[go: up one dir, main page]
More Web Proxy on the site http://driver.im/
You seem to have javascript disabled. Please note that many of the page functionalities won't work as expected without javascript enabled.
 
 
Sign in to use this feature.

Years

Between: -

Subjects

remove_circle_outline
remove_circle_outline
remove_circle_outline
remove_circle_outline
remove_circle_outline
remove_circle_outline
remove_circle_outline
remove_circle_outline
remove_circle_outline

Journals

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

Search Results (390)

Search Parameters:
Keywords = basidiomycetes

Order results
Result details
Results per page
Select all
Export citation of selected articles as:
20 pages, 2981 KiB  
Article
Purification and Biochemical Characterization of Trametes hirsuta CS5 Laccases and Its Potential in Decolorizing Textile Dyes as Emerging Contaminants
by Guadalupe Gutiérrez-Soto, Carlos Eduardo Hernández-Luna, Iosvany López-Sandin, Roberto Parra-Saldívar and Joel Horacio Elizondo-Luevano
Environments 2025, 12(1), 16; https://doi.org/10.3390/environments12010016 - 7 Jan 2025
Viewed by 479
Abstract
This study explores the purification, characterization, and application of laccases from Trametes hirsuta CS5 for degrading synthetic dyes as models of emerging contaminants. Purification involved ion exchange chromatography, molecular exclusion, and chromatofocusing, identifying th ree laccase isoforms: ThIa, ThIb, and ThII. Characterization included [...] Read more.
This study explores the purification, characterization, and application of laccases from Trametes hirsuta CS5 for degrading synthetic dyes as models of emerging contaminants. Purification involved ion exchange chromatography, molecular exclusion, and chromatofocusing, identifying th ree laccase isoforms: ThIa, ThIb, and ThII. Characterization included determining pH and temperature stability, kinetic parameters (Km, Kcat), and inhibition constants (Ki) for inhibitors like NaN3, SDS, TGA, EDTA, and DMSO, using 2,6-DMP and guaiacol as substrates. ThII exhibited the highest catalytic efficiency, with the lowest Km and highest Kcat. Optimal activity was observed at pH 3.5 and 55 °C. Decolorization tests with nine dyes showed that ThII and ThIa were particularly effective against Acid Red 44, Orange II, Indigo Blue, Brilliant Blue R, and Remazol Brilliant Blue R. ThIb displayed higher activity towards Crystal Violet and Acid Green 27. Among substrates, guaiacol showed the highest Kcat, while 2,6-DMP was preferred overall. Inhibitor studies revealed NaN3 as the most potent inhibitor. These results demonstrate the significant potential of T. hirsuta CS5 laccases, especially ThIa and ThII, as biocatalysts for degrading synthetic dyes and other xenobiotics. Their efficiency and stability under acidic and moderate temperature conditions position them as promising tools for sustainable wastewater treatment and environmental remediation. Full article
Show Figures

Figure 1

Figure 1
<p>SDS-PAGE analysis. Lane 1 corresponds to molecular weight markers, 2 corresponds to ThIa laccase, 3 to ThIb, and 4 to ThII isoform.</p>
Full article ">Figure 2
<p>Optimum pH determination.</p>
Full article ">Figure 3
<p>pH stability. (<b>A</b>) corresponds to ThIa isoform, (<b>B</b>) to ThIb laccase, and (<b>C</b>) to ThII isoform.</p>
Full article ">Figure 4
<p>Optimal temperature determination.</p>
Full article ">Figure 5
<p>Temperature stability.</p>
Full article ">Figure 6
<p>Effect of some inhibitors. (<b>A</b>) corresponds to sodium azide (NaN<sub>3</sub>), (<b>B</b>) to Sodium Dodecyl Sulfate (SDS), (<b>C</b>) to thioglycolic acid (TGA), (<b>D</b>) to ethylenediamine tetraacetic acid (EDTA), and (<b>E</b>) to dimethyl sulfoxide (DMSO).</p>
Full article ">Figure 7
<p>Decolorization kinetics graphs of laccase of <span class="html-italic">Trametes hirsuta</span> CS5 on nine synthetic dyes. (<b>A</b>) corresponds to Acid Red 44, (<b>B</b>) Orange II, (<b>C</b>) Reactive Black 5, (<b>D</b>) Blue Indigo, (<b>E</b>) Poly R-478, (<b>F</b>) Remazol Black B, (<b>G</b>) Violet Crystal, (<b>H</b>) Remazol Brilliant Blue R, and (<b>I</b>) Acid Green 27.</p>
Full article ">
34 pages, 2874 KiB  
Article
Annotated Checklist of Poroid Hymenochaetoid Fungi in Central Asia: Taxonomic Diversity, Ecological Roles, and Potential Distribution Patterns
by Yusufjon Gafforov, Manzura Yarasheva, Xue-Wei Wang, Milena Rašeta, Yelena Rakhimova, Lyazzat Kyzmetova, Kanaim Bavlankulova, Sylvie Rapior, Jia-Jia Chen, Ewald Langer, Burkhon Munnavarov, Zafar Aslonov, Bobozoda Bakokhoja and Li-Wei Zhou
J. Fungi 2025, 11(1), 37; https://doi.org/10.3390/jof11010037 - 5 Jan 2025
Viewed by 593
Abstract
Central Asia, located at the heart of Eurasia, is renowned for its varied climate and vertical vegetative distribution, which support diverse biomes and position it as a global biodiversity hotspot. Despite this ecological richness, Central Asia’s fungal diversity, particularly wood-inhabiting macrofungi, remains largely [...] Read more.
Central Asia, located at the heart of Eurasia, is renowned for its varied climate and vertical vegetative distribution, which support diverse biomes and position it as a global biodiversity hotspot. Despite this ecological richness, Central Asia’s fungal diversity, particularly wood-inhabiting macrofungi, remains largely unexplored. This study investigates the diversity, ecological roles, and potential distribution of poroid Hymenochaetoid fungi in the region. By conducting field surveys, collecting basidiomes, and reviewing the literature and herbarium records from five Central Asian countries, we compiled a comprehensive checklist of these fungi. In total, 43 Hymenochaetoid species belonging to 18 genera were identified, with Inonotus, Phellinus, and Phylloporia being the most species-rich. Notably, Inonotus hispidus and Phellinus igniarius were found to be the most widespread species. These macrofungi play essential ecological roles as saprotrophs and pathogens of various identified host plant families, aiding in lignin degradation and exhibiting diverse enzymatic activities. For the first time, we modelled the potential distribution patterns of Hymenochaetoid fungi in Central Asia, revealing that their distribution is strongly influenced by host plant availability and temperature-related factors. The three most critical variables were host plant density, annual temperature range (Bio7), and mean temperature of the warmest quarter (Bio10). The distribution of suitable habitats is uneven, with highly suitable areas (4.52%) concentrated in the mountainous border regions between Kazakhstan, Kyrgyzstan, Tajikistan, and Uzbekistan. These results underscore the significance of specific environmental conditions for the growth and survival of Hymenochaetoid fungi in this region. Our findings highlight the urgent need for continued mycological and host plant research and expanded conservation initiatives to document and preserve macrofungal and botanical biodiversity in this under-explored area. In light of climate change, the collected mycological and botanical data provide a valuable reference for promoting forest health management globally. Full article
(This article belongs to the Special Issue Diversity, Phylogeny and Ecology of Forest Fungi)
Show Figures

Figure 1

Figure 1
<p>Known geographic distribution of Hymenochaetoid fungi indicated by the red points in Central Asia.</p>
Full article ">Figure 2
<p>(A) <span class="html-italic">Coltricia perennis</span>; (B) <span class="html-italic">Fomitiporia hippophaeicola</span>; (C) <span class="html-italic">Inocutis rheades</span>; (D) <span class="html-italic">Coniferiporia uzbekistanensis</span>; (E) <span class="html-italic">Inonotus hispidus</span>; (F) <span class="html-italic">Inonotus obliquus</span>; (G) <span class="html-italic">Phellinopsis conchata</span>; (H) <span class="html-italic">Phellinus pomaceus</span>; (I) unidentified specimen; (J) <span class="html-italic">Phellinus igniarius</span>; (K) <span class="html-italic">Phylloporia yuchengii</span>; (L) <span class="html-italic">Sanghuangporus lonicerinus</span>. All photos credited to Yelena Rakhimova, Kanaim Bavlankulova, and Yusufjon Gafforov.</p>
Full article ">Figure 3
<p>Taxonomic composition of poroid Hymenochaetoid fungi in Central Asia.</p>
Full article ">Figure 4
<p>Distribution of poroid Hymenochaetoid genera in five Central Asian countries: Uzbekistan (UZ), Kazakhstan (KZ), Kyrgyzstan (KR), Tajikistan (TJ), and Turkmenistan (TR).</p>
Full article ">Figure 5
<p>Occurrence numbers of poroid Hymenochaetoid fungi on most representative host genera.</p>
Full article ">Figure 6
<p>Current potential distribution patterns of Hymenochaetoid fungi in Central Asia predicted by MaxEnt modeling. Green points represent known occurrences of Hymenochaetoid fungi, and colored regions indicate varying habitat suitability levels.</p>
Full article ">
24 pages, 8588 KiB  
Article
Saprotrophic Wood Decay Ability and Plant Cell Wall Degrading Enzyme System of the White Rot Fungus Crucibulum laeve: Secretome, Metabolome and Genome Investigations
by Alexander V. Shabaev, Olga S. Savinova, Konstantin V. Moiseenko, Olga A. Glazunova and Tatyana V. Fedorova
J. Fungi 2025, 11(1), 21; https://doi.org/10.3390/jof11010021 - 31 Dec 2024
Viewed by 565
Abstract
The basidiomycete Crucibulum laeve strain LE-BIN1700 (Agaricales, Nidulariaceae) is able to grow on agar media supplemented with individual components of lignocellulose such as lignin, cellulose, xylan, xyloglucan, arabinoxylan, starch and pectin, and also to effectively destroy and digest birch, alder and pine sawdust. [...] Read more.
The basidiomycete Crucibulum laeve strain LE-BIN1700 (Agaricales, Nidulariaceae) is able to grow on agar media supplemented with individual components of lignocellulose such as lignin, cellulose, xylan, xyloglucan, arabinoxylan, starch and pectin, and also to effectively destroy and digest birch, alder and pine sawdust. C. laeve produces a unique repertoire of proteins for the saccharification of the plant biomass, including predominantly oxidative enzymes such as laccases (family AA1_1 CAZymes), GMC oxidoreductases (family AA3_2 CAZymes), FAD-oligosaccharide oxidase (family AA7 CAZymes) and lytic polysaccharide monooxygenases (family LPMO X325), as well as accompanying acetyl esterases and loosenine-like expansins. Metabolomic analysis revealed that, specifically, monosaccharides and carboxylic acids were the key low molecular metabolites in the C. laeve culture liquids in the experimental conditions. The proportion of monosaccharides and polyols in the total pool of identified compounds increased on the sawdust-containing media. Multiple copies of the family AA1_1, AA3_2, AA7 and LPMOs CAZyme genes, as well as eight genes encoding proteins of the YvrE superfamily (COG3386), which includes sugar lactone lactonases, were predicted in the C. laeve genome. According to metabolic pathway analysis, the litter saprotroph C. laeve can catabolize D-gluconic and D-galacturonic acids, and possibly other aldonic acids, which seems to confer certain ecological advantages. Full article
(This article belongs to the Special Issue Fungal Metabolomics and Genomics)
Show Figures

Figure 1

Figure 1
<p>Growth rate (<b>A</b>), oxidative (<b>B</b>) and cellulolytic (<b>C</b>) activity during cultivation of <span class="html-italic">Crucibulum laeve</span> LE-BIN 1700 on MEA media containing various types of wood sawdust—birch (MEA-B), alder (MEA-A) and pine (MEA-P). The MEA medium without a sawdust addition was used as a control. Different letters indicate statistical differences between culture media at the same time point (<span class="html-italic">p</span> &lt; 0.05).</p>
Full article ">Figure 2
<p><span class="html-italic">Crucibulum laeve</span> LE-BIN 1700 growth rate on control medium (only bacteriological Agar) and medium supplemented with different lignocellulose compounds—lignin; birch xylan (Xylan B); larch xylan (Xylan L); carboxymetyl cellulose (CMC); starch and pectin. The same letters indicate that the values are not significantly different (<span class="html-italic">p</span> &gt; 0.05).</p>
Full article ">Figure 3
<p>(<b>A</b>) White rot fungus (<span class="html-italic">Crucibulum laeve</span> LE-BIN 1700) oxidase activity variation during stationary semi-solid cultivation on the control glucose–peptone (GP) medium and on the GP supplemented with birch (GP-B), alder (GP-A) and pine (GP-P) sawdust; (<b>B</b>) fungal biomass amount at the end of 30-day cultivation. Different letters indicate statistical differences between culture media at the same time point (<span class="html-italic">p</span> &lt; 0.05).</p>
Full article ">Figure 4
<p>Functional description and secretion conditions of proteins found in <span class="html-italic">C. laeve</span> LE-BIN 1700 exoproteomes during semi-solid cultivation on the control glucose–peptone (GP) medium and on the GP supplemented with birch (GP-B), alder (GP-A) and pine (GP-P) sawdust; aa—amino acid. The presence of proteins in the corresponding medium is indicated by a colored circle.</p>
Full article ">Figure 5
<p>Distribution profile of identified secreted proteins and the distribution of CAZymes (GHs, AAs and CEs classes) in the GP (glucose–peptone medium), GP-B (glucose–peptone with birch sawdust), GP-A (glucose–peptone with alder sawdust) and GP-P (glucose–peptone with pine sawdust) of the <span class="html-italic">Crucibulum laeve</span> LE-BIN 1700 secretomes. AA—auxiliary activity enzyme; CE—carbohydrate esterase; GH—glycoside hydrolase.</p>
Full article ">Figure 6
<p>Total pool of identified metabolites in the GP (glucose–peptone medium), GP-B (glucose–peptone with birch sawdust), GP-A (glucose–peptone with alder sawdust) and GP-P (glucose–peptone with pine sawdust) of the <span class="html-italic">Crucibulum laeve</span> LE-BIN 1700.</p>
Full article ">Figure 7
<p>Mass loss (% original mass) (<b>A</b>) and specific gravity (<b>B</b>) of wood substrates (birch, alder and pine sawdust) caused by <span class="html-italic">Trametes hirsuta</span> LE-BIN 072 (<b>top panel</b>) and <span class="html-italic">Crucibulum laeve</span> LE-BIN 1700 (<b>bottom panel</b>) after 2 and 4 months of fungal cultivation, respectively. Shaded columns—initial specific gravity of sawdust; unshaded columns—specific gravity of sawdust after the fungal cultivation.</p>
Full article ">Figure 8
<p>General scheme of the lignocellulose degradation enzyme machinery (<b>A</b>, <b>top panel</b>) and summary of proposed sugar catabolic pathways during lignocellulose degradation in WRF <span class="html-italic">Crucibulum laeve</span> (<b>B</b>, <b>bottom panel</b>). Enzymes that are absent from the <span class="html-italic">C. laeve</span> genome are marked in red. The question mark means that the genome of this fungus may contain analogues that are currently unknown.</p>
Full article ">
30 pages, 6492 KiB  
Review
Diversity, Distribution, and Evolution of Bioluminescent Fungi
by Brian A. Perry, Dennis E. Desjardin and Cassius V. Stevani
J. Fungi 2025, 11(1), 19; https://doi.org/10.3390/jof11010019 - 31 Dec 2024
Viewed by 1113
Abstract
All known bioluminescent fungi are basidiomycetes belonging to the Agaricales. They emit 520–530 nm wavelength light 24 h per day in a circadian rhythm. The number of known bioluminescent fungi has more than doubled in the past 15 years from 64 to 132 [...] Read more.
All known bioluminescent fungi are basidiomycetes belonging to the Agaricales. They emit 520–530 nm wavelength light 24 h per day in a circadian rhythm. The number of known bioluminescent fungi has more than doubled in the past 15 years from 64 to 132 species. We currently recognize five distinct lineages of bioluminescent Agaricales belonging to the Omphalotaceae (18 species), Physalacriaceae (14), Mycenaceae (96), Lucentipes lineage (3), and Cyphellopsidaceae (1). They are distributed across the globe with the highest diversity occurring on woody or leafy substrates in subtropical closed canopy forests with high plant diversity. With the caveat that most regions of the world have not been extensively sampled for bioluminescent fungi, the areas with the most known species are Japan (36), South America (30), North America (27), Malesia, South Asia, and Southeast Asia (26), Europe (23), Central America (21), China (13), Africa (10), Australasia, Papua New Guinea, and New Caledonia (11), and the Pacific Islands (5). Recent studies have elucidated the biochemical and genetic pathways of fungal bioluminescence and suggest the phenomenon originated a single time early in the evolution of the Agaricales. Multiple independent evolutionary losses explain the absence of luminescence in many species found within the five lineages and in the majority of Agaricales. Full article
(This article belongs to the Section Fungal Genomics, Genetics and Molecular Biology)
Show Figures

Figure 1

Figure 1
<p>Exemplary bioluminescent mushrooms from the five distinct evolutionary lineages present in Brazil: (<b>A</b>) <span class="html-italic">Neonothopanus gardneri</span> (<span class="html-italic">Omphalotus</span> lineage), (<b>B</b>) <span class="html-italic">Armillaria</span> sp. (<span class="html-italic">Armillaria</span> lineage), (<b>C</b>) <span class="html-italic">Mycena luxaeterna</span> (Mycenoid lineage), (<b>D</b>) <span class="html-italic">Mycena lucentipes</span> (Lucentipes lineage), and (<b>E</b>) <span class="html-italic">Eoscyphella luciurceolata</span> (<span class="html-italic">Eoscyphella</span> lineage).</p>
Full article ">Figure 2
<p>Known global distribution of bioluminescent fungi.</p>
Full article ">Figure 3
<p>Maximum likelihood phylogeny of order Agaricales based upon <span class="html-italic">nrLSU</span>, <span class="html-italic">rpb2</span>, and <span class="html-italic">ef1-a</span> sequence data, with bioluminescent lineages highlighted in green. Green circle indicates origin of bioluminescence in the MRCA to the Marasmiineae clade based upon relationships of bioluminescent lineages. Red circle represents hypothesized loss of bioluminescence that rendered remaining Agaricales non-luminescent. Please see <a href="#app1-jof-11-00019" class="html-app">Supplementary Materials</a> for details on data and phylogenetic reconstruction methods.</p>
Full article ">Figure 4
<p>Maximum likelihood phylogeny of Mycenceae based upon <span class="html-italic">ITS</span> and <span class="html-italic">nrLSU</span> sequence data, with bioluminescent species highlighted in green font. Values separated by/refer to ML bootstrap proportions and Bayesian posterior probabilities for values over 70/0.90, respectively. Sequences derived from type specimens are designated with an asterisk *. Please see <a href="#app1-jof-11-00019" class="html-app">Supplementary Materials</a> for details on data and phylogenetic reconstruction methods.</p>
Full article ">Figure 4 Cont.
<p>Maximum likelihood phylogeny of Mycenceae based upon <span class="html-italic">ITS</span> and <span class="html-italic">nrLSU</span> sequence data, with bioluminescent species highlighted in green font. Values separated by/refer to ML bootstrap proportions and Bayesian posterior probabilities for values over 70/0.90, respectively. Sequences derived from type specimens are designated with an asterisk *. Please see <a href="#app1-jof-11-00019" class="html-app">Supplementary Materials</a> for details on data and phylogenetic reconstruction methods.</p>
Full article ">Figure 4 Cont.
<p>Maximum likelihood phylogeny of Mycenceae based upon <span class="html-italic">ITS</span> and <span class="html-italic">nrLSU</span> sequence data, with bioluminescent species highlighted in green font. Values separated by/refer to ML bootstrap proportions and Bayesian posterior probabilities for values over 70/0.90, respectively. Sequences derived from type specimens are designated with an asterisk *. Please see <a href="#app1-jof-11-00019" class="html-app">Supplementary Materials</a> for details on data and phylogenetic reconstruction methods.</p>
Full article ">Figure 4 Cont.
<p>Maximum likelihood phylogeny of Mycenceae based upon <span class="html-italic">ITS</span> and <span class="html-italic">nrLSU</span> sequence data, with bioluminescent species highlighted in green font. Values separated by/refer to ML bootstrap proportions and Bayesian posterior probabilities for values over 70/0.90, respectively. Sequences derived from type specimens are designated with an asterisk *. Please see <a href="#app1-jof-11-00019" class="html-app">Supplementary Materials</a> for details on data and phylogenetic reconstruction methods.</p>
Full article ">Figure 5
<p>The Caffeic Acid Cycle (CAC), the biochemical pathway responsible for fungal bioluminescence.</p>
Full article ">
12 pages, 3120 KiB  
Article
Dispersion of Boletus-Type Spores Within and Beyond Beech Forest
by Magdalena Wójcik-Kanach and Idalia Kasprzyk
Forests 2024, 15(12), 2232; https://doi.org/10.3390/f15122232 - 19 Dec 2024
Viewed by 783
Abstract
Basidiomycetes produce huge numbers of spores that can disperse over various distances. High spore concentrations are typically observed near the fruiting bodies, with their abundance in the air influenced by the specific habitat in which the fungi grow. The aim of this study [...] Read more.
Basidiomycetes produce huge numbers of spores that can disperse over various distances. High spore concentrations are typically observed near the fruiting bodies, with their abundance in the air influenced by the specific habitat in which the fungi grow. The aim of this study was to investigate the concentration of airborne Boletus-type spores in a forest environment and evaluate the thesis that their dispersion beyond the forest is limited. Fungal spores were sampled in the summer and fall of 2022 in the forest, at its edge, and 1500 m away from the forest. The highest spore concentration was recorded in the forest interior, where approximately 76% of the total spores were detected. A noticeable decline in spore concentrations was observed at the forest edge, with a further significant reduction in open areas, including arable fields and mosaics of crops and heterogeneous plots, where spore presence accounted for only 12.1% of the total. During mushroom picking, airborne Boletus-type spores in the forest were unexpectedly high, so they could pose a threat to sensitized people. Full article
(This article belongs to the Section Forest Ecophysiology and Biology)
Show Figures

Figure 1

Figure 1
<p>Location of research stations (ST 1–ST 4) against the background of vegetation cover type.</p>
Full article ">Figure 2
<p>Homogeneous groups of sites (same letters) distinguished based on the comparison of spore concentrations using the Kruskal–Wallis test with multiple comparisons using Dunn’s test.</p>
Full article ">Figure 3
<p>Concentrations of Boletus-type spores at research sites (ST 1–ST 4) in August–November 2022.</p>
Full article ">Figure 4
<p>Dispersal of Boletus-type spores inside and outside forest.</p>
Full article ">Figure 5
<p>Boletus-type spore concentrations detected by stationary trap.</p>
Full article ">Figure 6
<p>The hourly concentrations (UTC+02:00) of airborne Boletus-type spores (spores in m<sup>3</sup>).</p>
Full article ">
19 pages, 4171 KiB  
Article
Characterisation of Itersonilia spp. from Parsnip and Other Hosts
by Lauren H. K. Chappell, Guy C. Barker and John P. Clarkson
J. Fungi 2024, 10(12), 873; https://doi.org/10.3390/jof10120873 - 16 Dec 2024
Viewed by 593
Abstract
Parsnips (Pastinaca sativa) are a speciality UK crop with an economic value of at least 31M GBP annually. Currently, the major constraints to production are losses associated with root canker disease due to a range of fungal pathogens, among which Itersonilia [...] Read more.
Parsnips (Pastinaca sativa) are a speciality UK crop with an economic value of at least 31M GBP annually. Currently, the major constraints to production are losses associated with root canker disease due to a range of fungal pathogens, among which Itersonilia pastinacae is of most concern to growers. With limited research conducted on this species, this work aimed to provide a much-needed characterisation of isolates from across the UK, continental Europe, and New Zealand. Previously, up to four separate Itersonilia species have been proposed based on the formation of chlamydospores and host specificity: I. pastinacae, I. perplexans, I. pyriformans, and I. pannonica. However, Itersonilia spp. isolates principally from parsnip, but also from a range of other hosts, which were found to infect both parsnip roots and leaves in pathogenicity tests. In growth rate assays, isolates were found to grow at temperatures of 0–25 °C and produce both chlamydospores and ballistospores across the same range of temperatures, although chlamydospore production was found to decrease as temperature increased. Following whole genome sequencing, specific primers were designed for the molecular characterisation of the isolates using six housekeeping genes and three highly variable functional genes. Phylogenetic analysis separated isolates into two and six clades, respectively, but the grouping was not associated with hosts or locations. Based on the results of this research, there was no evidence to support more than a single species of Itersonilia among the isolates studied. Full article
(This article belongs to the Section Fungal Evolution, Biodiversity and Systematics)
Show Figures

Figure 1

Figure 1
<p>(<b>A</b>) Symptoms of <span class="html-italic">Itersonilia</span> on inoculated parsnip roots (cv. Picador) after 21 days at 20 °C. Left: weakly virulent <span class="html-italic">Itersonilia</span> isolate IP15. Right: highly virulent <span class="html-italic">Itersonilia</span> isolate IP50. (<b>B</b>) Symptoms of <span class="html-italic">Itersonilia</span> on inoculated parsnip leaves (cv. Panache) after 7 days at 20 °C. Left: weakly virulent isolate IP8. Right: highly virulent isolate IP47.</p>
Full article ">Figure 2
<p>Mean lesion size (mm<sup>2</sup>) on parsnip roots (cv. Picador) for 48 different <span class="html-italic">Itersonilia</span> isolates. Error bars represent the standard error of the mean (SEM) for four independent replicates. Light grey bars represent isolates from parsnip hosts; dark grey indicates isolates from non-parsnip hosts (chrysanthemum, dill, fennel, and parsley [<a href="#jof-10-00873-t001" class="html-table">Table 1</a>]).</p>
Full article ">Figure 3
<p>Mean lesion size (mm<sup>2</sup>) on detached parsnip leaves (cv. Panache) for 48 different <span class="html-italic">Itersonilia</span> isolates. Data are plotted on a log scale; error bars represent the SEM for four independent replicates. LSD is indicated at the 5% level. Light grey bars indicate isolates from parsnip hosts; dark grey bars indicate isolates from non-parsnip hosts (chrysanthemum, dill, fennel, and parsley [<a href="#jof-10-00873-t001" class="html-table">Table 1</a>]).</p>
Full article ">Figure 4
<p>Effect of temperatures (<b>A</b>) 0 °C, (<b>B</b>) 5 °C, (<b>C</b>) 10 °C, (<b>D</b>) 15 °C, (<b>E</b>) 20 °C, and (<b>F</b>) 25 °C on the mean growth rate of <span class="html-italic">Itersonilia</span> isolates on MA. Error bars represent the SEM for four independent replicates. LSD is indicated at the 5% level. Light grey bars indicate isolates from parsnip hosts; dark grey bars indicate isolates from non-parsnip hosts (chrysanthemum, dill, fennel, and parsley [<a href="#jof-10-00873-t001" class="html-table">Table 1</a>]).</p>
Full article ">Figure 5
<p>Effect of temperature on the mean growth rate of <span class="html-italic">I. pastinacae</span> isolate IP10. Error bars represent the SEM for four independent replicates. The grey line represents a fitted Briere curve.</p>
Full article ">Figure 6
<p>Effect of temperatures (<b>A</b>) 0 °C, (<b>B</b>) 5 °C, (<b>C</b>) 10 °C, (<b>D</b>) 15 °C, (<b>E</b>) 20 °C, and (<b>F</b>) 25 °C on mean log<sub>10</sub> spore density (spores mm<sup>−2</sup>) for different <span class="html-italic">Itersonilia</span> isolates. Data points represent mean spore density for four replicates; error bars represent the SEM for four independent replicates. LSD is indicated at the 5% level. Lighter bars indicate isolates from parsnip hosts; darker bars indicate isolates from non-parsnip hosts (chrysanthemum, dill, fennel and parsley [<a href="#jof-10-00873-t001" class="html-table">Table 1</a>]).</p>
Full article ">Figure 7
<p>Maximum likelihood phylogenetic tree for <span class="html-italic">Itersonilia</span> spp. isolates based on the internal transcribed spacer region (ITS) of the rDNA (GenBank accession numbers MG198712-MG198760) alongside reference isolates for <span class="html-italic">Itersonilia pastinacae</span> (GenBank accession number CBS 356.64), <span class="html-italic">Itersonilia perplexans</span> (GenBank accession number CBS 144.68), and <span class="html-italic">Udenomyces pannonicus</span> (GenBank accession number AB072229.1). Numbers represent bootstrap values from 1000 replicates. Scale bar indicates 0.05 substitutions per site. The tree is rooted through <span class="html-italic">Cystofilobasidiales macerans</span> (GenBank Genome GCA_014825765.1) [<a href="#B25-jof-10-00873" class="html-bibr">25</a>]. ITS sequence for an <span class="html-italic">I. perplexans</span> reference isolate (IMI 264396) as taxonomically described by Ingold [<a href="#B9-jof-10-00873" class="html-bibr">9</a>] is also included (AB072233.1).</p>
Full article ">Figure 8
<p>Maximum likelihood phylogenetic tree for <span class="html-italic">Itersonilia</span> spp. isolates based on concatenated sequences for the internal transcribed spacer region (ITS) of the rDNA (GenBank accessions MG198712-MG198760), RNA polymerase II (<span class="html-italic">Rpb-II</span>), translation elongation factor (<span class="html-italic">EF-1α</span>), large ribosomal subunit (<span class="html-italic">LSU</span>) (GenBank accessions MG241126-MG241175), small ribosomal subunit (<span class="html-italic">SSU</span>) (GenBank accessions MG241176-MG241225), and beta-tubulin (<span class="html-italic">TUB2</span>). Numbers represent bootstrap values from 1000 replicates. Scale bar indicates 0.10 substitutions per site. The tree is rooted through <span class="html-italic">Cystofilobasidiales macerans</span> (GenBank Genome GCA_014825765.1) [<a href="#B25-jof-10-00873" class="html-bibr">25</a>].</p>
Full article ">Figure 9
<p>Maximum likelihood phylogenetic tree for <span class="html-italic">Itersonilia</span> spp. isolates based on the functional genes triosephosphate transporter family (<span class="html-italic">TTF</span>), tRNA methyl transferase (<span class="html-italic">tMT</span>), and cellobiose dehydrogenase (<span class="html-italic">CDH</span>). Numbers represent bootstrap values from 1000 replicates. Scale bar indicates 5 substitutions per site. The tree is rooted through <span class="html-italic">Cystofilobasidiales macerans</span> (GenBank Genome GCA_014825765.1) [<a href="#B25-jof-10-00873" class="html-bibr">25</a>].</p>
Full article ">
28 pages, 3496 KiB  
Article
The Diversity of Macrofungi in the Forests of Ningxia, Western China
by Xiaojuan Deng, Minqi Li, Yucheng Dai, Xuetai Zhu, Xingfu Yan, Zhaojun Wei and Yuan Yuan
Diversity 2024, 16(12), 725; https://doi.org/10.3390/d16120725 - 26 Nov 2024
Viewed by 560
Abstract
The diversity of macrofungi has been closely associated with forest diversity and stability. However, such a correlation has not been established for the forests of the Ningxia Autonomous Region due to the lack of systematic data on its macrofungal diversity. Therefore, for the [...] Read more.
The diversity of macrofungi has been closely associated with forest diversity and stability. However, such a correlation has not been established for the forests of the Ningxia Autonomous Region due to the lack of systematic data on its macrofungal diversity. Therefore, for the present study, we collected 3130 macrofungal specimens from the forests of the Helan Mts., Luo Mts., and Liupan Mts. in Ningxia and assessed them using morphological and molecular approaches. We identified 468 species belonging to 157 genera, 72 families, 18 orders, 11 classes, and 2 phyla. Among them, 31 species were ascomycetes, and 437 species were basidiomycetes. Tricholomataceae, with 96 species of 22 genera, was the most species-rich family, and Inocybe was the most species-rich genus (6.2%). The Jaccard similarity index measurement revealed the highest similarity in macrofungal species (16.15%) between the Helan and Liupan Mountains and the lowest (7.72%) between the Luo and Liupan Mountains. Further analyses of the macrofungal population of Ningxia showed that 206 species possess considerable potential for utilization, including 172 edible, 70 medicinal, and 36 edible–medicinal ones. Meanwhile, 54 species were identified as being poisonous. In these forests, saprophytic fungi were the most abundant, with 318 species (67.95%), followed by symbiotic fungi (31.62%) and parasitic fungi (0.04%). Grouping based on the geographical distribution indicated that the fungi of Ningxia are composed mainly of the cosmopolitan and north temperate types. These observations unveil the diversity and community structure of macrofungi in Ningxia forests. Full article
Show Figures

Figure 1

Figure 1
<p>The location of the Ningxia Province in China (<b>a</b>) and forest distribution in Ningxia Province (<b>d</b>). (<b>b</b>) Colors indicate the altitude of different sites in the Ningxia Province. (<b>c</b>) Geographic scale.</p>
Full article ">Figure 2
<p>An example of the field records of a collected specimen.</p>
Full article ">Figure 3
<p>Dominant macrofungal families (≥10 species) in the forests of the Ningxia Autonomous Region.</p>
Full article ">Figure 4
<p>Dominant macrofungal genera (≥5 species) in the forests of the Ningxia Autonomous Region. The ordinate shows the dominant genera, and the abscissa indicates the number of species in each genus.</p>
Full article ">Figure 5
<p>Proportion of geographical distribution patterns of macrofungal families in the forests of the Ningxia Autonomous Region.</p>
Full article ">Figure 6
<p>Proportion of geographical distribution patterns of macrofungal genera in the forests of the Ningxia Autonomous Region.</p>
Full article ">
11 pages, 2068 KiB  
Article
The Fungus Phlebiopsis flavidoalba’s Pathogenicity and Virulence Toward the Fluted Scale (Praelongorthezia acapulcoa) Pest of Rice and Sugarcane Crops
by Silvia Hernández-Hernández, Facundo Muñiz-Paredes, Guadalupe Peña-Chora and Víctor Manuel Hernández-Velázquez
Microbiol. Res. 2024, 15(4), 2414-2424; https://doi.org/10.3390/microbiolres15040162 - 26 Nov 2024
Viewed by 703
Abstract
Sugarcane is one of the main crops in Morelos State, Mexico. The presence of the insect pest Praelongorthezia acapulcoa (Morrison), commonly known as the fluted scale insect, was observed in sugarcane and rice crops, causing losses of up to 30% of production in [...] Read more.
Sugarcane is one of the main crops in Morelos State, Mexico. The presence of the insect pest Praelongorthezia acapulcoa (Morrison), commonly known as the fluted scale insect, was observed in sugarcane and rice crops, causing losses of up to 30% of production in both crops. In this work, a fungus isolated from the mycosic cadavers of P. acapulcoa was identified as the basidiomycete Phlebiopsis flavidoalba (Cooke) Hjortstam via morphological and molecular identification using the ITS, tef 1, and 28S regions. Its pathogenicity toward P. acapulcoa was verified in laboratory tests, causing a mortality rate higher than 60%. Its virulence toward P. acapulcoa, estimated as the mean lethal concentration (LC50), was 9.7 × 106 conidia mL−1. This work constitutes the first report about a basidiomycete with direct entomopathogenic activity and biological control of the fluted scale insect, P. acapulcoa. Full article
Show Figures

Figure 1

Figure 1
<p>Morphology of isolate SH52 at five days of growth: (<b>a</b>) growth on SDA medium; (<b>b</b>) septate hyphae; (<b>c</b>) fragmentation of hyphae into arthrospores. Crystals were present after 12 days of growth (<b>d</b>). All micrographs were observed at 100×.</p>
Full article ">Figure 2
<p>Phylogenetic tree based on concatenation of the ITS and 28S regions. The tree was constructed by the maximum likelihood method (1000 bootstrap replications). Isolate SH52 grouped with the <span class="html-italic">Phlebiopsis flavidoalba</span> clade. The numbers at branch points represent bootstrap support values. <span class="html-italic">Phlebia acerina</span> MY51 was used as the outgroup.</p>
Full article ">Figure 3
<p>Pathogenicity (<b>a</b>) and virulence (<b>b</b>) of <span class="html-italic">Phlebiopsis flavidoalba</span> toward <span class="html-italic">Praelongorthezia acapulcoa</span> at day 12. The LC<sub>50</sub> (9.7 × 10<sup>6</sup> conidia mL<sup>−1</sup>) was estimated via probit analysis. The asterisk above the bar represents a significant difference according to Student’s <span class="html-italic">t</span> test (<span class="html-italic">p</span> &lt; 0.001).</p>
Full article ">Figure 4
<p>Females of <span class="html-italic">Praelongorthezia acapulcoa:</span> (<b>a</b>) mycosic insect (eight days post-exposure to <span class="html-italic">Phlebiopsis flavidoalba</span>); (<b>b</b>) healthy specimen.</p>
Full article ">
22 pages, 8845 KiB  
Article
A Mycovirus Representing a Novel Lineage and a Mitovirus of Botrytis cinerea Co-Infect a Basidiomycetous Fungus, Schizophyllum commune
by Jie Duan, Anmeng Zhang, Yanping Fu, Yang Lin, Jiatao Xie, Jiasen Cheng, Tao Chen, Bo Li, Xiao Yu, Xueliang Lyu and Daohong Jiang
Viruses 2024, 16(11), 1767; https://doi.org/10.3390/v16111767 - 13 Nov 2024
Viewed by 1127
Abstract
Strain IBc-114 was isolated from a gray mold lesion and was identified as the fungus Schizophyllum commune. In this strain, two mycoviruses, Schizophyllum commune RNA virus 1 (ScRV1, C_AA053475.1) and Botrytis cinerea mitovirus 9 strain IBc-114 (BcMV9/IBc-114, C_AA053476.1), were isolated and characterized. ScRV1 [...] Read more.
Strain IBc-114 was isolated from a gray mold lesion and was identified as the fungus Schizophyllum commune. In this strain, two mycoviruses, Schizophyllum commune RNA virus 1 (ScRV1, C_AA053475.1) and Botrytis cinerea mitovirus 9 strain IBc-114 (BcMV9/IBc-114, C_AA053476.1), were isolated and characterized. ScRV1 has flexuous filamentous particles about 20 ± 2.1 nm in diameter and 1000 ± 94.2 nm in length. The genome of ScRV1 is 7370 nt in length and contains two open reading frames (ORFs) which encode a polyprotein and a coat protein, respectively. The polyprotein has 1967 aa, including a helicase domain and an RdRp domain which has the highest identity of 28.21% with that of Entomophthora benyvirus E (EbVE). The coat protein has 241 aa which is mostly phylogenetically close to the coat proteins of Alphatetraviridae. Based on the phylogenetic analysis of ScRV1 and viruses selected, ScRV1 might represent a new family (temporarily named Mycobenyviridae) of the order Hepelivirales. The genome of BcMV9/IBc-114 that infects S. commune is 2729 nt in length and has only one ORF encoding an RdRp protein with 719 aa. BcMV9/IBc-114 has the highest identity of 98.61% with Botrytis cinerea mitovirus 9 (BcMV9) (MT089704). ScRV1, but not BcMV9/IBc-114, has certain effects on the host growth of S. commune. Furthermore, BcMV9/IBc-114 has been demonstrated to replicate in the ascomycetous fungi Botrytis cinerea and Sclerotinia sclerotiorum, and it negatively affects the growth and pathogenicity of B. cinerea, but it does not affect S. sclerotiorum. This is the first report of mycoviruses in S. commune and cross-phyla transmission of mitovirus in nature. Full article
(This article belongs to the Collection Mycoviruses)
Show Figures

Figure 1

Figure 1
<p>Identification of strain IBc-114. (<b>A</b>) Colony morphology of strain IBc-114. All strains were cultured on plates with 20 mL of PDA medium at 20 °C. (<b>B</b>) Maximum-likelihood tree inferred from ITS, <span class="html-italic">Calmodulin</span>, <span class="html-italic">Ras</span> and <span class="html-italic">β-tubulin</span> sequences. Numbers above nodes indicate bootstrap percentage per 1000 replicates. Fungal species and associated CBS strains are listed in <a href="#app1-viruses-16-01767" class="html-app">Table S2</a>.</p>
Full article ">Figure 2
<p>Detection and information of viral contigs in strain IBc-114 using RT-PCR. (<b>A</b>) Detection of viral contigs in strain IBc-114 using RT-PCR. Lane M: DL2000, DNA marker; ACT1: <span class="html-italic">S. commune</span> actin 1 gene (442 bp); BC-1_Contig348: 801 bp; BC-3_Contig60: 798 bp; ACT1-H<sub>2</sub>O, BC-1_Contig348-H<sub>2</sub>O and BC-3_Contig60-H<sub>2</sub>O: H<sub>2</sub>O was used as a negative control. (<b>B</b>) Information of mycoviruses obtained from high-throughput sequencing (HTS) analysis of the <span class="html-italic">Schizophyllum commune</span> strain IBc-114.</p>
Full article ">Figure 3
<p>Diagrammatic sketch of the genome organization of ScRV1 and the alignment of RdRp domain and helicase (Hel) domains of ScRV1 with beny-like viruses. (<b>A</b>) The genome organization of SsBLV1, showing 5′ and 3′ untranslated regions and two ORF regions. (<b>B</b>) The polyprotein encoded by ORF1 of ScRV1 and conserved domains Hel and RdRp and the coat protein (CP) encoded by ORF2. The Arabic numerals show the amino acid position of each conserved domain. (<b>C</b>) The structure of the CP protein of ScRV1 obtained by prediction using AlphaFold2. (<b>D</b>) Multiple alignment of ScRV1 Hel motifs with those of selected beny-like viruses. Conserved motifs are marked by Roman numerals from I to VI. (<b>E</b>) Multiple alignment of ScRV1 RdRp motifs with those of selected beny-like viruses. Conserved motifs are marked by Roman numerals from I to VIII. Identical residues are indicated by asterisks and highlighted in black.</p>
Full article ">Figure 4
<p>Identification of virions from strain IBc-114. (<b>A</b>) The distribution of protein in the crude extract of virions was analyzed in a 10–60% sucrose gradient solution after ultracentrifugation. (<b>B</b>) RT-PCR detection of BcMV9/IBc-114 and ScRV1 in different sucrose concentration gradients. Lane M: DL5000, DNA marker; S-RNA: RT-PCR detection of BcMV9/IBc-114 in strain IBc-114; 10%, 20%, 30%, 40%, 50%, 60%: RT-PCR was performed on BcMV9/IBc-114 and ScRV1 in 10–60% sucrose solutions after ultracentrifugation, respectively; CE: ScRV1 in crude extract containing virions was detected by RT-PCR; H<sub>2</sub>O: ddH<sub>2</sub>O was used instead of RT products. (<b>C</b>) CP and RdRp peptide information of ScRV1 in 40% sucrose solution determined by 4DFastDIA-based proteomic. (<b>D</b>) A histogram of length and diameter of ScRV1 particles. Error bars indicate standard deviation (SD) from sample means. (<b>E</b>) Virions of ScRV1 observed under a TEM after negative staining.</p>
Full article ">Figure 5
<p>Phylogenetic analysis of ScRV1 based on the polyprotein using maximum likelihood with 1000 bootstrap replicates. Phylogenetic tree of ScRV1 based on the polyproteins constructed with the best-fit model Q.pfam + F + R5. The parameter of bootstrap was set as 1000 replicates, and bootstrap values over 50% are indicated on branches. The virus reported in this study is in red; the scale bar at the bottom left corresponds to the genetic distance.</p>
Full article ">Figure 6
<p>Phylogenetic analysis of ScRV1 based on the CP using maximum likelihood with 1000 bootstrap replicates. Phylogenetic tree of ScRV1 based on the CPs constructed with the best-fit model Q.pfam + F + I + G4. The parameter of bootstrap was set as 1000 replicates, and bootstrap values over 50% were indicated on branches. The virus reported in this study is in red; the scale bar at the bottom left corresponds to the genetic distance.</p>
Full article ">Figure 7
<p>Diagrammatic sketch of the genome organization of BcMV9/IBc-114 and sequence alignment of BcMV9/IBc-114, GaNV1 and BcMV9. (<b>A</b>) The genome organization of BcMV9/IBc-114, showing 5′ and 3′ UTR and ORF regions. (<b>B</b>) The RdRp encoded by the ORF of BcMV9/IBc-114 and conserved RdRp domain. Arabic numerals show the amino acid position of the conserved domain. (<b>C</b>) Nucleotide sequence alignment of BcMV9/IBc-114, GaNV1 and BcMV9. (<b>D</b>) Amino acid sequence alignment of BcMV9/IBc-114, GaNV1 and BcMV9.</p>
Full article ">Figure 8
<p>The biological characteristics of strains IBc-114, IBc-114-39, and IBc-114-46 and the virus detection using RT-PCR. (<b>A</b>) The colony morphology of strains IBc-114, IBc-114-39, and IBc-114-46 on PDA at 20 °C. (<b>B</b>) The growth rate of strains IBc-114, IBc-114-39, and IBc-114-46 on PDA at 20 °C. The growth rate was measured by recording the colony diameter at 24 h and 36 h. Error bars indicate standard deviation (SD) from sample means. Different uppercase letters on the top of each column indicate significant differences (<span class="html-italic">p</span> &lt; 0.05). (<b>C</b>) Detection of viruses in strains IBc-114, IBc-114-39, and IBc-114-46. Lane M: DL2000, DNA marker.</p>
Full article ">Figure 9
<p>Horizontal transmission of ScRV1 and BcMV9/IBc-114, the donor strain IBc-114 (D, donor; ScRV1<sup>+</sup>, BcMV9/IBc-114<sup>+</sup>), to the recipient strain IBc-114-39 (R, recipient; ScRV1<sup>−</sup>, BcMV9/IBc-114<sup>−</sup>) through hyphal contact in a pair culture. (<b>A</b>) A pair culture of R/D on PDA (20 °C, 12 days), two single cultures of isolates R and D, respectively, and six derivative isolates (R1, R2, R3, R4, R5 and R6) on PDA (20 °C, 5 days). “★” in the pair culture indicates the area where a mycelial agar plug was removed for generating a derivative of the recipient strain designated as R. (<b>B</b>) Average growth rates (<span class="html-italic">n</span> = 3) for IBc-114 and IBc-114-39, six derivative isolates. Different letters on the bars in each graph indicate significant difference (<span class="html-italic">p</span> &lt; 0.05) according to the least significant difference test. (<b>C</b>) Detection of <span class="html-italic">Actin</span> genes, ScRV1 and BcMV9/IBc-114 in IBc-114 and IBc-114-39, six derivative isolates by total RNA extraction and RT-PCR.</p>
Full article ">Figure 10
<p>Viruses and biological characteristics of <span class="html-italic">B. cinerea</span> and <span class="html-italic">S. sclerotiorum</span> transfectants. (<b>A</b>) Virus detection was conducted after transfecting <span class="html-italic">B. cinerea</span> B05.10 with the total RNA from strain IBc-114. (<b>B</b>) Virus detection was conducted after transfecting <span class="html-italic">S. sclerotiorum</span> Ep-1PNA367 with the total RNA from strain IBc-114. (<b>C</b>) The growth rate of strains B05.10, B05.10-P1-114, Ep-1PNA367 and Ep-1PNA367-P5-114 on PDA at 20 °C. (<b>D</b>) The colony morphology of strains B05.10, B05.10-P1-114, Ep-1PNA367 and Ep-1PNA367-P5-114 on PDA at 20 °C. (<b>E</b>) The area of lesions caused by strains B05.10, B05.10-P1-114, Ep-1PNA367 and Ep-1PNA367-P5-114 on detached tomato fruits at 4 dpi. The data are presented as means ± SD (<span class="html-italic">n</span> = 5). “***” indicates significant difference (<span class="html-italic">p</span> &lt; 0.001), while “ns” indicates no significant difference. (<b>F</b>) The lesions of strains B05.10, B05.10-P1-114, Ep-1PNA367 and Ep-1PNA367-P5-114 on the detached tomato fruits at 4 dpi.</p>
Full article ">
26 pages, 21773 KiB  
Article
The Diversity and Taxonomy of Thelephoraceae (Basidiomycota) with Descriptions of Four Species from Southwestern China
by Xiaojie Zhang, Fulei Shi, Sicheng Zhang, Md. Iqbal Hosen and Changlin Zhao
J. Fungi 2024, 10(11), 775; https://doi.org/10.3390/jof10110775 - 7 Nov 2024
Viewed by 1299
Abstract
Taxonomy plays a central role in understanding the diversity of life, translating the products of biological exploration and discovery specimens and observations into systems of names that settle a “classification home” to taxa. The ectomycorrhizal basidiomycetes family Thelephoraceae has been understudied in subtropical [...] Read more.
Taxonomy plays a central role in understanding the diversity of life, translating the products of biological exploration and discovery specimens and observations into systems of names that settle a “classification home” to taxa. The ectomycorrhizal basidiomycetes family Thelephoraceae has been understudied in subtropical ecosystems. Many species of Thelephoraceae are important edible and medicinal fungi, with substantial economic value. Four new species, Thelephora resupinata, T. subtropica, T. yunnanensis, and Tomentella tenuifarinacea, are proposed based on a combination of the morphological features and molecular evidence. Thelephora resupinata is characterized by the resupinate basidiomata having a tuberculate pileal surface hymenial, and the presence of the subglobose to globose basidiospores (9–12 × 7–9 µm). T. subtropica is solitary coriaceous infundibuliform gray-brown basidiomata with a presence of the subclavate basidia and subglobose to globose basidiospores (6–8 × 5–7 µm). T. yunnanensis is typical of the laterally stipitate basidiomata having a smooth, umber to coffee hymenial surface, a monomitic hyphal system with clamped generative hyphae, and the presence of the subglobose basidiospores (7–10 × 6–8 µm). Tomentella tenuifarinacea is typical of the arachnoid basidiomata having a smooth, gray, or dark gray hymenial surface, a monomitic hyphal system with clamped generative hyphae, and the presence of the subglobose to globose basidiospores (7–9 × 6–8 µm). Sequences of ITS+nLSU+mtSSU genes were used for the phylogentic analyses using maximum likelihood, maximum parsimony, and Bayesian inference methods. The three genes’ (ITS+nLSU+mtSSU) phylogenetic analysis showed that the genera Thelephora and Tomentella grouped together within the family Thelephoraceae and three new species were nested into the genus Thelephora, and one new species was nested into the genus Tomentella. Full article
(This article belongs to the Section Fungal Evolution, Biodiversity and Systematics)
Show Figures

Figure 1

Figure 1
<p>Maximum parsimony strict consensus tree illustrating the phylogeny of <span class="html-italic">Thelephora</span> and <span class="html-italic">Tomentella</span> and related genera in the family Thelephoraceae based on ITS+nLSU+mtSSU sequences. Branches are labeled with maximum likelihood bootstrap values ≥ 70%, parsimony bootstrap values ≥ 50%, and Bayesian posterior probabilities ≥ 0.95, respectively.</p>
Full article ">Figure 2
<p>Maximum parsimony strict consensus tree illustrating the phylogeny of the three new species and related species in <span class="html-italic">Thelephora</span> based on ITS sequences. Branches are labeled with maximum likelihood bootstrap values ≥ 70%, parsimony bootstrap values ≥ 50%, and Bayesian posterior probabilities ≥ 0.95, respectively. The new species are in bold.</p>
Full article ">Figure 3
<p>Maximum parsimony strict consensus tree illustrating the phylogeny of the one new species and related species in <span class="html-italic">Tomentella</span>, based on ITS+nLSU sequences. Branches are labeled with maximum likelihood bootstrap values ≥ 70%, parsimony bootstrap values ≥ 50%, and Bayesian posterior probabilities ≥ 0.95, respectively. The new species are in bold.</p>
Full article ">Figure 3 Cont.
<p>Maximum parsimony strict consensus tree illustrating the phylogeny of the one new species and related species in <span class="html-italic">Tomentella</span>, based on ITS+nLSU sequences. Branches are labeled with maximum likelihood bootstrap values ≥ 70%, parsimony bootstrap values ≥ 50%, and Bayesian posterior probabilities ≥ 0.95, respectively. The new species are in bold.</p>
Full article ">Figure 4
<p>Split graphs showing the results of PHI test for the ITS data of <span class="html-italic">Thelephora resupinata</span> and <span class="html-italic">T. subtropica</span> and closely related taxa using LogDet transformation and splits decomposition. PHI test results Φw ≤ 0.05 indicate that there is significant recombination within the dataset. New taxa are in red.</p>
Full article ">Figure 5
<p>Split graphs showing the results of PHI test for the ITS data of <span class="html-italic">Thelephora yunnanensis</span> and closely related taxa using LogDet transformation and splits decomposition. PHI test results Φw ≤ 0.05 indicate that there is significant recombination within the dataset. New taxa are in red.</p>
Full article ">Figure 6
<p>Split graphs showing the results of PHI test for the ITS data of <span class="html-italic">Tomentella tenuifarinacea</span> and closely related taxa using LogDet transformation and splits decomposition. PHI test results Φw ≤ 0.05 indicate that there is significant recombination within the dataset. New taxa are in red.</p>
Full article ">Figure 7
<p>Basidiomata of <span class="html-italic">Thelephora resupinata</span>: CLZhao 34538 (holotype). Basidiomata on the substrate (<b>A</b>), macroscopic characteristics of hymenophore (<b>B</b>). Bars: (<b>A</b>) = 1 cm; (<b>B</b>) = 1 mm.</p>
Full article ">Figure 8
<p>Microscopic structures of <span class="html-italic">Thelephora resupinata</span>: CLZhao 34538 (holotype). (<b>A</b>) Basidiospores, (<b>B</b>) basidia and basidioles, and (<b>C</b>) a section of hymenium. Bars: (<b>A</b>–<b>C</b>) = 10 µm.</p>
Full article ">Figure 9
<p>Basidiomata of <span class="html-italic">Thelephora subtropica</span>: CLZhao 30590 (holotype). Basidiomata on the substrate (<b>A</b>), macroscopic characteristics of hymenophore (<b>B</b>). Bars: (<b>A</b>) 1 cm; (<b>B</b>) 1 mm.</p>
Full article ">Figure 10
<p>Microscopic structures of <span class="html-italic">Thelephora subtropica</span>: CLZhao 30590 (holotype). (<b>A</b>) Basidiospores, (<b>B</b>) basidia and basidioles, and (<b>C</b>) a section of hymenium. Bars: (<b>A</b>–<b>C</b>) = 10 µm.</p>
Full article ">Figure 11
<p>Basidiomata of <span class="html-italic">Thelephora yunnanensis</span>: CLZhao 20929 (holotype). Basidiomata on the substrate (<b>A</b>), macroscopic characteristics of hymenophore (<b>B</b>). Bars: (<b>A</b>) 1 cm; (<b>B</b>) 1 mm.</p>
Full article ">Figure 12
<p>Microscopic structures of <span class="html-italic">Thelephora yunnanensis</span>: CLZhao 20929 (holotype). (<b>A</b>) Basidiospores, (<b>B</b>) basidia and basidioles, and (<b>C</b>) a section of hymenium. Bars: (<b>A</b>–<b>C</b>) = 10 µm.</p>
Full article ">Figure 13
<p>Basidiomata of <span class="html-italic">Tomentella tenuifarinacea</span>: CLZhao 31337 (holotype). Basidiomata on the substrate (<b>A</b>), macroscopic characteristics of hymenophore (<b>B</b>). Bars: (<b>A</b>) 1 cm; (<b>B</b>) 1 mm.</p>
Full article ">Figure 14
<p>Microscopic structures of <span class="html-italic">Tomentella tenuifarinacea</span>: CLZhao 31337 (holotype). (<b>A</b>) Basidiospores, (<b>B</b>) basidia and basidioles, (<b>C</b>) and a section of hymenium. Bars: (<b>A</b>–<b>C</b>) = 10 µm.</p>
Full article ">
11 pages, 524 KiB  
Article
Novel Basidiomycetous Alcohol Oxidase from Cerrena unicolor—Characterisation, Kinetics, and Proteolytic Modifications
by Sylwia Stefanek, Rafał Typek, Michał Dybowski, Dorota Wianowska, Magdalena Jaszek and Grzegorz Janusz
Int. J. Mol. Sci. 2024, 25(22), 11890; https://doi.org/10.3390/ijms252211890 - 5 Nov 2024
Viewed by 609
Abstract
Intracellular alcohol oxidase (AOX) was isolated from the basidiomycetous white rot fungus Cerrena unicolor FCL139. The enzyme was semi-purified (13-fold) using two-step chromatography with 30% activity recovery. The identity of the protein was confirmed by LC-MS/MS analysis, and its MW (72 kDa) and [...] Read more.
Intracellular alcohol oxidase (AOX) was isolated from the basidiomycetous white rot fungus Cerrena unicolor FCL139. The enzyme was semi-purified (13-fold) using two-step chromatography with 30% activity recovery. The identity of the protein was confirmed by LC-MS/MS analysis, and its MW (72 kDa) and pI (6.18) were also determined. The kinetics parameters of the AOX reaction towards various substrates were analysed, which proved that, in addition to methanol (4.36 ± 0.27% of the oxidised substrate), AOX most potently oxidises aromatic alcohols, such as 4-hydroxybenzyl alcohol (14.0 ± 0.8%), benzyl alcohol (4.2 ± 0.3%), anisyl alcohol (7.6 ± 0.4%), and veratryl alcohol (5.0 ± 0.3%). Moreover, the influence of selected commercially available proteases on the biocatalytic properties of AOX from C. unicolor was studied. It was proved that the digested enzyme lost its catalytic potential properties except when incubated with pepsin, which significantly boosted its activity up to 123%. Full article
Show Figures

Figure 1

Figure 1
<p>Estimation of the pH (<b>left</b>) and temperature (<b>right</b>) optima for the reaction of <span class="html-italic">C. unicolor</span> alcohol oxidase towards methanol.</p>
Full article ">
13 pages, 3118 KiB  
Article
Genome-Wide Identification and Analysis of Gene Family of Carbohydrate-Binding Modules in Ustilago crameri
by Dongyu Zhai, Deze Xu, Ting Xiang, Yu Zhang, Nianchen Wu, Fuqing Nie, Desuo Yin and Aijun Wang
Int. J. Mol. Sci. 2024, 25(21), 11790; https://doi.org/10.3390/ijms252111790 - 2 Nov 2024
Viewed by 845
Abstract
Ustilago crameri is a pathogenic basidiomycete fungus that causes foxtail millet kernel smut (FMKS), a devastating grain disease in most foxtail millet growing regions of the world. Carbohydrate-Binding Modules (CBMs) are one of the important families of carbohydrate-active enzymes (CAZymes) in fungi and [...] Read more.
Ustilago crameri is a pathogenic basidiomycete fungus that causes foxtail millet kernel smut (FMKS), a devastating grain disease in most foxtail millet growing regions of the world. Carbohydrate-Binding Modules (CBMs) are one of the important families of carbohydrate-active enzymes (CAZymes) in fungi and play a crucial role in fungal growth and development, as well as in pathogen infection. However, there is little information about the CBM family in U. crameri. Here, 11 CBM members were identified based on complete sequence analysis and functional annotation of the genome of U. crameri. According to phylogenetic analysis, they were divided into six groups. Gene structure and sequence composition analysis showed that these 11 UcCBM genes exhibit differences in gene structure and protein motifs. Furthermore, several cis-regulatory elements involved in plant hormones were detected in the promoter regions of these UcCBM genes. Gene ontology (GO) enrichment and protein–protein interaction (PPI) analysis showed that UcCBM proteins were involved in carbohydrate metabolism, and multiple partner protein interactions with UcCBM were also detected. The expression of UcCBM genes during U. crameri infection is further clarified, and the results indicate that several UcCBM genes were induced by U. crameri infection. These results provide valuable information for elucidating the features of U. crameri CBMs’ family proteins and lay a crucial foundation for further research into their roles in interactions between U. crameri and foxtail millet. Full article
(This article belongs to the Special Issue Plant–Microbe Interactions)
Show Figures

Figure 1

Figure 1
<p>A phylogenetic tree of the UcCBM proteins and their homologous proteins in other smut fungi, including <span class="html-italic">Sporisorium scitamineum</span>, <span class="html-italic">U. maydis</span>, and <span class="html-italic">U. hordei</span>. The phylogenetic tree was constructed with MEGA7 software using the Neighbor-Joining (NJ) method. Different groups are represented by branches and frames of different colors.</p>
Full article ">Figure 2
<p>Motif distribution, conserved domain, and gene structure analysis of UcCBMs. (<b>A</b>) Phylogenetic trees and conserved motifs of 11 UcCBMs. (<b>B</b>) nserved motif sequence logo of UcCBM proteins. (<b>C</b>) Domain analysis of UcCBMs. (<b>D</b>) Exon–intron structures of UcCBMs.</p>
Full article ">Figure 3
<p>Cis-acting elements of UcCBM family members in <span class="html-italic">U. crameri</span>. The 2000 bp promoter sequences of <span class="html-italic">U. crameri</span> UcCBM genes contain a variety of cis-acting elements, including hormone responsive elements, drought, low-temperature, and other response elements, as well as an MYBHv1 binding site.</p>
Full article ">Figure 4
<p>Three-dimensional (3D) modeling of the UcCBM proteins that were predicted, displayed at a confidence level &gt; 0.7.</p>
Full article ">Figure 5
<p>Gene Ontology enrichment analysis of <span class="html-italic">UcCBM</span> genes.</p>
Full article ">Figure 6
<p>Expression of UcCBM gene in <span class="html-italic">U. crameri</span> at different inoculation time points.</p>
Full article ">Figure 7
<p>Protein–protein interaction network of UcCBM family proteins, constructed based on smut fungi <span class="html-italic">U. hordei</span> genome. UHOR_06976: UcCBM1; UHOR_02432: UcCBM3; UHOR_02718: UcCBM2; UHOR_00509: UcCBM4 or UcCBM6; UHOR_00973: UcCBM5; UHOR_02067: UcCBM7; UHOR_04077: UcCBM8; UHOR_06273: UcCBM9; UHOR_03469: UcCBM10; UHOR_03282: UcCBM11.</p>
Full article ">
21 pages, 1656 KiB  
Review
Antibacterial and Antifungal Activity of Metabolites from Basidiomycetes: A Review
by Valeria Lysakova, Larissa Krasnopolskaya, Maria Yarina and Mayya Ziangirova
Antibiotics 2024, 13(11), 1026; https://doi.org/10.3390/antibiotics13111026 - 31 Oct 2024
Viewed by 1405
Abstract
Background/Objectives: The search for new antimicrobial molecules is important to expand the range of available drugs, as well as to overcome the drug resistance of pathogens. One of the promising sources of antibacterial and antifungal metabolites is basidial fungi, which have wide [...] Read more.
Background/Objectives: The search for new antimicrobial molecules is important to expand the range of available drugs, as well as to overcome the drug resistance of pathogens. One of the promising sources of antibacterial and antifungal metabolites is basidial fungi, which have wide biosynthetic capabilities. Methods: The review summarized the results of studying the antimicrobial activity of extracts and metabolites from basidiomycetes published from 2018–2023. Results: In all studies, testing for antibacterial and antifungal activity was carried out in in vitro experiments. To obtain the extracts, mainly the fruiting bodies of basidiomycetes, as well as their mycelia and culture liquid were used. Antimicrobial activity was found in aqueous, methanol, and ethanol extracts. Antimicrobial metabolites of basidiomycetes were isolated mainly from the submerged culture of basidiomycetes. Metabolites active against Gram-positive and Gram-negative bacteria and mycelial and yeast-like fungi were identified. Conclusions: Basidiomycete extracts and metabolites have shown activity against collectible strains of bacteria and fungi and multi-resistant and clinical strains of pathogenic bacteria. The minimum inhibitory concentration (MIC) values of the most active metabolites ranged from 1 to 16.7 µg/mL. Full article
(This article belongs to the Special Issue The Search for Antimicrobial Agents from Natural Products)
Show Figures

Figure 1

Figure 1
<p>The percentage of basidiomycete orders’ representatives in studies on the antimicrobial activity of their various extracts.</p>
Full article ">Figure 2
<p>The use of various solvents to produce extracts.</p>
Full article ">Figure 3
<p>(<b>A</b>) Sources of antimicrobial molecules of basidiomycetes. (<b>B</b>) The percentage of basidiomycete orders representatives in studies on the isolation of antimicrobial molecules.</p>
Full article ">Figure 4
<p>Basidiomycetes metabolites with antimicrobial activity.</p>
Full article ">
15 pages, 4435 KiB  
Article
Four New Species of Deconica (Strophariaceae, Agaricales) from Subtropical Regions of China
by Jun-Qing Yan, Sheng-Nan Wang, Ya-Ping Hu, Cheng-Feng Nie, Bin-Rong Ke, Zhi-Heng Zeng and Hui Zeng
J. Fungi 2024, 10(11), 745; https://doi.org/10.3390/jof10110745 - 29 Oct 2024
Viewed by 997
Abstract
Deconica is a relatively small genus, with only 90 names recorded in previous research. In this study, four new species of Deconica have been identified based on morphological and phylogenetic evidence from subtropical regions of China. This represents the first discovery of new [...] Read more.
Deconica is a relatively small genus, with only 90 names recorded in previous research. In this study, four new species of Deconica have been identified based on morphological and phylogenetic evidence from subtropical regions of China. This represents the first discovery of new species of Deconica in China. Morphologically, D. austrosinensis is characterized by medium-sized spores that are elliptical to elongated-ellipsoid in face view, and fusiform to sublageniform and slightly thick-walled pleurocystidia; D. furfuracea is identified by a well-developed and evanescent veil, medium-sized spores that are rhomboid to mitriform in face view, and fusiform to subclavate pleurocystidia that are rare and subacute at apex; D. fuscobrunnea is recognized by dark brown pileus, medium-sized spores that are rhomboid to mitriform in face view, an ixocutis pileipellis, lageniform cheilocystidia with a long neck and lacks pleurocystidia; D. ovispora is distinguished from other Deconica species by medium-sized spores that are ovoid in face view, an ixocutis pileipellis, lageniform cheilocystidia with a long to short neck, and lacks pleurocystidia. Their distinct taxonomic status is confirmed by the positions of the four new species in ITS + LSU phylogenetic trees. Detailed descriptions and morphological photographs of four new species are presented. 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>Phylogram of <span class="html-italic">Deconica</span> spp. generated with maximum likelihood (ML) analysis based on ITS and LSU, rooted with <span class="html-italic">Kuehneromyces brunneoalbescens</span>. Bayesian inference (BI-PP) ≥ 0.95 and ML bootstrap proportions (ML-BP) ≥ 75 are indicated as PP/BP. The new taxa are marked in bold.</p>
Full article ">Figure 2
<p>Morphological structures of <span class="html-italic">Deconica austrosinensis.</span> (<b>A</b>–<b>D</b>) Basidiomata. (<b>E</b>) Pileipellis. (<b>F</b>–<b>J</b>) Pleurocystidia. (<b>K</b>) Spores. (<b>L</b>) Cheilocystidia. Scale bars: (<b>A</b>–<b>D</b>) 10 mm, (<b>E</b>–<b>L</b>) 10 μm. All microstructures were observed in 5%KOH. Structures of (<b>F</b>–<b>I</b>,<b>L</b>) were stained with 1%Congo red.</p>
Full article ">Figure 3
<p>Morphological structures of <span class="html-italic">Deconica furfuracea.</span> (<b>A</b>–<b>D</b>) Basidiomata. (<b>E</b>) Spores. (<b>F</b>) Pileipellis. (<b>G</b>–<b>I</b>) Pleurocystidia. (<b>J</b>) Hymenium. (<b>K</b>) Cheilocystidia. Scale bars: (<b>A</b>–<b>D</b>) 10 mm, (<b>E</b>–<b>K</b>) 10 μm. All microstructures were observed in 5%KOH. Structures of (<b>G</b>–<b>K</b>) were stained with 1%Congo red.</p>
Full article ">Figure 4
<p>Morphological structures of <span class="html-italic">Deconica fuscobrunnea.</span> (<b>A</b>,<b>B</b>) Basidiomata. (<b>C</b>) Pileipellis. (<b>D</b>) Spores. (<b>E</b>,<b>F</b>) Cheilocystidia. Scale bars: (<b>A</b>,<b>B</b>) 10 mm, (<b>C</b>–<b>F</b>) 10 μm. All microstructures were observed in 5%KOH. Structures of (<b>E</b>,<b>F</b>) were stained with 1%Congo red.</p>
Full article ">Figure 5
<p>Morphological structures of <span class="html-italic">Deconica ovispora.</span> (<b>A</b>–<b>D</b>) Basidiomata. (<b>E</b>) Spores. (<b>F</b>) Pileipellis. (<b>G</b>,<b>H</b>) Cheilocystidia. Scale bars: (<b>A</b>–<b>D</b>) 10 mm, (<b>E</b>–<b>H</b>) 10 μm. All microstructures were observed in 5%KOH. Structures of (<b>F</b>–<b>H</b>) were stained with 1%Congo red.</p>
Full article ">
19 pages, 1933 KiB  
Article
The Effect of Flame Sterilization on the Microorganisms in Continuously Cultivated Soil and the Yield and Quality of Tobacco Leaves
by Xueying Han, Liang Wang, Ruyan Li and Qingli Han
Agriculture 2024, 14(11), 1868; https://doi.org/10.3390/agriculture14111868 - 23 Oct 2024
Viewed by 704
Abstract
Flame disinfection is a new technology that uses high temperatures to kill pathogens and control soil-borne diseases. In order to determine the feasibility of applying flame disinfection technology to flue-cured tobacco, a field experiment was conducted in Pianpo Village (test site I) and [...] Read more.
Flame disinfection is a new technology that uses high temperatures to kill pathogens and control soil-borne diseases. In order to determine the feasibility of applying flame disinfection technology to flue-cured tobacco, a field experiment was conducted in Pianpo Village (test site I) and Lühuai Village (test site II), Luquan County, Yunnan Province. The effects of flame disinfection on soil-borne disease control, flue-cured tobacco growth agronomic traits, the tobacco yield and quality, and the soil microbial community in the flue-cured tobacco field were investigated. The results were as follows. (1) After flame disinfection, the control rates of the four main soil-borne diseases—black root rot, root rot, wilt, and root knot nematodes—were all over 70%. (2) Samples were taken from the experimental site of Pianpo Village at 0 and 114 days after disinfection (tobacco boom period) to study the effects of soil microbial communities by high-throughput sequencing. Compared with the control group, after 0 days of flame disinfection, the abundance of bacterial actinobacteria, Nocardia, Streptomyces, and fungal ascomycetes decreased, while the abundance of bacterial Proteobacteria, Bacteroidetes, Arthrobacter, and mycospora increased. After 114 days of disinfection, the abundance of bacterial actinobacteria, Proteobacteria, chloromyces, and fungal ascomycetes decreased. The abundance of Mortierella was recovered, the abundance of Gibberella and Fusarium increased, and the abundance of Trichospora and Basidiomycetes decreased in both periods. (3) After flame disinfection treatment, the tobacco yield in the two experimental areas was increased by 50.80% and 54.70%, respectively, and the proportion of high-quality tobacco was also increased. In conclusion, flame disinfection before tobacco planting can improve the soil conditions, effectively control soil-borne tobacco diseases, and improve the quality and yield of tobacco leaves. Full article
(This article belongs to the Special Issue Integrated Management of Soil-Borne Diseases)
Show Figures

Figure 1

Figure 1
<p>Rarefaction curves of bacteria (<b>A</b>) and fungi (<b>B</b>) in flame disinfection and control soil samples.</p>
Full article ">Figure 2
<p>Numbers of specific OTUs of bacteria (<b>A</b>) and fungi (<b>B</b>) in flame disinfection treatment and control. (Note: Different colors in the figure represent different soil samples, and overlapping areas indicate the OTU shared by microorganisms among different soil samples).</p>
Full article ">Figure 3
<p>Principal coordinate analysis of flame disinfection and control bacterial and fungal communities.</p>
Full article ">Figure 4
<p>Hierarchical clustering analysis (based on Bray–Curtis method) of bacteria (<b>A</b>) and fungi (<b>B</b>) at the OTU level under different treatments.</p>
Full article ">Figure 5
<p>Taxonomic composition and relative abundance of the top 10 phyla of bacteria (<b>A</b>) and fungi (<b>B</b>).</p>
Full article ">Figure 5 Cont.
<p>Taxonomic composition and relative abundance of the top 10 phyla of bacteria (<b>A</b>) and fungi (<b>B</b>).</p>
Full article ">Figure 6
<p>Taxonomic composition and relative abundance of the top 10 genera of bacteria (<b>A</b>) and fungi (<b>B</b>).</p>
Full article ">
Back to TopTop