[go: up one dir, main page]
More Web Proxy on the site http://driver.im/
Next Issue
Volume 8, November
Previous Issue
Volume 8, September
You seem to have javascript disabled. Please note that many of the page functionalities won't work as expected without javascript enabled.
 
 

J. Fungi, Volume 8, Issue 10 (October 2022) – 132 articles

Cover Story (view full-size image): The formation of fruiting bodies is a highly regulated process that requires the concerted formation of various cell types. Though many developmental factors have been identified, a complete understanding of this process is still lacking. In this study, we analyze the sterile mutant pro34 of Sordaria macrospora. Genome sequencing revealed a deletion in the pro34 gene encoding a putative assembly factor of mitochondrial respiratory chain complex I. We showed that PRO34 is a mitochondrial protein and that the pro34 mutant lacks mitochondrial complex I. Inhibitor experiments revealed that pro34 respires via complexes III and IV, but also shows induction of alternative oxidase, a shunt pathway to bypass these complexes. We discuss the hypothesis that alternative oxidase is induced to prevent retrograde electron transport to complex I intermediates, thereby protecting from oxidative stress. View this paper
  • Issues are regarded as officially published after their release is announced to the table of contents alert mailing list.
  • You may sign up for e-mail alerts to receive table of contents of newly released issues.
  • PDF is the official format for papers published in both, html and pdf forms. To view the papers in pdf format, click on the "PDF Full-text" link, and use the free Adobe Reader to open them.
Order results
Result details
Section
Select all
Export citation of selected articles as:
16 pages, 2671 KiB  
Article
Resporulation of Calcium Alginate Encapsulated Metarhizium anisopliae on Metham®-Fumigated Soil and Infectivity on Larvae of Tenebrio molitor
by Sudhan Shah, Gavin J. Ash and Bree A. L. Wilson
J. Fungi 2022, 8(10), 1114; https://doi.org/10.3390/jof8101114 - 21 Oct 2022
Cited by 4 | Viewed by 2842
Abstract
Metarhizium anisopliae infects and kills a large range of insects and is a promising biocontrol agent to manage soil insects, such as wireworm in sweetpotato. The presence of other soil microbes, which exhibit competitive fungistasis, may inhibit the establishment of M. anisopliae in [...] Read more.
Metarhizium anisopliae infects and kills a large range of insects and is a promising biocontrol agent to manage soil insects, such as wireworm in sweetpotato. The presence of other soil microbes, which exhibit competitive fungistasis, may inhibit the establishment of M. anisopliae in soil. Microbially depleted soil, for example, sterilized soil, has been shown to improve the resporulation of the fungus from nutrient-fortified M. anisopliae. Prior to planting, sweetpotato plant beds can be disinfected with fumigants, such as Metham®, to control soil-borne pests and weeds. Metham® is a broad-spectrum soil microbial suppressant; however, its effect on Metarhizium spp. is unclear. In the research presented here, fungal resporulation was examined in Metham®-fumigated soil and the infectivity of the resulting granule sporulation was evaluated on mealworm, as a proxy for wireworm. The fungal granules grown on different soil treatments (fumigated, field and pasteurized soil) resporulated profusely (for example, 4.14 × 107 (±2.17 × 106) conidia per granule on fumigated soil), but the resporulation was not significantly different among the three soil treatments. However, the conidial germination of the resporulated granules on fumigated soil was >80%, which was significantly higher than those on pasteurized soil or field soil. The resporulated fungal granules were highly infective, causing 100% insect mortality 9 days after the inoculation, regardless of soil treatments. The results from this research show that the fungal granules applied to soils could be an infective inoculant in sweetpotato fields in conjunction with soil fumigation. Additional field studies are required to validate these results and to demonstrate integration with current farming practices. Full article
Show Figures

Figure 1

Figure 1
<p>Fungal (<span class="html-italic">M. anisopliae</span>) granules (CAG<sub>Ma+Cs+By</sub>) (<b>A</b>) characterized with Ø = 3.5 mm, weight 25 mg per granule, 9 × 10<sup>6</sup> conidia per granule and food (control) granules (CAG<sub>Cs+By</sub>) (<b>B</b>) as the control treatment.</p>
Full article ">Figure 2
<p>Resporulated <span class="html-italic">M. anisopliae</span> granules inoculated in field soil (<b>A</b>), fumigated soil (<b>B</b>) and pasteurized soil (<b>C</b>) 10 days post-incubation.</p>
Full article ">Figure 3
<p>The colony-forming units (CFUs) present on standard nutrient agar (NA: black) or potato dextrose agar (PDA: Grey) media soil sourced from field soil, pasteurized soil, fumigated soil with soil sampled from the top fumigant-treated area (fumigated 1) and fumigated soil with soil sampled from homogenized fumigated soil (fumigated 2).</p>
Full article ">Figure 4
<p>Sporulated cadavers were found in field soil samples sourced from the agricultural field at the USQ, Toowoomba (<b>A</b>); a fungal colony, suspected to be <span class="html-italic">Metarhizium brunneum</span> grown from a single colony isolated from an insect cadaver (<b>B</b>) and the fungal conidia obtained from the fungal colony (<b>C</b>).</p>
Full article ">Figure 5
<p>Resporulated fungal granules following 10 days incubation of fungal granules on field soil (<b>A</b>), fumigated soil (<b>B</b>) and pasteurized soil (<b>C</b>) and showing impact of soil treatments on conidial germination and germ-tube length (following a 14h incubation at 27 °C and a 12:12 h dark and light photoperiod) of the resporulated fungal granules grown on field soil (<b>D</b>), fumigated soil (<b>E</b>) and pasteurized soil (<b>F</b>).</p>
Full article ">Figure 6
<p>Percentage germination of extracted conidia from resporulated fungal granules applied on either field or fumigated or pasteurized soil for 10 days (<span class="html-italic">p &lt;</span> 0.05) (<b>A</b>) and measurement of germ-tube length from the conidia extracted from the resporulated fungal granules (<span class="html-italic">p</span> &lt; 0.05) (<b>B</b>).</p>
Full article ">Figure 7
<p>A dead larval mealworm recovered from fumigated soil containing resporulated fungal granules at 8 days post-inoculation with fungal inocula (<b>A</b>) a larval mealworm with profuse sporulation following placement in a moist chamber for 4 days (<b>B</b>).</p>
Full article ">
24 pages, 7419 KiB  
Article
Description of Four Novel Species in Pleosporales Associated with Coffee in Yunnan, China
by Li Lu, Samantha C. Karunarathna, Dong-Qin Dai, Yin-Ru Xiong, Nakarin Suwannarach, Steven L. Stephenson, Abdallah M. Elgorban, Salim Al-Rejaie, Ruvishika S. Jayawardena and Saowaluck Tibpromma
J. Fungi 2022, 8(10), 1113; https://doi.org/10.3390/jof8101113 - 21 Oct 2022
Cited by 17 | Viewed by 2843
Abstract
In Yunnan Province, the coffee-growing regions are mainly distributed in Pu’er and Xishuangbanna. During the surveys of microfungi associated with coffee in Yunnan Province, seven taxa were isolated from coffee samples. Based on molecular phylogenetic analyses of combined ITS, LSU, SSU, rpb2, [...] Read more.
In Yunnan Province, the coffee-growing regions are mainly distributed in Pu’er and Xishuangbanna. During the surveys of microfungi associated with coffee in Yunnan Province, seven taxa were isolated from coffee samples. Based on molecular phylogenetic analyses of combined ITS, LSU, SSU, rpb2, and tef1-α sequence data and morphological characteristics, four new species viz. Deniquelata yunnanensis, Paraconiothyrium yunnanensis, Pseudocoleophoma puerensis, and Pse. yunnanensis, and three new records viz. Austropleospora keteleeriae, Montagnula thailandica, and Xenocamarosporium acaciae in Pleosporales are introduced. In addition, Paracamarosporium fungicola was transferred back to Paraconiothyrium based on taxonomy and DNA sequences. Full descriptions, illustrations, and phylogenetic trees to show the placement of new and known taxa are provided. In addition, the morphological comparisons of new taxa with closely related taxa are given. Full article
(This article belongs to the Special Issue Polyphasic Identification of Fungi 2.0)
Show Figures

Figure 1

Figure 1
<p>The best-scoring RAxML tree was constructed from a concatenated SSU, LSU, ITS, <span class="html-italic">rpb</span>2, and <span class="html-italic">tef</span>1-α dataset. The tree is rooted with <span class="html-italic">Periconia pseudodigitata</span> (KT1395, KT1195A). Nodes were annotated if the maximum likelihood bootstrap support value was ≥ 60% (ML, left) or if the Bayesian posterior probability was ≥ 0.90 (BYPP, right). The newly described species are in red, new records are in green, and type strains are in bold. Red stars are used to indicate the two species that have uncertain placements and are discussed in the discussion.</p>
Full article ">Figure 1 Cont.
<p>The best-scoring RAxML tree was constructed from a concatenated SSU, LSU, ITS, <span class="html-italic">rpb</span>2, and <span class="html-italic">tef</span>1-α dataset. The tree is rooted with <span class="html-italic">Periconia pseudodigitata</span> (KT1395, KT1195A). Nodes were annotated if the maximum likelihood bootstrap support value was ≥ 60% (ML, left) or if the Bayesian posterior probability was ≥ 0.90 (BYPP, right). The newly described species are in red, new records are in green, and type strains are in bold. Red stars are used to indicate the two species that have uncertain placements and are discussed in the discussion.</p>
Full article ">Figure 2
<p><span class="html-italic">Pseudocoleophoma puerensis</span> (ZHKU 22-0118). (<b>a</b>,<b>b</b>) Ascomata on a decayed branch of <span class="html-italic">Coffea arabica</span>; (<b>c</b>) section of an ascoma; (<b>d</b>) peridium at side; (<b>e</b>) pseudoparaphyses; (<b>f</b>–<b>j</b>) immature and mature asci (arrows indicate the club-like pedicel); (<b>k</b>–<b>n</b>) ascospores; (<b>o</b>) germinated ascospore; (<b>p</b>) culture on PDA from above and reverse (60 days). Scale bars: (<b>c</b>,<b>e</b>) = 100 µm; (<b>f</b>–<b>j</b>) = 20 µm; (<b>k</b>–<b>n</b>) = 5 µm; (<b>o</b>) = 10 µm.</p>
Full article ">Figure 3
<p><span class="html-italic">Pseudocoleophoma yunnanensis</span> (ZHKU 22-0116). (<b>a</b>,<b>b</b>) Ascomata on a decayed branch of <span class="html-italic">Coffea</span> sp.; (<b>c</b>) section of an ascoma; (<b>d</b>) peridium at side; (<b>e</b>) pseudoparaphyses; (<b>f</b>–<b>i</b>) immature and mature asci (arrows indicate the club-shape pedicel); (<b>j</b>) germinated ascospore; (<b>k</b>–<b>m</b>) ascospores; (<b>n</b>,<b>o</b>) ascospore stained with Lugol’s iodine; (<b>p</b>,<b>q</b>) ascospore stained with Indian ink; (<b>r</b>) culture on PDA from above and reverse (60 days). Scale bars: (<b>c</b>) = 100 µm; (<b>d</b>–<b>j</b>) = 20 µm; (<b>k</b>–<b>q</b>) = 10 µm.</p>
Full article ">Figure 4
<p><span class="html-italic">Austropleospora</span><span class="html-italic">keteleeriae</span> (ZHKU 22-0120). (<b>a</b>,<b>b</b>) Appearance of the conidiomata on decaying branch of <span class="html-italic">Coffea</span> <span class="html-italic">arabica</span> var. <span class="html-italic">catimor</span>; (<b>c</b>) cross sections of a conidioma; (<b>d</b>) pycnidial wall; (<b>e</b>,<b>f</b>) conidia attached to conidiogenous cells; (<b>g</b>–<b>j</b>) conidia; (<b>k</b>) germinated conidium; (<b>l</b>) culture on PDA from above and reverse (60 days); scale bars: (<b>c</b>) = 100 µm; (<b>d</b>,<b>e</b>) = 20 µm; (<b>f</b>,<b>k</b>) = 10 µm; (<b>g</b>–<b>j</b>) = 5 µm.</p>
Full article ">Figure 5
<p><span class="html-italic">Deniquelata yunnanensis</span> (ZHKU 22-0115). (<b>a</b>,<b>b</b>) Ascomata on a decaying branch of <span class="html-italic">Coffea</span> sp.; (<b>c</b>) vertical section of an ascoma; (<b>d</b>) peridium; (<b>e</b>) pseudoparaphyses; (<b>f</b>–<b>k</b>) asci (arrows indicate the short furcate pedicel); (<b>l</b>–<b>n</b>) ascospores; (<b>o</b>) germinated ascospore; (<b>p</b>) culture on PDA from above and reverse (30 days). Scale bars: (<b>c</b>) = 100 µm; (<b>d</b>–<b>k</b>) = 15 µm; (<b>l</b>–<b>o</b>) = 5 µm.</p>
Full article ">Figure 6
<p><span class="html-italic">Montagnula thailandica</span> (ZHKU 22-0119). (<b>a</b>,<b>b</b>) Ascomata on a decaying branch of <span class="html-italic">Coffea</span> <span class="html-italic">arabica</span>; (<b>c</b>) vertical section of an ascoma; (<b>d</b>) peridium; (<b>e</b>) pseudoparaphyses; (<b>f</b>–<b>j</b>) asci (arrows indicate the club-shape pedicel); (<b>k</b>–<b>n</b>) ascospores; (<b>o</b>) germinated ascospore; (<b>p</b>) culture on PDA from above and reverse (30 days). Scale bars: (<b>c</b>) = 100 µm; (<b>d</b>–<b>j</b>) = 30 µm; (<b>k</b>–<b>n</b>) = 5 µm; (<b>o</b>) = 10 µm.</p>
Full article ">Figure 7
<p><span class="html-italic">Paraconiothyrium yunnanensis</span> (ZHKU 22-0114). (<b>a</b>,<b>b</b>) Ascomata on a decaying branch of <span class="html-italic">Coffea</span> sp.; (<b>c</b>) vertical section of an ascoma; (<b>d</b>) peridium; (<b>e</b>) pseudoparaphyses; (<b>f</b>–<b>i</b>) asci (arrows indicate the club-like pedicel); (<b>j</b>) germinated ascospore; (<b>k</b>–<b>o</b>) ascospores; (<b>p</b>) culture on PDA from above and reverse (30 days). Scale bars: (<b>c</b>) = 100 µm; (<b>d</b>–<b>i</b>) = 15 µm; (<b>j</b>–<b>o</b>) = 10 µm.</p>
Full article ">Figure 8
<p><span class="html-italic">Xenocamarosporium acaciae</span> (ZHKU 22-0117). (<b>a</b>,<b>b</b>) Appearance of ascomata on a decaying branch of <span class="html-italic">Coffea</span> sp.; (<b>c</b>,<b>d</b>) vertical section of an ascoma; (<b>e</b>) peridium; (<b>f</b>) pseudoparaphyses; (<b>g</b>–<b>j</b>) asci; (<b>k</b>) germinated ascospore; (<b>l</b>–<b>p</b>) ascospores; (<b>q</b>) culture on PDA from above and reverse (60 days). Scale bars: (<b>d</b>) = 100 µm; (<b>e</b>–<b>f</b>) = 15 µm; (<b>g</b>–<b>k</b>) = 20 µm; (<b>l</b>–<b>p</b>) = 5 µm.</p>
Full article ">
22 pages, 3895 KiB  
Article
Principal Drivers of Fungal Communities Associated with Needles, Shoots, Roots and Adjacent Soil of Pinus sylvestris
by Diana Marčiulynienė, Adas Marčiulynas, Valeriia Mishcherikova, Jūratė Lynikienė, Artūras Gedminas, Iva Franic and Audrius Menkis
J. Fungi 2022, 8(10), 1112; https://doi.org/10.3390/jof8101112 - 21 Oct 2022
Cited by 6 | Viewed by 2412
Abstract
The plant- and soil-associated microbial communities are critical to plant health and their resilience to stressors, such as drought, pathogens, and pest outbreaks. A better understanding of the structure of microbial communities and how they are affected by different environmental factors is needed [...] Read more.
The plant- and soil-associated microbial communities are critical to plant health and their resilience to stressors, such as drought, pathogens, and pest outbreaks. A better understanding of the structure of microbial communities and how they are affected by different environmental factors is needed to predict and manage ecosystem responses to climate change. In this study, we carried out a country-wide analysis of fungal communities associated with Pinus sylvestris growing under different environmental conditions. Needle, shoot, root, mineral, and organic soil samples were collected at 30 sites. By interconnecting the high-throughput sequencing data, environmental variables, and soil chemical properties, we were able to identify key factors that drive the diversity and composition of fungal communities associated with P. sylvestris. The fungal species richness and community composition were also found to be highly dependent on the site and the substrate they colonize. The results demonstrated that different functional tissues and the rhizosphere soil of P. sylvestris are associated with diverse fungal communities, which are driven by a combination of climatic (temperature and precipitation) and edaphic factors (soil pH), and stand characteristics. Full article
(This article belongs to the Special Issue Fungal Diversity in Europe)
Show Figures

Figure 1

Figure 1
<p>Map of Lithuania showing the sampling sites in <span class="html-italic">Pinus sylvestris</span> forest stands. The green color on the map shows the distribution of <span class="html-italic">P. sylvestris</span> stands and its abundance (%) in different areas.</p>
Full article ">Figure 2
<p>Differences in the Shannon diversity index of fungal OTUs among different substrates. In the Mann–Whitney test, <span class="html-italic">p</span>-values were: roots vs. shoots <span class="html-italic">p</span> &lt; 0.05, roots vs. needles <span class="html-italic">p</span> &lt; 0.05, roots vs. organic soil <span class="html-italic">p</span> &lt; 0.05, root vs. mineral soil <span class="html-italic">p</span> &lt; 0.05, shoot vs. needles <span class="html-italic">p</span> = 0.97, shoots vs. organic soil <span class="html-italic">p</span> = 0.06, shoots vs. mineral soil <span class="html-italic">p</span> = 0.23, needles vs. organic soil <span class="html-italic">p</span> = 0.06, needles vs. mineral soil <span class="html-italic">p</span> = 0.06, and organic soil vs. mineral soil <span class="html-italic">p</span> = 0.99.</p>
Full article ">Figure 3
<p>Venn diagram showing the diversity and overlap of fungal taxa in different sample types from <span class="html-italic">P. sylvestris</span> stands. Different colors represent different substrates: Yellow—Soil (organic and mineral combined), Orange—Needles, Blue—Shoots, and Gray—Roots.</p>
Full article ">Figure 4
<p>Species accumulation curves showing the relationship between the cumulative number of fungal OTUs and the number of ITS rRNA sequences shown as Yellow—Soil, Orange—Needles, Blue—Shoots, and Gray—Roots.</p>
Full article ">Figure 5
<p>Relative abundance of fungal classes in needles, shoots, roots, and the soil of <span class="html-italic">Pinus sylvestris</span>. Other represented fungal classes with a relative abundance of &lt;1%. The data from the 30 different sample sites are combined.</p>
Full article ">Figure 6
<p>Fungal functional groups in the needle, shoot, root, and soil samples from the 30 sampling sites of <span class="html-italic">Pinus sylvestris</span> are shown as a relative abundance (%) of the total number of fungal taxa.</p>
Full article ">Figure 7
<p>The relationship between the species richness of fungal OTUs in different substrates (needles, shoots, roots, mineral, and organic soil) of <span class="html-italic">Pinus sylvestris</span> and different stand (<b>A</b>,<b>B</b>) and climatic (<b>C</b>,<b>D</b>) parameters. The semitransparent field around each curve denotes the size of the deviation from the mean value.</p>
Full article ">Figure 8
<p>The relationship between the species richness of fungal OTUs in the different substrates (needles, shoots, roots, mineral, and organic soil) of <span class="html-italic">Pinus sylvestris</span> and different soil parameters (<b>A</b>–<b>D</b>). The semitransparent field around each curve denotes the size of the deviation from the mean value.</p>
Full article ">Figure 9
<p>Non-metric multidimensional scaling (NMDS) of fungal communities associated with different substrates (N—needles, R—roots, Sh—shoots, Sm—mineral soil, and So—organic soil) of <span class="html-italic">Pinus sylvestris</span>.</p>
Full article ">
10 pages, 286 KiB  
Review
COVID-19 Associated with Cryptococcosis: A New Challenge during the Pandemic
by Khee-Siang Chan, Chih-Cheng Lai, Wen-Liang Yu and Chien-Ming Chao
J. Fungi 2022, 8(10), 1111; https://doi.org/10.3390/jof8101111 - 21 Oct 2022
Cited by 7 | Viewed by 2569
Abstract
Coronavirus disease 2019 (COVID-19) caused by severe acute respiratory syndrome coronavirus 2 (SARS-CoV-2) is a great threat to global health. In addition to SARS-CoV-2 itself, clinicians should be alert to the possible occurrence of co-infection or secondary infection among patients with COVID-19. The [...] Read more.
Coronavirus disease 2019 (COVID-19) caused by severe acute respiratory syndrome coronavirus 2 (SARS-CoV-2) is a great threat to global health. In addition to SARS-CoV-2 itself, clinicians should be alert to the possible occurrence of co-infection or secondary infection among patients with COVID-19. The possible co-pathogens include bacteria, viruses, and fungi, but COVID-19-associated cryptococcosis is rarely reported. This review provided updated and comprehensive information about this rare clinical entity of COVID-19-associated cryptococcosis. Through an updated literature search till 23 August 2022, we identified a total of 18 culture-confirmed case reports with detailed information. Half (n = 9) of them were elderly. Fifteen (83.3%) of them had severe COVID-19 and ever received systemic corticosteroid. Disseminated infection with cryptococcemia was the most common type of cryptococcosis, followed by pulmonary and meningitis. Except one case of C. laurentii, all other cases are by C. neoformans. Liposomal amphotericin B and fluconazole were the most commonly used antifungal agents. The overall mortality was 61.1% (11/18) and four of them did not receive antifungal agents before death. Improving the poor outcome requires a physician’s high suspicion, early diagnosis, and prompt treatment. Full article
(This article belongs to the Special Issue Fungal Infections in COVID-19 Patients)
16 pages, 3743 KiB  
Article
Roles of BrlA and AbaA in Mediating Asexual and Insect Pathogenic Lifecycles of Metarhizium robertsii
by Jin-Guan Zhang, Si-Yuan Xu, Sheng-Hua Ying and Ming-Guang Feng
J. Fungi 2022, 8(10), 1110; https://doi.org/10.3390/jof8101110 - 21 Oct 2022
Cited by 11 | Viewed by 1868
Abstract
BrlA and AbaA are key activators of the central developmental pathway (CDP) that controls asexual development in Aspergillus but their roles remain insufficiently understood in hypocerealean insect pathogens. Here, regulatory roles of BrlA and AbaA orthologs in Metarhizium robertsii (Clavicipitaceae) were characterized [...] Read more.
BrlA and AbaA are key activators of the central developmental pathway (CDP) that controls asexual development in Aspergillus but their roles remain insufficiently understood in hypocerealean insect pathogens. Here, regulatory roles of BrlA and AbaA orthologs in Metarhizium robertsii (Clavicipitaceae) were characterized for comparison to those elucidated previously in Beauveria bassiana (Cordycipitaceae) at phenotypic and transcriptomic levels. Time-course transcription profiles of brlA, abaA, and the other CDP activator gene wetA revealed that they were not so sequentially activated in M. robertsii as learned in Aspergillus. Aerial conidiation essential for fungal infection and dispersal, submerged blastospore production mimicking yeast-like budding proliferation in insect hemocoel, and insect pathogenicity via cuticular penetration were all abolished as a consequence of brlA or abaA disruption, which had little impact on normal hyphal growth. The disruptants were severely compromised in virulence via cuticle-bypassing infection (intrahemocoel injection) and differentially impaired in cellular tolerance to oxidative and cell wall-perturbing stresses. The ΔbrlA and ΔabaA mutant shad 255 and 233 dysregulated genes (up/down ratios: 52:203 and 101:122) respectively, including 108 genes co-dysregulated. These counts were small compared with 1513 and 2869 dysregulated genes (up/down ratios: 707:806 and 1513:1356) identified in ΔbrlA and ΔabaA mutants of B. bassiana. Results revealed not only conserved roles for BrlA and AbaA in asexual developmental control but also their indispensable roles in fungal adaptation to the insect-pathogenic lifecycle and host habitats. Intriguingly, BrlA- or AbaA-controlled gene expression networks are largely different between the two insect pathogens, in which similar phenotypes were compromised in the absence of either brlA or abaA. Full article
(This article belongs to the Section Fungi in Agriculture and Biotechnology)
Show Figures

Figure 1

Figure 1
<p>Domain architectures, transcriptional profiles, and subcellular localization of BrlA and AbaA in <span class="html-italic">M. robertsii</span> (<span class="html-italic">Mr</span>). (<b>A</b>) Comparison of conserved domains and NLS motif predicted from amino acid sequences of BrlA and AbaA orthologs. <span class="html-italic">An</span>, <span class="html-italic">A. nidulans</span>. <span class="html-italic">Bb</span>, <span class="html-italic">B. bassiana</span>. (<b>B</b>) Relative transcript (RT) levels of three CDP genes in the <span class="html-italic">Mr</span> WT strain during a 7-day incubation on PDA at an optimal regime with respect to a standard on day 2. Error bars: standard deviations (SDs) of the means from three independent cDNA samples analyzed via qPCR. (<b>C</b>) LSCM images (scale bar: 5 μm) for subcellular localization of the BrlA-GFP and AbaA-GFP fusion proteins expressed in the WT strain. Images 1, 2, 3, and 4 are bright, expressed, DAPI-stained, and merged views of the same field, respectively. Hyphal nuclei are indicated by arrows.</p>
Full article ">Figure 2
<p>Impact of <span class="html-italic">brlA</span> or <span class="html-italic">abaA</span> disruption on radial growth and stress response of <span class="html-italic">M. robertsii</span>. (<b>A</b>,<b>B</b>) Images and diameters of fungal colonies grown at the optimal regime of 25 °C and L:D 12:12 for 7 days on nutrition-rich (PDA, SDAY, and 1/4 SDAY) and minimal (CDA) media. (<b>C</b>,<b>D</b>) Diameters of 7-day-old fungal colonies grown at the optimal regime on CDA amended with different amino acids as nitrogen sources. (<b>E</b>) Relative growth inhibition (RGI) percentages of fungal colonies grown at 25 °C for 7 days on CDA plates supplemented with H<sub>2</sub>O<sub>2</sub> (2 mM), menadione (MND, 0.03 mM), calcofluor white (CFW, 15 μg/mL), and Congo red (CGR, 1 mg/mL) respectively. All colonies were initiated with hyphal mass discs (<span class="html-italic">ϕ</span> = 5 mm) attached to plates. <span class="html-italic">p</span> &lt; 0.05 * or 0.001 *** in Tukey’s HSD tests (ns, no significance). Error bars: SDs from three independent replicates.</p>
Full article ">Figure 3
<p>Indispensability of either <span class="html-italic">brlA</span> or <span class="html-italic">abaA</span> for asexual spore production of <span class="html-italic">M. robertsii</span>. (<b>A</b>,<b>B</b>) Microscopic images (scale bar: 10 μm) of culture samples stained with calcofluor white after collected from 5-day-old cultures and images of 15-day-old cultures grown on PDA at the optimal regime of 25 °C and L:D 12:12. (<b>C</b>,<b>D</b>) Biomass levels and conidial yields of fungal cultures assessed during a 15-day incubation on PDA at the optimal regime. Each culture was initiated by spreading 100 μL of a fresh hyphal 50 mg/mL suspension. (<b>E</b>) Blastospore yields measured from the 3-day-old SDBY cultures initiated with fresh hyphal mass 1 mg/mL. * <span class="html-italic">p</span> &lt; 0.05 in Tukey’s HSD tests. Error bars: SDs from three replicates.</p>
Full article ">Figure 4
<p>Indispensability of <span class="html-italic">brlA</span> or <span class="html-italic">AbaA</span> for insect-pathogenic lifecycle of <span class="html-italic">M. robertsii</span>. (<b>A</b>) Survival trends of <span class="html-italic">G. mellonella</span> larvaeafter immersion in a fresh hyphal 100 mg/mL suspension for normal cuticle infection (NCI) and intrahemocoel injection of 5 μL fresh hyphal (10 mg/mL) suspension per larva for cuticle-bypassing infection (CBI). (<b>B</b>) LT<sub>50</sub> values estimated from time-mortality trends. Triangles indicate an LT<sub>50</sub> not accessible for Δ<span class="html-italic">brlA</span> and Δ<span class="html-italic">abaA</span> mutants. (<b>C</b>) Total activities of cuticle degrading enzymes (ECEs and Pr1 proteases) and biomass levels assessed from the 3-day-old submerged cultures generated by shaking incubation of a fresh hyphal 1 mg/mL suspension in CDB-BSA. (<b>D</b>) Microscopic images (scale bar: 20 μm) for status and abundance of fungal budding cells (marked with red stars) and insect hemocytes (HC) in hemolymph samples, which were taken from surviving larvae 6 days post-NCI and 3 days post-CBI and incubated at 25 °C for 36 h in SDBY. Error bars: SDs from three independent replicates.</p>
Full article ">Figure 5
<p>Effect of <span class="html-italic">brlA</span> or <span class="html-italic">abaA</span> disruption on gene expression networks of <span class="html-italic">M. robertsii</span>. (<b>A</b>) Counts of genes dysregulated and co-dysregulated in the Δ<span class="html-italic">brlA</span> and Δ<span class="html-italic">abaA</span>-1 mutants versus the WT strain. (<b>B</b>,<b>C</b>) Counts of dysregulated genes significantly enriched (<span class="html-italic">p</span> &lt; 0.05) to GO terms of three function classes in Δ<span class="html-italic">brlA</span> and Δ<span class="html-italic">abaA</span>-1, respectively. (<b>D</b>) Counts of dysregulaated genes significantly enriched (<span class="html-italic">p</span> &lt; 0.05) to KEGG pathways in the two mutants. The transcriptome was constructed based on three 4-day-old cultures (replicates) of the Δ<span class="html-italic">brlA</span>, Δ<span class="html-italic">abaA</span>-1,and WT strains. All dysregulated genes were identified at the significant levels of log<sub>2</sub> ratio (fold change) ≤−1 (downregulated, DnR) or ≥1 (upregulated, UpR) and <span class="html-italic">q</span> &lt; 0.05.</p>
Full article ">
2 pages, 215 KiB  
Editorial
Fungal Pigments: More Insights from Colorful Fungi
by Laurent Dufossé
J. Fungi 2022, 8(10), 1109; https://doi.org/10.3390/jof8101109 - 20 Oct 2022
Viewed by 1447
Abstract
Following the previous Journal of Fungi (ISSN 2309-608X) Fungal Pigments Special Issue edited and published in 2017 (weblink https://www [...] Full article
(This article belongs to the Special Issue Fungal Pigments 2021)
12 pages, 2010 KiB  
Article
Histoplasma capsulatum Activates Hematopoietic Stem Cells and Their Progenitors through a Mechanism Dependent on TLR2, TLR4, and Dectin-1
by Carolina Rodríguez-Echeverri, Beatriz L. Gómez and Ángel González
J. Fungi 2022, 8(10), 1108; https://doi.org/10.3390/jof8101108 - 20 Oct 2022
Cited by 4 | Viewed by 2008
Abstract
Hematopoietic stem cells (HSCs), a multipotent and self-renewing population responsible for the generation and maintenance of blood cells, have been the subject of numerous investigations due to their therapeutic potential. It has been shown that these cells are able to interact with pathogens [...] Read more.
Hematopoietic stem cells (HSCs), a multipotent and self-renewing population responsible for the generation and maintenance of blood cells, have been the subject of numerous investigations due to their therapeutic potential. It has been shown that these cells are able to interact with pathogens through the TLRs that they express on their surface, affecting the hematopoiesis process. However, the interaction between hematopoietic stem and progenitor cells (HSPC) with fungal pathogens such as Histoplasma capsulatum has not been studied. Therefore, the objective of the present study was to determine if the interaction of HSPCs with H. capsulatum yeasts affects the hematopoiesis, activation, or proliferation of these cells. The results indicate that HSPCs are able to adhere to and internalize H. capsulatum yeasts through a mechanism dependent on TLR2, TLR4, and Dectin-1; however, this process does not affect the survival of the fungus, and, on the contrary, such interaction induces a significant increase in the expression of IL-1β, IL-6, IL-10, IL-17, TNF-α, and TGF-β, as well as the immune mediators Arg-1 and iNOS. Moreover, H. capsulatum induces apoptosis and alters HSPC proliferation. These findings suggest that H. capsulatum directly modulates the immune response exerted by HPSC through PRRs, and this interaction could directly affect the process of hematopoiesis, a fact that could explain clinical manifestations such as anemia and pancytopenia in patients with severe histoplasmosis, especially in those with fungal spread to the bone marrow. Full article
(This article belongs to the Special Issue New Insights into Current Understanding of Host–Fungal Interactions)
Show Figures

Figure 1

Figure 1
<p>Characterization of the HSPC population. Expression of CD105<sup>+</sup> and Sca-1<sup>+</sup> surface antigens on mouse bone marrow long-term culture initiator cells (LT-CIC or LT-HSC). (<b>A</b>) Control of Lin<sup>−</sup> cells without positive selection; (<b>B</b>) CD105<sup>+</sup> Sca-1<sup>+</sup> cells corresponding to HSPC.</p>
Full article ">Figure 2
<p>Expression of TLR2, TLR4 and Dectin-1 receptors in HSPC. (<b>A</b>,<b>D</b>) TLR2 expression in unstimulated and stimulated HSPC with <span class="html-italic">H. capsulatum</span> yeasts; (<b>B</b>,<b>E</b>) TLR4 expression in unstimulated and stimulated HSPC with <span class="html-italic">H. capsulatum</span> yeasts; (<b>C</b>,<b>F</b>) Expression of Dectin-1 in unstimulated and stimulated HSPC with <span class="html-italic">H. capsulatum</span> yeasts. HSPCs were selected from the CD105<sup>+</sup>/Sca-1<sup>+</sup> population and the level of expression of receptors is expressed as mean fluorescence intensity (MFI). (<b>A</b>–<b>D</b>) Results are expressed as means ± SD of pooled data from three independent experiments. (<b>D</b>–<b>F</b>) Data represent the percentage of HSPC positive cells and are from a representative experiment of three replicates, ** <span class="html-italic">p</span> &lt; 0.01; *** <span class="html-italic">p</span> &lt; 0.0001.</p>
Full article ">Figure 3
<p>Phagocytosis of <span class="html-italic">H. capsulatum</span> yeasts by HSPC. Phagocytosis was analyzed by flow cytometry and the result is expressed as the percentage of FITC positive cells (% phagocytosis) of HSPC cells infected with <span class="html-italic">H. capsulatum</span> yeasts. (<b>A</b>) Control, HSPC (<b>B</b>) Control, percentage of phagocytosis in HSPC co-cultured with <span class="html-italic">H. capsulatum</span> and without treatment; (<b>C</b>) percentage of phagocytosis in HSPC treated with anti-TLR2; (<b>D</b>) percentage of phagocytosis in HSPC treated with anti-TLR4; (<b>E</b>) percentage of phagocytosis in HSPC treated with the peptide CLEC7A; (<b>F</b>) percentage of phagocytosis in HSPC treated with the anti-TLR2/anti-TLR4 combination; (<b>G</b>) percentage of phagocytosis in HSPC treated with the anti-TLR2/CLEC7A combination; (<b>H</b>) percentage of phagocytosis in HSPC treated with the combination of anti-TLR4/CLEC7A, and (<b>I</b>) percentage of phagocytosis in HSPC treated with the combination of anti-TLR2/anti-TLR4/CLEC7A. Histograms in gray correspond to HSPC previously treated with blocking antibodies for TLR or with a specific peptide for Dectin-1 (CLEC7A). Data are representative from an experiment of three replicates; *** <span class="html-italic">p</span> &lt; 0.0001.</p>
Full article ">Figure 4
<p>Fungicidal activity of HSPC against <span class="html-italic">H. capsulatum</span>. Colony forming units (CFUs) were recovered from HPSC infected with <span class="html-italic">H. capsulatum</span> yeasts after incubation for 24 h at 37 °C. Results are expressed as median and IQR of pooled data from three independent experiments.</p>
Full article ">Figure 5
<p>Expression of cytokines and inflammatory mediators in HSPC stimulated with <span class="html-italic">H. capsulatum</span> yeast. Analysis of mRNA expression of proinflammatory cytokines, arginase-1, and iNOS in HSPC stimulated or not stimulated with <span class="html-italic">H. capsulatum</span> yeast. (<b>A</b>) IL-6; (<b>B</b>) IL-17; (<b>C</b>) IL-1β; (<b>D</b>) IL-10; (<b>E</b>) TNF-α; (<b>F</b>) TGF-β1; (<b>G</b>) Arg-1; and (<b>H</b>) iNOS. HSPC, control, unstimulated cells; HSPC + Hc, cells stimulated with <span class="html-italic">H. capsulatum</span>; TLR, <span class="html-italic">Toll</span>-like receptor; CLEC7A, peptide blocker specific for Dectin-1. Results are expressed as means ± SD of pooled data from three independent experiments; * <span class="html-italic">p</span> &lt; 0.0001, comparisons were done between HSPCs + <span class="html-italic">Hc</span> vs. HSPCs, and <sup>#</sup> <span class="html-italic">p</span> &lt; 0.0001 HSPCs + <span class="html-italic">Hc</span> plus the different treatments vs. HSPCs + <span class="html-italic">Hc</span>.</p>
Full article ">Figure 6
<p><span class="html-italic">Histoplasma capsulatum</span> induces apoptosis and necrosis in HSPC. HSPCs were treated with Annexin V-FITC and propidium iodide as described in materials and methods. (<b>A</b>) Control, uninfected HSPC; (<b>B</b>) HSPC stimulated with <span class="html-italic">H. capsulatum</span> yeasts. Percentages represent the number of cells positive for FITC and propidium iodide. Similar results were obtained from three independent experiments.</p>
Full article ">Figure 7
<p><span class="html-italic">H. capsulatum</span> yeasts affect HSPC proliferation. (<b>A</b>) Control, unstimulated HSPC; (<b>B</b>) HSPC + Pam3CysOH; (<b>C</b>) HSPC + LPS; <b>(D</b>) HSPC + β-glucan; (<b>E</b>) HSPC + Pam3CysOH + <span class="html-italic">H. capsulatum</span>; (<b>F</b>) HSPC + LPS + <span class="html-italic">H. capsulatum</span>; (<b>G</b>) HSPC + β-glucan + <span class="html-italic">H. capsulatum</span>. Data represent the percentage of BrdU positive cells; results are from a representative experiment of three replicates.</p>
Full article ">
14 pages, 2371 KiB  
Article
Epidemic Identification of Fungal Diseases in Morchella Cultivation across China
by Xiaofei Shi, Dong Liu, Xinhua He, Wei Liu and Fuqiang Yu
J. Fungi 2022, 8(10), 1107; https://doi.org/10.3390/jof8101107 - 20 Oct 2022
Cited by 20 | Viewed by 5549
Abstract
True morels (Morchella, Pezizales) are world-renowned edible mushrooms (ascocarps) that are widely demanded in international markets. Morchella has been successfully artificially cultivated since 2012 in China and is rapidly becoming a new edible mushroom industry occupying up to 16,466 hectares in [...] Read more.
True morels (Morchella, Pezizales) are world-renowned edible mushrooms (ascocarps) that are widely demanded in international markets. Morchella has been successfully artificially cultivated since 2012 in China and is rapidly becoming a new edible mushroom industry occupying up to 16,466 hectares in the 2021–2022 season. However, nearly 25% of the total cultivation area has annually suffered from fungal diseases. While a variety of morel pathogenic fungi have been reported their epidemic characteristics are unknown, particularly in regional or national scales. In this paper, ITS amplicon sequencing and microscopic examination were concurrently performed on the morel ascocarp lesions from 32 sites in 18 provinces across China. Results showed that Diploöspora longispora (75.48%), Clonostachys solani (5.04%), Mortierella gamsii (0.83%), Mortierella amoeboidea (0.37%) and Penicillium kongii (0.15%) were the putative pathogenic fungi. The long, oval, septate conidia of D. longispora was observed on all ascocarps. Oval asexual spores and sporogenic structures, such as those of Clonostachys, were also detected in C. solani infected samples with high ITS read abundance. Seven isolates of D. longispora were isolated from seven selected ascocarps lesions. The microscopic characteristics of pure cultures of these isolates were consistent with the morphological characteristics of ascocarps lesions. Diploöspora longispora had the highest amplification abundance in 93.75% of the samples, while C. solani had the highest amplification abundance in six biological samples (6.25%) of the remaining two sampling sites. The results demonstrate that D. longispora is a major culprit of morel fungal diseases. Other low-abundance non-host fungi appear to be saprophytic fungi infecting after D. longispora. This study provides data supporting the morphological and molecular identification and prevention of fungal diseases of morel ascocarps. Full article
(This article belongs to the Special Issue Edible and Medicinal Macrofungi)
Show Figures

Figure 1

Figure 1
<p>Sampling distribution of fungal disease samples from main morel cultivation farms across China.</p>
Full article ">Figure 2
<p>Morphological characteristics of fungal diseases of <span class="html-italic">Morchella</span>, Occurrences of fungal disease in morel ascocarps under outdoor cultivation. The diseased wilts and shrinks, and white hairy disease spots mostly occur on the cap, including the stipe (<b>A</b>–<b>E</b>); the fungal diseases lesion after 10 days of pathogen inoculation [<a href="#B31-jof-08-01107" class="html-bibr">31</a>] on the healthy ascocarp (<b>E</b>,<b>F</b>); morphological characteristics of pure cultured pathogen cultured in a CYM medium at 24 °C for 14 days, with irregular, white, villous colonies and irregular concentric rings (<b>G</b>); morphological characteristics of chlamydospores and asexual spores of the D1092 isolate, chlamydospores are spherical and locate in the posterior wall, and asexual spores are oval shaped with 0–3 septum (<b>H</b>,<b>I</b>); morphological characteristics of pathogens in the focus of cap in D1012, D1095 and D1036 (<b>J</b>–<b>L</b>); the hollow arrow refers to the large elliptic ascospore of the host morel, the solid arrow refers to the long rod-shaped septate conida of <span class="html-italic">Diploöspora longispora</span> (<b>H</b>–<b>K</b>), the hollow triangle refers to the elliptic conida of <span class="html-italic">Clonostachys</span> (L), and the solid triangle refers to the conidiophore of <span class="html-italic">Clonostachys</span> (L).</p>
Full article ">Figure 3
<p>The number of the fungal OTUs (<b>A</b>) presented in the samples infected by morel diseases, (<b>B</b>) differences in community compositions as indicated by principial component analysis, and (<b>C</b>) rarefaction curves which can be used to compare species richness in samples with different amounts of sequencing data.</p>
Full article ">Figure 4
<p>Molecular phylogenetic analysis by maximum likelihood method using ITS sequence of seven fungal lesion isolates. The bootstrap consensus tree inferred from 1000 replicates was taken to represent the evolutionary history of the taxa analyzed. New sequences generated from this study in bold, and the GenBank accession number was followed by species name.</p>
Full article ">Figure 5
<p>Variations in abundance or occurrence of putative pathogens in <span class="html-italic">Morchella</span> lesions across the morel cultivation across China.</p>
Full article ">Figure 6
<p>Microbial community composition (<b>A</b>) and abundance distribution (<b>B</b>) of fungi pathogens in 32 fungal lesions of <span class="html-italic">Morchella</span>.</p>
Full article ">
15 pages, 284 KiB  
Article
EUCAST Ibrexafungerp MICs and Wild-Type Upper Limits for Contemporary Danish Yeast Isolates
by Karin M. Jørgensen, Karen M. T. Astvad, Rasmus K. Hare and Maiken C. Arendrup
J. Fungi 2022, 8(10), 1106; https://doi.org/10.3390/jof8101106 - 20 Oct 2022
Cited by 4 | Viewed by 2164
Abstract
Ibrexafungerp is a novel triterpenoid antifungal that inhibits glucan synthase and thus fungal cell wall synthesis. We examined the in vitro activity against contemporary clinical yeast, investigated inter-laboratory and intra-laboratory variability, suggested wild-type upper-limit values (WT-UL), and compared in vitro activity of ibrexafungerp [...] Read more.
Ibrexafungerp is a novel triterpenoid antifungal that inhibits glucan synthase and thus fungal cell wall synthesis. We examined the in vitro activity against contemporary clinical yeast, investigated inter-laboratory and intra-laboratory variability, suggested wild-type upper-limit values (WT-UL), and compared in vitro activity of ibrexafungerp to five licensed antifungals. Susceptibility to ibrexafungerp and comparators was investigated prospectively for 1965 isolates (11,790 MICs) and repetitively for three QC strains (1764 MICs) following the EUCAST E.Def 7.3.2 method. Elevated ibrexafungerp/echinocandin MICs prompted FKS sequencing. Published ibrexafungerp EUCAST MIC-distributions were retrieved and aggregated for WT-UL determinations following EUCAST principles. Ibrexafungerp MICs were ≤2 mg/L except against C. pararugosa, Cryptococcus and some rare yeasts. Modal MICs (mg/L) were 0.06/0.125/0.25/0.5/0.5/0.5/0.5/1/2 for C. albicans/C. dubliniensis/C. glabrata/C. krusei/C. parapsilosis/C. tropicalis/S. cerevisiae/C. guilliermondii/C. lusitaniae and aligned within ±1 dilution with published values. The MIC ranges for QC strains were: 0.06–0.25/0.5–1/0.125–0.5 for CNM-CL-F8555/ATCC6258/ATCC22019. The WT-UL (mg/L) were: 0.25/0.5/1/1/2 for C. albicans/C. glabrata/C. krusei/C. parapsilosis/C. tropicalis. Adopting these, non-wild-type rates were 0.3%/0.6%/0%/8%/3% for C. albicans/C. glabrata/C. krusei/C. parapsilosis/C. tropicalis and overall lower than for comparators except amphotericin B. Five/six non-wild-type C. albicans/C. glabrata were echinocandin and Fks non-wild-type (F641S, F659del or F659L). Eight C. parapsilosis and three C. tropicalis non-wild-type isolates were echinocandin and Fks wild-type. Partial inhibition near 50% in the supra-MIC range may explain variable MICs. Ibrexafungerp EUCAST MIC testing is robust, although the significance of paradoxical growth for some species requires further investigation. The spectrum is broad and will provide an oral option for the growing population with azole refractory infection. Full article
(This article belongs to the Special Issue Antifungal Resistance 2.0)
25 pages, 3122 KiB  
Article
Assessing the Various Antagonistic Mechanisms of Trichoderma Strains against the Brown Root Rot Pathogen Pyrrhoderma noxium Infecting Heritage Fig Trees
by Harrchun Panchalingam, Daniel Powell, Cherrihan Adra, Keith Foster, Russell Tomlin, Bonnie L. Quigley, Sharon Nyari, R. Andrew Hayes, Alison Shapcott and D. İpek Kurtböke
J. Fungi 2022, 8(10), 1105; https://doi.org/10.3390/jof8101105 - 19 Oct 2022
Cited by 14 | Viewed by 3532
Abstract
A wide range of phytopathogenic fungi exist causing various plant diseases, which can lead to devastating economic, environmental, and social impacts on a global scale. One such fungus is Pyrrhoderma noxium, causing brown root rot disease in over 200 plant species of [...] Read more.
A wide range of phytopathogenic fungi exist causing various plant diseases, which can lead to devastating economic, environmental, and social impacts on a global scale. One such fungus is Pyrrhoderma noxium, causing brown root rot disease in over 200 plant species of a variety of life forms mostly in the tropical and subtropical regions of the globe. The aim of this study was to discover the antagonistic abilities of two Trichoderma strains (#5001 and #5029) found to be closely related to Trichoderma reesei against P. noxium. The mycoparasitic mechanism of these Trichoderma strains against P. noxium involved coiling around the hyphae of the pathogen and producing appressorium like structures. Furthermore, a gene expression study identified an induced expression of the biological control activity associated genes in Trichoderma strains during the interaction with the pathogen. In addition, volatile and diffusible antifungal compounds produced by the Trichoderma strains were also effective in inhibiting the growth of the pathogen. The ability to produce Indole-3-acetic acid (IAA), siderophores and the volatile compounds related to plant growth promotion were also identified as added benefits to the performance of these Trichoderma strains as biological control agents. Overall, these results show promise for the possibility of using the Trichoderma strains as potential biological control agents to protect P. noxium infected trees as well as preventing new infections. Full article
(This article belongs to the Special Issue Advances in Trichoderma-Plant Beneficial Interactions)
Show Figures

Figure 1

Figure 1
<p>Maximum-likelihood phylogenetic tree of 15 <span class="html-italic">Trichoderma</span> strains constructed using 3059 single-copy genes present in all the analysed genomes, using the maximum-likelihood method and with <span class="html-italic">A. niger</span> as the outgroup.</p>
Full article ">Figure 2
<p>Mycoparasitism by the <span class="html-italic">Trichoderma</span> strain #5001 (T) on <span class="html-italic">P. noxium</span> strain B (P). (<b>a</b>) <span class="html-italic">Trichoderma coiled</span> around hyphae of <span class="html-italic">P. noxium</span> strain B (X 400), (<b>b</b>) appressorium-like structure (X 400) and (<b>c</b>) <span class="html-italic">Trichoderma</span> growth over <span class="html-italic">P. noxium</span> strain B hyphae (X 400). The red and black arrows indicate the <span class="html-italic">Trichoderma</span> and <span class="html-italic">P. noxium</span>, respectively. Circles indicate the coiling, appressorium-like structures and <span class="html-italic">Trichoderma</span> overgrowth on <span class="html-italic">P. noxium</span>.</p>
Full article ">Figure 3
<p>Growth inhibition % of <span class="html-italic">P. noxium</span> (strains: A, B, C, D, E and F) (<b>a</b>) during the confrontation with <span class="html-italic">Trichoderma</span> strains #5001 and #5029 in dual culture set up; (<b>b</b>) by the volatile compounds of <span class="html-italic">Trichoderma</span> strains #5001 and #5029; (<b>c</b>) growth inhibition % of <span class="html-italic">P. noxium</span> strain B by the crude extracts of the <span class="html-italic">Trichoderma</span> strains #5029 and #5001 at different concentrations. Different letters with the same color on the top of SE bars indicate significant differences in growth inhibition observed between <span class="html-italic">P. noxium</span> strains caused by the particular <span class="html-italic">Trichoderma</span> strain based on Tukey’s test, <span class="html-italic">p</span> &lt; 0.05.</p>
Full article ">Figure 4
<p>Overlay plot of (<b>a</b>) pure culture of strain #5001 on top and media control at bottom; (<b>b</b>) co−culture of strain #5001 with <span class="html-italic">P. noxium</span> on top and pure culture of <span class="html-italic">P. noxium</span> strain at bottom. Numbers refer to compounds listed in <a href="#jof-08-01105-t002" class="html-table">Table 2</a>.</p>
Full article ">Figure 5
<p>Overlay plot of (<b>a</b>) pure culture of strain #5029 on top and media control at bottom; (<b>b</b>) co−culture of strain #5029 with <span class="html-italic">P. noxium</span> on top and pure culture of <span class="html-italic">P. noxium</span> strain at bottom. Numbers refer to compounds listed in <a href="#jof-08-01105-t002" class="html-table">Table 2</a>.</p>
Full article ">Figure 6
<p>Expression patterns of biocontrol-associated genes after confrontation with <span class="html-italic">P. noxium</span> strain B in (<b>a</b>,<b>c</b>) <span class="html-italic">Trichoderma</span> strain #5001 and (<b>b</b>,<b>d</b>) <span class="html-italic">Trichoderma</span> strain #5029. Elongation factor 1−α and β−actin genes (<span class="html-italic">tef</span> and <span class="html-italic">act</span>, respectively) were used as references to normalize biological control associated genes expression in <span class="html-italic">Trichoderma</span> strain #5001 and #5029. Asterisks indicate significant differences by Student’s <span class="html-italic">t</span>-test (<span class="html-italic">p</span> &lt; 0.05).</p>
Full article ">Figure 6 Cont.
<p>Expression patterns of biocontrol-associated genes after confrontation with <span class="html-italic">P. noxium</span> strain B in (<b>a</b>,<b>c</b>) <span class="html-italic">Trichoderma</span> strain #5001 and (<b>b</b>,<b>d</b>) <span class="html-italic">Trichoderma</span> strain #5029. Elongation factor 1−α and β−actin genes (<span class="html-italic">tef</span> and <span class="html-italic">act</span>, respectively) were used as references to normalize biological control associated genes expression in <span class="html-italic">Trichoderma</span> strain #5001 and #5029. Asterisks indicate significant differences by Student’s <span class="html-italic">t</span>-test (<span class="html-italic">p</span> &lt; 0.05).</p>
Full article ">Figure 7
<p>Effect of volatiles produced by the <span class="html-italic">Trichoderma</span> strains on plant growth parameters: (<b>a</b>) shoot length (cm), (<b>b</b>) root length, (<b>c</b>) total chlorophyll content (mg/g) and (<b>d</b>) fresh weight. Different letters on the top of SE bars indicate significant differences between the treatments based on Tukey’s test, <span class="html-italic">p</span> &lt; 0.05.</p>
Full article ">Figure 8
<p>Production of siderophores screened on CAS agar medium (front view) (<b>a</b>) uninoculated CAS agar medium, (<b>b</b>) CAS medium inoculated with the strain #5001, (<b>c</b>) CAS medium inoculated with the strain #5029.</p>
Full article ">
26 pages, 742 KiB  
Article
The Role of Preharvest Natural Infection and Toxin Contamination in Food and Feed Safety in Maize, South-East Hungary, 2014–2021
by Akos Mesterhazy, Denes Szieberth, Eva Toldine Tóth, Zoltan Nagy, Balazs Szabó, Beata Herczig, Istvan Bors and Beata Tóth
J. Fungi 2022, 8(10), 1104; https://doi.org/10.3390/jof8101104 - 19 Oct 2022
Cited by 3 | Viewed by 2639
Abstract
Mycotoxins originating in the preharvest period represent a less studied research problem, even though they are of the utmost practical significance in maize production, determining marketability (within EU limits), and storage ability, competitiveness, and profit rate. In this study, 18–23 commercial hybrids were [...] Read more.
Mycotoxins originating in the preharvest period represent a less studied research problem, even though they are of the utmost practical significance in maize production, determining marketability (within EU limits), and storage ability, competitiveness, and profit rate. In this study, 18–23 commercial hybrids were tested between 2014 and 2021. Natural infection from Fusarium spp. was higher than 1.5%, and for Aspergillus spp. this was normally 0.01% or 0, much lower than would be considered as severe infection. In spite of this, many hybrids provided far higher toxin contamination than regulations allow. The maximum preharvest aflatoxin B1 was in 2020 (at 2286 μg/kg), and, in several cases, the value was higher than 1000 μg/kg. The hybrid differences were large. In Hungary, the presence of field-originated aflatoxin B1 was continuous, with three AFB1 epidemics in the 8 years. The highest DON contamination was in 2014 (at 27 mg/kg), and a detectable DON level was found in every hybrid. FUMB1+B2 were the highest in 2014 (at 45.78 mg/kg). At these low infection levels, correlations between visual symptoms and toxin contaminations were mostly non-significant, so it is not feasible to draw a conclusion about toxin contamination from ear rot coverage alone. The toxin contamination of hybrids for a percentage of visual infection is highly variable, and only toxin data can decide about food safety. Hybrids with no visual symptoms and high AFB1 contamination were also identified. Preharvest control, including breeding and variety registration, is therefore of the utmost importance to all three pathogens. Even natural ear rot and toxin data do not prove differences in resistance, so a high ear rot or toxin contamination level should be considered as a risk factor for hybrids. The toxin control of freshly harvested grain is vital for separating healthy and contaminated lots. In addition, proper growing and storage conditions must be ensured to protect the feed safety of the grain. Full article
(This article belongs to the Special Issue Plant Fungal Pathogenesis 2022)
Show Figures

Figure 1

Figure 1
<p><span class="html-italic">A. flavus</span>-contaminated symptomless grains (left) and colonies (middle and right) developing from them following surface sterilization by a NaClO 1% solution (5 min) (courtesy of Mesterhazy, 2012).</p>
Full article ">
12 pages, 1770 KiB  
Article
MALDI-TOF Mass Spectrometry Online Identification of Trichophyton indotineae Using the MSI-2 Application
by Anne-Cécile Normand, Alicia Moreno-Sabater, Arnaud Jabet, Samia Hamane, Geneviève Cremer, Françoise Foulet, Marion Blaize, Sarah Dellière, Christine Bonnal, Sébastien Imbert, Sophie Brun, Ann Packeu, Stéphane Bretagne and Renaud Piarroux
J. Fungi 2022, 8(10), 1103; https://doi.org/10.3390/jof8101103 - 19 Oct 2022
Cited by 26 | Viewed by 2896
Abstract
Trichophyton indotineae is an emerging pathogen which recently spread from India to Europe and that is more prone than other species of the Trichophyton mentagrophytes complex to show resistance to terbinafine, resulting in the necessity of rapid identification. Here, we improved the online [...] Read more.
Trichophyton indotineae is an emerging pathogen which recently spread from India to Europe and that is more prone than other species of the Trichophyton mentagrophytes complex to show resistance to terbinafine, resulting in the necessity of rapid identification. Here, we improved the online MSI-2 MALDI-TOF identification tool in order to identify T. indotineae. By multiplying the culture conditions (2 culture media and 6 stages of growth) prior to protein extractions for both test isolates and reference strains, we added 142 references corresponding to 12 strains inside the T. mentagrophytes complex in the online MSI-2 database, of which 3 are T. indotineae strains. The resulting database was tested with 1566 spectra of 67 isolates from the T. mentagrophytes complex, including 16 T. indotineae isolates. Using the newly improved MSI-2 database, we increased the identification rate of T. indotineae from 5% to 96%, with a sensitivity of 99.6%. We also identified specific peaks (6834/6845 daltons and 10,634/10,680 daltons) allowing for the distinction of T. indotineae from the other species of the complex. Our improved version of the MSI-2 application allows for the identification of T. indotineae. This will improve the epidemiological knowledge of the spread of this species throughout the world and will help to improve patient care. Full article
(This article belongs to the Special Issue Dermatophytes and Dermatophytoses)
Show Figures

Figure 1

Figure 1
<p>Study design [<a href="#B23-jof-08-01103" class="html-bibr">23</a>]. * For isolates that are in both the reference database and in the test panel, self-identification at the strain level is not allowed.</p>
Full article ">Figure 2
<p>Comparison of the aligned spectra obtained at the various stages of growth for <span class="html-italic">T. indotineae</span> isolates on both IDFP (<b>A</b>) and SDA-CG agar (<b>B</b>). To visualize increases or decreases in the intensities of some peaks, a zoomed-in view of two portions of the spectra (5000–16750 Da; 10,000–11,750 Da) is shown. The highest peaks in intensity are represented by the brightest colors.</p>
Full article ">Figure 3
<p>Distribution of the percentages of the intensities corresponding to one specific peak compared to the overall intensities of the spectrum at 10 days of growth. <span class="html-italic">T. indotineae</span> plots are colored blue, and <span class="html-italic">T. interdigitale</span> genotype II plots are colored green. (<b>A</b>) = peak at 6834 (±5) daltons (Da), (<b>B</b>) = peak at 6845 (±5) Da, (<b>C</b>) = peak at 10,634 (±10) Da, and (<b>D</b>) = peak at 10,680 (±10) Da.</p>
Full article ">
16 pages, 2711 KiB  
Article
Calcineurin Inhibitors Synergize with Manogepix to Kill Diverse Human Fungal Pathogens
by Sean D. Liston, Luke Whitesell, Mili Kapoor, Karen J. Shaw and Leah E. Cowen
J. Fungi 2022, 8(10), 1102; https://doi.org/10.3390/jof8101102 - 19 Oct 2022
Cited by 4 | Viewed by 2600
Abstract
Invasive fungal infections have mortality rates of 30–90%, depending on patient co-morbidities and the causative pathogen. The frequent emergence of drug resistance reduces the efficacy of currently approved treatment options, highlighting an urgent need for antifungals with new modes of action. Addressing this [...] Read more.
Invasive fungal infections have mortality rates of 30–90%, depending on patient co-morbidities and the causative pathogen. The frequent emergence of drug resistance reduces the efficacy of currently approved treatment options, highlighting an urgent need for antifungals with new modes of action. Addressing this need, fosmanogepix (N-phosphonooxymethylene prodrug of manogepix; MGX) is the first in a new class of gepix drugs, and acts as a broad-spectrum, orally bioavailable inhibitor of the essential fungal glycosylphosphatidylinositol (GPI) acyltransferase Gwt1. MGX inhibits the growth of diverse fungal pathogens and causes accumulation of immature GPI-anchored proteins in the fungal endoplasmic reticulum. Relevant to the ongoing clinical development of fosmanogepix, we report a synergistic, fungicidal interaction between MGX and inhibitors of the protein phosphatase calcineurin against important human fungal pathogens. To investigate this synergy further, we evaluated a library of 124 conditional expression mutants covering 95% of the genes encoding proteins involved in GPI-anchor biosynthesis or proteins predicted to be GPI-anchored. Strong negative chemical-genetic interactions between the calcineurin inhibitor FK506 and eleven GPI-anchor biosynthesis genes were identified, indicating that calcineurin signalling is required for fungal tolerance to not only MGX, but to inhibition of the GPI-anchor biosynthesis pathway more broadly. Depletion of these GPI-anchor biosynthesis genes, like MGX treatment, also exposed fungal cell wall (1→3)-β-D-glucans. Taken together, these findings suggest the increased risk of invasive fungal infections associated with use of calcineurin inhibitors as immunosuppressants may be mitigated by their synergistic fungicidal interaction with (fos)manogepix and its ability to enhance exposure of immunostimulatory glucans. Full article
(This article belongs to the Special Issue Antifungal Drug Discovery: Novel Therapies and Approaches)
Show Figures

Figure 1

Figure 1
<p>Gwt1 inhibitors synergize with calcineurin inhibitors to kill diverse human fungal pathogens. (<b>A</b>) FK506 and CsA synergize with MGX to inhibit growth of <span class="html-italic">C. albicans</span>. Two-fold dose response matrices for MGX and FK506 or MGX and CsA tested against <span class="html-italic">C. albicans</span>. Growth was determined by OD600 after 24 h incubation in YPD medium and normalized to vehicle control wells. Data are mean relative growth for three replicate wells from a representative experiment. The lowest FICI value is inset (white text), and the corresponding well marked with an asterisk. To assess viability after combination drug treatments, cells were transferred onto drug-free YPD agar and photographed after culture for an additional 48 h. A scale bar for all heat-map depictions presented in panels (<b>A</b>–<b>C</b>) is provided to the right of panel (<b>D</b>). (<b>B</b>) FK506 and CsA synergistically interact with MGX or the structural analog APX2039 to inhibit growth of <span class="html-italic">C. neoformans</span>. Assays were performed as in (<b>A</b>), except optical densities were determined after 48 h growth. (<b>C</b>) FK506 and CsA synergistically interact with MGX to inhibit growth of <span class="html-italic">A. fumigatus</span>. Assays were performed as in (<b>A</b>), except optical densities were determined at 48 h. Relative metabolic activity was assessed by measuring Alamar blue dye reduction. Data are mean relative fluorescence intensity normalized to vehicle control wells for three replicate wells from a representative experiment. (<b>D</b>) Merged phase contrast and fluorescence photomicrographs visualizing the dose-response assay wells outlined in orange in panel (<b>C</b>) after addition of propidium iodide (1 µg/mL) to identify dead cells. Scale bar,100 µm. Experiments were repeated at least two times with similar results.</p>
Full article ">Figure 2
<p>Calcineurin and Gwt1 are responsible for synergistic interaction between MGX and FK506 or CsA. (<b>A</b>) Dose-response matrices were generated for MGX and FK506 or MGX and Cyclosporin A against wild-type <span class="html-italic">C. albicans</span> and a <span class="html-italic">cna1Δ/Δ</span> mutant derivative. Growth was monitored by OD<sub>600</sub> after 24 h incubation in YPD medium and normalized to vehicle control wells. Data are mean relative growth for triplicate wells from a representative experiment. This experiment was performed in two additional replicates with similar results. To assess viability after combination treatments, cells were transferred onto drug-free YPD agar and photographed after 48 h growth. Contaminated wells are marked with a pound sign and surrounded with a yellow box. The lowest FICI value is inset (white text), and the corresponding well indicated with an asterisk. (<b>B</b>) Transcriptional repression of GWT1 sensitizes <span class="html-italic">C. albicans</span> to FK506 and CsA. Dose response assays were performed with <span class="html-italic">C. albicans</span> or tetO-GWT1/Δ mutant derivative in the presence or absence of DOX. Growth was monitored by OD<sub>600</sub> after 24 h incubation and normalized to vehicle control wells. Data are mean relative growth for three replicate wells from a representative experiment. This experiment was performed in two additional replicates with similar results.</p>
Full article ">Figure 3
<p>Transcriptional repression of diverse GPI-anchor biosynthesis genes sensitizes <span class="html-italic">C. albicans</span> to MGX and FK506. (<b>A</b>) Screen of GPI-anchor focused library. Conditional expression strains were grown in the presence or absence of FK506 (100 ng/mL) to inhibit calcineurin and in the presence or absence of DOX (5 µg/mL) to repress target gene expression. Growth was assessed by OD<sub>600</sub> after 24 h. Chemical genetic (CG) interaction scores are displayed as the difference between expected and observed relative growth according to a multiplicative model. Data are mean CG Score ± SEM for triplicate wells from a representative experiment. Strains with statistically significant negative CG scores are indicated (unpaired t-test, Holm–Šídák’s multiple comparison test, <span class="html-italic">p</span> &lt; 0.05) are indicated. Inset depicts values obtained for entire library. Gene names based on the respective <span class="html-italic">S. cerevisiae</span> homologues are indicated for <span class="html-italic">C1_04280</span> (<span class="html-italic">GPI18</span>), <span class="html-italic">C_505040</span> (<span class="html-italic">GPI10</span>) and <span class="html-italic">C503120</span> (<span class="html-italic">GAB1</span>). (<b>B</b>) MIC testing confirms transcriptional repression of multiple GPI anchor biosynthesis genes confers hypersensitivity to FK506. <span class="html-italic">C. albicans</span> or tetO-GENE/Δ mutant derivatives were grown in YPD at 30 °C in the presence or absence of 0.05 µg/mL DOX, then sub-cultured into YPD medium ± 5 µg/mL DOX and a two-fold dilution series of FK506. Growth was determined by OD<sub>600</sub> after 24 h and normalized to vehicle control wells. Data are mean relative growth for triplicate wells from a representative experiment. (<b>C</b>) Schematic depiction of GPI anchor biosynthesis. Legend depicts sugar residues of the GPI structure according to Symbol Nomenclature for Glycans [<a href="#B47-jof-08-01102" class="html-bibr">47</a>]. Genes with CG interactions with FK506 are colored in red. Genes determined be essential based on induction of a severe growth defect in the presence of DOX are colored in blue [<a href="#B30-jof-08-01102" class="html-bibr">30</a>,<a href="#B41-jof-08-01102" class="html-bibr">41</a>].</p>
Full article ">Figure 4
<p>MGX treatment increases chitin levels and exposes (1→3)-β-D-glucan at the fungal cell surface. (<b>A</b>) <span class="html-italic">C. albicans</span> was grown in YPD supplemented with MGX for 24 h at 30 °C then growth was measured by optical density. Left: Mean relative growth for duplicate wells from a representative experiment. Cells were formaldehyde fixed, then stained with CFW or labelled with a mouse monoclonal antibody to (1→3)-β-D-glucan and goat anti-mouse secondary antibody conjugated to AlexaFluor488 and examined by flow cytometry. Right: Median fluorescence intensity for CFW and AlexaFluor488 labelled cells normalized to untreated controls (&gt;20,000 events/sample). Fluorescence intensity for CFW staining saturated the flow cytometer at [MGX] &gt; 31.25 ng/mL. (<b>B</b>) Representative fluorescence micrographs of CFW-stained <span class="html-italic">C. albicans</span> (blue) prepared as in (<b>A</b>). Scale bar, 20 µm. (<b>C</b>) Representative (1→3)-β-D-glucan immunofluorescence (green) micrographs of <span class="html-italic">C. albicans</span> prepared as in A. Scale bar, 20 µm.</p>
Full article ">Figure 5
<p>Transcriptional repression of GPI-anchor biosynthesis genes exposes (1→3)-β-D-glucan at the fungal cell surface and increases chitin levels. Conditional expression strains were grown in the presence or absence of DOX to repress target gene expression. Glucan and CFW staining and flow cytometry were performed as in <a href="#jof-08-01102-f004" class="html-fig">Figure 4</a>. (<b>A</b>) Glucan Staining. Median fluorescence intensity (FITC-A) is plotted for each repressible mutant in the library showing &gt; two-fold change in median fluorescence intensity in screening and validation experiments upon repression of the indicated gene in shaking cultures. Values depicted represent a single determination (~30,000 events/sample). Inset: Screening data. Ratio of median fluorescence intensity of DOX-treated to untreated samples for all strains in the collection grown in static cultures. Values depicted represent a single determination (&gt;1000–10,000 events/sample). (<b>B</b>) CFW Staining. Median fluorescence intensity (PB450-A) is plotted for each repressible mutant in the library showing &gt; two-fold change in median fluorescence intensity in two experiments upon repression of the indicated gene in shaking cultures. Values depicted represent a single determination (~30,000 events/sample). Inset: Screening data. Ratio of median fluorescence intensity of DOX-treated to untreated samples for all strains in the collection grown in static culture. Values depicted represent a single determination (&gt;1000–10,000 events/sample).</p>
Full article ">
6 pages, 1008 KiB  
Brief Report
The Effects of Sex and Strain on Pneumocystis murina Fungal Burdens in Mice
by Nikeya L. Macioce-Tisdale, Alan Ashbaugh, Keeley Hendrix, Margaret S. Collins, Aleksey Porollo and Melanie T. Cushion
J. Fungi 2022, 8(10), 1101; https://doi.org/10.3390/jof8101101 - 19 Oct 2022
Cited by 1 | Viewed by 1511
Abstract
Many preclinical studies of infectious diseases have neglected experimental designs that evaluate potential differences related to sex with a concomitant over-reliance on male model systems. Hence, the NIH implemented a monitoring system for sex inclusion in preclinical studies. Methods: Per this mandate, we [...] Read more.
Many preclinical studies of infectious diseases have neglected experimental designs that evaluate potential differences related to sex with a concomitant over-reliance on male model systems. Hence, the NIH implemented a monitoring system for sex inclusion in preclinical studies. Methods: Per this mandate, we examined the lung burdens of Pneumocystis murina infection in three mouse strains in both male and female animals at early, mid, and late time points. Results: Females in each strain had higher infection burdens compared to males at the later time points. Conclusion: Females should be included in experimental models studying Pneumocystis spp. Full article
(This article belongs to the Special Issue Women in Mycology)
Show Figures

Figure 1

Figure 1
<p><b><span class="html-italic">Pneumocystis murina</span> burdens comparing males and females in 3 mouse strains.</b> At 4, 6, and 8 weeks, asci and nuclei (total fungal burden) of male (M) and female (F) Balb/c, C3H/HeN (C3H), and C57BL/6 (C57) were quantified by microscopic enumeration, log transformed, and expressed as the log<sub>10</sub> mean ± the standard deviation per lung (Y-axis). (*) Indicates statistical significance at <span class="html-italic">p</span> &lt; 0.05 using an unpaired <span class="html-italic">t</span>-test between male and female mice of the same strain. Panel <b>A</b>: Female vs Male nuclei counts in the 3 mouse strains over the 3 time points; Panel <b>B</b>: Female vs Male asci counts in the 3 mouse strains over the 3 time points.</p>
Full article ">Figure 2
<p><b><span class="html-italic">Pneumocystis murina</span> burdens comparing the female and male sexes of each mouse strain at the three time points and survival curve of the study</b>. The same data presented in <a href="#jof-08-01101-f001" class="html-fig">Figure 1</a> were parsed to examine differences among the sexes in each mouse strain. (<b>A</b>) Female nuclei count. * Indicates statistical significance of C57 mice nuclei burdens at 4 and 8 weeks compared to other strains. (<b>B</b>) Female asci count. * Indicates statistical significance of asci burdens of female C57 mice compared to the other strains at these time points. (<b>C</b>) Male nuclei count. * Indicates statistically significant differences in organism burdens. (<b>D</b>) Male asci count. * Indicates statistically significant asci burdens of Balb/c mice when compared to other strains at 6 and 8 weeks. All statistical significance was set at <span class="html-italic">p</span> &lt; 0.05 using one-way ANOVA and Newman–Keuls multiple comparison test. (<b>E</b>) Survival curves for all mouse strains prior to the terminal time point. Survival curves were analyzed using GraphPad Prism v.5. (*) Indicates statistical significance at <span class="html-italic">p</span> &lt; 0.0088 for the male C3H/HeN and <span class="html-italic">p</span> &lt; 0.0001 for the female C3H/HeN.</p>
Full article ">
14 pages, 2376 KiB  
Article
Discovery of Oleaginous Yeast from Mountain Forest Soil in Thailand
by Sirawich Sapsirisuk, Pirapan Polburee, Wanlapa Lorliam and Savitree Limtong
J. Fungi 2022, 8(10), 1100; https://doi.org/10.3390/jof8101100 - 18 Oct 2022
Cited by 9 | Viewed by 2989
Abstract
As an interesting alternative microbial platform for the sustainable synthesis of oleochemical building blocks and biofuels, oleaginous yeasts are increasing in both quantity and diversity. In this study, oleaginous yeast species from northern Thailand were discovered to add to the topology. A total [...] Read more.
As an interesting alternative microbial platform for the sustainable synthesis of oleochemical building blocks and biofuels, oleaginous yeasts are increasing in both quantity and diversity. In this study, oleaginous yeast species from northern Thailand were discovered to add to the topology. A total of 127 yeast strains were isolated from 22 forest soil samples collected from mountainous areas. They were identified by an analysis of the D1/D2 domain of the large subunit rRNA (LSU rRNA) gene sequences to be 13 species. The most frequently isolated species were Lipomyces tetrasporus and Lipomyces starkeyi. Based on the cellular lipid content determination, 78 strains of ten yeast species, and two potential new yeast that which accumulated over 20% of dry biomass, were found to be oleaginous yeast strains. Among the oleaginous species detected, Papiliotrema terrestris and Papiliotrema flavescens have never been reported as oleaginous yeast before. In addition, none of the species in the genera Piskurozyma and Hannaella were found to be oleaginous yeast. L. tetrasporus SWU-NGP 2-5 accumulated the highest lipid content of 74.26% dry biomass, whereas Lipomyces mesembrius SWU-NGP 14-6 revealed the highest lipid quantity at 5.20 ± 0.03 g L−1. The fatty acid profiles of the selected oleaginous yeasts varied depending on the strain and suitability for biodiesel production. Full article
Show Figures

Figure 1

Figure 1
<p>Phylogenetic tree (<b>A</b>) showing the positions of the strains SWU-YGP 11-1 and SWU-NAPS 5-1 and their related species constructed by the maximum-likelihood method based on the D1/D2 of the LSU rRNA gene sequences. <span class="html-italic">Dioszegia hungarica</span> CBS 4214<sup>T</sup> was used as an outgroup. The numerals represent the percentages from 1000 replicate bootstrap resampling (a frequency of less than 50% is not shown). The cells (<b>B</b>) and colony (<b>C</b>) morphology of SWU-YGP 11-1 grown on YM agar for 48 h. The bar represents as 10 µm.</p>
Full article ">Figure 2
<p>Phylogenetic tree (<b>A</b>) showing the positions of the strains SWU-NGP 14-3-2, SWU-NGP 14-3-4, SWU-NGTP 4-12, and their related species constructed by the maximum-likelihood method based on the D1/D2 of the LSU rRNA gene sequences. <span class="html-italic">Holtermannia corniformis</span> CBS 6979<sup>T</sup> was used as an outgroup. The numerals represent the percentages from 1000 replicate bootstrap resampling (a frequency of less than 50% is not shown). The cells (<b>B</b>) and colony (<b>C</b>) morphology of SWU-YGP 11-1 grown on YM agar for 48 h. The bar represents as 10 µm.</p>
Full article ">Figure 3
<p>The number of oleaginous yeasts in this study. The lipid accumulation in the rage of 20–25% of dry biomass (<span style="color:#2A4B90">■</span>), 25–30% of dry biomass (<span style="color:#2E5AB7">■</span>), 30–40% of dry biomass (<span style="color:#A4BAE9">■</span>) and &gt;40% of dry biomass (<span style="color:#D6DEEE">■</span>). X-axis is the percentage of lipid content and Y-axis is number of yeast strains.</p>
Full article ">Figure 4
<p>Photomicrographs of cell under light microscope (<b>A1</b>,<b>A2</b>) and lipid bodies by Nile red staining under fluorescence microscope (<b>B1</b>,<b>B2</b>) of <span class="html-italic">Lipomyces tetrasporus</span> SWU-NGP 2-5 (<b>A1</b>,<b>B1</b>) and <span class="html-italic">Lipomyces mesembrius</span> SWU-NGP 14-6 (<b>A2</b>,<b>B2</b>) grown in 2G2M broth for 7 days. The bar represents as 10 µm.</p>
Full article ">Figure 5
<p>Summary of 13 species isolated from forest soil on the mountains in Thailand. The range of the highest lipid content including average and standard deviation (blue bar), as well as the number of oleaginous yeasts (yellow bar) and the number of non-oleaginous yeasts (green bar), are shown.</p>
Full article ">
14 pages, 857 KiB  
Review
Fungal Contamination in Microalgal Cultivation: Biological and Biotechnological Aspects of Fungi-Microalgae Interaction
by Carmen Laezza, Giovanna Salbitani and Simona Carfagna
J. Fungi 2022, 8(10), 1099; https://doi.org/10.3390/jof8101099 - 18 Oct 2022
Cited by 21 | Viewed by 4990
Abstract
In the last few decades, the increasing interest in microalgae as sources of new biomolecules and environmental remediators stimulated scientists’ investigations and industrial applications. Nowadays, microalgae are exploited in different fields such as cosmeceuticals, nutraceuticals and as human and animal food supplements. Microalgae [...] Read more.
In the last few decades, the increasing interest in microalgae as sources of new biomolecules and environmental remediators stimulated scientists’ investigations and industrial applications. Nowadays, microalgae are exploited in different fields such as cosmeceuticals, nutraceuticals and as human and animal food supplements. Microalgae can be grown using various cultivation systems depending on their final application. One of the main problems in microalgae cultivations is the possible presence of biological contaminants. Fungi, among the main contaminants in microalgal cultures, are able to influence the production and quality of biomass significantly. Here, we describe fungal contamination considering both shortcomings and benefits of fungi-microalgae interactions, highlighting the biological aspects of this interaction and the possible biotechnological applications. Full article
(This article belongs to the Special Issue Plant and Fungal Interactions, 2nd Edition)
Show Figures

Figure 1

Figure 1
<p>Predominant genera in different aquatic habitats. Yellow: coastal and oceanic environments; Blue: deep sea and sub-sea floor; Purple: lakes; Red: rivers, streams, and ponds.</p>
Full article ">Figure 2
<p>(<b>a</b>). Interaction between microalga (negative charge) and fungus (positive charge); (<b>b</b>). microalga entrapped among fungal hyphae; (<b>c</b>). nutrients exchange between microalga and fungus.</p>
Full article ">
21 pages, 9370 KiB  
Review
Paracoccidioidomycosis: What We Know and What Is New in Epidemiology, Diagnosis, and Treatment
by Paulo Mendes Peçanha, Paula Massaroni Peçanha-Pietrobom, Tânia Regina Grão-Velloso, Marcos Rosa Júnior, Aloísio Falqueto and Sarah Santos Gonçalves
J. Fungi 2022, 8(10), 1098; https://doi.org/10.3390/jof8101098 - 18 Oct 2022
Cited by 29 | Viewed by 7217
Abstract
Paracoccidioidomycosis (PCM) is a systemic mycosis endemic to Latin America caused by thermodimorphic fungi of the genus Paracoccidioides. In the last two decades, enhanced understanding of the phylogenetic species concept and molecular variations has led to changes in this genus’ taxonomic classification. Although [...] Read more.
Paracoccidioidomycosis (PCM) is a systemic mycosis endemic to Latin America caused by thermodimorphic fungi of the genus Paracoccidioides. In the last two decades, enhanced understanding of the phylogenetic species concept and molecular variations has led to changes in this genus’ taxonomic classification. Although the impact of the new species on clinical presentation and treatment remains unclear, they can influence diagnosis when serological methods are employed. Further, although the infection is usually acquired in rural areas, the symptoms may manifest years or decades later when the patient might be living in the city or even in another country outside the endemic region. Brazil accounts for 80% of PCM cases worldwide, and its incidence is rising in the northern part of the country (Amazon region), owing to new settlements and deforestation, whereas it is decreasing in the south, owing to agriculture mechanization and urbanization. Clusters of the acute/subacute form are also emerging in areas with major human intervention and climate change. Advances in diagnostic methods (molecular and immunological techniques and biomarkers) remain scarce, and even the reference center’s diagnostics are based mainly on direct microscopic examination. Classical imaging findings in the lungs include interstitial bilateral infiltrates, and eventually, enlargement or calcification of adrenals and intraparenchymal central nervous system lesions are also present. Besides itraconazole, cotrimoxazole, and amphotericin B, new azoles may be an alternative when the previous ones are not tolerated, although few studies have investigated their use in treating PCM. Full article
Show Figures

Figure 1

Figure 1
<p>Geographic distribution of Paracoccidioidomycosis in relation to endemic areas and imported cases.</p>
Full article ">Figure 2
<p>(<b>A</b>) Chronic form: gingival and lingual frenulus “mulberry-like” ulcers with hemorrhagic dots; (<b>B</b>) chronic form: deep ulcerative lesion on the tongue with infiltrative borders, hemorrhagic dots, covered with fibrin; (<b>C</b>) acute/subacute form: multiple polymorphic (nodular, papular, and ulcerated) skin lesions and cervical inflammatory lymphadenopathy; (<b>D</b>) disseminated form in a patient with PCM and HIV co-infection; the following may be seen: cervical lymphadenopathy, exuberant ulcerated, crusted skin lesions, and larges subcutaneous abscesses in the thorax and abdomen.</p>
Full article ">Figure 3
<p>Laboratory diagnosis of Paracoccidioidomycosis. (<b>A</b>) <span class="html-italic">Paracoccidioides</span> spp. on Sabouraud agar slants for 14 days at 25 °C (left, filamentous phase) and 37 °C (right, yeast phase); (<b>B</b>) yeast form showing multiple budding (5 to 15 µm) seen upon direct examination of a lymph node aspirate, stained with KOH and Parker ink; Bars 10 µm (<b>C</b>) double immunodiffusion assay: <span class="html-italic">Paracoccidioides brasiliensis</span> exoantigen (Ag) is in the center well; sera sample from a patient with paracoccidioidomycosis is used in different titrations (wells 1:1 to 1:16) and positive control (PC); positive 1:2 (<b>left</b>) and negative (<b>right</b>); (<b>D</b>) Grocott’s methenamine silver stain showing <span class="html-italic">Paracoccidioides</span> yeast cells. Bars 10 µm.</p>
Full article ">Figure 4
<p>Computed tomography showing ground-glass opacities (arrows) with bilateral and symmetric distribution, occurring mainly in the periphery and the middle-third of the lungs (“butterfly wing” pattern), as shown from top to bottom in (<b>A</b>–<b>D</b>).</p>
Full article ">Figure 5
<p>Computed tomography showing nodular opacities bilaterally distributed (arrows in <b>A</b>–<b>D</b>), with cavitation (arrow in <b>C</b>), as shown from top to bottom in figures (<b>A</b>–<b>D</b>).</p>
Full article ">Figure 6
<p>Computed tomography axial without contrast (<b>A</b>) and coronal post-contrast (<b>B</b>) images showing asymmetric adrenal thickening without significant contrast enhancement. The right adrenal is calcified, whereas the left is enlarged. Computed tomography axial post-contrast (<b>C</b>) and coronal post-contrast (<b>D</b>) from another patient showing adrenal thickening.</p>
Full article ">Figure 7
<p>Magnetic resonance imaging showing a hypointense lesion on fluid-attenuated inversion recovery (FLAIR) (<b>A</b>) and T2-weighted imaging (T2WI) (<b>B</b>), with an annular enhancement on T1 post-contrast (<b>C</b>) located in the right occipital and temporal lobes.</p>
Full article ">
29 pages, 6682 KiB  
Article
The Epichloë festucae Antifungal Protein Efe-AfpA Protects Creeping Bentgrass (Agrostis stolonifera) from the Plant Pathogen Clarireedia jacksonii, the Causal Agent of Dollar Spot Disease
by Patrick A. Fardella, Zipeng Tian, Bruce B. Clarke and Faith C. Belanger
J. Fungi 2022, 8(10), 1097; https://doi.org/10.3390/jof8101097 - 18 Oct 2022
Cited by 3 | Viewed by 2270
Abstract
Dollar spot disease, caused by the fungal pathogen Clarireedia jacksonii, is a major problem in many turfgrass species, particularly creeping bentgrass (Agrostis stolonifera). It is well-established that strong creeping red fescue (Festuca rubra subsp. rubra) exhibits good dollar [...] Read more.
Dollar spot disease, caused by the fungal pathogen Clarireedia jacksonii, is a major problem in many turfgrass species, particularly creeping bentgrass (Agrostis stolonifera). It is well-established that strong creeping red fescue (Festuca rubra subsp. rubra) exhibits good dollar spot resistance when infected by the fungal endophyte Epichloë festucae. This endophyte-mediated disease resistance is unique to the fine fescues and has not been observed in other grass species infected with other Epichloë spp. The mechanism underlying the unique endophyte-mediated disease resistance in strong creeping red fescue has not yet been established. We pursued the possibility that it may be due to the presence of an abundant secreted antifungal protein produced by E. festucae. Here, we compare the activity of the antifungal protein expressed in Escherichia coli, Pichia pastoris, and Penicillium chrysogenum. Active protein was recovered from all systems, with the best activity being from Pe. chrysogenum. In greenhouse assays, topical application of the purified antifungal protein to creeping bentgrass and endophyte-free strong creeping red fescue protected the plants from developing severe symptoms caused by C. jacksonii. These results support the hypothesis that Efe-AfpA is a major contributor to the dollar spot resistance observed with E. festucae-infected strong creeping red fescue in the field, and that this protein could be developed as an alternative or complement to fungicides for the management of this disease on turfgrasses. Full article
(This article belongs to the Special Issue Fungal Endophytes of Grasses)
Show Figures

Figure 1

Figure 1
<p>Comparison of activity of bacterially produced <span class="html-italic">Efe</span>-AfpA with varying N-terminal amino acids. Increasing concentrations of the modified <span class="html-italic">Efe</span>-AfpA proteins were assayed for activity in the <span class="html-italic">N. crassa</span> growth assay with 1 × 10<sup>6</sup> conidia mL <sup>−1</sup>. The data presented are the means and standard deviations of three replicates. For each concentration, columns with different letters indicate a significant difference in activity (<span class="html-italic">p</span> ≤ 0.05, two-way ANOVA).</p>
Full article ">Figure 2
<p>Comparison of activity of <span class="html-italic">Efe</span>-AfpA and PAF against <span class="html-italic">Pe</span>. <span class="html-italic">chrysogenum</span> Δ<span class="html-italic">paf</span> conidial growth. Five µg mL<sup>−1</sup> of either <span class="html-italic">Efe</span>-AfpA or PAF was assayed for activity against 1 × 10<sup>4</sup> conidia mL<sup>−1</sup>. The data presented are the means and standard deviations of three replicates. Columns with different letters indicate a significant difference in activity (<span class="html-italic">p</span> ≤ 0.05, one-way ANOVA).</p>
Full article ">Figure 3
<p>SDS polyacrylamide gel of purification of the <span class="html-italic">E. festucae</span> antifungal protein produced in <span class="html-italic">Pe. chrysogenum</span>. Lane 1, Bio-Rad Precision Plus Protein Dual Xtra Standards, size of markers in kD given on left; Lane 2, Crude culture filtrate of <span class="html-italic">Pe. chrysogenum</span> expressing the antifungal protein; Lane 3, 1 µg purified <span class="html-italic">Efe</span>-AfpA, indicated by the arrow.</p>
Full article ">Figure 4
<p>Comparison of activities of PAF purified from <span class="html-italic">Pe. chrysogenum</span> and <span class="html-italic">Efe</span>-AfpA purified from <span class="html-italic">E. coli</span>, <span class="html-italic">Pi. pastoris</span>, and <span class="html-italic">Pe. chrysogenum</span>. Increasing concentrations of the antifungal proteins were assayed for activity in the <span class="html-italic">N. crassa</span> growth assay with 1 × 10<sup>6</sup> conidia mL<sup>−1</sup>. The data presented are the means and standard deviations of three replicates. For each concentration, columns with different letters indicate a significant difference in activity (<span class="html-italic">p</span> ≤ 0.05, two-way ANOVA).</p>
Full article ">Figure 5
<p>Comparison of the activity of <span class="html-italic">Efe</span>-AfpA expressed in <span class="html-italic">Pe. chrysogenum</span> with PAF, a similar antifungal protein from <span class="html-italic">Pe. chrysogenum</span>, against <span class="html-italic">C. jacksonii</span> in culture. In the upper panels (<b>A</b>) water (10 µL), PAF (300 ng), or purified <span class="html-italic">Efe-</span>AfpA (300 ng), was placed on the right side of a plug of <span class="html-italic">C. jacksonii</span>. The lower panels (<b>B</b>) show the <span class="html-italic">C. jacksonii</span> hyphae from the upper panels treated with Evans blue. Bars in the lower panels are 750 μm.</p>
Full article ">Figure 6
<p>Effect of <span class="html-italic">Efe</span>-AfpA (<b>A</b>) or PAF (<b>B</b>) on <span class="html-italic">C. jacksonii</span> mycelial growth. <span class="html-italic">C. jacksonii</span> mycelial plugs were subcultured onto PDA plates amended with increasing concentrations of either <span class="html-italic">Efe</span>-AfpA or PAF. The colony diameters were measured daily. The data presented are the means and standard deviations of three replicates. Photographs of the plates are shown in <a href="#app1-jof-08-01097" class="html-app">Figures S2 and S3</a>.</p>
Full article ">Figure 7
<p><span class="html-italic">Efe</span>-AfpA prevented severe symptoms of dollar spot disease when endophyte-free strong creeping red fescue plants were inoculated with an 8 mm plug of <span class="html-italic">C. jacksonii</span>. Plants were sprayed daily for 10 days with either water (<b>A</b>,<b>B</b>) or 100 μg mL<sup>−1</sup> of <span class="html-italic">Efe</span>-AfpA (<b>C</b>). Photos within a row are replicates of the labeled treatment.</p>
Full article ">Figure 8
<p>Endophyte infected strong creeping red fescue plants exhibited no dollar spot disease symptoms when inoculated with an 8 mm plug of <span class="html-italic">C. jacksonii</span>. Plants were sprayed daily for 10 days with either water (<b>A</b>,<b>B</b>) or 100 μg mL<sup>−1</sup> of <span class="html-italic">Efe</span>-AfpA (<b>C</b>). Photos within a row are replicates of the labeled treatment.</p>
Full article ">Figure 9
<p>Effect of <span class="html-italic">Efe</span>-AfpA concentration on expression of dollar spot disease symptoms when creeping bentgrass cv. ‘Crenshaw’ plants were inoculated with an 8 mm plug of <span class="html-italic">C. jacksonii</span>. Plants were sprayed daily for 7 days with either water (<b>A</b>,<b>B</b>) or different concentrations of <span class="html-italic">Efe</span>-AfpA (<b>C</b>–<b>E</b>). Photos within a row are replicates of the labeled treatment.</p>
Full article ">Figure 10
<p>Effect of the number of treatments of <span class="html-italic">Efe</span>-AfpA applications on expression of dollar spot disease symptoms after creeping bentgrass cv. ‘Crenshaw’ plants were inoculated with an 8 mm plug of <span class="html-italic">C. jacksonii</span>. Plants were treated water (<b>A</b>,<b>B</b>) or with 100 μg mL <sup>−</sup><sup>1</sup> of <span class="html-italic">Efe</span>-AfpA as indicated (<b>C</b>–<b>E</b>) and photographed 7 days post inoculation. Photos within a row are replicates of the labeled treatment.</p>
Full article ">Figure 11
<p><span class="html-italic">Efe</span>-AfpA (<b>A</b>) and PAF (<b>B</b>) activity against <span class="html-italic">N. crassa</span> wild type and glucosylceramide pathway mutants. Increasing concentrations of the proteins were assayed for activity in the <span class="html-italic">N. crassa</span> growth assay with 1 × 10<sup>4</sup> conidia mL<sup>−1</sup>. The data presented are the means and standard deviations of three replicates. For each concentration, columns with different letters indicate a significant difference in activity (<span class="html-italic">p</span> ≤ 0.05, two-way ANOVA).</p>
Full article ">Figure 12
<p>Effect of <span class="html-italic">Efe</span>-AfpA and PAF on conidia germination of <span class="html-italic">N. crassa</span> wild type (<b>A</b>) and glucosylceramide pathway mutants (<b>B</b>,<b>C</b>). Conidia (100 μL of 1 × 10<sup>5</sup> conidia mL<sup>−</sup><sup>1</sup>) were treated for six hours with either water or an antifungal protein at 10 μg mL<sup>−</sup><sup>1</sup>. One-hundred conidia were counted as either germinated or ungerminated. The data presented are the means and standard deviations of three replicates. For each <span class="html-italic">N. crassa</span> strain, columns with different letters indicate a significant difference in activity (<span class="html-italic">p</span> ≤ 0.05, one-way ANOVA).</p>
Full article ">Figure 13
<p>Rooted 50% majority rule maximum parsimony phylogenetic tree of the antifungal protein amino acid sequences. The <span class="html-italic">A. giganteus</span> and <span class="html-italic">A. clavatus</span> sequences were designated as the outgroups for rooting the tree. The numbers at the nodes are the bootstrap percentages based on 1000 replications. The tree was based upon 101 total characters, of which 14 were constant, 6 variable characters were parsimony uninformative, and 81 variable characters were parsimony informative. The NCBI accession numbers are given following the species names. The <span class="html-italic">Efe</span>-AfpA, PAF, NFAP, and PAFB sequences are identified in parentheses. The clades previously designated as Class A and Class B [<a href="#B49-jof-08-01097" class="html-bibr">49</a>] are indicated. Genera abbreviations are F., <span class="html-italic">Fusarium</span>; E., <span class="html-italic">Epichloë</span>; Po., <span class="html-italic">Pochonia</span>; Pe., <span class="html-italic">Penicillium</span>; C., <span class="html-italic">Cordyceps</span>; B., <span class="html-italic">Beauveria</span>; A., <span class="html-italic">Aspergillus</span>; N., <span class="html-italic">Neosartorya</span>; M., <span class="html-italic">Monascus</span>.</p>
Full article ">
15 pages, 19940 KiB  
Article
Transcription Factors in Aureobasidium spp.: Classification, Regulation and a Newly Built Database
by Guang Yang, Yuhan Wang, Yaowei Fang, Hongjuan Mo, Zhihong Hu, Xiaoyue Hou, Shu Liu, Zhongwei Chen and Shulei Jia
J. Fungi 2022, 8(10), 1096; https://doi.org/10.3390/jof8101096 - 17 Oct 2022
Viewed by 2095
Abstract
Transcription factors (TFs) can regulate the synthesis of secondary metabolites through different metabolic pathways in Aureobasidium spp. In this study, a set of 16 superfamilies, 45 PFAM families of TFs with the DNA-binding domains, seven zinc finger families and eight categories of the [...] Read more.
Transcription factors (TFs) can regulate the synthesis of secondary metabolites through different metabolic pathways in Aureobasidium spp. In this study, a set of 16 superfamilies, 45 PFAM families of TFs with the DNA-binding domains, seven zinc finger families and eight categories of the C2H2 TFs have been identified in Aureobasidium spp. Among all the identified TFs, four superfamilies and six PFAM families are the fungal-specific types in this lineage. The Zn2Cys6 and fungal-specific domain regulators are found to be overwhelmingly predominated, while the C2H2 zinc finger class comprises a smaller regulator class. Since there are currently no databases that allow for easy exploration of the TFs in Aureobasidium spp., based on over 50 references and 2405 homologous TFs, the first TFs pipeline—the Aureobasidium Transcription Factor Database (ATFDB)—has been developed to accelerate the identification of metabolic regulation in various Aureobasidium species. It would be useful to investigate the mechanisms behind the wide adaptability and metabolite diversity of Aureobasidium spp. Full article
Show Figures

Figure 1

Figure 1
<p>Flowchart of major steps of ATFDB construction. First, a seed database was constructed for selected genes by retrieving protein sequences from the NCBI database using the accession numbers. Second, target genes from the NCBI and their homologs were identified and integrated to construct the core database. At last, all data were uploaded to the online database for visualization, and the trained HMM models were developed to conduct species-specific TFs profiling of <span class="html-italic">Aureobasidium</span> spp. via the sequencing data.</p>
Full article ">Figure 2
<p>Distribution of PFAM families in different fungi. (<b>A</b>) <span class="html-italic">A. melanogenum</span> CBS110374; (<b>B</b>) <span class="html-italic">A. pullulans</span> EXF-150; (<b>C</b>) <span class="html-italic">A. subglaciale</span> EXF-2481; (<b>D</b>) <span class="html-italic">A. namibiae</span> CBS 147.97; (<b>E</b>) <span class="html-italic">Aspergillus</span> spp.: <span class="html-italic">A. aculeatus</span>, <span class="html-italic">A. niger</span>, <span class="html-italic">A. fumigatis</span> and <span class="html-italic">A. oryzae</span> RIB40; (<b>F</b>) <span class="html-italic">S. cerevisiae</span>: <span class="html-italic">S. cerevisiae</span> S288C, <span class="html-italic">S. cerevisiae</span> YJM244, <span class="html-italic">S. cerevisiae</span> YJM450, <span class="html-italic">S. cerevisiae</span> YJM993 and <span class="html-italic">S. cerevisiae</span> YJM1078. The average number of the regulators for each species is indicated underneath each pie chart. The number of genomes analyzed in all species is indicated in parentheses.</p>
Full article ">Figure 3
<p>Conserved functional domains of the KilA-N type protein Swi4.</p>
Full article ">Figure 4
<p>Phylogenetic tree of the C2H2 transcription factors (<b>A</b>) and the Cys2His2 zinc finger motif, consisting of α helix and an antiparallel β sheet (<b>B</b>). The zinc ion (green) is coordinated by two histidine residues and two cysteine residues.</p>
Full article ">Figure 5
<p>Transcription factor Cmr1 and the target protein Pks1 in different fungi.</p>
Full article ">Figure 6
<p>Workflow of the <span class="html-italic">Aureobasidium</span> Transcription Factor Database.</p>
Full article ">Figure 7
<p>Comparison of the TFs numbers of <span class="html-italic">Aureobasidium</span> spp. in the NCBI (GenBank) and the ATFDB databases. (<b>A</b>) Comparison of the NCBI (GenBank) and the ATFDB databases at type-level of TFs classifications. (<b>B</b>) Comparison of the NCBI (GenBank) and the ATFDB databases at the metabolism-level of TFs classifications.</p>
Full article ">
13 pages, 1156 KiB  
Article
Design and Validation of qPCR-Specific Primers for Quantification of the Marketed Terfezia claveryi and Terfezia crassiverrucosa in Soil
by Francisco Arenas, Asunción Morte and Alfonso Navarro-Ródenas
J. Fungi 2022, 8(10), 1095; https://doi.org/10.3390/jof8101095 - 17 Oct 2022
Cited by 1 | Viewed by 2502
Abstract
Desert truffle crop is a pioneer in southeastern Spain, a region where native edible hypogeous fungi are adapted to the semiarid areas with low annual rainfall. Terfezia claveryi Chatin was the first species of desert truffle to be cultivated, and has been increasing [...] Read more.
Desert truffle crop is a pioneer in southeastern Spain, a region where native edible hypogeous fungi are adapted to the semiarid areas with low annual rainfall. Terfezia claveryi Chatin was the first species of desert truffle to be cultivated, and has been increasing in recent years as an alternative rainfed crop in the Iberian Peninsula. However, its behaviour in the field has yet not been investigated. For this purpose, specific primers were designed for the soil DNA quantification of both T. claveryi and Terfezia crassiverrucosa and a real-time qPCR protocol was developed, using the ITS rDNA region as a target. Moreover, a young desert truffle orchard was sampled for environmental validation. The results showed the highest efficiency for the TerclaF3/TerclaR1 primers pair, 89%, and the minimal fungal biomass that could be reliable detected was set at 4.23 µg mycelium/g soil. The spatial distribution of fungal biomass was heterogeneous, and there was not a direct relationship between the quantity of winter soil mycelium and the location/productivity of desert truffles. This protocol could be applied to tracking these species in soil and understand their mycelial dynamics in plantations and wild areas. Full article
(This article belongs to the Topic Mycorrhizal Fungi Mediated Sustainable Crop Production)
Show Figures

Figure 1

Figure 1
<p>Diagram of sampling points (pyramid marks) in different years of plantation establishment.</p>
Full article ">Figure 2
<p>Real-time qPCR standard curve for <span class="html-italic">T. claveryi</span> DNA quantification in soil. The curve was generated by plotting the Ct values obtained from 10-fold serial dilutions of DNA standard sample against the logarithm of the quantity of mycelium in soil (μg/g). Efficiency for the primer set TerclaF3/R1 was 89%.</p>
Full article ">Figure 3
<p>Distribution of <span class="html-italic">turmas</span> mycelium in the different planting years (2019 in black, 2018 in grey, and 2016 in white raster) across sampling points (1–6).</p>
Full article ">
25 pages, 4960 KiB  
Article
Taxonomy and Phylogenetic Appraisal of Dothideomycetous Fungi Associated with Magnolia, Lilium longiflorum and Hedychium coronarium
by Nimali I. de Silva, Kasun M. Thambugala, Danushka S. Tennakoon, Samantha C. Karunarathna, Jaturong Kumla, Nakarin Suwannarach and Saisamorn Lumyong
J. Fungi 2022, 8(10), 1094; https://doi.org/10.3390/jof8101094 - 17 Oct 2022
Cited by 7 | Viewed by 2559
Abstract
This paper highlights the taxonomy of some interesting saprobic microfungi associated with dead plant materials of Hedychium coronarium, Lilium longiflorum, and Magnolia species. The taxa reported in this study belong to the orders Pleosporales and Kirschsteiniotheliales (Dothideomycetes). These taxa [...] Read more.
This paper highlights the taxonomy of some interesting saprobic microfungi associated with dead plant materials of Hedychium coronarium, Lilium longiflorum, and Magnolia species. The taxa reported in this study belong to the orders Pleosporales and Kirschsteiniotheliales (Dothideomycetes). These taxa were identified based on multi-locus phylogeny of nuclear ribosomal DNA (rDNA) (LSU, SSU, and ITS) and protein-coding genes (tef1-α and rpb2), together with comprehensive morphological characterization. Two novel saprobic species, Leptoparies magnoliae sp. nov. and Neobambusicola magnoliae sp. nov., are introduced from Magnolia species in Thailand. Another new species, Asymmetrispora zingiberacearum sp. nov., is also described from dead stems of H. coronarium, which is the first asexual morph species of the genus Asymmetrispora. In addition, Ramusculicola thailandica and Kirschsteiniothelia thailandica are reported as new host records from dead twigs of Magnolia species. Sphaerellopsis paraphysata is reported as a new host record from L. longiflorum. Newly described taxa are compared with other similar species and detailed descriptions, micrographs, and phylogenetic trees to show the positions are provided. Full article
(This article belongs to the Topic Fungal Diversity)
Show Figures

Figure 1

Figure 1
<p>Phylogram generated from maximum likelihood analysis is based on combined LSU, SSU, and ITS sequence data. Related sequences of <span class="html-italic">Leptosphaeriaceae</span> were obtained from Doilom et al. [<a href="#B33-jof-08-01094" class="html-bibr">33</a>]. ML bootstrap values equal to or greater than 75% and Bayesian posterior probabilities (BYPP) equal to or greater than 0.95 are indicated above the branches. The tree was rooted to <span class="html-italic">Didymella exigua</span> (CBS 183.55). The newly generated sequences are indicated in red. Type and ex-type strains are in bold.</p>
Full article ">Figure 2
<p>Phylogram generated from maximum likelihood analysis is based on combined LSU, SSU, ITS, <span class="html-italic">tef1-α</span>, and <span class="html-italic">rpb2</span> sequence data. Related sequences of <span class="html-italic">Leptoparies</span> and closely related genera in <span class="html-italic">Lophiostomataceae</span> were obtained from Andreasen et al. [<a href="#B34-jof-08-01094" class="html-bibr">34</a>]. ML bootstrap values equal to or greater than 75% and Bayesian posterior probabilities (BYPP) equal to or greater than 0.95 are indicated above the branches. The tree was rooted to <span class="html-italic">Teichospora trabicola</span> (C134). The newly generated sequences are indicated in red. Type and ex-type strains are in bold.</p>
Full article ">Figure 3
<p>Phylogram generated from maximum likelihood analysis is based on combined LSU, SSU, ITS, and <span class="html-italic">tef1-α</span> sequence data. Related sequences of <span class="html-italic">Sulcatisporaceae</span> were obtained from Phukhamsakda et al. [<a href="#B35-jof-08-01094" class="html-bibr">35</a>,<a href="#B36-jof-08-01094" class="html-bibr">36</a>]. ML bootstrap values equal to or greater than 75% and Bayesian posterior probabilities (BYPP) equal to or greater than 0.95 are indicated above the branches. The tree was rooted to <span class="html-italic">Didymosphaeria rubi-ulmifolii</span> (MFLUCC 14-0024). The newly generated sequences are indicated in red. Type and ex-type strains are in bold.</p>
Full article ">Figure 4
<p>Phylogram generated from maximum likelihood analysis is based on combined LSU, ITS, SSU, <span class="html-italic">tef1-α</span>, and <span class="html-italic">rpb2</span> sequence data. Related sequences of <span class="html-italic">Teichosporaceae</span> were obtained from Tennakoon et al. [<a href="#B37-jof-08-01094" class="html-bibr">37</a>]. ML bootstrap values equal to or greater than 75% and Bayesian posterior probabilities (BYPP) equal to or greater than 0.95 are indicated above the branches. The tree was rooted to <span class="html-italic">Hermatomyces tectonae</span> (MFLUCC 14-1140) and <span class="html-italic">H. thailandica</span> (MFLUCC 14-1143). The newly generated sequences are indicated in red. Type and ex-type strains are in bold.</p>
Full article ">Figure 5
<p>Phylogram generated from maximum likelihood analysis is based on combined LSU, SSU, and ITS sequence data. Related sequences of <span class="html-italic">Kirschsteiniothelia</span> species were obtained from Sun et al. [<a href="#B38-jof-08-01094" class="html-bibr">38</a>]. ML bootstrap values equal to or greater than 75% and Bayesian posterior probabilities (BYPP) equal to or greater than 0.95 are indicated above the branches. The tree was rooted to <span class="html-italic">Pseudorobillarda eucalypti</span> (MFLUCC 12-0422) and <span class="html-italic">P. phragmitis</span> (CBS 398.61). The newly generated sequences are indicated in red. Type and ex-type strains are in bold.</p>
Full article ">Figure 6
<p><span class="html-italic">Sphaerellopsis paraphysata</span> (MFLU 19-2774, new host record). (<b>a</b>,<b>b</b>) Conidiomata on host. (<b>c</b>) Close-up of conidiomata on host. (<b>d</b>) Section through conidioma. (<b>e</b>) Conidiomatal wall. (<b>f</b>) Conidiogenous cells with developing conidia. (<b>g</b>–<b>k</b>) Conidia. (<b>l</b>) A germinating conidium. (<b>m</b>) Colony from above (on PDA). (<b>n</b>) Colony from below (on PDA). Scale bars: (<b>d</b>) = 50 µm, (<b>e</b>,<b>f</b>) = 5 µm, (<b>g</b>–<b>l</b>) = 8 µm.</p>
Full article ">Figure 7
<p><span class="html-italic">Leptoparies magnoliae</span> (MFLU 18-1291, holotype). (<b>a</b>) The specimen. (<b>b</b>,<b>c</b>) Appearance of ascomata on the host substrate. (<b>d</b>,<b>e</b>) Vertical sections through ascoma. (<b>f</b>) Peridium. (<b>g</b>,<b>h</b>) Pseudoparaphyses and asci. (<b>i</b>) Ascus. (<b>j</b>–<b>m</b>) Ascospores. Scale bars: (<b>a</b>) = 500 μm, (<b>b</b>,<b>c</b>) = 200 μm, (<b>d</b>,<b>e</b>) = 50 µm, (<b>g</b>–<b>i</b>) = 10 µm, (<b>f</b>,<b>j</b>–<b>m</b>) = 5 µm.</p>
Full article ">Figure 8
<p><span class="html-italic">Neobambusicola magnoliae</span> (HKAS 107122, holotype). (<b>a</b>–<b>c</b>) Appearance of conidiomata on substrate. (<b>d</b>,<b>e</b>) Vertical sections through conidiomata. (<b>f</b>) Conidiomatal wall. (<b>g</b>,<b>h</b>) Conidiogenous cells and developing conidia. (<b>i</b>–<b>m</b>) Conidia. Scale bars: (<b>c</b>) = 200 μm, (<b>d</b>,<b>e</b>) = 20 μm, (<b>f</b>–<b>m</b>) = 5 μm.</p>
Full article ">Figure 9
<p><span class="html-italic">Asymmetrispora zingiberacearum</span> (MFLU 19-2813, holotype). (<b>a</b>,<b>b</b>) Conidiomata on the host. (<b>c</b>) Section through conidioma. (<b>d</b>) Conidiomatal wall. (<b>e</b>,<b>f</b>) Conidiogenous cells and developing conidia. (<b>g</b>,<b>h</b>) Conidia. (<b>i</b>) A germinated conidium. (<b>j</b>) Colony from below (on PDA). (<b>k</b>) Colony from above (on PDA). Scale bars: (<b>c</b>) = 50 µm, (<b>d</b>) = 10 µm, (<b>e</b>–<b>i</b>) = 5 µm.</p>
Full article ">Figure 10
<p><span class="html-italic">Ramusculicola thailandica</span> (HKAS 107136, new host record). (<b>a</b>) The specimen. (<b>b</b>,<b>c</b>) Appearance of ascomata on substrate. (<b>d</b>,<b>e</b>) Vertical sections through ascoma. (<b>f</b>) Peridium. (<b>g</b>) Ascus and pseudoparaphyses. (<b>h</b>,<b>i</b>) Asci. (<b>j</b>–<b>m</b>) Ascospores. Scale bars: (<b>d</b>,<b>e</b>) = 50 μm, (<b>f</b>) = 10 μm, (<b>g</b>–<b>i</b>) = 20 μm, (<b>j</b>–<b>m</b>) = 10 μm.</p>
Full article ">Figure 11
<p><span class="html-italic">Kirschsteiniothelia thailandica</span> (HKAS 107110, new host record). (<b>a</b>–<b>c</b>) Colonies on natural substrate. (<b>d</b>) Conidiophore and Conidiogenous cell. (<b>e</b>–<b>g</b>) Conidia. (<b>h</b>) Conidiophores. (<b>i</b>–<b>k</b>) Conidia. Scale bars: (<b>c</b>) = 100 μm, (<b>d</b>–<b>k</b>) = 20 μm.</p>
Full article ">
16 pages, 1611 KiB  
Review
Oral Chronic Hyperplastic Candidiasis and Its Potential Risk of Malignant Transformation: A Systematic Review and Prevalence Meta-Analysis
by Alejandro I. Lorenzo-Pouso, Alba Pérez-Jardón, Vito Carlo Alberto Caponio, Francesca Spirito, Cintia M. Chamorro-Petronacci, Óscar Álvarez-Calderón-Iglesias, Pilar Gándara-Vila, Lorenzo Lo Muzio and Mario Pérez-Sayáns
J. Fungi 2022, 8(10), 1093; https://doi.org/10.3390/jof8101093 - 17 Oct 2022
Cited by 22 | Viewed by 4588
Abstract
Chronic hyperplastic candidiasis (CHC) is a prototypical oral lesion caused by chronic Candida infection. A major controversy surrounding CHC is whether this oral lesion owns malignant transformation (MT) potential. The aim of the present study was to evaluate current evidence on the MT [...] Read more.
Chronic hyperplastic candidiasis (CHC) is a prototypical oral lesion caused by chronic Candida infection. A major controversy surrounding CHC is whether this oral lesion owns malignant transformation (MT) potential. The aim of the present study was to evaluate current evidence on the MT of CHC and to determine the variables which have the greatest influence on cancer development. Bibliographical searches included PubMed, Embase, Web of Science, Scopus and LILACS. The cohort studies and case series used to investigate the MT of CHC were deemed suitable for inclusion. The quality of the enrolled studies was measured by the Joanna Briggs Institute scale. Moreover, we undertook subgroup analyses, assessed small study effects, and conducted sensitivity analyses. From 338 studies, nine were finally included for qualitative/quantitative analysis. The overall MT rate for CHC across all studies was 12.1% (95% confidential interval, 4.1–19.8%). Subgroup analysis showed that the MT rate increased when pooled analysis was restricted to poor quality studies. It remains complex to affirm whether CHC is an individual and oral, potentially malignant disorder according to the retrieved evidence. Prospective cohort studies to define the natural history of CHC and a consensus statement to clarify a proper set of diagnostic criteria are strongly needed. PROSPERO ID: CRD42022319572. Full article
(This article belongs to the Special Issue Diagnosis and Treatments of Invasive Fungal Diseases)
Show Figures

Figure 1

Figure 1
<p>Prisma flow diagram of the searching processes.</p>
Full article ">Figure 2
<p>Quality plot graphically depicting the risk of bias among individual studies, assessed using Joanna Briggs Institute Critical Appraisal Checklist designed for systematic reviews addressing prevalence, cumulative incidence questions, and/or for proportion meta-analyses [<a href="#B6-jof-08-01093" class="html-bibr">6</a>,<a href="#B7-jof-08-01093" class="html-bibr">7</a>,<a href="#B20-jof-08-01093" class="html-bibr">20</a>,<a href="#B21-jof-08-01093" class="html-bibr">21</a>,<a href="#B22-jof-08-01093" class="html-bibr">22</a>,<a href="#B23-jof-08-01093" class="html-bibr">23</a>,<a href="#B24-jof-08-01093" class="html-bibr">24</a>,<a href="#B25-jof-08-01093" class="html-bibr">25</a>,<a href="#B26-jof-08-01093" class="html-bibr">26</a>].</p>
Full article ">Figure 3
<p>(<b>A</b>) Forest plot. Overall malignant transformation of chronic hyperplastic candidiasis. RE (random effects) and weight of each study; (<b>B</b>) Funnel plot assessing the publication bias [<a href="#B6-jof-08-01093" class="html-bibr">6</a>,<a href="#B7-jof-08-01093" class="html-bibr">7</a>,<a href="#B20-jof-08-01093" class="html-bibr">20</a>,<a href="#B21-jof-08-01093" class="html-bibr">21</a>,<a href="#B22-jof-08-01093" class="html-bibr">22</a>,<a href="#B23-jof-08-01093" class="html-bibr">23</a>,<a href="#B24-jof-08-01093" class="html-bibr">24</a>,<a href="#B25-jof-08-01093" class="html-bibr">25</a>,<a href="#B26-jof-08-01093" class="html-bibr">26</a>].</p>
Full article ">Figure A1
<p>Bubble plot of meta-regression on the chronic hyperplastic candidiasis malignant transformation rate plotted against the year of study.</p>
Full article ">
21 pages, 2882 KiB  
Article
Comprehensive Analysis of Transcriptome and Metabolome Elucidates the Molecular Regulatory Mechanism of Salt Resistance in Roots of Achnatherum inebrians Mediated by Epichloë gansuensis
by Chao Wang, Rong Huang, Jianfeng Wang, Jie Jin, Kamran Malik, Xueli Niu, Rong Tang, Wenpeng Hou, Chen Cheng, Yinglong Liu and Jie Liu
J. Fungi 2022, 8(10), 1092; https://doi.org/10.3390/jof8101092 - 17 Oct 2022
Cited by 5 | Viewed by 2577
Abstract
Salinization of soil is a major environmental risk factor to plant functions, leading to a reduction of productivity of crops and forage. Epichloë gansuensis, seed-borne endophytic fungi, establishes a mutualistic symbiotic relationship with Achnatherum inebrians and confers salt tolerance in the host [...] Read more.
Salinization of soil is a major environmental risk factor to plant functions, leading to a reduction of productivity of crops and forage. Epichloë gansuensis, seed-borne endophytic fungi, establishes a mutualistic symbiotic relationship with Achnatherum inebrians and confers salt tolerance in the host plants. In this study, analysis of transcriptome and metabolome was used to explore the potential molecular mechanism underlying the salt-adaptation of A. inebrians roots mediated by E. gansuensis. We found that E. gansuensis played an important role in the gene expression of the host’s roots and regulated multiple pathways involved in amino acid metabolism, carbohydrate metabolism, TCA cycle, secondary metabolism, and lipid metabolism in the roots of A. inebrians. Importantly, E. gansuensis significantly induced the biological processes, including exocytosis, glycolytic process, fructose metabolic process, and potassium ion transport in roots of host plants at transcriptional levels, and altered the pathways, including inositol phosphate metabolism, galactose metabolism, starch, and sucrose metabolism at metabolite levels under NaCl stress. These findings provided insight into the molecular mechanism of salt resistance in roots of A. inebrians mediated by E. gansuensis and could drive progress in the cultivation of new salt-resistance breeds with endophytes. Full article
(This article belongs to the Special Issue Fungal Endophytes of Grasses)
Show Figures

Figure 1

Figure 1
<p>Principal component analysis (PCA) score plot showing the differences of transcriptome in roots between <span class="html-italic">E. gansuensis</span>-infected (E+) and <span class="html-italic">E. gansuensis</span>-free (E−) <span class="html-italic">A. inebrians</span> under 0 and 200 mM NaCl concentrations. 0RE+/−: roots of <span class="html-italic">A.inebrians</span> with or without <span class="html-italic">E. gansuensis</span> at 0 mM NaCl; 200RE+/−: roots of <span class="html-italic">A.inebrians</span> with or without <span class="html-italic">E. gansuensis</span> at 200 mM NaCl.</p>
Full article ">Figure 2
<p>The number of up- and down-regulated DEGs in comparative groups of 0RE+ versus 0RE− (abbreviated as 0RE+ vs. 0RE−), 200RE+ vs. 200RE−, 200RE+ vs. 0RE+, and 200RE− vs. 0RE−.</p>
Full article ">Figure 3
<p>Gene ontology (GO) and KEGG pathway enrichment analysis of DEGs that identified between E+ and E− host roots at 0 and 200 mM NaCl concentrations. The GO terms of up-regulated (<b>A</b>) and down-regulated (<b>B</b>) DEGs between 0RE+ and 0RE−, The GO terms of up-regulated (<b>C</b>) and down-regulated (<b>D</b>) DEGs between 200RE+ and 200RE−. “Rich factor” represents the ratio of the enriched DEGs to the total genes annotated in the corresponding pathway. KEGG pathway enrichment analysis of up-regulated (<b>E</b>) and down-regulated (<b>F</b>) DEGs identified between roots of E+ and E− plants under 0 and 200 mM NaCl concentrations. Color panels show the <span class="html-italic">p</span>-value of KEGG pathway enrichment among the comparisons.</p>
Full article ">Figure 4
<p>GO terms and KEGG pathway enrichment analysis based on the NaCl-induced DEGs in roots of both E+ and E− <span class="html-italic">A. inebrians</span>. The GO terms of up-regulated (<b>A</b>) and down-regulated (<b>B</b>) DEGs between 200RE+ and 0RE+. The GO terms of up-regulated (<b>C</b>) and down-regulated (<b>D</b>) DEGs between 200RE− and 0RE−. KEGG pathway enrichment analysis of up-regulated (<b>E</b>) and down-regulated (<b>F</b>) DEGs between 200RE+/0RE+ and 200RE−/0RE−.</p>
Full article ">Figure 5
<p>Clustering of DEGs associated with amino acid metabolism (<b>A</b>), MAPK signaling pathway (<b>B</b>), starch and sucrose metabolism (<b>C</b>) and circadian rhythm (<b>D</b>) between 0RE+ and 0RE− and between 200RE+ and 200RE−. Color panels stand for log<sub>2</sub>(fold change). Asterisks indicate statistically significant changes with FDR &lt; 0.01 and fold change ≥ 2.</p>
Full article ">Figure 6
<p>Impact of 200 mM NaCl on amino acid metabolism (<b>A</b>), starch and sucrose metabolism (<b>B</b>), fructose and mannose metabolism (<b>C</b>), citrate cycle (<b>D</b>), flavonoid biosynthesis (<b>E</b>) and circadian rhythm (<b>F</b>) in roots of E+ and E− <span class="html-italic">A. inebrians</span>. Color panels stand for log<sub>2</sub>(fold change). Asterisks indicate statistically significant changes with FDR &lt; 0.01 and fold change ≥ 2.</p>
Full article ">Figure 7
<p>Bubble plots showing the enrichment analysis of altered metabolic pathways in roots. (<b>A</b>,<b>B</b>) represent the differential pathways between RE+ and RE− under 0 mM and 200 mM NaCl concentrations, respectively. (<b>C</b>,<b>D</b>) represent the differential pathways between 200 mM and 0 mM NaCl treatment in RE+ and RE-, respectively. The plots were organized by pathway enrichment analysis (log<sub>10</sub>(<span class="html-italic">p</span>-values)) and pathway topology analysis (pathway impact).</p>
Full article ">Figure 8
<p>Integration of changes in transcription and metabolite levels mapped to the central metabolic pathways in roots of E+ and E− <span class="html-italic">A. inebrians</span> at 0 and 200 mM NaCl concentrations. Boxes with red and green frames denote the gene expression and metabolites, respectively. The straight arrows represent direct paths, and the dotted arrows represent indirect paths. Asterisks represent statistically significant changes (<span class="html-italic">p</span> ≤ 0.05).</p>
Full article ">
14 pages, 1482 KiB  
Article
A Whole Genome Sequencing-Based Approach to Track down Genomic Variants in Itraconazole-Resistant Species of Aspergillus from Iran
by Sanaz Nargesi, Reza Valadan, Mahdi Abastabar, Saeed Kaboli, Jose Thekkiniath and Mohammad Taghi Hedayati
J. Fungi 2022, 8(10), 1091; https://doi.org/10.3390/jof8101091 - 17 Oct 2022
Cited by 3 | Viewed by 2636
Abstract
The antifungal resistance in non-fumigatus Aspergillus spp., as well as Aspergillus fumigatus, poses a major therapeutic challenge which affects the entire healthcare community. Mutation occurrence of cyp51 gene paralogs is the major cause of azole resistance in Aspergillus spp. To obtain [...] Read more.
The antifungal resistance in non-fumigatus Aspergillus spp., as well as Aspergillus fumigatus, poses a major therapeutic challenge which affects the entire healthcare community. Mutation occurrence of cyp51 gene paralogs is the major cause of azole resistance in Aspergillus spp. To obtain a full map of genomic changes, an accurate scan of the entire length of the Aspergillus genome is necessary. In this study, using whole genome sequencing (WGS) technique, we evaluated the mutation in cyp51A, cyp51B, Cdr1B, AtrR, Hmg1, HapE and FfmA genes in different clinical isolates of Aspergillus fumigatus, Aspergillus niger, Aspergillus tubingensis, Aspergillus welwitschiae and Aspergillus terreus which responded to minimum inhibitory concentrations of itraconazole above 16 µg mL−1. We found different nonsynonymous mutations in the cyp51A, cyp51B, Cdr1B, AtrR, Hmg1, HapE and FfmA gene loci. According to our findings, Aspergillus species isolated from different parts of the world may represent different pattern of resistance mechanisms which may be revealed by WGS. Full article
Show Figures

Figure 1

Figure 1
<p>The steps of project implementation, from sample collection to WGS. The estimated time to complete each of the key stages of the project indicates that after the sampling phase, which spans a period of one year, within two months, it is possible to analyze azole-resistant <span class="html-italic">Aspergillus</span> spp. through whole genome sequencing to identify all SNPs in the entire genomic content.</p>
Full article ">Figure 2
<p>(<b>A</b>) The distribution of minimum inhibitory concentrations (MICs) of different <span class="html-italic">Aspergillus</span> isolates against tested triazoles compared to epidemilogical Cutoff values (ECVs). (<b>B</b>) The MICs of different <span class="html-italic">Aspergillus</span> isolates without defined ECV against tested triazoles.</p>
Full article ">Figure 3
<p>Homology modeling of CYP51A of the selected Aspergillus species and mutation mapping on the structures. Each structure was shown in two different presentations. (<b>A</b>) (<span class="html-italic">A. terreus</span>), (<b>B</b>) (<span class="html-italic">A. tubingensis</span>), (<b>C</b>) (<span class="html-italic">A. niger</span>) (<b>D</b>) (<span class="html-italic">A. fumigatus</span>).</p>
Full article ">
14 pages, 2879 KiB  
Article
FgLEU1 Is Involved in Leucine Biosynthesis, Sexual Reproduction, and Full Virulence in Fusarium graminearum
by Shaohua Sun, Mingyu Wang, Chunjie Liu, Yilin Tao, Tian Wang, Yuancun Liang, Li Zhang and Jinfeng Yu
J. Fungi 2022, 8(10), 1090; https://doi.org/10.3390/jof8101090 - 17 Oct 2022
Cited by 6 | Viewed by 2217
Abstract
Fusarium head blight (FHB) caused by Fusarium graminearum is a significant disease among cereal crops. In F. graminearum, biosynthesis of leucine, which is a branched chain amino acid, is achieved by converting α-isopropylmalate to β-isopropylmalate catalyzed by isopropylmalate isomerase encoded by LEU1 [...] Read more.
Fusarium head blight (FHB) caused by Fusarium graminearum is a significant disease among cereal crops. In F. graminearum, biosynthesis of leucine, which is a branched chain amino acid, is achieved by converting α-isopropylmalate to β-isopropylmalate catalyzed by isopropylmalate isomerase encoded by LEU1. Considering the potential for targeting this pathway by fungicides, we characterized the gene FgLEU1 (FGSG-09589) in the Fusarium graminearum genome using bioinformatics methods. For functional characterization, we constructed a deletion mutant of FgLEU1LEU1) through homologous recombination. Compared with the wild-type strain PH-1, ΔLEU1 showed slower colony growth and fewer aerial mycelia. Leucine addition was needed to ensure proper mutant growth. Further, ΔLEU1 showed decreased conidial production and germination rates, and could not produce ascospores. Moreover, ΔLEU1 showed complete loss of pathogenicity and reduced ability to produce deoxynivalenol (DON) and aurofusarin. Upstream and downstream genes of FgLEU1 were significantly upregulated in ΔLEU1. Contrary to previous reports, the deletion mutant was more resistant to osmotic stress and cell wall-damaging agents than the wild-type. Taken together, FgLEU1 plays a crucial role in leucine synthesis, aerial mycelial growth, sexual and asexual reproduction, pathogenicity, virulence, and pigmentation in Fusarium graminearum, indicating its potential as a target for novel antifungal agents. Full article
(This article belongs to the Special Issue Plant Fungal Pathogenesis 2022)
Show Figures

Figure 1

Figure 1
<p>Sequence and phylogenetic analysis of LEU1. (<b>A</b>) Sequence and phylogenetic analysis of LEU1 in <span class="html-italic">Fusarium graminearum</span> (Fg), <span class="html-italic">Magnaporthe oryzae</span> (Mo), and <span class="html-italic">Saccharomyces cerevisiae</span> (Sc). Aconitase: An enzyme that catalyzes the stereo-specific isomerization of citrate to isocitrate via cis-aconitate in the tricarboxylic acid cycle through a non-redox-active process. Aconitase C: The Aconitase C-terminal domain, which undergoes conformational change in the enzyme mechanism. (<b>B</b>) A phylogenetic tree of LEU1 with homologues from other fungal species was constructed using the neighbor-joining method in MEGA version 7.0. Numbers at the nodes in the rooted tree represent the bootstrap value after 1000 replications. GenBank accession numbers are presented after the fungal species names. The bar indicates 0.05 distance units.</p>
Full article ">Figure 2
<p>Hyphal growth of PH-1, Δ<span class="html-italic">LEU1</span>, and Δ<span class="html-italic">LEU1</span>-C strains. (<b>A</b>) Colonies of the wild type (PH-1), <span class="html-italic">FgLEU1</span> deletion mutants, and complemented strains were cultured on PDA, FGA, and YEPD agar plates after 4 days at 25 °C. (<b>B</b>) Height of aerial mycelium in test tubes containing PDA medium after 4 days at 25 °C. (<b>C</b>) Colony diameters of strains cultured on PDA, FGA, and YEPD plates. (<b>D</b>) Colony height of strains cultured in tubes with PDA. (<b>E</b>) Hyphal quality of strains cultured in potato dextrose broth (PDB) at 25 °C for 5 days at 180 rpm. Different letters on the bars for each treatment indicate significant differences at <span class="html-italic">p</span> &lt; 0.05 by Duncan’s multiple range test. Measurements represent the average of three independent experiments. All experiments were repeated three times with three replicates at each time.</p>
Full article ">Figure 3
<p>Conidial germination and sexual reproduction of <span class="html-italic">FgLEU1</span> deletion mutants. (<b>A</b>) All strains were incubated in liquid YEPD for 4 h at 25 °C with 180 rpm and examined for germination. Scale bar = 20 μm. (<b>B</b>) Sexual development of the strains. Perithecia were visualized as black structures on carrot agar plates. Photographs were taken at 21 days post self-fertilization. All experiments were repeated three times with three replicates at each time.</p>
Full article ">Figure 4
<p>Virulence of the WT strain and <span class="html-italic">FgLEU1</span> deletion mutants on wheat heads and corn silks. (<b>A</b>) Disease symptom on flowering wheat heads caused by PH-1, ∆<span class="html-italic">LEU1</span>, and ∆<span class="html-italic">LEU1</span>-C strains after injection with conidial suspension (10 μL, 4 × 10<sup>5</sup> conidia/mL). Images were taken at 14 days post-inoculation (dpi). (<b>B</b>) Brown necrosis caused by PH-1, ∆<span class="html-italic">LEU1</span>, and ∆<span class="html-italic">LEU1</span>-C strain mycelial plugs on corn silks. Images were taken at 5 dpi at 25 °C. (<b>C</b>) Disease indices of <span class="html-italic">F. graminearum</span> determined at 14 dpi. At least 20 wheat heads were examined in each replicate. Error bars represent the standard errors of the means. Different letters on the bars for each treatment indicate significant differences at <span class="html-italic">p</span> &lt; 0.05 by Duncan’s multiple range test. (<b>D</b>) The length of brown necrotic tissue upon infection was determined at 5 dpi. At least 20 corn silks were examined in each replicate. Different letters on the bars for each treatment indicate significant differences at <span class="html-italic">p</span> &lt; 0.05 by Duncan’s multiple range test.</p>
Full article ">Figure 5
<p>Pigment production and relative expression level of aurofusarin biosynthesis genes by the wild-type strain PH-1 and <span class="html-italic">FgLEU1</span> deletion mutants. (<b>A</b>) Red pigment aurofusarin formation by the wild-type PH-1, <span class="html-italic">FgLEU1</span> deletion mutants, and complementation strains cultured on PDA at 25 °C for 4 days. (<b>B</b>) Red pigment aurofusarin formation by the wild-type PH-1, <span class="html-italic">FgLEU1</span> deletion mutants, and complementation strains cultured in liquid medium in flasks containing 50 mL of PDB at 180 rpm, 25 °C for 5 days. (<b>C</b>) Relative expression levels of aurofusarin biosynthesis-related genes. <span class="html-italic">GAPDH</span> was used as an internal control. Gene expression in PH-1 was set to 1.0 (<span class="html-italic">p</span> &lt; 0.05).</p>
Full article ">Figure 6
<p>The mycelial biomass of wild-type, Δ<span class="html-italic">LEU1</span>, and ∆<span class="html-italic">LEU1</span>-C strains with different concentrations of leucine. Different concentrations of leucine (final concentrations 0, 0.01, 0.05, 0.25, 1.25, and 6.25 mM) were added to fructose gelatin broth medium and the strains were cultured at 25 °C, 180 rpm, for 6 days. Mycelium was collected from the medium, dried at 65 °C and weighed after two days. Error bars represent the standard errors of the means. Different letters on the bars for each treatment indicate significant differences at <span class="html-italic">p</span> &lt; 0.05 by Duncan’s multiple range test.</p>
Full article ">Figure 7
<p>Role of <span class="html-italic">FgLEU1</span> in resistance to cell wall-damaging agents and cell wall-degrading enzymes. (<b>A</b>) Comparison of PH-1, Δ<span class="html-italic">LEU1</span>, and Δ<span class="html-italic">LEU1</span>-C after 4 days of incubation at 25 °C on CM plates with or without 300 mg/mL Congo Red (CR). The plates were photographed at 4 days post cultivation. (<b>B</b>) Mean rates of mycelial growth inhibition by CR on PH-1, Δ<span class="html-italic">LEU1</span>, or Δ<span class="html-italic">LEU1</span>-C strains at 4 days post cultivation. Different letters on the bars for each treatment indicate significant differences at <span class="html-italic">p</span> &lt; 0.05 by Duncan’s multiple range test. Bar = 20 µm. Bars denote standard errors from three experiments. Values on bars followed by the same letter for each treatment indicate absence of a significant difference at <span class="html-italic">p</span> &lt; 0.05. (<b>C</b>) After treatment with lysozyme, driselase, and snailase for 4 h at 30 °C, the mycelia of PH-1 and Δ<span class="html-italic">LEU1</span>-C mutants were well digested and numerous protoplasts were released, but this was not observed for Δ<span class="html-italic">LEU1</span>.</p>
Full article ">
12 pages, 1611 KiB  
Article
Genetic Diversity and Population Structure of Fusarium oxysporum f. sp. conglutinans Race 1 in Northern China Samples
by Jian Ling, Xin Dong, Xingxing Ping, Yan Li, Yuhong Yang, Jianlong Zhao, Xiaofei Lu, Bingyan Xie and Zhenchuan Mao
J. Fungi 2022, 8(10), 1089; https://doi.org/10.3390/jof8101089 - 16 Oct 2022
Cited by 1 | Viewed by 2360
Abstract
Fusarium oxysporum f. sp. conglutinans (FOC), the causal agent of cabbage fusarium wilt, is a serious threat to cabbage production in northern China, and most Chinese FOC isolates were identified as FOC race 1 (FOC1). To better understand the genetic diversity of FOC1 [...] Read more.
Fusarium oxysporum f. sp. conglutinans (FOC), the causal agent of cabbage fusarium wilt, is a serious threat to cabbage production in northern China, and most Chinese FOC isolates were identified as FOC race 1 (FOC1). To better understand the genetic diversity of FOC1 in northern China, we collected FOC isolates from five provinces in northern China and identified them as FOC1 through pathogenicity and race test. To evaluate the genome-level diversity of FOC1, we performed a genome assembly for a FOC1 isolate (FoYQ-1) collected from Yanqing, Beijing, where cabbage fusarium wilt was first reported in China. Using resequencing data of FOC1 isolates, we conducted a genome-wide SNP (single nucleotide polymorphism) analysis to investigate the genetic diversity and population structure of FOC1 isolates in northern China. Our study indicated that Chinese FOC1 can be grouped into four populations and revealed that the genetic diversity of FOC1 were closely associated with geographical locations. Our study further suggests that genetic differentiation occurred when FOC1 spread to the northwest provinces from Beijing Province in China. The FOC1 genetic diversity based on whole-genome SNPs could deepen our understanding of FOC1 variation and provide clues for the control of cabbage fusarium wilt in China. Full article
(This article belongs to the Special Issue Plant-Pathogenic Fusarium Species)
Show Figures

Figure 1

Figure 1
<p>Morphological identification of FOC1 isolates from diseased plants. (<b>a</b>) The diseased cabbage plants show typical symptoms of fusarium wilt including stunts, wilts, and vascular necrosis; (<b>b</b>) observed colony morphology of FOC1 isolates; (<b>c</b>,<b>d</b>) microscopic observation on mycelia and conidia of FOC1 isolates; (<b>e</b>) different cabbage cultivars were used to FoYQ-1 pathogenicity and race tests. FOC1 race1 (52,557) and FOC race2 (58,385) were used as race test control. The upper two lines were FoYQ-1 pathogenicity test, and the middle two lines and the lower two lines were 52,557 and 58,385 pathogenicity test, respectively. 1 Cultivar “Xiaqiang” was resistant (R) to FOC1 but susceptible (S) to FOC2. 2 Cultivar “Zhonggan18” was R to FOC1 but S to FOC2. 3 Cultivar “Huifeng6” was R to FOC1 and R to FOC2. 4 Cultivar “Lutailang” was R to FOC1 and R to FOC2. 5 Cultivar “Zhenqi” was R to FOC1 and R to FOC2. 6 Cultivar “Huifeng7” was R to FOC1 but S to FOC2. 7 Cultivar “Zhonggan96” was R to FOC1 and R to FOC2. 8 Cultivar “Hanchun1” was S to FOC1 and S to FOC2. 9 Cultivar “Zhonggan21” was S to FOC1 and S to FOC2. The pathogenicity of FoYQ-1 isolates was similar to FOC1.</p>
Full article ">Figure 2
<p>Genome feature of FoYQ-1 and the syntenic relationships with the published chromosome-level FOC genome GCA_014839635.1 (<b>a</b>) Genome feature of FoYQ-1. The outermost circle is the contigs. The bar chart from outside to inside in turn is secondary metabolites gene clusters (black), secreted proteins (orange), density of repetitive sequence (blue), and gene density (dark red); (<b>b</b>) the syntenic relationships between two genomes. Y-axis represents FoYQ-1 contigs, and x-axis represents GCA_014839635.1 chromosomes. The axis tick values represent genome size (×10 Mb). The red dot or line represents forward matches, and the blue dot or line represents reverse matches between two genomes.</p>
Full article ">Figure 3
<p>Phylogenetic tree constructed by of EF-1a (<b>a</b>) and ITS (<b>b</b>) gene sequences using the neighbor-joining (NJ) tree method. Red, blue, yellow, black, and green circle represent the FOC1 isolates collected from Beijing, Hebei, Shanxi, Gansu and Shaanxi Provinces, respectively.</p>
Full article ">Figure 4
<p>Population structure of FOC1 in northern China. Red, blue, yellow, black, and green circles represent the FOC1 isolates collected from Beijing, Hebei, Shanxi, Gansu, and Shaanxi Provinces, respectively. (<b>a</b>) Population structure of 26 FOC1 isolates and other six FO using the admixture software. K is the delta K value for population structure; (<b>b</b>) delta K value estimation of sequenced FO; (<b>c</b>) neighbor-joining phylogenetic tree constructed by SNP of the 26 FOC1 isolates and the other 6 FO.</p>
Full article ">
17 pages, 3921 KiB  
Article
The First Telomere-to-Telomere Chromosome-Level Genome Assembly of Stagonospora tainanensis Causing Sugarcane Leaf Blight
by Fu Xu, Xiuxiu Li, Hui Ren, Rensen Zeng, Zhoutao Wang, Hongli Hu, Jiandong Bao and Youxiong Que
J. Fungi 2022, 8(10), 1088; https://doi.org/10.3390/jof8101088 - 16 Oct 2022
Cited by 6 | Viewed by 3766
Abstract
The sexual morph Leptosphaeria taiwanensis Yen and Chi and its asexual morph Stagonospora tainanensis W. H. Hsieh is an important necrotrophic fungal phytopathogen, which causes sugarcane leaf blight, resulting in loss of cane tonnage and sucrose in susceptible sugarcane varieties. Decoding the genome [...] Read more.
The sexual morph Leptosphaeria taiwanensis Yen and Chi and its asexual morph Stagonospora tainanensis W. H. Hsieh is an important necrotrophic fungal phytopathogen, which causes sugarcane leaf blight, resulting in loss of cane tonnage and sucrose in susceptible sugarcane varieties. Decoding the genome and understanding of the basis of virulence is vitally important for devising effective disease control strategies. Here, we present a 38.25-Mb high-quality genome assembly of S. tainanensis strain StFZ01, denovo assembled with 10.19 Gb Nanopore sequencing long reads (~267×) and 3.82 Gb Illumina short reads (~100×). The genome assembly consists of 12 contigs with N50 of 2.86 Mb of which 5 belong to the telomere to telomere (T2T) chromosome. It contains 13.20% repeat sequences, 12,543 proteins, and 12,206 protein-coding genes with the BUSCO completeness 99.18% at fungi (n = 758) and 99.87% at ascomycota (n = 1706), indicating the high accuracy and completeness of our gene annotations. The virulence analysis in silico revealed the presence of 2379 PHIs, 599 CAZys, 248 membrane transport proteins, 191 cytochrome P450 enzymes, 609 putative secreted proteins, and 333 effectors in the StFZ01 genome. The genomic resources presented here will not only be helpful for development of specific molecular marker and diagnosis technique, population genetics, molecular taxonomy, and disease managements, it can also provide a significant precise genomic reference for investigating the ascomycetous genome, the necrotrophic lifestyle, and pathogenicity in the future. Full article
(This article belongs to the Special Issue Genomics of Fungal Plant Pathogens)
Show Figures

Figure 1

Figure 1
<p>The species of strain StFZ01 was verified as <span class="html-italic">Stagonospora tainanensis</span> by maximum likelihood phylogenetic tree analysis (bootstrap = 1000) conducted by MEGA v11 (<a href="https://www.megasoftware.net/" target="_blank">https://www.megasoftware.net/</a>) (accessed on 17 August 2022) based on ITS sequences with high similarity collected from NCBI. <a href="#app1-jof-08-01088" class="html-app">Table S1</a> Summary of sequencing reads.</p>
Full article ">Figure 2
<p>The phenotypic symptoms of sugarcane leaf blight (SLB) on sugarcane cultivar Yuetang93-159 initiated by <span class="html-italic">Stagonospora tainanensis</span> and the morphological characteristics of the strain StFZ01. (<b>A</b>) The phenotypic symptoms of SLB on plants. (<b>B</b>) Development of the lesions of SLB on leaves. (<b>C</b>) The pathogenic colony and mycelia. (<b>D</b>) The pathogenic conidia. (<b>E</b>) The sexual asci and ascospores.</p>
Full article ">Figure 3
<p>Genome features of <span class="html-italic">Stagonospora tainanensis</span> strain StFZ01. (<b>A</b>) Summary of ONT reads. (<b>B</b>) Genome size estimation with NGS genomic reads; (<b>C</b>) Circos plot of genome assembly features. Circles from outside to inside present contigs (1st circle, the smallest contig ctg12 was not shown), distribution of protein-coding genes (2nd), TEs (3rd), and putative secreted proteins (4rd) per 50 kb window size (color blue to red means number from low to high). The lines in the center of circle show the synteny blocks (≥10 kb) between different contigs.</p>
Full article ">Figure 4
<p>Genome completeness assessment of <span class="html-italic">Stagonospora tainanensis</span> strain StFZ01. (<b>A</b>) BUSCO completeness assessment of genome assembly and annotated proteins with the lineage dataset of fungi_odb10 and ascomycota_odb10. G means genome and P means proteins. (<b>B</b>) Genome completeness valued by mapping rate of different reads.</p>
Full article ">Figure 5
<p>General gene functional annotation. (<b>A</b>) KOG annotation and (<b>B</b>) Pfam annotation (top20). Red color indicates the top terms associated with candidate pathogenicity-related genes.</p>
Full article ">Figure 6
<p>Summary of pathogenicity-related gene annotations. (<b>A</b>) CAZys, (<b>B</b>) Membrane transport proteins (top10), (<b>C</b>) PHIs, and (<b>D</b>) Putative secreted proteins.</p>
Full article ">Figure 7
<p>Profiles of secondary metabolite biosynthetic gene clusters. (<b>A</b>) Type of SMBGCs. (<b>B</b>) Contig distribution of SMBGCs. (<b>C</b>) Most similarity with known SMBGCs (similarity ≥ 50%).</p>
Full article ">Figure 8
<p>Comparative genomics analysis with close species. (<b>A</b>) Genes clustered in orthogroups. (<b>B</b>) Summary of orthogroups among species. Species specific genes (white), total orthogroups (black, in the red dotted circle), and core orthogroups (red) were shown from outside to inside of flower. (<b>C</b>) Phylogenetic tree inferred with alignment of single-copy core orthogroups.</p>
Full article ">
15 pages, 1486 KiB  
Review
Management of Plant Beneficial Fungal Endophytes to Improve the Performance of Agroecological Practices
by Bouchra Nasslahsen, Yves Prin, Hicham Ferhout, Abdelaziz Smouni and Robin Duponnois
J. Fungi 2022, 8(10), 1087; https://doi.org/10.3390/jof8101087 - 15 Oct 2022
Cited by 4 | Viewed by 2742
Abstract
By dint of the development of agroecological practices and organic farming, stakeholders are becoming more and more aware of the importance of soil life and banning a growing number of pesticide molecules, promoting the use of plant bio-stimulants. To justify and promote the [...] Read more.
By dint of the development of agroecological practices and organic farming, stakeholders are becoming more and more aware of the importance of soil life and banning a growing number of pesticide molecules, promoting the use of plant bio-stimulants. To justify and promote the use of microbes in agroecological practices and sustainable agriculture, a number of functions or services often are invoked: (i) soil health, (ii) plant growth promotion, (iii) biocontrol, (iv) nutrient acquiring, (v) soil carbon storage, etc. In this paper, a review and a hierarchical classification of plant fungal partners according to their ecosystemic potential with regard to the available technologies aiming at field uses will be discussed with a particular focus on interactive microbial associations and functions such as Mycorrhiza Helper Bacteria (MHB) and nurse plants. Full article
(This article belongs to the Special Issue New Perspectives on Fungal Endophytes Research)
Show Figures

Figure 1

Figure 1
<p>Strategies for managing the mycorrhizal infection potential (MIP) according to the extent of degradation (resilience threshold) of the environment to be remediated. Holistic approach: increased MIP via biological vectors (cover plants, nursery plants, etc.). Reductionist approach: mass introduction of mycorrhizal or MHB inoculant into the environment to be remediated (controlled mycorrhization technique).</p>
Full article ">Figure 2
<p>Growing hyphae at the periphery of a Tuber melanosporum colony, ensheathed in <span class="html-italic">Rhodopseudomonas</span> sp. colonies [<a href="#B9-jof-08-01087" class="html-bibr">9</a>].</p>
Full article ">Figure 3
<p>(<b>A</b>) Crude extract of spores of <span class="html-italic">Glomerycota</span> from a soil under argan tree in Morrocco. Extraction procedure using a saccharose gradient according to Brundrett et al. [<a href="#B18-jof-08-01087" class="html-bibr">18</a>]. (<b>B</b>–<b>E</b>) Some of the spores of the crude extract obtained after sorting under a stereomicroscope.</p>
Full article ">Figure 4
<p>Longitudinal section of a <span class="html-italic">Piptadenia</span> sp. nitrogen-fixing nodule with numerous spores and hyphae [<a href="#B51-jof-08-01087" class="html-bibr">51</a>].</p>
Full article ">
19 pages, 2260 KiB  
Article
Genetic Diversity of Epichloë Endophytes Associated with Brachypodium and Calamagrostis Host Grass Genera including Two New Species
by Adrian Leuchtmann and Christopher L. Schardl
J. Fungi 2022, 8(10), 1086; https://doi.org/10.3390/jof8101086 - 15 Oct 2022
Cited by 3 | Viewed by 2225
Abstract
Fungi of genus Epichloë (Ascomycota, Clavicipitaceae) are common endophytic symbionts of Poaceae, including wild and agronomically important cool-season grass species (subfam. Poöideae). Here, we examined the genetic diversity of Epichloë from three European species of Brachypodium (B. sylvaticum, B. pinnatum and B. [...] Read more.
Fungi of genus Epichloë (Ascomycota, Clavicipitaceae) are common endophytic symbionts of Poaceae, including wild and agronomically important cool-season grass species (subfam. Poöideae). Here, we examined the genetic diversity of Epichloë from three European species of Brachypodium (B. sylvaticum, B. pinnatum and B. phoenicoides) and three species of Calamagrostis (C. arundinacea, C. purpurea and C. villosa), using DNA sequences of tubB and tefA genes. In addition, microsatellite markers were obtained from a larger set of isolates from B. sylvaticum sampled across Europe. Based on phylogenetic analyses the isolates from Brachypodium hosts were placed in three different subclades within the Epichloë typhina complex (ETC) but did not strictly group according to host grass species, suggesting that the host does not always select for particular endophyte genotypes. Analysis of microsatellite markers confirmed the presence of genetically distinct lineages of Epichloësylvatica on B. sylvaticum, which appeared to be tied to different modes of reproduction (sexual or asexual). Among isolates from Calamagrostis hosts, two subclades were detected which were placed outside ETC. These endophyte lineages are recognized as distinct species for which we propose the names E. calamagrostidis Leuchtm. & Schardl, sp. nov. and E. ftanensis Leuchtm. & A.D. Treindl, sp. nov. This study extends knowledge of the phylogeny and evolutionary diversification of Epichloë endophytes that are symbionts of wild Brachypodium and Calamagrostis host grasses. Full article
(This article belongs to the Special Issue Fungal Endophytes of Grasses)
Show Figures

Figure 1

Figure 1
<p>Location of collection sites in Europe of <span class="html-italic">Epichloë</span>-infected <span class="html-italic">Brachypodium</span> species: <span class="html-italic">B. sylvaticum</span> (Bs), <span class="html-italic">B. pinnatum</span> (Bp) and <span class="html-italic">B. phoenicoides</span> (Bph). Symbols distinguish between host, and stroma-forming (sexual) or symptomless (asexual) infections.</p>
Full article ">Figure 2
<p>Principal coordinate analysis (PCoA) based on 16 microsatellite loci of <span class="html-italic">Epichloë</span> isolates from <span class="html-italic">Brachypodium</span> hosts: <span class="html-italic">B. sylvaticum</span> (Bss, sexual isolates; Bsa, asexual isolates), <span class="html-italic">B. pinnatum</span> (Bp) and <span class="html-italic">B. phoenicoides</span> (Bph).</p>
Full article ">Figure 3
<p>Phylogenetic tree inferred from maximum likelihood (ML) analysis of partial <span class="html-italic">tubB</span> gene sequences including introns 1–3 of <span class="html-italic">Epichloë</span> isolates obtained from <span class="html-italic">Brachypodium</span> and <span class="html-italic">Calamagrostis</span> host grasses: <span class="html-italic">B. sylvaticum</span> (Bs), <span class="html-italic">B. pinnatum</span> (Bp), <span class="html-italic">B. phoenicoides</span> (Bph), <span class="html-italic">C. arundinacea</span> (Ca), <span class="html-italic">C. purpurea</span> (Cp) and <span class="html-italic">C. villosa</span> (Cv). Asexual isolates are indicated with filled squares. representative sequences of all other sexual <span class="html-italic">Epichloë</span> species or subspecies described to date are included. Dashed lines with Roman numerals denote distinct subclades. The tree is midpoint rooted at the left edge. Branch support values were estimated by 100 ML bootstrap replicates and are indicated if above 50%.</p>
Full article ">Figure 4
<p>Phylogenetic tree inferred from maximum likelihood (ML) analysis of partial <span class="html-italic">tefA</span> gene sequences including introns 1–4 of <span class="html-italic">Epichloë</span> isolates obtained from <span class="html-italic">Brachypodium</span> and <span class="html-italic">Calamagrostis</span> host grasses: <span class="html-italic">B. sylvaticum</span> (Bs), <span class="html-italic">B. pinnatum</span> (Bp), <span class="html-italic">B. phoenicoides</span> (Bph), <span class="html-italic">C. arundinacea</span> (Ca), <span class="html-italic">C. purpurea</span> (Cp) and <span class="html-italic">C. villosa</span> (Cv). Asexual isolates are indicated with filled squares. Included are representative sequences of all other sexual <span class="html-italic">Epichloë</span> species or subspecies described to date. Dashed lines with Roman numerals denote distinct subclades. The tree is midpoint rooted at the left edge. Branch support values were estimated by 100 ML bootstrap replicates and are indicated if above 50%.</p>
Full article ">Figure 5
<p><span class="html-italic">Epichloë calamagrostidis</span>. (<b>A</b>). Colony grown on PDA for 21 days at 24 °C. (<b>B</b>). Conidiogenous cell with emerging conidium. (<b>C</b>,<b>D</b>). Conidia. (<b>E</b>). Stroma with mature perithecia on <span class="html-italic">C. varia</span> (coll. H. Seitter). (<b>F</b>). Sporulating hypha in culture.</p>
Full article ">Figure 6
<p><span class="html-italic">Epichloë ftanensis</span>. (<b>A</b>) Colony grown on PDA for 21 days at 24 °C. (<b>B</b>) Conidiogenous cell with emerging conidium. (<b>C</b>–<b>D</b>) Conidia. (<b>E</b>) Stromata with mature perithecia (holotype).</p>
Full article ">
31 pages, 6583 KiB  
Review
Guanidine-Containing Antifungal Agents against Human-Relevant Fungal Pathogens (2004–2022)—A Review
by Simon D. P. Baugh
J. Fungi 2022, 8(10), 1085; https://doi.org/10.3390/jof8101085 - 15 Oct 2022
Cited by 3 | Viewed by 2935
Abstract
The guanidine moiety is typically a highly basic group, and can be found in a wide variety of drugs, such as zanamivir (Relenza) and metformin (Fortamet), as well as in biologically active compounds for numerous disease areas, including central nervous system (CNS) diseases [...] Read more.
The guanidine moiety is typically a highly basic group, and can be found in a wide variety of drugs, such as zanamivir (Relenza) and metformin (Fortamet), as well as in biologically active compounds for numerous disease areas, including central nervous system (CNS) diseases and chemotherapeutics. This review will focus on antifungal agents which contain at least one guanidine group, for the treatment of human-related fungal pathogens, described in the literature between 2004 and 2022. These compounds include small molecules, steroids, polymers, metal complexes, sesquiterpenes, natural products, and polypeptides. It shall be made clear that a diverse range of guanidine-containing derivatives have been published in the literature and have antifungal activity, including efficacy in in vivo experiments. Full article
(This article belongs to the Special Issue Antifungal Drug Discovery: Novel Therapies and Approaches)
Show Figures

Figure 1

Figure 1
<p>Structures of representative guanidine-containing approved drugs. zanamivir (Relenza) (<b>1</b>), famotidine (Pepcid), (<b>2</b>), and tizanidine (Zanaflex) (<b>3</b>).</p>
Full article ">Figure 2
<p>Structure of mirabilin B (<b>4</b>).</p>
Full article ">Figure 3
<p>Phenyl guanidinium salt antifungal agents (<b>5</b>) and (<b>6</b>).</p>
Full article ">Figure 4
<p>Pyrrole-diguanide antifungal derivative (<b>7</b>).</p>
Full article ">Figure 5
<p>Guanidine-containing small molecules (<b>8</b>) and (<b>9</b>) separated from guazatine.</p>
Full article ">Figure 6
<p>Cyclized derivatives (<b>10</b>) and (<b>11</b>) related to guazatine.</p>
Full article ">Figure 7
<p>Fluorescent probe derivative (<b>12</b>) based on a cyclized analog of guazatine.</p>
Full article ">Figure 8
<p>Cyclic guazatine-related analogs (<b>13</b>), (<b>14</b>), and (<b>15</b>).</p>
Full article ">Figure 9
<p>Arylated derivatives (<b>16</b>), (<b>17</b>), and (<b>18</b>) of cyclized guazatine-derived analog (<b>13</b>).</p>
Full article ">Figure 10
<p>Drimenol-derived antifungal compound (<b>19</b>).</p>
Full article ">Figure 11
<p>Structure of abafungin (<b>20</b>).</p>
Full article ">Figure 12
<p>Fungistatic and fungicidal piperazine-1-carboxamidines (<b>21</b>) and (<b>22</b>).</p>
Full article ">Figure 13
<p>Structure of polyhexamethylene-guanidine hydrochloride (PHMGH) (<b>23</b>).</p>
Full article ">Figure 14
<p>Fungicidal benzylsulfanyl-phenylguanidines (<b>24</b>) and (<b>25</b>).</p>
Full article ">Figure 15
<p>Mannich base antifungal derivative (<b>26</b>).</p>
Full article ">Figure 16
<p>Copper (II) complex antifungal agent (<b>27</b>).</p>
Full article ">Figure 17
<p>Antifungal derivatives (<b>28</b>) and (<b>29</b>) of batzelladine K.</p>
Full article ">Figure 18
<p>Structure of cabanillasin (<b>30</b>).</p>
Full article ">Figure 19
<p>Ferrocene-based guanidine-containing antifungal agents (<b>31</b>), (<b>32</b>), and (<b>33</b>).</p>
Full article ">Figure 20
<p>Short cationic antimicrobial peptide (LTX-109, AMC-109) (<b>34</b>).</p>
Full article ">Figure 21
<p>Lead thiophene-based bis-guanylhydrazone antifungal compound (<b>35</b>).</p>
Full article ">Figure 22
<p>Furan-based bis-guanylhydrazone antifungal derivative (<b>36</b>).</p>
Full article ">Figure 23
<p>Lead guanidine-containing benzothiazole derivatives (<b>37</b>) and (<b>38</b>).</p>
Full article ">Figure 24
<p>α-Mangostin derived antifungal analog (<b>39</b>).</p>
Full article ">Figure 25
<p>Structure of alexidine dihydrochloride (<b>40</b>).</p>
Full article ">Figure 26
<p>Structure of metformin (<b>41</b>).</p>
Full article ">Figure 27
<p>Bis-guanidine containing oxadiazole analog (<b>42</b>).</p>
Full article ">Figure 28
<p>1,2,4-triazole bis-guanidine-containing antifungal derivatives (<b>43</b>) and (<b>44</b>).</p>
Full article ">Figure 29
<p>Lead thiazole-aminoguanidine analog (<b>45</b>).</p>
Full article ">Figure 30
<p>Aminoguanidine-containing antifungal compounds (<b>46</b>) and (<b>47</b>).</p>
Full article ">Figure 31
<p>Guanidine-containing GlcN-6-P synthase inhibitor (<b>48</b>).</p>
Full article ">Figure 32
<p>Guanidine-containing betulinic acid derived antifungal analogs (<b>49</b>) and (<b>50</b>).</p>
Full article ">Figure 33
<p>Surfactants (<b>51</b>) and (<b>52</b>) containing guanidine groups as antifungal agents.</p>
Full article ">Figure 34
<p>Guanidine-containing bis-thiazole analog antifungal agent (<b>53</b>).</p>
Full article ">Figure 35
<p>Optimized antifungal polypeptide K-oLBF127 (<b>54</b>).</p>
Full article ">Figure 36
<p>Peptide-heterocycle conjugate antifungal derivative (<b>55</b>).</p>
Full article ">Figure 37
<p>28-Mer antifungal polymer (<b>56</b>).</p>
Full article ">Figure 38
<p>Antifungal cyclic pentapeptide (<b>57</b>).</p>
Full article ">Figure 39
<p>Structure of crambescidic acid-671 (<b>58</b>).</p>
Full article ">Figure 40
<p>General structure of guanylated lysine-based polymer (<b>59</b>).</p>
Full article ">
Previous Issue
Next Issue
Back to TopTop