blocks of PRSs integrated into essential tRNA genes of <span class="html-italic">S. meliloti.</span> Designations: there are sequences of PRSs around the circle; PRS signature: the strain that contained a PRS, with the ordinal number of the tRNA gene following the dash (see <a href="#app1-ijms-25-10421" class="html-app">Table S2</a>, <a href="#ijms-25-10421-f002" class="html-fig">Figure 2</a>). Type of PRS indicated by color: gray—GI, green—int-Ph, blue—q-Ph, red—inc-Ph (see text). The red lines inside the circle connect matching sequences between pairs of PRSs (see <a href="#sec4-ijms-25-10421" class="html-sec">Section 4</a> Materials and Methods; <a href="#app1-ijms-25-10421" class="html-app">Table S3B</a>). Detailed blastn PRS alignment results are in <a href="#app1-ijms-25-10421" class="html-app">Table S3A</a>. *, **, ***—PRS similar to phages phiLM21, 16-3, and RR1-A, respectively (see text). USDA1157-054, BL225C-054, and USDA1021-013—PRSs localized on the megaplasmid pSymB.</p>
Full article ">Figure 6
<p>Phylogenetic tree constructed using the amino acid sequences of integrases of PRS integrated into tRNA genes of <span class="html-italic">S. meliloti</span>. Designations: <sup>1</sup>—PI means a point of integration, it combines the strain and an ordinal number of tRNA genes through a dash (see <a href="#ijms-25-10421-t003" class="html-table">Table 3</a>); <sup>2</sup>—GI—genomic island; int-Ph—intact phage; inc-Ph—incomplete phage sequence; q-Ph—questionable phage sequence; <sup>3</sup>—ordinal numbers of tRNA genes (see <a href="#ijms-25-10421-t003" class="html-table">Table 3</a>); <sup>4</sup>—PRS localized on pSymB in case of the corresponding strains (see text and <a href="#app1-ijms-25-10421" class="html-app">Table S2</a>).</p>
Full article ">Figure 7
<p>The alignment of the 3′ ends of essential tRNA genes and PRS attachment sequences (<span class="html-italic">attR</span> sites). Designations: (<b>a</b>)—principal scheme of PRSs integrated into tRNA genes; (<b>b</b>)—the alignment of the 3′ ends of tRNA genes of bacteria and attachment sites of PRS detected in <span class="html-italic">S. meliloti</span> strains; and (<b>c</b>)—phages whose integrases were similar to those of the PRSs. The similarity in phage integrases and PRS integrases was determined by analyzing their amino acid sequences (see text).</p>
Full article ">
16 pages, 5970 KiB  
Article
Molecular Markers for Analyses of Genetic Diversity within the Anastrepha fraterculus Complex with Emphasis on Argentine Populations
by Ludvik M. Gomulski, María Teresa Vera, Silvia B. Lanzavecchia, Riccardo Piccinno, Giulia Fiorenza, Daniel De Luca, Beatriz N. Carrizo, Juan Pedro R. Bouvet, Valeria A. Viana, Carlos Cárceres, Walther Enkerlin, Anna R. Malacrida and Giuliano Gasperi
Insects 2024, 15(10), 748; https://doi.org/10.3390/insects15100748 - 27 Sep 2024
Viewed by 461
Abstract
The South American fruit fly Anastrepha fraterculus (Wiedmann) has a vast range extending from northern Mexico, through Central America, to South America where it is an extremely polyphagous pest of wild and cultivated fruits. It is a complex of cryptic species currently composed [...] Read more.
The South American fruit fly Anastrepha fraterculus (Wiedmann) has a vast range extending from northern Mexico, through Central America, to South America where it is an extremely polyphagous pest of wild and cultivated fruits. It is a complex of cryptic species currently composed of eight recognised morphotypes: “Mexican”, “Venezuelan”, “Andean”, “Peruvian”, “Ecuadorian”, and the three Brazilian morphotypes “Brazilian-1”, “Brazilian-2”, and “Brazilian-3”. Molecular markers that can identify the member species of the complex are crucial for the implementation of effective pest control measures, such as the sterile insect technique. The object of this study was to evaluate the use of the internal transcribed spacer 2 (ITS2) sequence for discriminating several members of the A. fraterculus cryptic species complex (Mexican, Peruvian, and Brazilian-1) and a related species, Anastrepha schultzi Blanchard. The analysis highlighted significant genetic differentiation between the evaluated morphotypes, allowed their discrimination within the A. fraterculus cryptic species complex, and provided new insights into their genetic relationships. The ITS2 marker provides a basis for the development of taxonomic keys for the discrimination of the cryptic taxa within the A. fraterculus cryptic species complex. ITS2 also represents an important marker for the poorly studied species A. schultzi. Full article
(This article belongs to the Section Insect Molecular Biology and Genomics)
Show Figures

Figure 1

Figure 1
<p>Geographic locations of the sites of collections that gave rise to the laboratory strains and of the wild samples considered in the study.</p>
Full article ">Figure 2
<p>Scheme of the characterisation of the ribosomal DNA region in <span class="html-italic">Anastrepha fraterculus</span> using sequences of <span class="html-italic">A. fraterculus</span> (red) and <span class="html-italic">A. suspensa</span> (green) available in GenBank. The 18S, ITS1, 5.8S, ITS2A, 2S, ITS2, and 28S regions and the positions of the primers and the amplification products (blue) are shown.</p>
Full article ">Figure 3
<p>Principal component analysis on the ITS2 sequences from the different samples. PC1 vs. PC2 on the left and PC2 vs. PC3 on the right.</p>
Full article ">Figure 4
<p>Maximum Likelihood tree of the ITS2 sequences from the different samples. Bootstrap numbers (percentage) of 100,000 replications are shown. The presence of haplotype sharing is indicated by strain/population-specific symbols.</p>
Full article ">Figure 5
<p>Bayesian analysis tree of the ITS2 sequences from the different samples. Values at the nodes are the posterior probabilities of each partition. The presence of haplotype sharing is indicated by strain/population-specific symbols.</p>
Full article ">
21 pages, 10375 KiB  
Article
Sturnidae sensu lato Mitogenomics: Novel Insights into Codon Aversion, Selection, and Phylogeny
by Shiyun Han, Hengwu Ding, Hui Peng, Chenwei Dai, Sijia Zhang, Jianke Yang, Jinming Gao and Xianzhao Kan
Animals 2024, 14(19), 2777; https://doi.org/10.3390/ani14192777 - 26 Sep 2024
Viewed by 598
Abstract
The Sturnidae family comprises 123 recognized species in 35 genera. The taxa Mimidae and Buphagidae were formerly treated as subfamilies within Sturnidae. The phylogenetic relationships among the Sturnidae and related taxa (Sturnidae sensu lato) remain unresolved due to high rates of morphological [...] Read more.
The Sturnidae family comprises 123 recognized species in 35 genera. The taxa Mimidae and Buphagidae were formerly treated as subfamilies within Sturnidae. The phylogenetic relationships among the Sturnidae and related taxa (Sturnidae sensu lato) remain unresolved due to high rates of morphological change and concomitant morphological homoplasy. This study presents five new mitogenomes of Sturnidae sensu lato and comprehensive mitogenomic analyses. The investigated mitogenomes exhibit an identical gene composition of 37 genes—including 13 protein-coding genes (PCGs), 2 rRNA genes, and 22 tRNA genes—and one control region (CR). The most important finding of this study is drawn from CAM analyses. The surprisingly unique motifs for each species provide a new direction for the molecular species identification of avian. Furthermore, the pervasiveness of the natural selection of PCGs is found in all examined species when analyzing their nucleotide composition and codon usage. We also determine the structures of mt-tRNA, mt-rRNA, and CR structures of Sturnidae sensu lato. Lastly, our phylogenetic analyses not only well support the monophyly of Sturnidae, Mimidae, and Buphagidae, but also define nine stable subclades. Taken together, our findings will enable the further elucidation of the evolutionary relationships within Sturnidae sensu lato. Full article
(This article belongs to the Section Animal Genetics and Genomics)
Show Figures

Figure 1

Figure 1
<p>(<b>a</b>) Circular map of mitogenomes of Sturnidae <span class="html-italic">sensu lato</span>. The different colors represent the BLAST-identical percentages. The mitogenomes from outside to inside are as follows (labeled 1 to 10, respectively): <span class="html-italic">A. cristatellus</span>, <span class="html-italic">A. tristis</span>, <span class="html-italic">S. cineraceus</span>, <span class="html-italic">S. sericeus</span>, <span class="html-italic">L. rothschildi</span>, <span class="html-italic">G. nigricollis</span>, <span class="html-italic">S. vulgaris</span>, <span class="html-italic">T. redivivum</span>, <span class="html-italic">M. polyglottos</span>, and <span class="html-italic">B. erythrorynchus</span>. (<b>b</b>) Phylogenetic tree of the relationships among the 10 Sturnidae <span class="html-italic">sensu lato</span> species based on 13 mitochondrial PCGs, with two Muscicapidae outgroups. The support values of each node are indicated in the order of ML, MP, and BI inferences, and “*” indicates full support.</p>
Full article ">Figure 2
<p>(<b>a</b>) RSCU heatmap of overall mitochondrial PCGs of Sturnidae <span class="html-italic">sensu lato</span>. (<b>b</b>) PR2 plots of each mitochondrial PCG of Sturnidae <span class="html-italic">sensu lato</span>. (<b>c</b>) PCA analysis based on the ENC values of each mitochondrial PCG. (<b>d</b>) The comparison of the ENC vs. GC3s curve of the PCGs in the mitogenomes of Sturnidae <span class="html-italic">sensu lato</span>. The continuous red line represents the expected ENC curve.</p>
Full article ">Figure 3
<p>(<b>a</b>) Codon aversion motifs of ten investigated Sturnidae <span class="html-italic">sensu lato</span> mitochondrial PCGs. (<b>b</b>) Codon aversion numbers of the three families.</p>
Full article ">Figure 4
<p>The presumed secondary structures of the tRNAs in <span class="html-italic">A. cristatellus</span>. The new form of the TΨC stem in MT-TF is displayed using an orange box.</p>
Full article ">Figure 5
<p>The predicted secondary structures of MT-RNR1 in <span class="html-italic">A. cristatellus</span>.</p>
Full article ">Figure 6
<p>The predicted secondary structures of MT-RNR2 in <span class="html-italic">A. cristatellus</span>.</p>
Full article ">Figure 7
<p>The structure of the control region in the mitogenome of <span class="html-italic">A. cristatellus</span>. Note that ETAS denotes extended termination-associated sequences, CSB denotes conserved sequence block, and HSP denotes heavy-strand transcription promoter.</p>
Full article ">Figure 8
<p>Nucleotide-based phylogenetic tree of 132 Sturnidae <span class="html-italic">sensu lato</span> taxa with two Muscicapidae outgroups. This analysis utilized a combined multilocus dataset (five mitochondrial genes and eight nuclear genes). The support value for each node is denoted using different line types.</p>
Full article ">
33 pages, 49504 KiB  
Article
The Late Early–Middle Pleistocene Mammal Fauna from the Megalopolis Basin (Peloponnese, Greece) and Its Importance for Biostratigraphy and Paleoenvironment
by George E. Konidaris, Athanassios Athanassiou, Vangelis Tourloukis, Krystalia Chitoglou, Thijs van Kolfschoten, Domenico Giusti, Nicholas Thompson, Georgia Tsartsidou, Effrosyni Roditi, Eleni Panagopoulou, Panagiotis Karkanas and Katerina Harvati
Quaternary 2024, 7(4), 41; https://doi.org/10.3390/quat7040041 - 24 Sep 2024
Viewed by 1659
Abstract
Recent investigations in the upper Lower–Middle Pleistocene deposits of the Megalopolis Basin (Greece) led to the discovery of several sites/findspots with abundant faunal material. Here, we provide an updated overview including new results on the micro- and macro-mammal fauna. Important new discoveries comprise [...] Read more.
Recent investigations in the upper Lower–Middle Pleistocene deposits of the Megalopolis Basin (Greece) led to the discovery of several sites/findspots with abundant faunal material. Here, we provide an updated overview including new results on the micro- and macro-mammal fauna. Important new discoveries comprise partial hippopotamus skeletons from Marathousa 1 and the new Lower Pleistocene site Choremi 6, as well as a second partial elephant skeleton from Marathousa 1, including a complete tusk and the rarely found stylohyoideum. Based on the first results from the newly collected micromammals, we discuss age constraints of the sites, and we provide biostratigraphic/biochronologic remarks on key mammal taxa for the Middle Pleistocene of Greece and southeastern Europe. The presence of mammals highly dependent on freshwater for their survival, together with temperate-adapted ones in several stratigraphic layers of the basin, including those correlated with glacial stages, when conditions were colder and/or drier, indicate the capacity of the basin to retain perennial freshwater bodies under milder climatic conditions, even during the harsher glacial periods of the European Middle Pleistocene, and further support its refugial status. Yet, the smaller dimensions of the Megalopolis hippopotamuses may represent a response to the changing environmental conditions of the epoch, not optimal for hippopotamuses. Overall, the Megalopolis Basin comprises a unique fossil record for southeastern Europe and provides valuable insights into the Middle Pleistocene terrestrial ecosystems of Europe, and hominin adaptations in particular. Full article
Show Figures

Figure 1

Figure 1
<p>(<b>a</b>) Geographic position of the Megalopolis Basin (maps taken from Copernicus Land Monitoring Service: <a href="https://land.copernicus.eu/" target="_blank">https://land.copernicus.eu/</a> and <a href="https://maps-for-free.com/" target="_blank">https://maps-for-free.com/</a> (accessed on 30 April 2024)). (<b>b</b>) Geographic position of the investigated sites within the lignite mines (numbered 1–4) of the Megalopolis Basin (satellite image from Google Earth). (<b>c</b>) Panoramic view of the eastern quarry of the Marathousa mine, indicating the Marathousa and Megalopolis Members of the Choremi Formation, and the lignite seams. (<b>d</b>) Simplified stratigraphic column of the Marathousa Member showing the lignite seams and the intercalating detrital intervals, their correlation to the Marine Isotope Stages (MIS) following the age model proposed by Tourloukis et al. [<a href="#B3-quaternary-07-00041" class="html-bibr">3</a>], and the stratigraphic position of the sites and survey units following Karkanas et al. [<a href="#B25-quaternary-07-00041" class="html-bibr">25</a>,<a href="#B28-quaternary-07-00041" class="html-bibr">28</a>].</p>
Full article ">Figure 2
<p>(<b>a</b>) Panoramic view of the site Choremi 6. (<b>b</b>–<b>e</b>) Fossils belonging to the <span class="html-italic">Hippopotamus antiquus</span> skeleton in situ at Choremi 6. (<b>f</b>) Distribution map showing the position of the hippopotamus remains (recorded with the use of a Differential GPS—WGS84 datum).</p>
Full article ">Figure 3
<p><span class="html-italic">Hippopotamus antiquus</span> remains from Choremi 6. (<b>a</b>) Lower molar (m1?) in occlusal view; (<b>b</b>) series of caudal vertebrae (CHO-6-36, 37, 38, 39, 40, and 41); (<b>c</b>,<b>d</b>) left scapula (CHO-6-49) in lateral (<b>c</b>) and distal (<b>d</b>) view; (<b>e</b>) right patella (CHO-6-10) in caudal view; (<b>f</b>) left tibia (CHO-6-42) in dorsal view (note the presence of a limonite concretion on the distal end); (<b>g</b>) left fibula (CHO-6-43) in medial view; (<b>h</b>) left astragalus (CHO-6-7) in dorsal view; (<b>i</b>) left navicular (central tarsal, CHO-6-15) in lateral view; (<b>j</b>) left calcaneus (CHO-6-8) in medial view; (<b>k</b>) left ectocuneiform (tarsal I, CHO-6-17) in proximal view; (<b>l</b>) left mesocuneiform (tarsal II, CHO-6-25) in proximal view; (<b>m</b>) left entocuneiform (tarsal III, CHO-6-11) in distal view; (<b>n</b>) left cuboid (tarsal IV, CHO-6-13) in medial view; (<b>o</b>) left Mt II (CHO-6-20) in dorsal view; (<b>p</b>) left MT V (CHO-6-24) in dorsal view; (<b>q</b>) right MT III (CHO-6-21) in dorsal view.</p>
Full article ">Figure 4
<p>Metrical comparison of <span class="html-italic">Hippopotamus</span> postcranial specimens from Megalopolis Basin with <span class="html-italic">Hippopotamus antiquus</span>, <span class="html-italic">Hippopotamus tiberinus</span> (<span class="html-italic">H</span>. ex gr. <span class="html-italic">antiquus</span>), and <span class="html-italic">Hippopotamus amphibius</span>. Data from [<a href="#B12-quaternary-07-00041" class="html-bibr">12</a>,<a href="#B20-quaternary-07-00041" class="html-bibr">20</a>,<a href="#B33-quaternary-07-00041" class="html-bibr">33</a>,<a href="#B34-quaternary-07-00041" class="html-bibr">34</a>,<a href="#B35-quaternary-07-00041" class="html-bibr">35</a>,<a href="#B36-quaternary-07-00041" class="html-bibr">36</a>,<a href="#B37-quaternary-07-00041" class="html-bibr">37</a>]. The blue and green lines correspond to the ranges of <span class="html-italic">H. amphibius</span> and <span class="html-italic">H. antiquus</span> given by [<a href="#B38-quaternary-07-00041" class="html-bibr">38</a>]. Abbreviations: a, articular facet; d, distal; DAP, anteroposterior diameter; DT, transverse diameter; H, height; L, length; med, medial; p, proximal.</p>
Full article ">Figure 5
<p>Small mammals from several sites and findspots of the Megalopolis Basin. (<b>a</b>,<b>b</b>) Right M2 (KYP-4-MM008) of <span class="html-italic">Mimomys</span> sp. (smaller size) from Kyparissia 4 in occlusal (<b>a</b>) and lingual (<b>b</b>) view; (<b>c</b>,<b>d</b>) left m1 (KYP-4-MM017) of <span class="html-italic">Mimomys</span> cf. <span class="html-italic">savini</span> from Kyparissia 4 in occlusal (<b>c</b>) and lingual (<b>d</b>) view; (<b>e</b>,<b>f</b>) left M3 (KYP-4-MM015) of <span class="html-italic">Pliomys</span> sp. from Kyparissia 4 in occlusal (<b>e</b>) and lingual (<b>f</b>) view; (<b>g</b>,<b>h</b>) left m2 (KYP-3-MM001) of <span class="html-italic">Mimomys</span> sp. (smaller size) from Kyparissia 3 in occlusal (<b>g</b>) and lingual (<b>h</b>) view; (<b>i</b>) right m1 (KYP SU 6-MM002) of <span class="html-italic">Pliomys</span> cf. <span class="html-italic">episcopalis</span> from Kyparrisia SU 6 in occlusal view; (<b>j</b>,<b>k</b>) left m1 (CHO-7-MM135) of <span class="html-italic">Arvicola mosbachensis</span> from Choremi 7 in occlusal (<b>j</b>) and lingual (<b>k</b>) view; (<b>l</b>) left m1 (CHO-7-MM66) of <span class="html-italic">Microtus</span> sp. (cf. <span class="html-italic">Microtus arvalis</span>) from Choremi 7 in occlusal view; (<b>m</b>) right m1 (CHO-7-MM214) of <span class="html-italic">Microtus</span> (<span class="html-italic">Terricola</span>) <span class="html-italic">subterraneus</span> from Choremi 7 in occlusal view.</p>
Full article ">Figure 6
<p>(<b>a</b>) Panoramic view of the Choremi 4 section showing the fossiliferous layer. (<b>b</b>) A tooth of <span class="html-italic">Castor fiber</span> in situ within a shell-rich layer. (<b>c</b>) In situ concentration of hippopotamus bones. (<b>d</b>,<b>e</b>) Right astragalus of <span class="html-italic">Hippopotamus antiquus</span> in dorsal (<b>d</b>) and plantar (<b>e</b>) view. (<b>f</b>) Right entocuneiform (tarsal III) in proximal view. (<b>g</b>,<b>h</b>) Atlas of <span class="html-italic">H. antiquus</span> in cranial (<b>g</b>) and caudal (<b>h</b>) view. (<b>i</b>–<b>l</b>) Right astragalus of <span class="html-italic">Bison</span> sp. found at the surface near the Choremi 4 section in dorsal (<b>i</b>), plantar (<b>j</b>), medial (<b>k</b>), and lateral (<b>l</b>) view.</p>
Full article ">Figure 7
<p>(<b>a</b>) Panoramic view of the northern profile in the Marathousa mine, indicating the Marathousa and Megalopolis Members, the lignite seams, and the location of the MAR SU 7. (<b>b</b>) Details of the profile showing the MAR SU 7 section. (<b>c</b>,<b>d</b>) Left calcaneus of <span class="html-italic">Castor fiber</span> in medial (<b>c</b>) and plantar (<b>d</b>) view. (<b>e</b>–<b>h</b>) Left humerus of <span class="html-italic">Lutra simplicidens</span> in cranial (<b>e</b>), caudal (<b>f</b>), medial (<b>g</b>), and lateral (<b>h</b>) view. (<b>i</b>,<b>j</b>) Left radius of a <span class="html-italic">Capreolus</span>-sized cervid in dorsal (<b>i</b>) and volar (<b>j</b>) view. (<b>k</b>,<b>l</b>) Metacarpal of a <span class="html-italic">Capreolus</span>-sized cervid in dorsal (<b>k</b>) and volar (<b>l</b>) view.</p>
Full article ">Figure 8
<p><span class="html-italic">Hippopotamus antiquus</span> skeleton from Marathousa 1 (Area B; UB2b). (<b>a</b>) Distribution map showing the position (recorded with the use of a total station) of the hippopotamus remains; (<b>b</b>) hippopotamus fossils in situ; (<b>c</b>) sacrum (MAR-1B-926/587-17) in ventral view; (<b>d</b>) right radius (MAR-1B-1) in dorsal view; (<b>e</b>) right patella (MAR-1B-925/587-2) in caudal view; (<b>f</b>) right femur (MAR-1B-925/587-1) in cranial view; (<b>g</b>) right tibia (MAR-1B-926/587-10) in dorsal view; (<b>h</b>) right astragalus (MAR-1B-926/587-4) in dorsal view; (<b>i</b>) right calcaneus (MAR-1B-2) in medial view; (<b>j</b>) right navicular (central tarsal, MAR-1B-924/587-3) in lateral view; (<b>k</b>) right MT II (MAR-1B-4) in dorsal view; (<b>l</b>) right MT V (MAR-1B-926/586-1) in dorsal view; (<b>m</b>) lumbar vertebra (MAR-1B-926/587-9) in cranial view.</p>
Full article ">Figure 9
<p>(<b>a</b>,<b>b</b>) Left m2 (MAR-1B-8) of <span class="html-italic">Hippopotamus antiquus</span> from Marathousa 1 (Area B), most possibly belonging to the hippopotamus skeleton, in occlusal (<b>a</b>) and buccal (<b>b</b>) view. (<b>c</b>) Biplot comparing <span class="html-italic">Hippopotamus</span> m2 from various localities. (<b>d</b>,<b>e</b>) Left M3 (MAR-1B-7) of <span class="html-italic">H. antiquus</span> from Marathousa 1 (Area B), most possibly belonging to the hippopotamus skeleton, in occlusal (<b>d</b>) and buccal (<b>e</b>) view. (<b>f</b>) Biplot comparing <span class="html-italic">Hippopotamus</span> M3 from various localities. (<b>g</b>) Left M2 of <span class="html-italic">H. antiquus</span> from MAR SU 1 in occlusal view. (<b>h</b>) Left M2 of <span class="html-italic">H. antiquus</span> from MAR SU 2 in occlusal view. (<b>i</b>) Biplot comparing <span class="html-italic">Hippopotamus</span> M2 from various localities. Biplots with 95% confidence ellipses for <span class="html-italic">Hippopotamus amphibius</span> and <span class="html-italic">H. antiquus</span>, performed with the software package PAST v. 4.16 [<a href="#B63-quaternary-07-00041" class="html-bibr">63</a>]. Data are from [<a href="#B33-quaternary-07-00041" class="html-bibr">33</a>,<a href="#B64-quaternary-07-00041" class="html-bibr">64</a>,<a href="#B65-quaternary-07-00041" class="html-bibr">65</a>,<a href="#B66-quaternary-07-00041" class="html-bibr">66</a>,<a href="#B67-quaternary-07-00041" class="html-bibr">67</a>,<a href="#B68-quaternary-07-00041" class="html-bibr">68</a>].</p>
Full article ">Figure 10
<p><span class="html-italic">Palaeoloxodon antiquus</span> skeleton from Marathousa 1 (Area B; UB4c). (<b>a</b>) Tusk and cervical vertebrae (including the axis) in situ. (<b>b</b>) Distribution map of the southwestern part of the trench at Area B showing the position (recorded with the use of a total station) of the elephant remains, stratigraphically and spatially associated with lithic artefacts and other faunal remains. (<b>c</b>,<b>d</b>) Axis (MAR-1B-928/588-39) in cranial (<b>c</b>) and right lateral (<b>d</b>) view; (<b>e</b>) cervical vertebra (MAR-1B-928/586-16) in cranial view; (<b>f</b>–<b>h</b>) left stylohyoid (MAR-1B-928/587-32) in medial (<b>f</b>), lateral (<b>g</b>), and anterior (<b>h</b>) view.</p>
Full article ">Figure 11
<p>Mammal remains from Marathousa 1. (<b>a</b>,<b>b</b>) Left hemimandible with the incisor and p4–m3 of <span class="html-italic">Castor fiber</span> (surface find at Area B) in medial (<b>a</b>) and dorsal (<b>b</b>) view. (<b>c</b>) Left hemimandible (MAR-1A-941/677-41) of <span class="html-italic">Castor fiber</span> in situ at Area A. (<b>d</b>,<b>e</b>) Second phalanx of the front limb (MAR-1B-930/593-46) of <span class="html-italic">Bison</span> sp. in dorsal (<b>d</b>) and lateral (<b>e</b>) view. (<b>f</b>,<b>g</b>) Third phalanx of the front limb (MAR-1B-931/594-42) of <span class="html-italic">Bison</span> sp. in lateral (<b>f</b>) and medial (<b>g</b>) view. (<b>h</b>) Refitting mandibular fragments (MAR-1B-934/594-71, 932/598-51, 933/595-60, 931/593-38, and 931/596-39) of the same individual of <span class="html-italic">Dama</span> sp., above the right hemimandible in medial view, and below the left hemimandible in lateral view. (<b>i</b>,<b>j</b>) Left scapula fragment (MAR-1B-934/597-35) of <span class="html-italic">Dama</span> sp. in lateral (<b>i</b>) and distal (<b>j</b>) view. (<b>k</b>) Left calcaneus (MAR-1B-928/588-21) of <span class="html-italic">Dama</span> sp. in plantar view. (<b>l</b>,<b>m</b>) Left scapula fragment (MAR-1A-935/672-20) of <span class="html-italic">Cervus elaphus</span> in lateral (<b>l</b>) and distal (<b>m</b>) view. (<b>n</b>) Distal fragment of left humerus (MAR-1B-931/564-58) of <span class="html-italic">C. elaphus</span> in cranial view. (<b>o</b>) Proximal fragment of left radius (MAR-1B-926-587-22) in dorsal view. (<b>p</b>) Distal fragment of left tibia (MAR-1B-925/586-7) in dorsal view. (<b>q</b>) Right astragalus (MAR-1B-926/586-4) of <span class="html-italic">C. elaphus</span> in dorsal view. (<b>r</b>) Left calcaneus (MAR-1-939/635-1) of <span class="html-italic">C. elaphus</span> in medial view. (<b>s</b>) Left MT IV (MAR-1A-941/672-44) of <span class="html-italic">Canis</span> sp. in dorsal view. Silhouette images from PhyloPic, phylopic.org.</p>
Full article ">Figure 12
<p>Mammal remains from Tripotamos 4 (TRP-4) and Choremi 7 (CHO-7). (<b>a</b>) Antler fragment (TRP-4-F2) of <span class="html-italic">Cervus elaphus</span> in lateral view; (<b>b</b>) distal metacarpal (TRP-4-F57) of <span class="html-italic">Dama</span> sp. in dorsal view; (<b>c</b>) distal metacarpal (TRP-4-F47) of <span class="html-italic">Dama</span> sp. in dorsal view; (<b>d</b>–<b>f</b>) lower molar (TRP-4-F2, surface) of Bovini in lingual (<b>d</b>), occlusal (<b>e</b>), and buccal (<b>f</b>) view; (<b>g</b>) fragment of upper canine (TRP-4-F70) of <span class="html-italic">Hippopotamus</span> sp.; (<b>h</b>) lamellar fragment (TRP-4-F36, F44, F58, and F73) of Elephantidae indet.; (<b>i</b>–<b>k</b>) right hemimandible fragment with m1–m3 (CHO-7-F125) of <span class="html-italic">Cervus elaphus</span> in lateral (<b>i</b>), lingual (<b>j</b>), and occlusal (<b>k</b>) view; (<b>l</b>–<b>n</b>) right m3 (CHO-7-F147) of <span class="html-italic">Bos</span> sp. in occlusal (<b>l</b>), lingual (<b>m</b>), and buccal (<b>n</b>); (<b>o</b>–<b>q</b>) upper molar (CHO-7-F68) of <span class="html-italic">Bos</span> sp. in occlusal (<b>o</b>), buccal (<b>p</b>), and lingual (<b>q</b>) view.</p>
Full article ">Figure 13
<p>(<b>a</b>,<b>b</b>) Tusk fragment of <span class="html-italic">Palaeoloxodon antiquus</span> from CHO SU 11 in cross-sectional (<b>a</b>) and lateral (<b>b</b>) view. (<b>c</b>–<b>e</b>) Left upper molar of cf. <span class="html-italic">Bos</span> from Choremi 5 in occlusal (<b>c</b>), lingual (<b>d</b>), and buccal (<b>e</b>) view. (<b>f</b>) Distal metacarpal of <span class="html-italic">Bos</span> sp. from CHO SU 11 in dorsal view. (<b>g</b>,<b>h</b>) Right upper premolar (P3/P4) of <span class="html-italic">Equus</span> sp. from CHO SU 10, close to the site Choremi 6, in lingual (<b>g</b>) and occlusal (<b>h</b>) view. (<b>i</b>) Left astragalus of <span class="html-italic">Equus</span> sp. from MAR SU North in dorsal view. (<b>j</b>) Left astragalus of <span class="html-italic">Equus</span> sp. from CHO SU 7 in dorsal view. (<b>k</b>) Left radius and ulna of <span class="html-italic">Hippopotamus antiquus</span> from the Marathousa mine in lateral view.</p>
Full article ">Figure 14
<p>Simplified stratigraphic column of the Marathousa Member showing the lignite seams and the intercalating detrital intervals, their correlation to the Marine Isotope Stages (MIS) following the age model proposed by Tourloukis et al. [<a href="#B3-quaternary-07-00041" class="html-bibr">3</a>], the stratigraphic position of the sites and survey units following Karkanas et al. [<a href="#B25-quaternary-07-00041" class="html-bibr">25</a>,<a href="#B28-quaternary-07-00041" class="html-bibr">28</a>], and the biostratigraphic range of selected micro- and macro-mammals from the Megalopolis Basin, indicating their First and Last Occurrences in the Greek fossil record. Silhouette images from PhyloPic, phylopic.org.</p>
Full article ">
Back to TopTop