[go: up one dir, main page]

 
 
Sign in to use this feature.

Years

Between: -

Subjects

remove_circle_outline
remove_circle_outline
remove_circle_outline
remove_circle_outline
remove_circle_outline
remove_circle_outline
remove_circle_outline
remove_circle_outline
remove_circle_outline

Journals

remove_circle_outline
remove_circle_outline
remove_circle_outline
remove_circle_outline
remove_circle_outline
remove_circle_outline
remove_circle_outline
remove_circle_outline
remove_circle_outline
remove_circle_outline
remove_circle_outline
remove_circle_outline
remove_circle_outline
remove_circle_outline
remove_circle_outline
remove_circle_outline
remove_circle_outline
remove_circle_outline
remove_circle_outline
remove_circle_outline
remove_circle_outline
remove_circle_outline
remove_circle_outline
remove_circle_outline
remove_circle_outline
remove_circle_outline
remove_circle_outline
remove_circle_outline
remove_circle_outline

Article Types

Countries / Regions

remove_circle_outline
remove_circle_outline
remove_circle_outline
remove_circle_outline
remove_circle_outline
remove_circle_outline
remove_circle_outline
remove_circle_outline
remove_circle_outline
remove_circle_outline
remove_circle_outline

Search Results (6,843)

Search Parameters:
Keywords = antifungal

Order results
Result details
Results per page
Select all
Export citation of selected articles as:
17 pages, 2183 KiB  
Review
Polyene-Based Derivatives with Antifungal Activities
by Kwanele Ngece, Thabisa L. Ntondini, Vuyolwethu Khwaza, Athandwe M. Paca and Blessing A. Aderibigbe
Pharmaceutics 2024, 16(8), 1065; https://doi.org/10.3390/pharmaceutics16081065 (registering DOI) - 14 Aug 2024
Abstract
Polyenes are a class of organic compounds well known for their potent antifungal properties. They are effective due to their ability to target and disrupt fungal cell membranes by binding to ergosterol and forming pores. Despite their effectiveness as antifungal drugs, polyenes have [...] Read more.
Polyenes are a class of organic compounds well known for their potent antifungal properties. They are effective due to their ability to target and disrupt fungal cell membranes by binding to ergosterol and forming pores. Despite their effectiveness as antifungal drugs, polyenes have several limitations, such as high toxicity to the host cell and poor solubility in water. This has prompted ongoing research to develop safer and more efficient derivatives to overcome such limitations while enhancing their antifungal activity. In this review article, we present a thorough analysis of polyene derivatives, their structural modifications, and their influence on their therapeutic effects against various fungal strains. Key studies are discussed, illustrating how structural modifications have led to improved antifungal properties. By evaluating the latest advancements in the synthesis of polyene derivatives, we highlight that incorporating amide linkers at the carboxylic moiety of polyene molecules notably improves their antifungal properties, as evidenced by derivatives 4, 5, 6G, and 18. This review can help in the design and development of novel polyene-based compounds with potent antifungal activities. Full article
(This article belongs to the Special Issue Emerging Pharmaceutical Strategies against Infectious Diseases)
Show Figures

Graphical abstract

Graphical abstract
Full article ">Figure 1
<p>Chemical structures of polyene antibiotics: <b>AmB</b>(<b>1</b>), <b>nystatin A1</b>(<b>2</b>), and <b>natamycin</b>(<b>3</b>).</p>
Full article ">Figure 2
<p>Structure of dual modified derivative C16-DMAE-amide (compounds <b>4</b> and <b>5</b>).</p>
Full article ">Figure 3
<p>Structure of compound <b>6</b> (amphamide) and its salt <b>6G</b>.</p>
Full article ">Figure 4
<p>Structure of AmB derivatives containing a salicylic acid moiety.</p>
Full article ">Figure 5
<p>Structure of compound <b>8</b>, an <b>AmB</b> derivative.</p>
Full article ">Figure 6
<p>Structures of nystatin derivatives (<b>Nyt 9–11</b>).</p>
Full article ">Figure 7
<p>Structure of nystatin derivative <b>12</b>.</p>
Full article ">Figure 8
<p>Structures of promising <b>AmB</b> (<b>13</b>) and nystatin (<b>14</b>) derivatives.</p>
Full article ">Figure 9
<p>Structures of benzofuroxane hybrids and polyene antibiotics (<b>AmB</b> and <b>Nys</b>).</p>
Full article ">Figure 10
<p>Structure of natamycin derivative compound <b>17</b>.</p>
Full article ">Figure 11
<p>Structure of amide derivatives of <b>natamycin</b>.</p>
Full article ">
15 pages, 2249 KiB  
Article
Isolation of Antagonistic Endophytic Fungi from Postharvest Chestnuts and Their Biocontrol on Host Fungal Pathogens
by Yunmin Wen, Meng Li, Shuzhen Yang, Litao Peng, Gang Fan and Huilin Kang
J. Fungi 2024, 10(8), 573; https://doi.org/10.3390/jof10080573 - 14 Aug 2024
Viewed by 47
Abstract
In this study, antagonistic endophytic fungi were isolated from postharvest chestnut fruits; endophytic antagonistic fungi and their combination of inhibitory effects on the fungal pathogen Neofusicoccum parvum were evaluated. A total of 612 endophytic fungi were isolated from 300 healthy chestnut kernels, and [...] Read more.
In this study, antagonistic endophytic fungi were isolated from postharvest chestnut fruits; endophytic antagonistic fungi and their combination of inhibitory effects on the fungal pathogen Neofusicoccum parvum were evaluated. A total of 612 endophytic fungi were isolated from 300 healthy chestnut kernels, and 6 strains out of them including NS-3, NS-11, NS-38, NS-43, NS-56, and NS-58 were confirmed as antagonistic endophytic fungi against Neofusicoccum parvum; these were separately identified as Penicillium chermesinum, Penicillium italicum, Penicillium decaturense, Penicillium oxalicum, Talarmyces siamensis, and Penicillium guanacastense. Some mixed antagonistic endophytic fungi, such as NS-3-38, NS-11-38, NS-43-56, and NS-56-58-38, exhibited a much stronger antifungal activity against N. parvum than that applied individually. Among them, the mixture of NS-3-38 showed the highest antifungal activity, and the inhibition rate was up to 86.67%. The fermentation broth of NS-3, NS-38, and their combinations exhibited an obvious antifungal activity against N. parvum, and the ethyl acetate phase extract of NS-3-38 had the strongest antifungal activity, for which the inhibitory rate was up to 90.19%. The NS-3-38 fermentation broth combined with a chitosan coating significantly reduced N. parvum incidence in chestnuts from 100% to 19%. Furthermore, the fruit decay and weight loss of chestnuts during storage were significantly decreased by the NS-3-38 fermentation broth mixture along with a chitosan coating. Therefore, a mixture of P. chermesinum and P. decaturense could be used as a potential complex biocontrol agent to control postharvest fruit decay in chestnuts. Full article
(This article belongs to the Special Issue Control of Postharvest Fungal Diseases)
Show Figures

Figure 1

Figure 1
<p>Pathogenicity assay of endophytic fungi from postharvest chestnuts using artificial inoculation. CK stands for PDA inoculation only, without endophytic fungi.</p>
Full article ">Figure 2
<p>Morphology analyses of colony, hypha, and spores from endophytic antagonistic fungi from postharvest chestnuts. (<b>A</b>) NS-3; (<b>B</b>) NS-11; (<b>C</b>) NS-38; (<b>D</b>) NS-43; (<b>E</b>) NS-56; (<b>F</b>) NS-58. Microscope with 40× magnification.</p>
Full article ">Figure 3
<p>Phylogenetic tree constructed based on ITS and BenA gene sequences. (<b>A</b>) NS-3; (<b>B</b>) NS-11; (<b>C</b>) NS-38; (<b>D</b>) NS-43; (<b>E</b>) NS-56; (<b>F</b>) NS-58.</p>
Full article ">Figure 4
<p>Antagonistic activity and inhibitory rate of the culture filtrate from endophytic fungi and its fractions against <span class="html-italic">N. parvum</span>. (<b>A</b>,<b>B</b>) culture filtrate; (<b>C</b>,<b>D</b>) fraction extracted using petroleum ether extract treatment; (<b>E</b>,<b>F</b>) fraction extracted using ethyl acetate extract treatment; (<b>G</b>,<b>H</b>) fraction extracted using 1-butanol. Different letters (a, b, c) above the columns indicate significant difference between the groups (<span class="html-italic">p</span> &lt; 0.05).</p>
Full article ">Figure 5
<p>Effects of endophytic antagonistic fungi culture filtrate combined with chitosan coating on disease index of chestnut fruits caused by <span class="html-italic">N. parvum</span>. (<b>A</b>) Appearance of chestnut fruits inoculated with <span class="html-italic">N. parvum</span>. (<b>B</b>) Disease index of chestnut fruits.Different letters (a, b, c, d and e) above the columns indicate significant difference between the groups (<span class="html-italic">p</span> &lt; 0.05).</p>
Full article ">Figure 6
<p>Effects of endophytic antagonistic fungi culture filtrate combined with chitosan coating on decay incidence and weight loss rate of chestnut during storage. (<b>A</b>) Decay incidence; (<b>B</b>) weight loss rate.Different letters (a, b, c, d) above the columns indicate significant difference between the groups (<span class="html-italic">p</span> &lt; 0.05).</p>
Full article ">
27 pages, 4952 KiB  
Review
Green Innovation and Synthesis of Honeybee Products-Mediated Nanoparticles: Potential Approaches and Wide Applications
by Shaden A. M. Khalifa, Aya A. Shetaia, Nehal Eid, Aida A. Abd El-Wahed, Tariq Z. Abolibda, Abdelfatteh El Omri, Qiang Yu, Mohamed A. Shenashen, Hidayat Hussain, Mohamed F. Salem, Zhiming Guo, Abdulaziz M. Alanazi and Hesham R. El-Seedi
Bioengineering 2024, 11(8), 829; https://doi.org/10.3390/bioengineering11080829 - 14 Aug 2024
Viewed by 84
Abstract
Bee products, abundant in bioactive ingredients, have been utilized in both traditional and contemporary medicine. Their antioxidant, antimicrobial, and anti-inflammatory properties make them valuable for food, preservation, and cosmetics applications. Honeybees are a vast reservoir of potentially beneficial products such as honey, bee [...] Read more.
Bee products, abundant in bioactive ingredients, have been utilized in both traditional and contemporary medicine. Their antioxidant, antimicrobial, and anti-inflammatory properties make them valuable for food, preservation, and cosmetics applications. Honeybees are a vast reservoir of potentially beneficial products such as honey, bee pollen, bee bread, beeswax, bee venom, and royal jelly. These products are rich in metabolites vital to human health, including proteins, amino acids, peptides, enzymes, sugars, vitamins, polyphenols, flavonoids, and minerals. The advancement of nanotechnology has led to a continuous search for new natural sources that can facilitate the easy, low-cost, and eco-friendly synthesis of nanomaterials. Nanoparticles (NPs) are actively synthesized using honeybee products, which serve dual purposes in preventive and interceptive treatment strategies due to their richness in essential metabolites. This review aims to highlight the potential role of bee products in this line and their applications as catalysts and food preservatives and to point out their anticancer, antibacterial, antifungal, and antioxidant underlying impacts. The research used several online databases, namely Google Scholar, Science Direct, and Sci Finder. The overall findings suggest that these bee-derived substances exhibit remarkable properties, making them promising candidates for the economical and eco-friendly production of NPs. Full article
Show Figures

Graphical abstract

Graphical abstract
Full article ">Figure 1
<p>Flowchart of the systematic literature review strategy. n: Number of published papers.</p>
Full article ">Figure 2
<p>Deciphering the role of bee product metabolites in eco-friendly nanoparticle synthesis.</p>
Full article ">Figure 3
<p>Honey mediated nanoparticles and their possible application (figure used with permission) [<a href="#B31-bioengineering-11-00829" class="html-bibr">31</a>].</p>
Full article ">Figure 4
<p>Diagrammatic illustration of the eco-friendly production of AgNPs using royal jelly (<b>A</b>), and the possible antibacterial action mechanism of AgNPs against both Gram-negative and Gram-positive bacteria (<b>B</b>) (Figure used with permission) [<a href="#B97-bioengineering-11-00829" class="html-bibr">97</a>].</p>
Full article ">Figure 5
<p>(<b>A</b>): Synthesis of AgNPs using <span class="html-italic">Cupressus sempervirens</span> pollen extract [<a href="#B55-bioengineering-11-00829" class="html-bibr">55</a>] and (<b>B</b>): ZnO synthesis using royal jelly (figure used with permission) [<a href="#B98-bioengineering-11-00829" class="html-bibr">98</a>].</p>
Full article ">Figure 6
<p>(<b>A</b>) The overall diagrammatic preparation of CuNPs derived from Honey, and (<b>B</b>) illustrates the anti-cancer properties of the synthesized CuNPs (Figure used with permission) [<a href="#B114-bioengineering-11-00829" class="html-bibr">114</a>].</p>
Full article ">Figure 7
<p>Highlighting the diverse biological impacts of nanoparticles derived from honeybee products.</p>
Full article ">
12 pages, 1555 KiB  
Article
Invasive Pulmonary Aspergillosis in Patients with HBV-Related Acute on Chronic Liver Failure
by Man Yuan, Ning Han, Duoduo Lv, Wei Huang, Mengjie Zhou, Libo Yan and Hong Tang
J. Fungi 2024, 10(8), 571; https://doi.org/10.3390/jof10080571 - 14 Aug 2024
Viewed by 108
Abstract
Background: We aim to investigate the characteristics of invasive pulmonary aspergillosis (IPA) in patients with HBV-related acute on chronic liver failure (HBV-ACLF). Methods: A total of 44 patients with probable IPA were selected as the case group, and another 88 patients without lung [...] Read more.
Background: We aim to investigate the characteristics of invasive pulmonary aspergillosis (IPA) in patients with HBV-related acute on chronic liver failure (HBV-ACLF). Methods: A total of 44 patients with probable IPA were selected as the case group, and another 88 patients without lung infections were chosen as the control group. Results: HBV-ACLF patients with probable IPA had more significant 90-day mortality (38.6% vs. 15.9%, p = 0.0022) than those without. The white blood cell (WBC) count was the independent factor attributed to the IPA development [odds ratio (OR) 1.468, p = 0.027]. Respiratory failure was associated with the mortality of HBV-ACLF patients with IPA [OR 26, p = 0.000]. Twenty-seven patients received voriconazole or voriconazole plus as an antifungal treatment. Plasma voriconazole concentration measurements were performed as therapeutic drug monitoring in 55.6% (15/27) of the patients. The drug concentrations exceeded the safe range with a reduced dosage. Conclusions: The WBC count might be used to monitor patients’ progress with HBV-ACLF and IPA. The presence of IPA increases the 90-day mortality of HBV-ACLF patients mainly due to respiratory failure. An optimal voriconazole regimen is needed for such critical patients, and voriconazole should be assessed by closely monitoring blood levels. Full article
(This article belongs to the Special Issue Diagnosis of Invasive Fungal Diseases)
Show Figures

Figure 1

Figure 1
<p>IPA criteria.</p>
Full article ">Figure 2
<p>Flow chart of patient selection.</p>
Full article ">Figure 3
<p>Impact of IPA on mortality. Cumulative 90-day mortality of patients with or without IPA. Log-rank test (38.6% vs. 15.9%, <span class="html-italic">p</span> = 0.0022).</p>
Full article ">Figure 4
<p>Voriconazole regimen for IPA. (<b>a</b>) A cumulative 90-day survival of patients with or without voriconazole regimen. Log-rank test (77.8% vs. 30.8%, <span class="html-italic">p</span> = 0.0022). (<b>b</b>) Loading doses of voriconazole in HBV-ACLF patients. The loading dose ranged from 100 to 400 mg twice daily. (<b>c</b>–<b>f</b>) Trough voriconazole concentrations comparing patients treated with different regimens. Nine patients underwent plasma voriconazole concentration monitoring more than twice; the maintenance dose was based on the plasma voriconazole concentration. (<b>c</b>) A standard voriconazole regimen. (<b>d</b>–<b>f</b>) A reduced voriconazole regimen.</p>
Full article ">
14 pages, 3474 KiB  
Article
A Real-World Data Observational Analysis of the Impact of Liposomal Amphotericin B on Renal Function Using Machine Learning in Critically Ill Patients
by Ignasi Sacanella, Erika Esteve-Pitarch, Jessica Guevara-Chaux, Julen Berrueta, Alejandro García-Martínez, Josep Gómez, Cecilia Casarino, Florencia Alés, Laura Canadell, Ignacio Martín-Loeches, Santiago Grau, Francisco Javier Candel, María Bodí and Alejandro Rodríguez
Antibiotics 2024, 13(8), 760; https://doi.org/10.3390/antibiotics13080760 - 12 Aug 2024
Viewed by 235
Abstract
Background: Liposomal amphotericin B (L-AmB) has become the mainstay of treatment for severe invasive fungal infections. However, the potential for renal toxicity must be considered. Aims: To evaluate the incidence of acute kidney injury (AKI) in critically ill patients receiving L-AmB for more [...] Read more.
Background: Liposomal amphotericin B (L-AmB) has become the mainstay of treatment for severe invasive fungal infections. However, the potential for renal toxicity must be considered. Aims: To evaluate the incidence of acute kidney injury (AKI) in critically ill patients receiving L-AmB for more than 48 h. Methods: Retrospective, observational, single-center study. Clinical, demographic and laboratory variables were obtained automatically from the electronic medical record. AKI incidence was analyzed in the entire population and in patients with a “low” or “high” risk of AKI based on their creatinine levels at the outset of the study. Factors associated with the development of AKI were studied using random forest models. Results: Finally, 67 patients with a median age of 61 (53–71) years, 67% male, a median SOFA of 4 (3–6.5) and a crude mortality of 34.3% were included. No variations in serum creatinine were observed during the observation period, except for a decrease in the high-risk subgroup. A total of 26.8% (total population), 25% (low risk) and 13% (high risk) of patients developed AKI. Norepinephrine, the SOFA score, furosemide (general model), potassium, C-reactive protein and procalcitonin (low-risk subgroup) were the variables identified by the random forest models as important contributing factors to the development of AKI other than L-AmB administration. Conclusions: The development of AKI is multifactorial and the administration of L-AmB appears to be safe in this group of patients. Full article
(This article belongs to the Special Issue Infection Diagnostics and Antimicrobial Therapy for Critical Patient)
Show Figures

Figure 1

Figure 1
<p>Flow chart of included patients.</p>
Full article ">Figure 2
<p>Contribution of each confounding variable according to the random forest (RF) model for variables associated with the development of AKI with the total dose of L-AmB administered at day 3 (Model 1) or the total duration of L-AmB administration (Model 2). This RF model was applied to the whole population of patients.</p>
Full article ">Figure 3
<p>Contribution of each confounding variable according to the random forest (RF) model for variables associated with the development of AKI, considering either the total dose of L-AmB administered on day 3 (model 1) or the total duration of L-AmB administration (model 2). This RF model was applied to patients with a lower risk of AKI. Note: the total duration of L-AmB was not deemed important by the model and thus does not appear in the graph.</p>
Full article ">
16 pages, 2682 KiB  
Article
Bacillus coagulans LMG S-24828 Impairs Candida Virulence and Protects Vaginal Epithelial Cells against Candida Infection In Vitro
by Luca Spaggiari, Andrea Ardizzoni, Natalia Pedretti, Ramona Iseppi, Carla Sabia, Rosario Russo, Samyr Kenno, Francesco De Seta and Eva Pericolini
Microorganisms 2024, 12(8), 1634; https://doi.org/10.3390/microorganisms12081634 - 10 Aug 2024
Viewed by 330
Abstract
Probiotics are living microbes that provide benefits to the host. The growing data on health promotion, following probiotics administration, increased interest among researchers and pharmaceutical companies. Infections of the lower genital tract in females, caused by a wide range of pathogens, represent one [...] Read more.
Probiotics are living microbes that provide benefits to the host. The growing data on health promotion, following probiotics administration, increased interest among researchers and pharmaceutical companies. Infections of the lower genital tract in females, caused by a wide range of pathogens, represent one of the main areas for the use of probiotics and postbiotics. Vulvovaginal candidiasis (VVC) affects 75% of women of reproductive age at least once during their lifetime, with 5–8% developing the recurrent form (RVVC). The disease is triggered by the overgrowth of Candida on the vaginal mucosa. Here, in order to establish its probiotic potential in the context of VVC, we evaluated the anti-fungal effects of the spore-producing Bacillus coagulans LMG S-24828 against C. albicans and C. parapsilosis as well as its beneficial effects in counteracting Candida vaginal infection in vitro. Our results show that both live B. coagulans and its Cell-Free Supernatant (CFS) exerted antifungal activity against both fungi. Moreover, live B. coagulans reduced hyphal formation, inhibited C. albicans adhesion to vaginal epithelial cells, showed co-aggregation capacity, and exerted a protective effect on vaginal epithelial cells infected with C. albicans. These data suggest that B. coagulans LMG S-24828 may provide benefits in the context of Candida vaginal infections. Full article
(This article belongs to the Special Issue Microbe–Host Interactions in Human Infections)
Show Figures

Figure 1

Figure 1
<p>Anti-<span class="html-italic">Candida</span> effect exerted by <span class="html-italic">B. coagulans</span>. (<b>A</b>) Mono-cultures and co-cultures of <span class="html-italic">C. albicans</span> (Ca), <span class="html-italic">C. parapsilosis</span> (Cp), and <span class="html-italic">B. coagulans</span> (Bc) pH values after 24 h of incubation at 37 °C. The chart reports the mean values of pH ± SEM from three different experiments. The range highlighted in light green represents the mean pH levels of the healthy vaginal environment. Statistical analysis was performed by the one-way ANOVA test followed by the uncorrected Fisher’s LSD test. Ca vs. Bc + Ca **** <span class="html-italic">p</span> &lt; 0.0001. Cp vs. Bc + Cp **** <span class="html-italic">p</span> &lt; 0.0001. Effect of <span class="html-italic">B. coagulans</span> (Bc) on <span class="html-italic">C. albicans</span> (Ca) (<b>B</b>) and <span class="html-italic">C. parapsilosis</span> (Cp) (<b>C</b>) growth capacity and acidification contribution upon 24 h of incubation at 37 °C. The graph shows the mean <span class="html-italic">C. albicans</span> (CFU × 10<sup>5</sup>/mL) ± SEM from three different experiments. Statistical analysis was performed by the one-way ANOVA test followed by the uncorrected Fisher’s LSD test. Ca vs. Bc + Ca **** <span class="html-italic">p</span> &lt; 0.0001. Cp vs. Bc + Cp *** <span class="html-italic">p</span> &lt; 0.001. ns = not significant.</p>
Full article ">Figure 2
<p>Kinetic measurement of <span class="html-italic">C. albicans</span> (Ca) (<b>A</b>) or <span class="html-italic">C. parapsilosis</span> (<b>B</b>) growth when cultivated with <span class="html-italic">B. coagulans</span> CFS (Bc CFS) or a sterile medium at 37 °C. Culture OD at 570 nm wavelength was automatically detected every 120 min for a total of 18 h. The graphs report the mean OD values ± SEM from triplicate samples of three different experiments. The Area Under the Curve (AUC) analysis was performed on kinetic data from <span class="html-italic">C. albicans</span> (<b>C</b>) and <span class="html-italic">C. parapsilosis</span> (<b>D</b>) samples. Each line represents one single experiment. Statistical analysis was performed on AUC values through the unpaired two-tailed Student’s <span class="html-italic">t</span>-test. Ca vs. Bc CFS + Ca and Cp vs. Bc CFS + Cp * <span class="html-italic">p</span> &lt; 0.05.</p>
Full article ">Figure 3
<p>Evaluation of <span class="html-italic">C. albicans</span> (Ca) adhesion capacity to a vaginal epithelial cell monolayer in the presence of <span class="html-italic">B. coagulans</span> (Bc) (<b>A</b>). The histogram graph shows the average % ± SEM of fungal adhesion inhibition exerted by <span class="html-italic">B. coagulans</span> (Bc). Data are from three independent experiments. (<b>B</b>) Assessment of <span class="html-italic">B. coagulans</span> (Bc) capacity to co-aggregate with <span class="html-italic">C. albicans</span> (Ca) after 1 h of co-incubation. Boxes in the heatmap represent the score assigned to each sample in three independent experiments—0: no aggregation; 1: aggregates with small clusters; 2: aggregates with larger numbers of yeasts; 3: clumps visible with the naked eye containing large numbers of yeast cells; 4: maximum score for large clumps visible with the naked eye in the well center. (<b>C</b>–<b>E</b>) Effect of <span class="html-italic">B. coagulans</span> (Bc) on <span class="html-italic">C. albicans</span> (Ca) hyphal formation upon 4 h of co-incubation. Hyphal fragments were optically counted by fluorescent microscopy imaging. The fungal cell wall was stained with Uvitex 2B fluorescent dye. (<b>C</b>) The bars chart reports the mean percentage ± SEM of hyphal fragments counted in three different fields from three independent experiments. Statistical analysis was performed by the unpaired, two-tailed Student <span class="html-italic">t</span>-test. Ca vs. Bc + Ca * <span class="html-italic">p</span> &lt; 0.05. (<b>D</b>) The heatmap shows the % of hyphal fragments counted in each field; the squares’ color indicates the abundance of hyphae in the field (red: high % hyphal fragments; green: low % hyphal fragments). (<b>E</b>) Representative images from fluorescence microscopy analysis are shown from <span class="html-italic">C. albicans</span> (Ca) or <span class="html-italic">C. albicans</span> plus <span class="html-italic">B. coagulans</span> (Bc + Ca) taken at 40× magnification.</p>
Full article ">Figure 4
<p>(<b>A</b>) Percentage of vaginal cell damage pre-colonized or not by <span class="html-italic">B. coagulans</span> (Bc) for 6 h and infected for further 18 h with <span class="html-italic">C. albicans</span> (Ca). The chart reports the average percentage of cell damage ± SEM of triplicate samples from three different experiments. Statistical analysis was performed by the unpaired, two-tailed Student <span class="html-italic">t</span>-test. Ca vs. Bc + Ca * <span class="html-italic">p</span> &lt; 0.05. (<b>B</b>) Production of β-defensin-2 by vaginal epithelial cells pre-colonized or not by <span class="html-italic">B. coagulans</span> (Bc) for 6 h and infected for further 18 h with <span class="html-italic">C. albicans</span> (Ca). Uninfected cells (Ctrl) and cells colonized by the bacterium without <span class="html-italic">C. albicans</span> were also included in the experiments. The graph reports the mean ± SEM from three independent experiments. Statistical analysis was performed by the one-way ANOVA test followed by the uncorrected Fisher’s LSD test. Untreated cells vs. Bc pre-colonized cells * <span class="html-italic">p</span> &lt; 0.05.</p>
Full article ">Figure 5
<p>Evaluation of antifungal effect permanency upon <span class="html-italic">B. coagulans</span> removal. <span class="html-italic">C. albicans</span> (Ca) (<b>A</b>) and <span class="html-italic">C. parapsilosis</span> (Cp) (<b>B</b>) metabolic activity quantification after being incubated with <span class="html-italic">B. coagulans</span> or sterile medium for 24 h and subsequent fungal isolation and cultivation for 24 h in the lack of bacteria. The graphs show the mean OD at 492 nm wavelength ± SEM from triplicate sample of three different experiments. Statistical analysis was performed by the one-way ANOVA test followed by the uncorrected Fisher’s LSD test. ns = not significant. (<b>C</b>) Capacity of <span class="html-italic">B. coagulans</span> spores to germinate on intestinal epithelial cells CaCo-2. Bacterial spores were seeded on an intestinal epithelial cell monolayer of CaCo-2 and incubated at 37 °C + 5% CO<sub>2</sub> for 24 h. After incubation, monolayers were photographed (upper images) and subsequently lysed. A Gram staining was then performed to visualize the presence of germinated <span class="html-italic">B. coagulans</span> (Bc) (lower images).</p>
Full article ">Figure 5 Cont.
<p>Evaluation of antifungal effect permanency upon <span class="html-italic">B. coagulans</span> removal. <span class="html-italic">C. albicans</span> (Ca) (<b>A</b>) and <span class="html-italic">C. parapsilosis</span> (Cp) (<b>B</b>) metabolic activity quantification after being incubated with <span class="html-italic">B. coagulans</span> or sterile medium for 24 h and subsequent fungal isolation and cultivation for 24 h in the lack of bacteria. The graphs show the mean OD at 492 nm wavelength ± SEM from triplicate sample of three different experiments. Statistical analysis was performed by the one-way ANOVA test followed by the uncorrected Fisher’s LSD test. ns = not significant. (<b>C</b>) Capacity of <span class="html-italic">B. coagulans</span> spores to germinate on intestinal epithelial cells CaCo-2. Bacterial spores were seeded on an intestinal epithelial cell monolayer of CaCo-2 and incubated at 37 °C + 5% CO<sub>2</sub> for 24 h. After incubation, monolayers were photographed (upper images) and subsequently lysed. A Gram staining was then performed to visualize the presence of germinated <span class="html-italic">B. coagulans</span> (Bc) (lower images).</p>
Full article ">Figure 6
<p>Schematic representation of biological activities of <span class="html-italic">B. coagulans</span> LMG S-24828 against <span class="html-italic">Candida</span>. Created with <a href="http://BioRender.com" target="_blank">BioRender.com</a>.</p>
Full article ">
34 pages, 4429 KiB  
Review
Synthetic and Natural Antifungal Substances in Cereal Grain Protection: A Review of Bright and Dark Sides
by Tomasz Szczygieł, Anna Koziróg and Anna Otlewska
Molecules 2024, 29(16), 3780; https://doi.org/10.3390/molecules29163780 - 9 Aug 2024
Viewed by 235
Abstract
Molds pose a severe challenge to agriculture because they cause very large crop losses. For this reason, synthetic fungicides have been used for a long time. Without adequate protection against pests and various pathogens, crop losses could be as high as 30–40%. However, [...] Read more.
Molds pose a severe challenge to agriculture because they cause very large crop losses. For this reason, synthetic fungicides have been used for a long time. Without adequate protection against pests and various pathogens, crop losses could be as high as 30–40%. However, concerns mainly about the environmental impact of synthetic antifungals and human health risk have prompted a search for natural alternatives. But do natural remedies only have advantages? This article reviews the current state of knowledge on the use of antifungal substances in agriculture to protect seeds against phytopathogens. The advantages and disadvantages of using both synthetic and natural fungicides to protect cereal grains were discussed, indicating specific examples and mechanisms of action. The possibilities of an integrated control approach, combining cultural, biological, and chemical methods are described, constituting a holistic strategy for sustainable mold management in the grain industry. Full article
Show Figures

Figure 1

Figure 1
<p>Term map based on keywords ‘fungicides’ and ‘cereal grains’. The colors on the map indicate terms belonging to different clusters (<b>A</b>) or the year of publication (<b>B</b>). The lengths of the lines correspond to the interrelationships between the terms. Bubble size presents the number of papers in the database. Bubble proximity presents frequency of co-occurrence of phrases in the same papers.</p>
Full article ">Figure 2
<p>Term map based on keywords ‘biocontrol’ and ‘cereal grains’. The colors on the map indicate terms belonging to different clusters (<b>A</b>) or the year of publication (<b>B</b>). The lengths of the lines correspond to the interrelationships between the terms. Bubble size presents the number of papers in the database. Bubble proximity presents frequency of co-occurrence of phrases in the same papers.</p>
Full article ">Figure 3
<p>Fungicide mechanisms of action in mold cells.</p>
Full article ">Figure 4
<p>Fungicides and their impacts on various environmental zones.</p>
Full article ">
22 pages, 1395 KiB  
Article
Exploring the Antifungal Activity of Moroccan Bacterial and Fungal Isolates and a Strobilurin Fungicide in the Control of Cladosporium fulvum, the Causal Agent of Tomato Leaf Mold Disease
by Zineb Belabess, Bilale Gajjout, Ikram Legrifi, Essaid Ait Barka and Rachid Lahlali
Plants 2024, 13(16), 2213; https://doi.org/10.3390/plants13162213 - 9 Aug 2024
Viewed by 276
Abstract
The causal agent of tomato leaf mold, Cladosporium fulvum, is prevalent in greenhouses worldwide, especially under high humidity conditions. Despite its economic impact, studies on antifungal agents targeting C. fulvum remain limited. This study evaluates biocontrol agents (BCAs) as alternatives to chemical [...] Read more.
The causal agent of tomato leaf mold, Cladosporium fulvum, is prevalent in greenhouses worldwide, especially under high humidity conditions. Despite its economic impact, studies on antifungal agents targeting C. fulvum remain limited. This study evaluates biocontrol agents (BCAs) as alternatives to chemical controls for managing this disease, alongside the strobilurin fungicide azoxystrobin. From a Moroccan collection of potential BCAs, five bacterial isolates (Alcaligenes faecalis ACBC1, Pantoea agglomerans ACBC2, ACBP1, ACBP2, and Bacillus amyloliquefaciens SF14) and three fungal isolates (Trichoderma spp. OT1, AT2, and BT3) were selected and tested. The in vitro results demonstrated that P. agglomerans isolates reduced mycelial growth by over 60% at 12 days post-inoculation (dpi), while Trichoderma isolates achieved 100% inhibition in just 5 dpi. All bacterial isolates produced volatile organic compounds (VOCs) with mycelial inhibition rates ranging from 38.8% to 57.4%. Likewise, bacterial cell-free filtrates significantly inhibited the pathogen’s mycelial growth. Greenhouse tests validated these findings, showing that all the tested isolates were effective in reducing disease incidence and severity. Azoxystrobin effectively impeded C. fulvum growth, particularly in protective treatments. Fourier transform infrared spectroscopy (FTIR) analysis revealed significant biochemical changes in the treated plants, indicating fungal activity. This study provides valuable insights into the efficacy of these BCAs and azoxystrobin, contributing to integrated management strategies for tomato leaf mold disease. Full article
(This article belongs to the Special Issue Fungus and Plant Interactions: Volume II)
Show Figures

Figure 1

Figure 1
<p>Spore germination inhibition rate (%) of <span class="html-italic">Cladosporium fulvum</span> in a total of 100 spores after incubation for 24 h, by cell-free filtrates of antagonistic bacteria at 100% concentration (<span class="html-italic">Bacillus amyloliquefaciens</span> SF14, <span class="html-italic">Alcaligenes faecalis</span> ACBC1, <span class="html-italic">Pantoea agglomerans</span> ACBC2, <span class="html-italic">P. agglomerans</span> ACBP1, <span class="html-italic">P. agglomerans</span> ACBP2, <span class="html-italic">Bacillus subtilis</span> Y1336, and <span class="html-italic">P. agglomerans</span> P10c). Data representing the average inhibition rate, with the same letter, are not significantly different according to the SNK test (<span class="html-italic">p</span> ≤ 0.05).</p>
Full article ">Figure 2
<p>Observation of disease severity and incidence on the leaf of tomato plants treated with the seven tested bacterial (<span class="html-italic">Bacillus amyloliquefaciens</span> SF14, <span class="html-italic">Alcaligenes faecalis</span> ACBC1, <span class="html-italic">Pantoea agglomerans</span> ACBC2, <span class="html-italic">P. agglomerans</span> ACBP1, <span class="html-italic">P. agglomerans</span> ACBP2, <span class="html-italic">Bacillus subtilis</span> Y1336, and <span class="html-italic">P. agglomerans</span> P10c) suspensions (10<sup>8</sup> CFU/mL) and inoculated with <span class="html-italic">Cladosporium fulvum</span>, after 30 days of incubation at 25 °C within greenhouse conditions. Control: positive control (pathogen only; <span class="html-italic">C. fulvum</span>). Asoxystrobin: plants treated with fungicide. Bar charts represent the mean value of disease severity of two trials over time with four replicates. Values of plant incidence and severity with the same letter (uppercase for incidence: A, B, etc.; lowercase for severity: a, b, etc.) were not significantly different according to the SNK test (<span class="html-italic">p</span> ≤ 0.05).</p>
Full article ">Figure 3
<p>Linear regression of severity % (in vivo) and inhibition rate % (in vitro) of the seven tested bacteria (<span class="html-italic">Bacillus amyloliquefaciens</span> SF14, <span class="html-italic">Alcaligenes faecalis</span> ACBC1, <span class="html-italic">Pantoea agglomerans</span> ACBC2, <span class="html-italic">P. agglomerans</span> ACBP1, <span class="html-italic">P. agglomerans</span> ACBP2, <span class="html-italic">Bacillus subtilis</span> Y1336, and <span class="html-italic">P. agglomerans</span> P10c) and the commercial fungicide against <span class="html-italic">Cladosporium fulvum</span>.</p>
Full article ">Figure 4
<p>The average percentage of infected tomato plants (%) for different treatments (<span class="html-italic">Bacillus amyloliquefaciens</span> SF14, <span class="html-italic">Alcaligenes faecalis</span> ACBC1, <span class="html-italic">Pantoea agglomerans</span> ACBC2, <span class="html-italic">P. agglomerans</span> ACBP1, <span class="html-italic">P. agglomerans</span> ACBP2, <span class="html-italic">Bacillus subtilis</span> Y1336, <span class="html-italic">P. agglomerans</span> P10c, and the commercial fungicide against <span class="html-italic">C. fulvum</span>) recorded after 15 and 30 days of artificial inoculation incubated at 25 °C and 85% HR. Treatments with the same letter are not significantly different according to the SNK test (<span class="html-italic">p</span> ≤ 0.05).</p>
Full article ">
15 pages, 7966 KiB  
Article
Anticandidal Properties of Launaea sarmentosa among the Salt Marsh Plants Collected from Palk Bay and the Gulf of Mannar Coast, Southeastern India
by Smriti Das, Karuppannagounder Rajan Priyanka, Kolandhasamy Prabhu, Ramachandran Vinayagam, Rajendran Rajaram and Sang Gu Kang
Antibiotics 2024, 13(8), 748; https://doi.org/10.3390/antibiotics13080748 - 9 Aug 2024
Viewed by 424
Abstract
Tidal wetlands, commonly known as salt marshes, are highly productive ecosystems in temperate regions worldwide. These environments constitute a unique flora composed primarily of salt-tolerant herbs, grasses, and shrubs. This study investigated the therapeutic properties of ten salt marsh plants collected mainly from [...] Read more.
Tidal wetlands, commonly known as salt marshes, are highly productive ecosystems in temperate regions worldwide. These environments constitute a unique flora composed primarily of salt-tolerant herbs, grasses, and shrubs. This study investigated the therapeutic properties of ten salt marsh plants collected mainly from Palk Bay and Mannar Gulf against Candida disease. This study examined the changes in natural plant products associated with their anti-Candida growth activity during two distinct seasonal changes—monsoon and summer. The potential of the salt marshes to inhibit the growth of five different Candida strains was assessed using four solvents. In phytochemical analysis, the extracts obtained from a Launaea sarmentosa exhibited the highest results compared to the other plant extracts. Fourier transform infrared spectroscopy revealed 12 peaks with alkane, aldehyde, amine, aromatic ester, phenol, secondary alcohol, and 1,2,3,4-tetrasubstituted. Gas-chromatography–mass spectrometry detected 30 compounds. Cyclotetracosane, lupeol, β-amyrin, and 12-oleanen-3-yl acetate showed the highest peak range. In particular, plant samples collected during the monsoon season were more effective in preventing Canda growth than the summer plant samples. In the monsoon season, the salt marsh plant extracted with ethyl acetate showed a high anti-Candida growth activity, while in the summer, the acetone extract exhibited a higher anti-Candida growth activity than the other solvents. The hexane extract of L. sarmentosa showed the highest inhibition zone against all Candidal strains. Furthermore, compounds, such as β-amyrin, lupeol, and oxirane, from the hexane extract of L. sarmentosa play a vital role in anti-Candida activity. This paper reports the potential of tidal marsh plant extracts for developing new antifungal agents for Candida infections. Full article
Show Figures

Figure 1

Figure 1
<p>Anticandial properties of salt marsh plants exhibited as the zone of inhibition against Candidal strains during the monsoon season ((<b>a</b>)—Acetone; (<b>b</b>)—Ethyl Acetate; (<b>c</b>)—Methonal and (<b>d</b>)—Hexane).</p>
Full article ">Figure 2
<p>Anticandidal properties of salt marsh plants exhibiting zone of inhibition against Candidal strains during the summer season ((<b>a</b>)—Acetone; (<b>b</b>)—Ethyl acetate; (<b>c</b>)—Methanol; and (<b>d</b>)—Hexane).</p>
Full article ">Figure 3
<p>Anticandidal activity of saltmarsh plant <span class="html-italic">Launaea sarmentosa</span> extract exhibits the highest inhibition (CA—<span class="html-italic">Candida albicans</span>; CR—<span class="html-italic">Candida kefyr</span>; CKr—<span class="html-italic">Candida krusei;</span> CT—<span class="html-italic">Candida tropicalis</span>; CP—<span class="html-italic">Candida parapsilosis</span>).</p>
Full article ">Figure 4
<p>FT−IR spectrum showing the peaks obtained from the hexane extract of salt marsh <span class="html-italic">Launaea sarmentosa</span>.</p>
Full article ">Figure 5
<p>GC-MS showing the peaks obtained from the hexane extract of saltmarsh <span class="html-italic">Launaea sarmentosa</span>.</p>
Full article ">Figure 6
<p>Sampling sites of salt marsh plants collected from Palk Bay and the Gulf of Mannar.</p>
Full article ">Figure 7
<p>Salt marsh plants collected from Palk Bay and the Gulf of Mannar ((<b>A</b>) <span class="html-italic">Ipomoea pes-caprae</span>, (<b>B</b>) <span class="html-italic">Suaeda maritima</span>, (<b>C</b>) <span class="html-italic">Sesuvium portulacastrum</span>, (<b>D</b>) <span class="html-italic">Heliotropium curassavicum</span>, (<b>E</b>) <span class="html-italic">Launaea sarmentosa</span>, (<b>F</b>) <span class="html-italic">Bulbostylis barbata</span>, (<b>G</b>) <span class="html-italic">Salicornia brachiata</span>, (<b>H</b>) <span class="html-italic">Spinifex littoreus</span>, (<b>I</b>) Fim<span class="html-italic">bristylis spathacea</span>, and (<b>J</b>) <span class="html-italic">Artiplex halimus</span>).</p>
Full article ">
22 pages, 3224 KiB  
Article
Phenotypic and Genotypic Characterization of Resistance and Virulence Markers in Candida spp. Isolated from Community-Acquired Infections in Bucharest, and the Impact of AgNPs on the Highly Resistant Isolates
by Viorica Maria Corbu, Ana-Maria Georgescu, Ioana Cristina Marinas, Radu Pericleanu, Denisa Vasilica Mogos, Andreea Ștefania Dumbravă, Liliana Marinescu, Ionut Pecete, Tatiana Vassu-Dimov, Ilda Czobor Barbu, Ortansa Csutak, Denisa Ficai and Irina Gheorghe-Barbu
J. Fungi 2024, 10(8), 563; https://doi.org/10.3390/jof10080563 - 9 Aug 2024
Viewed by 312
Abstract
Background: This study aimed to determine, at the phenotypic and molecular levels, resistance and virulence markers in Candida spp. isolated from community-acquired infections in Bucharest outpatients during 2021, and to demonstrate the efficiency of alternative solutions against them based on silver nanoparticles (AgNPs). [...] Read more.
Background: This study aimed to determine, at the phenotypic and molecular levels, resistance and virulence markers in Candida spp. isolated from community-acquired infections in Bucharest outpatients during 2021, and to demonstrate the efficiency of alternative solutions against them based on silver nanoparticles (AgNPs). Methods: A total of 62 Candida spp. strains were isolated from dermatomycoses and identified using chromogenic culture media and MALDI-TOF MS, and then investigated for their antimicrobial resistance and virulence markers (VMs), as well as for metabolic enzymes using enzymatic tests for the expression of soluble virulence factors, their biofilm formation and adherence capacity on HeLa cells, and PCR assays for the detection of virulence markers and the antimicrobial activity of alternative solutions based on AgNPs. Results: Of the total of 62 strains, 45.16% were Candida parapsilosis; 29.03% Candida albicans; 9.67% Candida guilliermondii; 3.22% Candida lusitaniae, Candia pararugosa, and Candida tropicalis; and 1.66% Candida kefyr, Candida famata, Candida haemulonii, and Candida metapsilosis. Aesculin hydrolysis, caseinase, and amylase production were detected in the analyzed strains. The strains exhibited different indices of adherence to HeLa cells and were positive in decreasing frequency order for the LIP1, HWP1, and ALS1,3 genes (C. tropicalis/C. albicans). An inhibitory effect on microbial growth, adherence capacity, and on the production of virulence factors was obtained using AgNPs. Conclusions: The obtained results in C. albicans and Candida non-albicans circulating in Bucharest outpatients were characterized by moderate-to-high potential to produce VMs, necessitating epidemiological surveillance measures to minimize the chances of severe invasive infections. Full article
(This article belongs to the Special Issue Fungal Biofilms, 2nd Edition)
Show Figures

Figure 1

Figure 1
<p>The experimental design (created with Biorender.com; accessed on 11 June 2024).</p>
Full article ">Figure 2
<p>The distribution of arbitrary units by isolation sources of <span class="html-italic">Candida</span> spp. strains. Legend: AU1 = 1 arbitrary unit, AU2 = 2 arbitrary units, and AU0 = 0 arbitrary units.</p>
Full article ">Figure 3
<p>Average MIC values for <span class="html-italic">Candida</span> spp. strains.</p>
Full article ">Figure 4
<p>Adherence inhibition percentage (PICA%) values for AgNPs compared to the <span class="html-italic">Candida</span> spp. strains (* <span class="html-italic">p</span> &lt; 0.05, ** <span class="html-italic">p</span> &lt; 0.001, **** <span class="html-italic">p</span> &lt; 0.0001) (Dunnett’s multiple comparisons test).</p>
Full article ">Figure 5
<p>AgNPs’ effects on <span class="html-italic">Candida</span> spp. strains’ virulence factors (* <span class="html-italic">p</span> &lt; 0.05, ** <span class="html-italic">p</span> &lt; 0.001, *** <span class="html-italic">p</span> &lt; 0.001, **** <span class="html-italic">p</span> &lt; 0.0001) (Dunnett’s multiple comparisons test). (<b>A</b>) Caseinase production of <span class="html-italic">Candida</span> sp. strains, (<b>B</b>) Hemolysis production of <span class="html-italic">Candida</span> sp. strains, (<b>C</b>) Amylase production of <span class="html-italic">Candida</span> sp. strains, (<b>D</b>) Esculin Hydrolysis production of <span class="html-italic">Candida</span> sp. strains.</p>
Full article ">Figure 6
<p>Extracellular NO content determined by the Griess reaction for AgNPs in the presence of <span class="html-italic">C. albicans</span> strains (Tukey’s method, * <span class="html-italic">p</span> &lt; 0.05, ** <span class="html-italic">p</span> &lt; 0.01, *** <span class="html-italic">p</span> &lt; 0.001, **** <span class="html-italic">p</span> &lt; 0.0001).</p>
Full article ">Figure 7
<p>Pearson correlation among extracellular NO content, adhesion inhibition percentage (PICA%), caseinase activity (%), amylase activity (%), and hemolysin (%) for <span class="html-italic">C. albicans</span> (<b>A</b>) and <span class="html-italic">C. parapsilosis</span> (<b>B</b>).</p>
Full article ">
25 pages, 17779 KiB  
Article
Geraniol Potentiates the Effect of Fluconazole against Planktonic and Sessile Cells of Azole-Resistant Candida tropicalis: In Vitro and In Vivo Analyses
by Gislaine Silva-Rodrigues, Isabela Madeira de Castro, Paulo Henrique Guilherme Borges, Helena Tiemi Suzukawa, Joyce Marinho de Souza, Guilherme Bartolomeu-Gonçalves, Marsileni Pelisson, Cássio Ilan Soares Medeiros, Marcelle de Lima Ferreira Bispo, Ricardo Sérgio Couto de Almeida, Kelly Ishida, Eliandro Reis Tavares, Lucy Megumi Yamauchi and Sueli Fumie Yamada-Ogatta
Pharmaceutics 2024, 16(8), 1053; https://doi.org/10.3390/pharmaceutics16081053 - 9 Aug 2024
Viewed by 339
Abstract
Candida tropicalis is regarded as an opportunistic pathogen, causing diseases ranging from superficial infections to life-threatening disseminated infections. The ability of this yeast to form biofilms and develop resistance to antifungals represents a significant therapeutic challenge. Herein, the effect of geraniol (GER), alone [...] Read more.
Candida tropicalis is regarded as an opportunistic pathogen, causing diseases ranging from superficial infections to life-threatening disseminated infections. The ability of this yeast to form biofilms and develop resistance to antifungals represents a significant therapeutic challenge. Herein, the effect of geraniol (GER), alone and combined with fluconazole (FLZ), was evaluated in the planktonic and sessile cells of azole-resistant C. tropicalis. GER showed a time-dependent fungicidal effect on the planktonic cells, impairing the cell membrane integrity. Additionally, GER inhibited the rhodamine 6G efflux, and the molecular docking analyzes supported the binding affinity of GER to the C. tropicalis Cdr1 protein. GER exhibited a synergism with FLZ against the planktonic and sessile cells, inhibiting the adhesion of the yeast cells and the viability of the 48-h biofilms formed on abiotic surfaces. C. tropicalis biofilms treated with GER, alone or combined with FLZ, displayed morphological and ultrastructural alterations, including a decrease in the stacking layers and the presence of wilted cells. Moreover, neither GER alone nor combined with FLZ caused toxicity, and both treatments prolonged the survival of the Galleria mellonella larvae infected with azole-resistant C. tropicalis. These findings indicate that the combination of GER and FLZ may be a promising strategy to control azole-resistant C. tropicalis infections. Full article
Show Figures

Figure 1

Figure 1
<p>Antifungal activity of geraniol (GER) in <span class="html-italic">Candida tropicalis</span>. Time-kill kinetics of <span class="html-italic">C. tropicalis</span> ATCC 28707 (<b>a</b>); <span class="html-italic">C. tropicalis</span> CTR1 (<b>b</b>); <span class="html-italic">C. tropicalis</span> CTR2 (<b>c</b>); <span class="html-italic">C. tropicalis</span> CTR3 (<b>d</b>) incubated with the minimum inhibitory (MIC) and fungicidal (MFC) concentrations of GER. The log<sub>10</sub> CFU/mL values are the mean and the standard deviation from three independent experiments. The dotted lines represent the 99.9% (3 log10) reduction in the CFU/mL counting. AmB: amphotericin B.</p>
Full article ">Figure 2
<p>Plasma membrane integrity analysis of <span class="html-italic">Candida tropicalis</span> ATCC 2870. Planktonic cells were incubated with or without 0.25 × MIC, 0.5 × MIC, and MIC of geraniol. The intracellular content absorbing at 260/280 nm (<b>a</b>–<b>d</b>) or 280 nm (<b>e</b>–<b>h</b>) was determined after the specified time intervals. The values are the means and standard deviations from three independent experiments. (<b>a</b>,<b>e</b>) <span class="html-italic">C. tropicalis</span> ATCC 28707; (<b>b</b>,<b>f</b>) <span class="html-italic">C. tropicalis</span> CTR1; (<b>c</b>,<b>g</b>) <span class="html-italic">C. tropicalis</span> CTR2; (<b>d</b>,<b>h</b>) <span class="html-italic">C. tropicalis</span> CTR3. ** <span class="html-italic">p</span> &lt;0.01, **** <span class="html-italic">p</span> &lt; 0.0001 compared with the untreated fungal cells.</p>
Full article ">Figure 3
<p>Effect of geraniol on Rhodamine 6G efflux by the planktonic cells of <span class="html-italic">Candida tropicalis</span>. Planktonic cells were incubated with 0.25 × MIC of geraniol for 5 h. The energy-dependent R6G efflux was initiated by adding 2% glucose (arrow) and quantified by measuring the absorbance of the supernatant at 530 nm. (<b>a</b>) <span class="html-italic">C. tropicalis</span> ATCC 28707; (<b>b</b>) <span class="html-italic">C. tropicalis</span> CTR1; (<b>c</b>) <span class="html-italic">C. tropicalis</span> CTR2; (<b>d</b>) <span class="html-italic">C. tropicalis</span> CTR3. The values are the means and standard deviations from three independent experiments. CUR: curcumin.</p>
Full article ">Figure 4
<p>3D structure prediction and validation of <span class="html-italic">Candida tropicalis</span> resistance protein 1 (<span class="html-italic">Ct</span>Cdr1). Predicted pLDDT (<b>a</b>) and amino acid residue position coverage (<b>b</b>) of <span class="html-italic">Ct</span>Cdr1. Validation of the maximum score model using the PROCHECK Ramachandran plot (<b>c</b>) and the MolProbity Ramachandran plot (<b>d</b>). Alignment of the five 3D structures of <span class="html-italic">Ct</span>Cdr1 predicted using AlphaFold2 (<b>e</b>).</p>
Full article ">Figure 5
<p>Molecular interactions of geraniol (GER), curcumin (CUR), and farnesol (FAR) with <span class="html-italic">Candida tropicalis</span> resistance protein 1 (<span class="html-italic">Ct</span>Cdr1). The main types of binding of GER (<b>a</b>), CUR (<b>b</b>), and FAR (<b>c</b>) to the binding site of the <span class="html-italic">Ct</span>Cdr1p) in 2D. The 3D distribution and chemical binding distances of GER (blue) with the amino acids (green) of the <span class="html-italic">Ct</span>Cdr1p binding site (<b>d</b>). 3D surface models showing the regions of the ligand (GER) with higher or lower degrees of hydrophobicity and the hydrogen donor and acceptor sites (<b>e</b>,<b>f</b>).</p>
Full article ">Figure 6
<p>Antifungal activity of geraniol (GER) alone and combined with fluconazole (FLZ) against the planktonic cells of <span class="html-italic">Candida tropicalis</span>. Time-kill kinetics of <span class="html-italic">C. tropicalis</span> ATCC 28707 (<b>a</b>); <span class="html-italic">C. tropicalis</span> CTR1 (<b>b</b>); <span class="html-italic">C. tropicalis</span> CTR2 (<b>c</b>); <span class="html-italic">C. tropicalis</span> CTR3 (<b>d</b>) incubated with GER (256 µg/mL) and FLZ (1 µg/mL) alone or in combination (GER/FLZ, 256/1 µg/mL). The dotted lines represent the 99.9% (3 log10) reduction in the CFU/mL counting. The log<sub>10</sub> CFU/mL values are the mean and the standard deviation from three independent experiments. **** <span class="html-italic">p</span> &lt; 0.0001 compared with the untreated fungal cells.</p>
Full article ">Figure 7
<p>Effect of geraniol (GER) alone or combined with fluconazole (GER/FLZ) on <span class="html-italic">Candida tropicalis</span> adhesion (<b>a</b>) and 48-h biofilms (<b>b</b>) formed on polystyrene surface. The effect of GER alone or the GER/FLZ combination was evaluated using colony forming units counting, and the values were converted into percentages, considering the untreated groups as controls. Values are the mean and standard deviation from three independent experiments. **** <span class="html-italic">p</span> &lt; 0.0001 compared with untreated fungal cells.</p>
Full article ">Figure 8
<p>Effect of geraniol (GER) alone or combined with fluconazole (FLZ) on the morphology and ultrastructure of <span class="html-italic">Candida tropicalis</span> ATCC 28707 biofilms. Scanning electron microscopy (SEM) images of biofilms on polystyrene during 48 h of incubation. (<b>a</b>,<b>b</b>) Untreated control; (<b>c</b>,<b>d</b>) treated with 128 µg/mL FLZ; (<b>e</b>,<b>f</b>) treated with 1024 µg/mL GER; (<b>g</b>,<b>h</b>) treated with 128/0.5 µg/mL GER/FLZ. Viability of sessile cells after treatment with GER alone or GER/FLZ combination was evaluated using colony forming units counting (<b>i</b>). Values are the mean and standard deviation from three independent experiments. **** <span class="html-italic">p</span> &lt; 0.0001.</p>
Full article ">Figure 9
<p>Kaplan–Meier plots of survival curves of <span class="html-italic">Galleria mellonella</span> larvae. (<b>a</b>) The larvae were inoculated with geraniol (GER), fluconazole (FLZ), and the GER/FLZ combination for the compounds toxicity analysis. (<b>b</b>) The larvae were infected with different <span class="html-italic">C. tropicalis</span> ATCC 28707 cell densities for determination of the lethal inoculum. (<b>c</b>) The larvae were infected with fungal cells (1 × 10<sup>6</sup>) and treated with (GER), fluconazole (FLZ), and the GER/FLZ combination. All the groups were compared with infected and untreated larvae. (<b>d</b>) Fungal load in the hemolymph of the larvae untreated and treated with the compounds determined using the colony forming unit (CFU) counts. The analysis of the <span class="html-italic">G. mellonella</span> survival data was performed using the log-rank (Mantel–Cox) of one representative experiment. The asterisks indicate a significant reduction in the fungal load of the infected treated group compared with the infected untreated group (ns, not significant; *** <span class="html-italic">p</span> &lt; 0.001; **** <span class="html-italic">p</span> &lt; 0.0001).</p>
Full article ">
14 pages, 736 KiB  
Review
QTc Interval Prolongation as an Adverse Event of Azole Antifungal Drugs: Case Report and Literature Review
by Shiori Kitaya, Makoto Nakano, Yukio Katori, Satoshi Yasuda and Hajime Kanamori
Microorganisms 2024, 12(8), 1619; https://doi.org/10.3390/microorganisms12081619 - 8 Aug 2024
Viewed by 311
Abstract
QTc prolongation and torsade de pointes (TdP) are significant adverse events linked to azole antifungals. Reports on QTc interval prolongation caused by these agents are limited. In this study, we report a case of a 77-year-old male with cardiovascular disease who experienced QTc [...] Read more.
QTc prolongation and torsade de pointes (TdP) are significant adverse events linked to azole antifungals. Reports on QTc interval prolongation caused by these agents are limited. In this study, we report a case of a 77-year-old male with cardiovascular disease who experienced QTc prolongation and subsequent TdP while being treated with fluconazole for Candida albicans-induced knee arthritis. Additionally, a literature review was conducted on cases where QTc prolongation and TdP were triggered as adverse events of azole antifungal drugs. The case study detailed the patient’s experience, whereas the literature review analyzed cases from May 1997 to February 2023, focusing on patient demographics, underlying diseases, antifungal regimens, concurrent medications, QTc changes, and outcomes. The review identified 16 cases, mainly in younger individuals (median age of 29) and women (75%). Fluconazole (63%) and voriconazole (37%) were the most common agents. Concurrent medications were present in 75% of cases, and TdP occurred in 81%. Management typically involved discontinuing or switching antifungals and correcting electrolytes, with all patients surviving. Risk assessment and concurrent medication review are essential before starting azole therapy. High-risk patients require careful electrocardiogram monitoring to prevent arrhythmias. Remote monitoring may enhance safety for patients with implanted devices. Further studies are needed to understand risk factors and management strategies. Full article
Show Figures

Figure 1

Figure 1
<p>Changes in QTc interval, serum potassium concentration, and treatment course during hospitalization. CFZ, cefazolin; CRT-D, cardiac resynchronization therapy-defibrillator; FLC, fluconazole; K, potassium; KCl, potassium chloride; LVX, levofloxacin; MFG, micafungin; MgSO4, magnesium sulfate; QTc, QT interval corrected for heart rate.</p>
Full article ">Figure 2
<p>Electrocardiogram findings (<b>a</b>) at admission; (<b>b</b>) occurrence of torsade de pointes and subsequent defibrillation using a cardiac resynchronization therapy-defibrillator; (<b>c</b>) at discharge. The light blue arrow indicates the area of the electrocardiogram where the torsade de pointes occurred, and the red arrow indicates the defibrillation activation timing. Torsade de pointes is a specific form of polymorphic ventricular tachycardia observed in patients with QTc prolongation. It is characterized by rapid and irregular QRS complexes that appear to twist around the baseline on the electrocardiogram.</p>
Full article ">
8 pages, 1951 KiB  
Proceeding Paper
Physiological Analysis (Monitoring of Germinal and Biometric Parameters) of Abelmoschus esculentus L. Seeds under the Effect of 100 mM and 150 mM NaCl
by Khedidja Dahli and Houria Medjdoub
Biol. Life Sci. Forum 2024, 36(1), 5; https://doi.org/10.3390/blsf2024036005 - 8 Aug 2024
Viewed by 113
Abstract
Abelmoschus esculentus L. is a plant of legume origin; it is used in several areas of nutrition, industry and medicine. This study was proposed in order to understand the adaptation mechanisms of the seeds of this species under the effect of NaCl. Germination [...] Read more.
Abelmoschus esculentus L. is a plant of legume origin; it is used in several areas of nutrition, industry and medicine. This study was proposed in order to understand the adaptation mechanisms of the seeds of this species under the effect of NaCl. Germination was monitored by varying the concentrations of NaCl saline solution (100 mM and 150 mM).The physiological parameters were determined by monitoring the capacity and speed of seed germination at a temperature of 25 °C. The monitoring of biometric parameters was carried out as follows: Calculation of the water content of sprouted okra seeds after one week and-evaluation of the radical length carried out by up-to-date measurements of the radical length of germinated seeds for one week. From the results obtained, it was noted that the cumulative rate of germinated seeds is estimated to be from 17.27% to 14.55% under the effect of 100 mM and 150 mM, respectively, as compared to 47.27% for the control seeds. In addition, it was observed that the treatment with 100 mM NaCl causes a decrease in the germination speed (23.58%) and a slight prolongation in time as compared to the control. On the other hand, the treatment with 150 mM induces a remarkable decrease in germination speed (13.96% versus 24.39%) and average time as compared to the control (2.9 days versus 4.1 days). Furthermore, the monitoring of biometric parameters highlights two essential points: the NaCl treatment limits the growth of seed radicals and reduces the water content of okra seeds as compared to the control. In light of these results, it is possible to conclude that treatment with sodium chloride has a detrimental effect on the germination and growth of Abelmoschus esculentus L. seeds. Full article
Show Figures

Figure 1

Figure 1
<p>Control seeds of <span class="html-italic">Abelmoschus esculentus</span> L. after a few days of germination (<b>a</b>). Seeds stressed at 100 mM of <span class="html-italic">Abelmoschus esculentus</span> L. after a few days of germination (<b>b</b>). Seeds stressed at 150 mM of <span class="html-italic">Abelmoschus esculentus</span> L. after a few days of germination (<b>c</b>).</p>
Full article ">Figure 2
<p>Length of the radical of control seeds of <span class="html-italic">Abelmoschus esculentus</span> L. after one week of sowing (<b>a</b>). Length of the radical seeds stressed at 100 mM of <span class="html-italic">Abelmoschus esculentus</span> L. after one week of sowing (<b>b</b>). Length of the radical seeds stressed at 150 mM of <span class="html-italic">Abelmoschus esculentus</span> L. after one week of sowing (<b>c</b>).</p>
Full article ">Figure 3
<p>Precocity of germination (% first germinated seeds) of <span class="html-italic">Abelmoschus esculentus</span> L. seeds under the effect of NaCl (100 mM, 150 mM).</p>
Full article ">Figure 4
<p>Kinetics of <span class="html-italic">Abelmoschus esculentus</span> L. seed germination under the effect of NaCl at 100 mM and 150 mM.</p>
Full article ">Figure 5
<p>Velocity coefficient (Vc %) and average time (day) of germination of <span class="html-italic">Abelmoschus esculentus</span> L. seeds under the effect of NaCl at 100 mM and 150 mM.</p>
Full article ">Figure 6
<p>The humidity of <span class="html-italic">Abelmoschus esculentus</span> L. seeds under the effect of NaCl at 100 mM and 150 mM.</p>
Full article ">Figure 7
<p>Length of radicals of <span class="html-italic">Abelmoschus esculentus</span> L. seeds under the effect of NaCl at 100 mM and 150 mM.</p>
Full article ">
13 pages, 2269 KiB  
Article
Antifungal Effect of Poly(methyl methacrylate) with Farnesol and Undecylenic Acid against Candida albicans Biofilm Formation
by Milica Išljamović, Debora Bonvin, Milena Milojević, Simona Stojanović, Milan Spasić, Branislava Stojković, Predrag Janošević, Suzana Otašević and Marijana Mionić Ebersold
Materials 2024, 17(16), 3936; https://doi.org/10.3390/ma17163936 - 8 Aug 2024
Viewed by 298
Abstract
The control of Candida albicans biofilm formation on dentures made of poly(methyl methacrylate) (PMMA) is an important challenge due to the high resistance to antifungal drugs. Interestingly, the natural compounds undecylenic acid (UDA) and farnesol (FAR) both prevent C. albicans biofilm formation and [...] Read more.
The control of Candida albicans biofilm formation on dentures made of poly(methyl methacrylate) (PMMA) is an important challenge due to the high resistance to antifungal drugs. Interestingly, the natural compounds undecylenic acid (UDA) and farnesol (FAR) both prevent C. albicans biofilm formation and could have a synergetic effect. We therefore modified PMMA with a combination of UDA and FAR (UDA+FAR), aiming to obtain the antifungal PMMA_UDA+FAR composites. Equal concentrations of FAR and UDA were added to PMMA to reach 3%, 6%, and 9% in total of both compounds in composites. The physico-chemical properties of the composites were characterized by Fourier-transform infrared spectroscopy and water contact angle measurement. The antifungal activity of the composites was tested on both biofilm and planktonic cells with an XTT test 0 and 6 days after the composites’ preparation. The effect of the UDA+FAR combination on C. albicans filamentation was studied in agar containing 0.0125% and 0.4% UDA+FAR after 24 h and 48 h of incubation. The results showed the presence of UDA and FAR on the composite and decreases in the water contact angle and metabolic activity of both the biofilm and planktonic cells at both time points at non-toxic UDA+FAR concentrations. Thus, the modification of PMMA with a combination of UDA+FAR reduces C. albicans biofilm formation on dentures and could be a promising anti-Candida strategy. Full article
Show Figures

Figure 1

Figure 1
<p>Fourier-transform infrared spectra of undecylenic acid (UDA), farnesol (FAR), poly(methyl methacrylate) (0% UDA+FAR), and composites with 3%, 6%, and 9% UDA+FAR.</p>
Full article ">Figure 2
<p>Water contact angle on the surface of the composites with 0%, 3%, 6%, and 9% of the total concentration of both antimicrobial compounds undecylenic acid (UDA) and farnesol (FAR). Asterisks above the columns show significant differences among the groups and compared to the controls (<span class="html-italic">p</span> &lt; 0.05, Tukey’s test). The results are presented as means ± SDs.</p>
Full article ">Figure 3
<p>Comparison of a metabolically active <span class="html-italic">Candida</span> (<span class="html-italic">C.</span>) <span class="html-italic">albicans</span> biofilm (<b>a</b>) and planktonic (<b>b</b>) cells after incubation with poly(methyl methacrylate) (PMMA) modified with the combination of equal concentrations of tundecylenic acid (UDA) and farnesol (FAR) (0%, 3%, 6%, and 9% UDA+FAR in total) in two points: 0 days (T0) and 6 days (T6) after PMMA_UDA+FAR composite preparation. The asterisks denote significant differences (<span class="html-italic">p</span> &lt; 0.05) compared to the PMMA control.</p>
Full article ">Figure 4
<p>The effect of the combination of undecylenic acid (UDA) and farnesol (FAR) on <span class="html-italic">C. albicans</span> growth after 24 h of incubation on the agar surface loaded with different UDA+FAR concentrations. The results are given as a mean ± standard deviation of optical density reading at a wavelength of 620 nm.</p>
Full article ">Figure 5
<p>Representative optical micrographs showing morphologic changes of <span class="html-italic">C. albicans</span> cells embedded in the agar without the combination of undecylenic acid (UDA) + farnesol (FAR) (control, 0% UDA+FAR, (<b>a</b>,<b>b</b>)), with 0.0125% UDA+FAR (<b>c</b>,<b>d</b>) and 0.4% UDA+FAR (<b>e</b>,<b>f</b>) after 24 h (<b>a</b>,<b>c</b>,<b>e</b>) and 48 h (<b>b</b>,<b>d</b>,<b>f</b>) of incubation. The arrows show spindle-shaped colonies with sporadic hyphal/pseudohyphal forms on the periphery and lateral yeasts in agar loaded with 0.0125% UDA+FAR in total (<b>c</b>,<b>d</b>). The images were taken with a magnification of 40×.</p>
Full article ">Figure 6
<p>The viability of human A549 cells incubated with media after the immersion of PMMA with the combination of undecylenic acid (UDA) and farnesol (FAR) in different concentrations, (in total, 3%, 6%, and 9% <span class="html-italic">w</span>/<span class="html-italic">w</span> UDA+FAR, more precisely per 1.5%, 3%, and 4.5% <span class="html-italic">w</span>/<span class="html-italic">w</span> of UDA and FAR, respectively), measured by the MTS test in two studied time points, T0—0 days—and T6—6 days after the PMMA_UDA+FAR disc preparation. The PMMA_UDA+FAR combinations were immersed for 24 h and six days longer for the T0 and T6 time points.</p>
Full article ">
23 pages, 3779 KiB  
Article
Vancomycin-Conjugated Polyethyleneimine-Stabilized Gold Nanoparticles Attenuate Germination and Show Potent Antifungal Activity against Aspergillus spp.
by Aishwarya Nikhil, Atul Kumar Tiwari, Ragini Tilak, Saroj Kumar, Prahlad Singh Bharti, Prem C. Pandey, Roger J. Narayan and Munesh Kumar Gupta
Appl. Sci. 2024, 14(16), 6926; https://doi.org/10.3390/app14166926 - 7 Aug 2024
Viewed by 353
Abstract
Antifungal drug resistance in filamentous fungi, particularly Aspergillus species, is increasing worldwide. Therefore, new antifungal drugs or combinations of drugs are urgently required to overcome this public health situation. In the present study, we examined the antifungal activity of vancomycin-functionalized AuNPs. These functionalized [...] Read more.
Antifungal drug resistance in filamentous fungi, particularly Aspergillus species, is increasing worldwide. Therefore, new antifungal drugs or combinations of drugs are urgently required to overcome this public health situation. In the present study, we examined the antifungal activity of vancomycin-functionalized AuNPs. These functionalized AuNPs were characterized, and their antifungal activity and associated killing mechanism were investigated using conventional methodologies against the conidia of A. fumigatus and A. flavus. The differential antifungal activity of vancomycin-functionalized Au-NPs against the conidia of Aspergillus species is dependent on structural differences in the conidial cell wall. The results demonstrated potent fungicidal activity against A. fumigatus, with a MIC value of 4.68 µg/mL, 93% germination inhibition, and 38.4% killing rate within 8 h of exposure. However, the activity against A. flavus was fungistatic; a MIC value of 18.7 µg/mL and 35% conidial germination inhibition, followed by 28.4% killing rate, were noted under similar conditions. Furthermore, endogenous reactive oxygen species (ROS) accumulation was 37.4 and 23.1% in conidial populations of A. fumigatus and A. flavus, respectively. Raman spectroscopy analysis confirmed the possible (but not confirmed) binding of functionalized AuNPs with the chitin and galactomannan components of the cell wall. A potential strategy that involves the exploration of antibacterial drugs using AuNPs as efficient drug carriers may also be appropriate for countering emerging drug resistance in filamentous fungi. Full article
(This article belongs to the Section Nanotechnology and Applied Nanosciences)
Show Figures

Figure 1

Figure 1
<p>(<b>A</b>) Physical characterization of PEI-AuNP@Van nanoparticles. (<b>a</b>) UV-Vis spectrum, (<b>b</b>) TEM micrograph, (<b>c</b>) high resolution image of nanoparticles, (<b>d</b>) mean nanoparticle size, (<b>e</b>) XRD diffractogram, and (<b>f</b>) dynamic light scattering data. (<b>B</b>) Zeta potential distribution histograms for non-functionalized gold nanoparticles (PEI-AuNPs) and vancomycin-functionalized gold nanoparticles (PEI-AuNP@Van).</p>
Full article ">Figure 1 Cont.
<p>(<b>A</b>) Physical characterization of PEI-AuNP@Van nanoparticles. (<b>a</b>) UV-Vis spectrum, (<b>b</b>) TEM micrograph, (<b>c</b>) high resolution image of nanoparticles, (<b>d</b>) mean nanoparticle size, (<b>e</b>) XRD diffractogram, and (<b>f</b>) dynamic light scattering data. (<b>B</b>) Zeta potential distribution histograms for non-functionalized gold nanoparticles (PEI-AuNPs) and vancomycin-functionalized gold nanoparticles (PEI-AuNP@Van).</p>
Full article ">Figure 2
<p>Fluorescence spectroscopic confirmation of vancomycin functionalization of PEI-stabilized gold nanoparticles. Adopted and modified under CC BY (2023) [<a href="#B33-applsci-14-06926" class="html-bibr">33</a>].</p>
Full article ">Figure 3
<p>Antifungal evaluation plates (<b>a</b>) against <span class="html-italic">A. fumigatus</span> and (<b>b</b>) <span class="html-italic">A. flavus</span> (NC = PEI-AuNPs control; NP = PEI-AuNP@Van; PC = positive control). (<b>c</b>) MIC values of non-functionalized AuNPs (PEI-AuNPs), vancomycin-functionalized AuNPs (PEI-AuNP@Van), positive control (voriconazole), vancomycin, and negative control (DW) against <span class="html-italic">A. fumigatus</span> and <span class="html-italic">A. flavus</span>.</p>
Full article ">Figure 4
<p>Conidial germination inhibition assay.</p>
Full article ">Figure 5
<p>Confocal micrograph of PEI-AuNP@Van treated conidia of <span class="html-italic">Aspergillus</span>. (<b>a</b>) Bright field image of <span class="html-italic">A. fumigatus</span> treated with PEI-AuNP@Van and (<b>b</b>) stained with PI; (<b>c</b>) blue panel showing surface adsorbed functionalized nanoparticles, and (<b>d</b>) merged panel; (<b>e</b>) bright field image of <span class="html-italic">A. flavus</span> treated with PEI-AuNP@Van and (<b>f</b>) stained with PI; (<b>g</b>) blue panel showing surface adsorbed functionalized nanoparticles, and (<b>h</b>) merged panel.</p>
Full article ">Figure 6
<p>Histograms of dead conidia after 8 h of exposure to PEI-AuNP@Van. (<b>a</b>,<b>c</b>) represent the untreated controls of <span class="html-italic">A. fumigatus</span> and <span class="html-italic">A. flavus</span>, respectively; (<b>b</b>,<b>d</b>) represent PEI-AuNP@Van-treated conidia of <span class="html-italic">A. fumigatus</span> and <span class="html-italic">A. flavus</span>.</p>
Full article ">Figure 7
<p>Endogenous ROS accumulation in PEI-AuNP@Van-treated conidia. (<b>a</b>,<b>c</b>) show the untreated control conidia of <span class="html-italic">A. fumigatus</span> and <span class="html-italic">A. flavus</span>, respectively; (<b>b</b>,<b>d</b>) represent the PEI-AuNP@Van-treated conidia.</p>
Full article ">Figure 8
<p>Raman spectrums of voriconazole and PEI-AuNP@Van-treated conidia of <span class="html-italic">A. flavus</span> and <span class="html-italic">A. fumigatus</span> along with an untreated control. (<b>ai</b>) Untreated conidia of <span class="html-italic">A. flavus</span>, (<b>aii</b>) treated with voriconazole, (<b>aiii</b>) treated with PEI-AuNP@Van, (<b>bi</b>) untreated conidia of <span class="html-italic">A. fumigatus</span>, (<b>bii</b>) treated with voriconazole, and (<b>biii</b>) treated with PEI-AuNP@Van.</p>
Full article ">Figure 9
<p>TEM micrograph of PEI-AuNP@Van-treated conidia of <span class="html-italic">A. fumigatus</span> and <span class="html-italic">A. flavus</span>: (<b>a</b>) represents the untreated conidia of <span class="html-italic">A. fumigatus</span>, and (<b>b</b>) treated with PEI-AuNP@Van; (<b>c</b>) represents the untreated conidia of <span class="html-italic">A. flavus</span>, and (<b>d</b>) treated with PEI-AuNP@Van.</p>
Full article ">
Back to TopTop