[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

Article Types

Countries / Regions

remove_circle_outline
remove_circle_outline
remove_circle_outline
remove_circle_outline
remove_circle_outline

Search Results (616)

Search Parameters:
Keywords = azoles

Order results
Result details
Results per page
Select all
Export citation of selected articles as:
20 pages, 1823 KiB  
Article
Synthesis, Copper(II) Binding, and Antifungal Activity of Tertiary N-Alkylamine Azole Derivatives
by Teresa Pissarro, Cláudia Malta-Luís, Luana Ferreira, Catarina Pimentel and Luís M. P. Lima
Inorganics 2024, 12(9), 242; https://doi.org/10.3390/inorganics12090242 - 5 Sep 2024
Viewed by 288
Abstract
The rise in antifungal resistance among medically important fungi causing severe infectious diseases has underscored the urgent need for developing more effective antifungal agents. Growing evidence suggests that compounds combining functional antifungal groups with metals are promising candidates and may well be the [...] Read more.
The rise in antifungal resistance among medically important fungi causing severe infectious diseases has underscored the urgent need for developing more effective antifungal agents. Growing evidence suggests that compounds combining functional antifungal groups with metals are promising candidates and may well be the key to addressing this global challenge. In this work, a range of new azole-containing tertiary amine compounds were prepared from three N-alkylamine azole skeletons appended with a 2,4-dihalogenobenzene function and one of the five different metal-binding motifs pyridine, quinoline, 8-hydroxyquinoline, 2-methoxyphenol, and 4-bromophenol. The copper(II) binding of these azole compounds was studied by spectrophotometric titrations in buffered aqueous medium to determine the metal binding equilibria and to comparatively characterize the copper(II)-binding ability of the compounds. The activity of all compounds against the opportunistic fungal pathogen Candida glabrata was also evaluated, allowing us to draw important conclusions about structure–activity relationships that will guide the future design of more effective metal-binding antifungal compounds. Full article
(This article belongs to the Special Issue Metal-Based Compounds: Relevance for the Biomedical Field)
Show Figures

Graphical abstract

Graphical abstract
Full article ">Figure 1
<p>Graphical representation of the pCu values of the studied compounds at pH = 7 and 100 μM concentration of total copper and ligand. Each color represents a metal-binding function (blue—pyridine, cyan—quinoline, green—8-hydroxyquinoline, red—2-methoxyphenol, and pink—4-bromophenol).</p>
Full article ">Figure 2
<p>Biological properties of the compounds <b>12c</b>, <b>16c</b>, and <b>19c</b> containing an 8-hydroxyquinoline metal-binding function. The ergosterol levels (<b>A</b>) of <span class="html-italic">Candida glabrata</span> cells left untreated or treated with the MIC concentration of the compounds in the presence or absence of Cu(II) or with 104.5 μM of fluconazole (Fluc) was measured by HPLC. Values are the means of three biological replicates and asterisks * or *** indicate <span class="html-italic">p</span> ≤ 0.05 or <span class="html-italic">p</span> ≤ 0.001, respectively. The intracellular levels of iron (<b>B</b>) and copper (<b>C</b>) in <span class="html-italic">Candida glabrata</span> cells left untreated or treated with the MIC concentration of the compounds were measured by ICP-AES. Values are the means of four biological replicates and asterisks *** indicates <span class="html-italic">p</span> ≤ 0.001 and **** <span class="html-italic">p</span> ≤ 0.0001. The cytotoxicity of the compounds in HeLa cells, in the absence (<b>D</b>) or presence (<b>E</b>) of Cu(II), was assessed using the MTT method. Values represent the mean of four biological replicates and are expressed as the percentage of viability relative to the control condition (cells treated with 0.5% DMSO).</p>
Full article ">Scheme 1
<p>Design of target copper(II)-binding azole compounds based on fluconazole. The pharmacophore fragments are represented in blue, and the metal-binding function in red.</p>
Full article ">Scheme 2
<p>Synthesis of intermediate <span class="html-italic">N</span>-alkylamine triazole skeletons.</p>
Full article ">Scheme 3
<p>Synthesis of target azole compounds.</p>
Full article ">
10 pages, 268 KiB  
Communication
Clinical Characteristics of Candidemia Due to Candida parapsilosis with Serial Episodes: Insights from 5-Year Data Collection at a Tertiary Hospital in Korea
by Eun Jeong Won, Heungsup Sung and Mi-Na Kim
J. Fungi 2024, 10(9), 624; https://doi.org/10.3390/jof10090624 - 1 Sep 2024
Viewed by 415
Abstract
Candida parapsilosis is a common cause of non-albicans Candida species causing candidemia, particularly invasive candidiasis. This study aimed to characterize candidemia due to the C. parapsilosis complex with serial episodes, including clinical and mycological features. Methods: Blood isolates of the C. parapsilosis complex [...] Read more.
Candida parapsilosis is a common cause of non-albicans Candida species causing candidemia, particularly invasive candidiasis. This study aimed to characterize candidemia due to the C. parapsilosis complex with serial episodes, including clinical and mycological features. Methods: Blood isolates of the C. parapsilosis complex were collected from February 2019 to January 2023 at a tertiary Korean hospital. Species identification was performed using Vitek 2 or matrix-assisted laser desorption/ionization time-of-flight mass spectrometry, and antifungal susceptibility testing was performed using the Sensititre YeastOne® system. Clinical information was collected, and characteristics were analyzed according to single or serial isolates. Results: A total of 586 blood isolates of the C. parapsilosis complex were recovered from 68 candidemia patients during the study period. Of them, only the first isolate per patient was investigated. The only two isolates were resistant to fluconazole and no isolate was resistant to echinocandins, amphotericin B, or 5-FC. A single episode of candidemia occurred in 35 patients, while serial episodes occurred in 33 patients. Underlying liver diseases, use of vasopressors, ICU admission, severe sepsis, and CVC use were more frequent in patients with serial episodes. There was no significant difference in the median MIC values of antifungal agents or the use of azoles or amphotericin B between single and serial episodes. However, patients with serial episodes more frequently received echinocandin therapy. Overall, there was no significant difference in the 30-day mortality rate between patients with single and serial episodes. Conclusion: Our data indicate that several factors related to the underlying conditions of the patients are associated with C. parapsilosis candidemia with serial episodes, rather than the characteristics of Candida itself. Full article
(This article belongs to the Collection Invasive Candidiasis)
17 pages, 351 KiB  
Review
Insight into Virulence and Mechanisms of Amphotericin B Resistance in the Candida haemulonii Complex
by Yuyan Huang, Yanyu Su, Xinfei Chen, Meng Xiao and Yingchun Xu
J. Fungi 2024, 10(9), 615; https://doi.org/10.3390/jof10090615 - 28 Aug 2024
Viewed by 624
Abstract
The Candida haemulonii complex includes emerging opportunistic human fungal pathogens with documented multidrug-resistance profiles. It comprises Candida haemulonii sensu stricto, Candida haemulonii var. vulnera, Candida duobushaemulonii, Candida pseudohaemulonii, and Candida vulturna. In recent years, rates of clinical isolation of [...] Read more.
The Candida haemulonii complex includes emerging opportunistic human fungal pathogens with documented multidrug-resistance profiles. It comprises Candida haemulonii sensu stricto, Candida haemulonii var. vulnera, Candida duobushaemulonii, Candida pseudohaemulonii, and Candida vulturna. In recent years, rates of clinical isolation of strains from this complex have increased in multiple countries, including China, Malaysia, and Brazil. Biofilm formation, hydrolytic enzymes, surface interaction properties, phenotype switching and cell aggregation abilities, extracellular vesicles production, stress response, and immune evasion help these fungi to infect the host and exert pathological effects. Multidrug resistance profiles also enhance the threat they pose; they exhibit low susceptibility to echinocandins and azoles and an intrinsic resistance to amphotericin B (AMB), the first fungal-specific antibiotic. AMB is commonly employed in antifungal treatments, and it acts via several known mechanisms. Given the propensity of clinical Candida species to initiate bloodstream infections, clarifying how C. haemulonii resists AMB is of critical clinical importance. This review outlines our present understanding of the C. haemulonii complex’s virulence factors, the mechanisms of action of AMB, and the mechanisms underlying AMB resistance. Full article
15 pages, 728 KiB  
Article
Exploring the Antibiofilm Effect of Sertraline in Synergy with Cinnamomum verum Essential Oil to Counteract Candida Species
by Alexia Barbarossa, Antonio Rosato, Antonio Carrieri, Luciana Fumarola, Roberta Tardugno, Filomena Corbo, Giuseppe Fracchiolla and Alessia Carocci
Pharmaceuticals 2024, 17(9), 1109; https://doi.org/10.3390/ph17091109 - 23 Aug 2024
Viewed by 475
Abstract
The emergence and spread of drug-resistant pathogens, resulting in antimicrobial resistance, continue to compromise our capability to handle commonly occurring infectious diseases. The rapid global spread of multi-drug-resistant pathogens, particularly systemic fungal infections, presents a significant concern, as existing antimicrobial drugs are becoming [...] Read more.
The emergence and spread of drug-resistant pathogens, resulting in antimicrobial resistance, continue to compromise our capability to handle commonly occurring infectious diseases. The rapid global spread of multi-drug-resistant pathogens, particularly systemic fungal infections, presents a significant concern, as existing antimicrobial drugs are becoming ineffective against them. In recent decades, there has been a notable increase in systemic fungal infections, primarily caused by Candida species, which are progressively developing resistance to azoles. Moreover, Candida species biofilms are among the most common in clinical settings. In particular, they adhere to biomedical devices, growing as a resilient biofilm capable of withstanding extraordinarily high antifungal concentrations. In recent years, many research programs have concentrated on the development of novel compounds with possible antimicrobial effects to address this issue, and new sources, such as plant-derived antimicrobial compounds, have been thoroughly investigated. Essential oils (EOs), among their numerous pharmacological properties, exhibit antifungal, antibacterial, and antiviral activities and have been examined at a global scale as the possible origin of novel antimicrobial compounds. A recent work carried out by our research group concerned the synergistic antibacterial activities of commercially available and chemically characterized Cinnamomum verum L. essential oil (C. verum EO) in association with sertraline, a selective serotonin reuptake inhibitor whose repositioning as a non-antibiotic drug has been explored over the years with encouraging results. The aim of this work was to explore the synergistic effects of C. verum EO with sertraline on both planktonic and sessile Candida species cells. Susceptibility testing and testing of the synergism of sertraline and C. verum EO against planktonic and sessile cells were performed using a broth microdilution assay and checkerboard methods. A synergistic effect was evident in both the planktonic cells and mature biofilms, with significant reductions in fungal viability. Indeed, the fractional inhibitory concentration index (FICI) was lower than 0.5 for all the associations, thus indicating significant synergism of the associations with the Candida strains examined. Moreover, the concentrations of sertraline able to inhibit Candida spp. strain growth and biofilm formation significantly decreased when it was used in combination with C. verum EO for all the strains considered, with a reduction percentage in the amount of each associated component ranging from 87.5% to 97%. Full article
(This article belongs to the Special Issue Natural Anti-Biofilm Agents)
Show Figures

Figure 1

Figure 1
<p>Binding pose of (<span class="html-italic">E</span>)-cinnamaldehyde to CYP51.</p>
Full article ">
51 pages, 24336 KiB  
Review
An Insight into Rational Drug Design: The Development of In-House Azole Compounds with Antimicrobial Activity
by Daniel Ungureanu, Ovidiu Oniga, Cristina Moldovan, Ioana Ionuț, Gabriel Marc, Anca Stana, Raluca Pele, Mihaela Duma and Brîndușa Tiperciuc
Antibiotics 2024, 13(8), 763; https://doi.org/10.3390/antibiotics13080763 - 13 Aug 2024
Viewed by 869
Abstract
Antimicrobial resistance poses a major threat to global health as the number of efficient antimicrobials decreases and the number of resistant pathogens rises. Our research group has been actively involved in the design of novel antimicrobial drugs. The blueprints of these compounds were [...] Read more.
Antimicrobial resistance poses a major threat to global health as the number of efficient antimicrobials decreases and the number of resistant pathogens rises. Our research group has been actively involved in the design of novel antimicrobial drugs. The blueprints of these compounds were azolic heterocycles, particularly thiazole. Starting with oxadiazolines, our research group explored, one by one, the other five-membered heterocycles, developing more or less potent compounds. An overview of this research activity conducted by our research group allowed us to observe an evolution in the methodology used (from inhibition zone diameters to minimal inhibitory concentrations and antibiofilm potential determination) correlated with the design of azole compounds based on results obtained from molecular modeling. The purpose of this review is to present the development of in-house azole compounds with antimicrobial activity, designed over the years by this research group from the departments of Pharmaceutical and Therapeutical Chemistry in Cluj-Napoca. Full article
(This article belongs to the Special Issue Discovery and Design of New Antimicrobial Agents)
Show Figures

Figure 1

Figure 1
<p>Structures of some of the currently FDA-approved azole drugs with antibacterial and antifungal activities.</p>
Full article ">Figure 2
<p>The general structures of the antimicrobial aryl and hetaryl-1,3,4-oxadiazoline compounds. The additional aromatic or heteroaromatic structures grafted on the 1,3,4-oxadiazoline ring were: 4-chloro-phenoxymethyl (<b>a</b>), pyridyl (<b>b</b>), 7-oxy-chromenyl (<b>c</b>), 2,4-bisthiazoles (<b>d</b>), 2-acetylamino-thiazole (<b>e</b>), and 2-aryl-thiazole (<b>f</b>).</p>
Full article ">Figure 3
<p>Structures of the chromenyl-acyl-hydrazones (<b>1a</b>–<b>m</b>) and chromenyl-1,3,4-oxadiazoline (<b>2a</b>–<b>m</b>) compounds.</p>
Full article ">Figure 4
<p>Structures of the antimicrobial 2,4-bisthiazolyl-1,3,4-oxadiazoline (<b>3a</b>–<b>i</b>) and <span class="html-italic">N</span>-acetyl-thiazolyl-1,3,4-oxadiazoline (<b>4a</b>–<b>i</b>) compounds.</p>
Full article ">Figure 5
<p>Structures of the antimicrobial 2-pyridyl-thiazolyl-1,3,4-oxadiazoline (<b>6a</b>–<b>g</b>) compounds and their acylhydrazones (<b>5a</b>–<b>g</b>).</p>
Full article ">Figure 6
<p>The general structures of the antimicrobial aryl and hetaryl-1,3,4-thiadiazoline compounds. The additional aromatic or heteroaromatic structures grafted on the 1,3,4-thiadiazoline ring were: chromone moieties (<b>a</b>), substituted phenyl rings (<b>b</b>), and aryl-thiazoles (<b>c</b>).</p>
Full article ">Figure 7
<p>Structures of the antimicrobial <span class="html-italic">N</span>-(4-acetyl-5-aryl-4,5-dihydro-1,3,4-thiadiazol-2-yl)-acetamides (<b>8a</b>–<b>h</b>), <span class="html-italic">N</span>-(4-acetyl-5-(2-arylthiazol-4-yl)-4,5-dihydro-1,3,4-thiadiazol-2-yl) (<b>8i</b>–<b>j</b>), and their corresponding <span class="html-italic">N<sup>1</sup></span>-arylidene-thiosemicarbazones (<b>7a</b>–<b>j</b>).</p>
Full article ">Figure 8
<p>The development of the antimicrobial oxadiazoline and thiadiazoline compounds. (-) means low or no activity against a strain, while (+) means activity against a strain. If the number of (+) increases, it means that the activity is better on certain strains. Similarly for the colors scheme: red means no or low activity, orange-yellow means low or moderate activity, while green means good to excellent activity.</p>
Full article ">Figure 9
<p>SAR studies in the antimicrobial 1,3,4-thiadiazolyl-thioethers (<b>9a</b>–<b>f</b>, <b>10a</b>–<b>b</b>, and <b>11a</b>–<b>e</b>) and Schiff bases (<b>12a</b>–<b>d</b>, <b>13a</b>–<b>d</b>, and <b>14a</b>–<b>b</b>). (+) means activity against a strain. If the number of (+) increases, it means that the activity is better on certain strains. Similarly for the colors scheme (reffering to the heat bars): red means no or low activity, orange-yellow means low or moderate activity, while green means good to excellent activity.</p>
Full article ">Figure 10
<p>Structures of the antimicrobial alkylidene-hydrazinyl-thiazole compounds.</p>
Full article ">Figure 11
<p>Structures of the antimicrobial 2-aryl-methylene-hydrazinyl-thiazolin-4-one derivatives.</p>
Full article ">Figure 12
<p>The unfavorable attempt to replace the thiazole ring with a thiazolin-4-one ring supplementary substituted in the fifth position. (+) means activity against a strain. If the number of (+) increases, it means that the activity is better on certain strains. Similarly for the colors scheme (reffering to the heat bars): red means no or low activity, orange-yellow means low or moderate activity, while green means good to excellent activity.</p>
Full article ">Figure 13
<p>The development of antimicrobial 5-arylidene-thiazolin-4-one derivatives as potential tryptophanyl-tRNA inhibitors.</p>
Full article ">Figure 14
<p>Structure-activity relationships in the antimicrobial 5-arylidene-thiazolin-4-one derivatives. (+) means activity against a strain. If the number of (+) increases, it means that the activity is better on certain strains. Similarly for the colors scheme (reffering to the heat bars): red means no or low activity, orange-yellow means low or moderate activity, while green means good to excellent activity.</p>
Full article ">Figure 15
<p>SAR studies on the antibacterial activity of the 3,5-disubstituted-thiazolidine-2,4-diones. The colors scheme (reffering to the heat bars): red means no or low activity, orange-yellow means low or moderate activity, while green means good to excellent activity.</p>
Full article ">Figure 16
<p>SAR studies in the antifungal <span class="html-italic">N</span>-substituted-5-hydroxyarylidene-thiazolidine-2,4-diones [<a href="#B139-antibiotics-13-00763" class="html-bibr">139</a>,<a href="#B140-antibiotics-13-00763" class="html-bibr">140</a>].</p>
Full article ">Figure 17
<p>SAR studies of the antifungal thiazolyl-1,2,4-triazole Schiff bases [<a href="#B148-antibiotics-13-00763" class="html-bibr">148</a>].</p>
Full article ">Figure 18
<p>SAR studies of the antibacterial thiazolyl-1,2,4-triazole Schiff bases [<a href="#B149-antibiotics-13-00763" class="html-bibr">149</a>].</p>
Full article ">Figure 19
<p>Structures of the antimicrobial 1,4-phenylene-bisthiazoles [<a href="#B156-antibiotics-13-00763" class="html-bibr">156</a>,<a href="#B157-antibiotics-13-00763" class="html-bibr">157</a>,<a href="#B158-antibiotics-13-00763" class="html-bibr">158</a>].</p>
Full article ">Figure 20
<p>SAR studies of antifungal 1,4-phenylene-bisthiazole acylhydrazone derivatives [<a href="#B158-antibiotics-13-00763" class="html-bibr">158</a>].</p>
Full article ">Figure 21
<p>Structures of the antimicrobial 4-(5-salicylamide)-thiazole derivatives [<a href="#B155-antibiotics-13-00763" class="html-bibr">155</a>,<a href="#B159-antibiotics-13-00763" class="html-bibr">159</a>].</p>
Full article ">Figure 22
<p>Structures of the antimicrobial 4,5′-bisthiazole derivatives [<a href="#B160-antibiotics-13-00763" class="html-bibr">160</a>].</p>
Full article ">Figure 23
<p>The development process of the antifungal thymolyl-thiazoles (HF = hydrophobic fragment; HBA = hydrogen bond acceptor; HBD = hydrogen bond donor) [<a href="#B158-antibiotics-13-00763" class="html-bibr">158</a>,<a href="#B161-antibiotics-13-00763" class="html-bibr">161</a>,<a href="#B162-antibiotics-13-00763" class="html-bibr">162</a>].</p>
Full article ">Figure 24
<p>Structures of the antifungal thymolyl-triazole derivatives [<a href="#B163-antibiotics-13-00763" class="html-bibr">163</a>]. Legend: <b>MeOH</b>—methanol; <b>DMF</b>—dimethylformamide; <b>rt</b>—room temperature.</p>
Full article ">Figure 25
<p>Structures of the antibiofilm 2-(3,4,5-trimethoxyphenyl)-4-Ar-5-R-thiazoles [<a href="#B186-antibiotics-13-00763" class="html-bibr">186</a>].</p>
Full article ">Figure 26
<p>Structures of the antibiofilm 1,4-phenylene-(2-phenyl)-bisthiazoles [<a href="#B187-antibiotics-13-00763" class="html-bibr">187</a>].</p>
Full article ">Figure 27
<p>The structures of antibiofilm pyridyl-thiazolyl-oxadiazoline derivatives [<a href="#B188-antibiotics-13-00763" class="html-bibr">188</a>].</p>
Full article ">Scheme 1
<p>The general synthetic route for the 4-acetyl-4,5-dihydro-1,3,4-oxadiazol-2-yl derivatives.</p>
Full article ">Scheme 2
<p>The general synthetic route for the 4-acetyl-4,5-dihydro-1,3,4-thiadiazol-2-yl derivatives. Legend: <b>abs.</b> = absolute; <b>EtOH</b> = ethanol; <b>rt</b> = room temperature; <b>t</b> = temperature; <b>Py</b> = pyridine.</p>
Full article ">Scheme 3
<p>The general synthetic routes for the antimicrobial 1,3,4-thiadiazolyl-thioethers (<b>9a</b>–<b>f</b>, <b>10a</b>–<b>b</b>, and <b>11a</b>–<b>e</b>) and Schiff bases (<b>12a</b>–<b>d</b>, <b>13a</b>–<b>d</b>, and <b>14a</b>–<b>b</b>). Legend: <b>EtOH</b> = ethanol; <b>AcOH</b> = acetic acid; <b>MW</b> = microwave; <b>W</b> = watts; <b>t</b> = temperature.</p>
Full article ">Scheme 4
<p>The general synthetic route for the alkylidene-hydrazinyl-thiazole derivatives.</p>
Full article ">Scheme 5
<p>The general synthetic route for the alkylidene- and arylidene-hydrazinyl-thiazolin-4-one derivatives. Legend: <b>abs.</b>—absolute; <b>anh.</b>—anhydrous.</p>
Full article ">Scheme 6
<p>The general synthetic pathways for the antimicrobial aryl- and hetaryl-thiazolidine-2,4-dione derivatives. Legend: <b>t</b>—temperature; <b>DMF</b>—dimethylformamide; <b>rt</b>—room temperature; <b>EtOH</b>—ethanol; <b>MW</b>—microwave; <b>W</b>—watts.</p>
Full article ">Scheme 7
<p>The synthetic pathway for the antimicrobial 3,5-disubstituted thiazolidinediones containing a PABA moiety [<a href="#B135-antibiotics-13-00763" class="html-bibr">135</a>]. Legend: <b>rt</b>—room temperature; <b>MW</b>—microwave; <b>Et<sub>3</sub>N</b>—triethylamine; <b>THF</b>—tetrahydrofuran; <b>DMF</b>—dimethylformamide; <b>W</b>—watts; <b>t</b>—temperature.</p>
Full article ">Scheme 8
<p>The general synthetic route for the piperazin-4-yl-(acetyl-thiazolidine-2,4-dione) norfloxacin analogues [<a href="#B141-antibiotics-13-00763" class="html-bibr">141</a>]. Legend: <b>Et<sub>3</sub>N</b>—triethylamine; <b>THF</b>—tetrahydrofuran; <b>MW</b>—microwave; <b>W</b>—watts; <b>t</b>—temperature.</p>
Full article ">Scheme 9
<p>The synthetic pathway for the antimicrobial thiazolyl-1,2,4-triazole Schiff bases [<a href="#B148-antibiotics-13-00763" class="html-bibr">148</a>,<a href="#B149-antibiotics-13-00763" class="html-bibr">149</a>]. Legend: <b>EtOH</b>—ethanol; <b>rt</b>—room temperature.</p>
Full article ">Scheme 10
<p>The general synthetic route for the in-house thiazoles obtained through Hantzsch condensation [<a href="#B155-antibiotics-13-00763" class="html-bibr">155</a>,<a href="#B156-antibiotics-13-00763" class="html-bibr">156</a>,<a href="#B157-antibiotics-13-00763" class="html-bibr">157</a>,<a href="#B158-antibiotics-13-00763" class="html-bibr">158</a>,<a href="#B159-antibiotics-13-00763" class="html-bibr">159</a>,<a href="#B160-antibiotics-13-00763" class="html-bibr">160</a>,<a href="#B161-antibiotics-13-00763" class="html-bibr">161</a>,<a href="#B162-antibiotics-13-00763" class="html-bibr">162</a>,<a href="#B163-antibiotics-13-00763" class="html-bibr">163</a>].</p>
Full article ">Scheme 11
<p>The general synthetic route and the structures of the antifungal thymolyl-thiazole derivatives [<a href="#B161-antibiotics-13-00763" class="html-bibr">161</a>,<a href="#B162-antibiotics-13-00763" class="html-bibr">162</a>]. Thymol was derivatized into two thioamide components (<b>A5</b> and <b>A6</b>), which were then treated with variously substituted 2-bromoacetophenones to yield series <b>52</b>–<b>55</b>. Legend: <b>rt</b>—room temperature; <b>t</b>—temperature; <b>EtOH</b>—ethanol.</p>
Full article ">Scheme 12
<p>The chemical synthesis route of the antibiofilm <span class="html-italic">N</span>-(oxazolylmethyl)-thiazolidinedione (<b>64</b>–<b>67: a</b>–<b>d</b>) [<a href="#B189-antibiotics-13-00763" class="html-bibr">189</a>]. Legend: <b>DMSO</b>—dimethyl sulfoxide; <b>t</b>—temperature; <b>rt</b>—room temperature.</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 678
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 605
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 ">
13 pages, 4237 KiB  
Article
Design of Marine Cyclodepsipeptide Analogues Targeting Candida albicans Efflux Pump CaCdr1p
by Ricardo Ribeiro, Sara Fortes, Lia Costa, Andreia Palmeira, Eugénia Pinto, Emília Sousa and Carla Fernandes
Drugs Drug Candidates 2024, 3(3), 537-549; https://doi.org/10.3390/ddc3030031 - 1 Aug 2024
Viewed by 585
Abstract
Fungal infections are a significant threat to human health and the environment. The emergence of multidrug-resistant strains of fungi and the growing prevalence of azole resistance in invasive fungal infections exacerbate the problem, with efflux pumps being a major cause of antifungal resistance [...] Read more.
Fungal infections are a significant threat to human health and the environment. The emergence of multidrug-resistant strains of fungi and the growing prevalence of azole resistance in invasive fungal infections exacerbate the problem, with efflux pumps being a major cause of antifungal resistance and a prime target for several counteractive strategies. In Candida albicans, the ATP-binding cassette superfamily transporter CaCdr1p is the predominant efflux pump involved in azole resistance. Marine organisms have unique phenotypic characteristics to survive in challenging environments, resulting in biologically active compounds. The cyclodepsipeptides unnarmicin A and C have shown promising results as inhibitors of rhodamine 6G efflux in cells expressing CaCdr1p. Herein, a series of unnarmicin analogues were designed and docked against a CaCdr1p efflux pump based on the cryogenic electron microscopy structure available to select the most promising compounds. Analogue 33 was predicted to be the best considering its high affinity for the efflux pump and pharmacokinetic profile. These results pave the way for further synthesis and in vitro biological studies of novel unnarmicins seeking a synergistic effect with fluconazole. Full article
(This article belongs to the Section Medicinal Chemistry and Preliminary Screening)
Show Figures

Figure 1

Figure 1
<p>Schematic view of CaCdr1p showing its two transmembrane domains (TMDs), two nucleotide-binding domains (NBDs), six extracellular loops (ECLs), four intracellular loops (ICLs), and twelve transmembrane helices (TMHs).</p>
Full article ">Figure 2
<p>Structures of unnarmicin A (<b>1</b>) and unnarmicin C (<b>2</b>).</p>
Full article ">Figure 3
<p>Structures of unnarmicin analogues <b>3</b>–<b>35</b>. The diverse structural moieties are highlighted at different color.</p>
Full article ">Figure 4
<p>(<b>A</b>) Unique conformation of compound <b>33</b> filling the active site, resembling a cannon-like structure. (<b>B</b>) The CaCdr1p active site showing all the Trp analogues (<b>17</b>–<b>20</b> and <b>33</b>). Notice the indole ring downward conformation of <b>33</b> highlighted in pink. (<b>C</b>) Conformations of the other Trp analogues (<b>17</b>–<b>20</b>) in the active site. (<b>D</b>) Conformation of unnarmicin C (<b>2</b>) in the active site.</p>
Full article ">Figure 5
<p>(<b>A</b>) Interaction of FLC (blue) (<b>47</b>) and analogue <b>33</b> (pink) in the same binding site of CaCdr1p efflux pump. Notice the perfect blockage of the FLC (<b>47</b>) by <b>33</b>. (<b>B</b>) View of the interaction of FLC (blue) (<b>47</b>) and <b>33</b> (pink) in the cannon-like structure of the active site. (<b>C</b>) View of the interaction of FLC (blue) (<b>47</b>) and other Trp analogues (<b>17</b>–<b>20</b>) in the cannon-like structure of the active site. (<b>D</b>) View of the interaction of FLC (blue) (<b>47</b>) and the unnarmicin C (orange) (<b>2</b>) in the cannon-like structure of the active site.</p>
Full article ">Scheme 1
<p>Retrosynthesis proposal for analogue <b>33</b>.</p>
Full article ">
44 pages, 14415 KiB  
Review
Towards Construction of the “Periodic Table” of 1-Methylbenzotriazole
by Christina Stamou, Zoi G. Lada, Sophia Paschalidou, Christos T. Chasapis and Spyros P. Perlepes
Inorganics 2024, 12(8), 208; https://doi.org/10.3390/inorganics12080208 - 30 Jul 2024
Viewed by 429
Abstract
Metal complexes of benzotriazole-type ligands continue to attract the intense interest of many inorganic chemistry groups around the world for a variety of reasons, including their aesthetically beautiful structures, physical properties and applications. 1-methylbenzotriazole (Mebta) is the N-substituted archetype of the parent [...] Read more.
Metal complexes of benzotriazole-type ligands continue to attract the intense interest of many inorganic chemistry groups around the world for a variety of reasons, including their aesthetically beautiful structures, physical properties and applications. 1-methylbenzotriazole (Mebta) is the N-substituted archetype of the parent 1H-benzotriazole. The first attempt to build a “periodic table” of Mebta, which includes its complexes with several metal ions, is described in this work. This, at first glance, trivial ligand has led to interesting results in terms of the chemistry, structures and properties of its metal complexes. This work reviews the to-date published coordination chemistry of Mebta with Mn(II), Fe(II), Fe(III), Co(II), Ni(II), Cu(I), Cu(II), Zn(II), Pd(II), Au(I) and {UVIO2}2+, with emphasis on their preparations, reactivity, structures and properties. Unpublished results from our group comprising other Co(II), Ni(II), Cu(II) and Zn(II) complexes, as well as Cd(II), Hg(II), Ag(I), In(III) and Sn(IV) ones are briefly reported. Mebta can also provide access to 1D and 3D heterometallic thiocyanato-bridged Co(II)/Hg(II) and Ni(II)/Hg(II) compounds. In almost all cases, Mebta behaves as a monodentate ligand with the nitrogen of position 3 of the azole ring as the donor atom. However, there are two copper complexes in which this molecule adopts a bidentate bridging coordination behavior. Our efforts to complete the “periodic table” of Mebta are continued. Full article
Show Figures

Figure 1

Figure 1
<p>(<b>Left</b>) The structural formula 1<span class="html-italic">H</span>-benzotriazole (btaH) and its derivatives. (<b>Right</b>) The structural formula of 1-methylbenzotriazole (Mebta), the subject of this review.</p>
Full article ">Figure 2
<p>The two tautomeric forms of btaH.</p>
Full article ">Figure 3
<p>Two Cu(II) complexes with closely related benzotriazole ligands with low (<b>1</b>) and high (<b>2</b>) in vitro antitumor activities; see text for discussion.</p>
Full article ">Figure 4
<p>The benzotriazole derivatives 5-amino-4-nitro-1<span class="html-italic">H</span>-benzo[1,2,3]triazole (<b>left</b>), 5-azido-4-nitro-1<span class="html-italic">H</span>-benzo[1,2,3]triazole (<b>middle</b>) and 5,7-dinitro-1-(2′,4′,6′-trinitrophenyl)benzo[1,2,3]triazole (<b>right</b>) which are efficient energetic materials.</p>
Full article ">Figure 5
<p>The to-date crystallographically observed coordination modes of the deprotonated C-substituted benzotriazoles and the Harris notation that describes these modes; R’ is a non-donor group. The coordination bonds are drawn with bold lines. The dashed lines indicate delocalization and the dotted line represents a H bond. The (btaHbta)<sup>−</sup> anion, illustrated in the bottom right of the figure, can be formally described as a neutral (btaH) and an anionic (bta<sup>−</sup>) benzotriazole and acts as a N, N’-bidentate “chelating” ligand (pseudo chelating), considering that the N∙∙∙H interaction is part of a 6-membered ring.</p>
Full article ">Figure 6
<p>Demethylation reactions of the stable nitrenium salt (Me)<sub>2</sub>btaI with strong nucleophiles to give crystalline Mebta.</p>
Full article ">Figure 7
<p>The to-date crystallographically confirmed coordination modes of Mebta in its metal complexes and the Harris notation that describes these modes. The 1.1<sub>3</sub>0<sub>1</sub>0<sub>2</sub> mode (<b>left</b>) has been observed in almost all of its complexes, and the 2.1<sub>2</sub>1<sub>3</sub>0<sub>1</sub> (<b>right</b>) in only two cases. The coordination bonds are drawn with bold lines.</p>
Full article ">Figure 8
<p>The structural formula of 1-methyl-1,2,3-triazole and its abbreviation.</p>
Full article ">Figure 9
<p>A view of <b>3</b> emphasizing the coordination sphere of one Mn<sup>II</sup> atom.</p>
Full article ">Figure 10
<p>The molecular structure of the dinuclear complex <b>5</b>.</p>
Full article ">Figure 11
<p>Plot of the molecular structure of <b>6</b>.</p>
Full article ">Figure 12
<p>The molecular structure of <b>7</b>.</p>
Full article ">Figure 13
<p>The molecular structure of <b>8</b>.</p>
Full article ">Figure 14
<p>The molecular structure of <b>9</b>.</p>
Full article ">Figure 15
<p>The molecular structure of <b>10</b>.</p>
Full article ">Figure 16
<p>The transformations in the 1:2 Ni(NO<sub>3</sub>)<sub>2</sub>/Mebta reaction mixtures; for more details, see text.</p>
Full article ">Figure 17
<p>The molecular structure of <b>12</b>.</p>
Full article ">Figure 18
<p>Plot of the [Ni(H<sub>2</sub>O)<sub>4</sub>(Mebta)<sub>2</sub>]<sup>2+</sup> cation that is present in the crystal structure of the salt <b>12b</b>.</p>
Full article ">Figure 19
<p>The molecular structure of <b>13</b>.</p>
Full article ">Figure 20
<p>The proposed <span class="html-italic">cis</span> square structure of <b>14</b> and <b>15</b> (M = Pd, Pt). The coordination bonds are represented by bold lines.</p>
Full article ">Figure 21
<p>The molecular structure of the cation [Cu(Mebta)<sub>4</sub>(H<sub>2</sub>O)]<sup>2+</sup> that is present in the crystal structure of <b>16</b>.</p>
Full article ">Figure 22
<p>The molecular structure of <b>17</b>.</p>
Full article ">Figure 23
<p>The molecular structure of <b>18</b>.</p>
Full article ">Figure 24
<p>A small portion of one chain that is present in the crystal structure of <b>19</b>.</p>
Full article ">Figure 25
<p>The molecular structure of <b>20</b>.</p>
Full article ">Figure 26
<p>The molecular structure of the cation [Cu<sub>2</sub>Cl<sub>2</sub>(Mebta)<sub>6</sub>]<sup>2+</sup> that exists in the crystal structure of its perchlorate salt <b>21</b>.</p>
Full article ">Figure 27
<p>The molecular structure of <b>22</b>.</p>
Full article ">Figure 28
<p>The molecular structure of <b>23</b>.</p>
Full article ">Figure 29
<p>The molecular structure of one of the two crystallographically independent [CuBr<sub>2</sub>(Mebta)<sub>3</sub>] molecules that exist in the crystal structure of <b>24</b>.</p>
Full article ">Figure 30
<p>A view of <b>25</b> emphasizing the coordination sphere of an Cu<sup>II</sup> atom in the linear chain.</p>
Full article ">Figure 31
<p>Schematic illustration of a small portion of one linear chain that is present in the crystal structure of <b>26</b>.</p>
Full article ">Figure 32
<p>A small portion of one double chain of <b>27</b>.</p>
Full article ">Figure 33
<p>The asymmetric unit of the 1D polymer <b>28</b>.</p>
Full article ">Figure 34
<p>A small part of a 1D chain containing corrugated double-stranded stairlike {CuI}<sub>n</sub> chain; this model was derived from a poor structural solution of <b>29</b>. N-N represents the N(2)-N(3) part of Mebta that contains the two donor atoms of the Mebta ligands (the whole ligands are not drawn for clarity reasons). The Cu<sup>I</sup> atoms are tetrahedral, with a coordination sphere comprising three iodo groups and one N atom of a bidentate bridging ligand. The terminal Cu<sup>II</sup> atoms appear as three-coordinate, but they are actually four-coordinate.</p>
Full article ">Figure 35
<p>The molecular structures of <b>30</b> (<b>up</b>) and <b>31</b> (<b>bottom</b>).</p>
Full article ">Figure 36
<p>The molecular structures of <b>32</b> (<b>up</b>) and <b>33</b> (<b>bottom</b>).</p>
Full article ">Figure 37
<p>The molecular structure of the cation [Zn(Mebta)<sub>4</sub>]<sup>2+</sup> that is present in the crystal structure of complex <b>34</b>.</p>
Full article ">Figure 38
<p>Complex <b>35</b>-catalyzed cross-coupling of terminal alkynes with ethynylbenzoiodioxole hypervalent iodine reagents (R<sup>1</sup>, R<sup>2</sup> = various organic groups).</p>
Full article ">Figure 39
<p>The molecular structure of the cation [Au(Ph<sub>3</sub>P)(Mebta)]<sup>+</sup> that is present in the crystal structure of <b>35</b>.</p>
Full article ">Figure 40
<p>The molecular structure of <b>36</b>.</p>
Full article ">Figure 41
<p>The molecular structure of the cation [Cu<sub>3</sub>(OH)<sub>2</sub>(Mebta)<sub>10</sub>]<sup>4+</sup> that is present in the crystal structure of <b>38</b>.</p>
Full article ">Figure 42
<p>The molecular structures of <b>42</b> (<b>up</b>) and <b>43</b> (<b>bottom</b>).</p>
Full article ">Figure 43
<p>The molecular structure of <b>46</b>.</p>
Full article ">Figure 44
<p>A small portion of one linear chain that is present in the crystal structure of <b>50</b>.</p>
Full article ">Figure 45
<p>The molecular structure of <b>53</b>.</p>
Full article ">Figure 46
<p>The molecular structure of <b>55.</b></p>
Full article ">Figure 47
<p>The molecular structure of <b>59.</b></p>
Full article ">Figure 48
<p>The molecular structure of <b>57</b>.</p>
Full article ">Figure 49
<p>A small portion of one chain that is present in the crystal structure of <b>66</b>.</p>
Full article ">Figure 50
<p>The present form of the “periodic table” of Mebta. Colour code: Red; metals whose complexes with Mebta have been published. Yellow; metals with unpublished coordination chemistry of Mebta.</p>
Full article ">
17 pages, 3909 KiB  
Article
Harnessing Machine Learning to Uncover Hidden Patterns in Azole-Resistant CYP51/ERG11 Proteins
by Otávio Guilherme Gonçalves de Almeida and Marcia Regina von Zeska Kress
Microorganisms 2024, 12(8), 1525; https://doi.org/10.3390/microorganisms12081525 - 25 Jul 2024
Viewed by 510
Abstract
Fungal resistance is a public health concern due to the limited availability of antifungal resources and the complexities associated with treating persistent fungal infections. Azoles are thus far the primary line of defense against fungi. Specifically, azoles inhibit the conversion of lanosterol to [...] Read more.
Fungal resistance is a public health concern due to the limited availability of antifungal resources and the complexities associated with treating persistent fungal infections. Azoles are thus far the primary line of defense against fungi. Specifically, azoles inhibit the conversion of lanosterol to ergosterol, producing defective sterols and impairing fluidity in fungal plasmatic membranes. Studies on azole resistance have emphasized specific point mutations in CYP51/ERG11 proteins linked to resistance. Although very insightful, the traditional approach to studying azole resistance is time-consuming and prone to errors during meticulous alignment evaluation. It relies on a reference-based method using a specific protein sequence obtained from a wild-type (WT) phenotype. Therefore, this study introduces a machine learning (ML)-based approach utilizing molecular descriptors representing the physiochemical attributes of CYP51/ERG11 protein isoforms. This approach aims to unravel hidden patterns associated with azole resistance. The results highlight that descriptors related to amino acid composition and their combination of hydrophobicity and hydrophilicity effectively explain the slight differences between the resistant non-wild-type (NWT) and WT (nonresistant) protein sequences. This study underscores the potential of ML to unravel nuanced patterns in CYP51/ERG11 sequences, providing valuable molecular signatures that could inform future endeavors in drug development and computational screening of resistant and nonresistant fungal lineages. Full article
(This article belongs to the Special Issue Healthcare-Associated Infections and Antimicrobial Therapy)
Show Figures

Figure 1

Figure 1
<p>Experimental design for CYP51/ERG11 isoforms obtained from public databases and machine learning modeling.</p>
Full article ">Figure 2
<p>Modeling metrics. (<b>A</b>) Performance of scaling methods. (<b>B</b>) Accuracy of machine learning algorithms using Uniform, the best scaling method. (<b>C</b>) ROC–AUC of machine learning algorithms using Uniform, the best scaling method. *, <span class="html-italic">p</span>-value of ≤0.05; ns, <span class="html-italic">p</span>-value of &gt;0.05 (Mann–Whitney multiple groups comparison with Bonferroni correction). (<b>D</b>) MCC of machine learning algorithms using Uniform, the best scaling method. (<b>E</b>) Comparison of accuracy metric between training and test datasets among machine learning algorithms.</p>
Full article ">Figure 3
<p>Metrics of supervised learning. (<b>A</b>) The confusion matrix shows correct and incorrect predictions for WT and NWT phenotypes, and (<b>B</b>) heatmap resuming the main classification metrics.</p>
Full article ">Figure 4
<p>Most important (top 20) features’ attributes. (<b>A</b>) Main attributes related to WT and NWT phenotypes used in random forest classifier. Permutation importance analysis on (<b>B</b>) train and (<b>C</b>) test datasets.</p>
Full article ">Figure 5
<p>The protein logo shows the conservation degree of motifs shared by CYP51 and ERG11 amino acids’ sequences.</p>
Full article ">
15 pages, 2697 KiB  
Article
Yeast Particle Encapsulation of Azole Fungicides for Enhanced Treatment of Azole-Resistant Candida albicans
by Ernesto R. Soto, Florentina Rus and Gary R. Ostroff
J. Funct. Biomater. 2024, 15(8), 203; https://doi.org/10.3390/jfb15080203 - 23 Jul 2024
Cited by 1 | Viewed by 688
Abstract
Addressing the growing problem of antifungal resistance in medicine and agriculture requires the development of new drugs and strategies to preserve the efficacy of existing fungicides. One approach is to utilize delivery technologies. Yeast particles (YPs) are 3–5 µm porous, hollow microspheres, a [...] Read more.
Addressing the growing problem of antifungal resistance in medicine and agriculture requires the development of new drugs and strategies to preserve the efficacy of existing fungicides. One approach is to utilize delivery technologies. Yeast particles (YPs) are 3–5 µm porous, hollow microspheres, a byproduct of food-grade Saccharomyces cerevisiae yeast extract manufacturing processes and an efficient and flexible drug delivery platform. Here, we report the use of YPs for encapsulation of tetraconazole (TET) and prothioconazole (PRO) with high payload capacity and stability. The YP PRO samples were active against both sensitive and azole-resistant strains of Candida albicans. The higher efficacy of YP PRO versus free PRO is due to interactions between PRO and saponifiable lipids in the YPs. Encapsulation of PRO in glucan lipid particles (GLPs), a highly purified form of YPs that do not contain saponifiable lipids, did not result in enhanced PRO activity. We evaluated the co-encapsulation of PRO with a mixture of the terpenes: geraniol, eugenol, and thymol. Samples co-encapsulating PRO and terpenes in YPs or GLPs were active on both sensitive and azole-resistant C. albicans. These approaches could lead to the development of more effective drug combinations co-encapsulated in YPs for agricultural or GLPs for pharmaceutical applications. Full article
(This article belongs to the Special Issue Active Biomedical Materials and Their Applications)
Show Figures

Figure 1

Figure 1
<p>Schematics of two methods for loading of fungicides in yeast particles: (<b>A</b>) solvent-free loading method and (<b>B</b>) loading of payload in an organic solvent.</p>
Full article ">Figure 2
<p>Encapsulation efficiency and microscopy images showing Nile-red-stained fungicides inside the cavity of YPs: (<b>A</b>) tetraconazole and (<b>B</b>) prothioconazole. Both compounds were loaded at a target weight ratio of 1:1 fungicide:YP.</p>
Full article ">Figure 3
<p>Kinetics of free and YP-encapsulated TET (Tetraconazole (<b>A</b>)) and PRO (prothioconazole (<b>B</b>)) released over time following dilution of YP fungicide samples in water at different fungicide concentrations and incubation at 23 °C.</p>
Full article ">Figure 4
<p>Effect of YP encapsulation on antifungal activity of azole fungicides reported as the ratio of the MIC 75% of unencapsulated and YP-encapsulated samples (statistically significant results were obtained between the paired samples, * <span class="html-italic">p</span> &lt; 0.1, ** <span class="html-italic">p</span> &lt; 0.05).</p>
Full article ">Figure 5
<p><span class="html-italic">Candida albicans</span> growth response curves with unencapsulated PRO and PRO encapsulated in YPs or GLPs.</p>
Full article ">Figure 6
<p>Heat maps showing effect of unencapsulated GET 212 and unencapsulated PRO (heat maps (<b>A</b>–<b>C</b>)), and YP GET 212 and YP PRO (heat maps (<b>D</b>–<b>F</b>)) on growth inhibition of sensitive and azole-resistant <span class="html-italic">C. albicans</span> strains.</p>
Full article ">Figure 7
<p>Two-dimensional contour plots showing synergy scores of unencapsulated GET212 and PRO (top (<b>A</b>–<b>C</b>)), and YP GET212 and YP PRO (bottom (<b>D</b>–<b>F</b>)). ZIP synergy scoring and 2D contour plots were generated using SynergyFinder R online tool (<a href="https://synergyfinder.org" target="_blank">https://synergyfinder.org</a>, accessed on 14 March 2024).</p>
Full article ">Figure 8
<p>(<b>A</b>) Schematics of GET 212 and PRO co-encapsulation in YPs and GLPs, (<b>B</b>) GET212 and PRO encapsulation efficiency in YPs and GLPs in samples prepared at 0.055:1.1:1 PRO:GET 212:YP or GLP weight ratio, and (<b>C</b>) PRO release in water at 37 °C from control YP and GLP particles containing only PRO and samples co-encapsulating PRO and GET; samples were diluted at 0.05 mg PRO/mL (~2.3× higher concentration than maximum solubility of PRO in water) and at 0.01 mg PRO/mL (~2.2× lower concentration than maximum solubility of PRO in water).</p>
Full article ">Figure 9
<p>Effect of YP or GLP encapsulation on activity of prothioconazole, reported as the ratio of the MIC 75% of unencapsulated PRO and YP- or GLP-encapsulated PRO samples. Statistically significant results were obtained between the paired samples indicated with letter superscripts (A–E <span class="html-italic">p</span> &lt; 0.005, F–H <span class="html-italic">p</span> &lt; 0.01, I–R <span class="html-italic">p</span> &lt; 0.05).</p>
Full article ">
23 pages, 8826 KiB  
Article
Synthesis, Characterization, DNA, Fluorescence, Molecular Docking, and Antimicrobial Evaluation of Novel Pd(II) Complex Containing O, S Donor Schiff Base Ligand and Azole Derivative
by Maged S. Al-Fakeh, Muneera Alrasheedi, Ard Elshifa M. E. Mohammed, Ahmed B. M. Ibrahim, Sadeq M. Al-Hazmy, Ibrahim A. Alhagri and Sabri Messaoudi
Inorganics 2024, 12(7), 189; https://doi.org/10.3390/inorganics12070189 - 11 Jul 2024
Viewed by 683
Abstract
Pd(II) with the Schiff base ligand 2-Hydroxy-3-Methoxy Benzaldehyde-Thiosemicarbazone (HMBATSC) (L2) and 2-aminobenzothiazole (2-ABZ) (L1) was synthesized. The Schiff base ligand and the Palladium(II) complex were characterized by C.H.N.S, FT-IR, conductance studies, magnetic susceptibility, XRD, and TGA. From the elemental analysis and spectral data, [...] Read more.
Pd(II) with the Schiff base ligand 2-Hydroxy-3-Methoxy Benzaldehyde-Thiosemicarbazone (HMBATSC) (L2) and 2-aminobenzothiazole (2-ABZ) (L1) was synthesized. The Schiff base ligand and the Palladium(II) complex were characterized by C.H.N.S, FT-IR, conductance studies, magnetic susceptibility, XRD, and TGA. From the elemental analysis and spectral data, the complex was proposed to have the formula [Pd(HMBATSC)(2-ABZ)H2O]. The interaction between the Pd(II) complex and DNA was examined through various methods, including UV–Vis spectroscopy, fluorescence techniques, and DNA viscosity titrations. The findings provided strong evidence that the interaction between the Pd(II) complex and DNA occurs through the intercalation mode. The analysis yielded the following values: a Stern–Volmer quenching constant (ksv) of 1.67 × 104 M−1, a quenching rate constant (kq) of 8.35 × 1011 M−1 s−1, a binding constant (kb) of 5.20 × 105 M−1, and a number of binding the sites (n) of 1.392. DFT studies suggest that the azole derivative may act as an electron donor through pyridine nitrogen, while the Schiff base ligand may act as an electron donor via oxygen and sulfur atoms. TDDFT calculations indicate that the intramolecular charge transfer from the Schiff base to Pd(II) is responsible for the complex’s fluorescence quenching. The powder X-ray diffraction data revealed that the complex is arranged in a monoclinic system. The resulting Pd(II) complex was investigated for its antimicrobial activity and demonstrated antibacterial efficiency. Interestingly, it showed potent activity against E. coli and E. niger that was found to be more powerful than that recorded for Neomycin. Full article
(This article belongs to the Special Issue Noble Metals in Medicinal Inorganic Chemistry)
Show Figures

Figure 1

Figure 1
<p>Structure of ligands.</p>
Full article ">Figure 2
<p>FT-IR of Pd(II) compound.</p>
Full article ">Figure 3
<p>Absorption spectra of Pd(II) compound in ethanol (5 × 10<sup>−5</sup> M).</p>
Full article ">Figure 4
<p>Absorption spectra of 5 × 10<sup>−5</sup> M L1, L2, and Pd(II) complex in ethanol.</p>
Full article ">Figure 5
<p>Absorption and fluorescence λ<sub>ex</sub> = 320 nm spectra of Pd(II) complex in ethanol (5 × 10<sup>−5</sup> M).</p>
Full article ">Figure 6
<p>Absorption, fluorescence, λ<sub>ex</sub> = 250 nm, and excitation, λ<sub>em max</sub> = 420 nm spectra of Pd(II) complex in ethanol (5 × 10<sup>−5</sup> M).</p>
Full article ">Figure 7
<p>Normalized fluorescence spectra of 5 × 10<sup>−5</sup> M L1, L2, and Pd(II) complex λ<sub>ex</sub> = λ<sub>ab</sub>. max.</p>
Full article ">Figure 8
<p>Suggested structures of Pd(II) complex containing HMBATSC and ABZ ligands.</p>
Full article ">Figure 9
<p>T.G, T.G.A, and D.T.A of Pd(II) compound.</p>
Full article ">Figure 10
<p>X-ray diffraction pattern of Pd(II) complex. <span style="color:red">_____Y obs</span>; <span style="color:blue">_____ Y cal</span>.</p>
Full article ">Figure 11
<p>Microbiological screening of Pd(II) complex against <span class="html-italic">S. aureus</span>.</p>
Full article ">Figure 12
<p>Microbiological screening of Pd(II) complex against <span class="html-italic">E. coli</span>.</p>
Full article ">Figure 13
<p>Microbiological screening of Pd(II) complex against <span class="html-italic">C. albicans</span>.</p>
Full article ">Figure 14
<p>Microbiological screening of Pd(II) complex against <span class="html-italic">A. niger</span>.</p>
Full article ">Figure 15
<p>UV–Vis absorption spectra of Pd(II) complex (2.4 × 10<sup>−5</sup> M) in absence and presence of DNA. The concentrations of DNA from 0 to 7 were (1) 0.0, (2) 2 × 10<sup>−7</sup> M, (3) 6 × 10<sup>−7</sup> M, (4) 1 × 10<sup>−6</sup> M, (5) 1.6 × 10<sup>−6</sup> M, (6) 2 × 10<sup>−6</sup> M, and (7) 2.4 × 10<sup>−6</sup> M.</p>
Full article ">Figure 16
<p>The plot of log (<span class="html-italic">A<sub>o</sub></span>/<span class="html-italic">A<sub>o</sub></span> − <span class="html-italic">A</span>) versus log 1/<span class="html-italic">C<sub>DNA</sub></span>.</p>
Full article ">Figure 17
<p>Emission intensity of 2.4 × 10<sup>−5</sup> M EB, 7.2 × 10<sup>−5</sup> M DNA, 2.4 × 10<sup>−5</sup> M Pd(II) complex, and a mixture of 2.4 × 10<sup>−5</sup> M EB and 7.2 × 10<sup>−5</sup> M DNA, at λ<sub>ex</sub> = 350 nm.</p>
Full article ">Figure 18
<p>Fluorescence spectra of EB bound to DNA in absence and presence of Pd(II) complex. C<sub>EB</sub> = 2.4 × 10<sup>−5</sup> M; C<sub>DNA</sub> = 7.2 × 10<sup>−5</sup> M; C<sub>Pd(II)</sub> complex = (1) 0.00, (2) 1.2 × 10<sup>−5</sup>, (3) 2.4 × 10<sup>−5</sup>, (4) 3.6 × 10<sup>−5</sup>, (5) 4.8 × 10<sup>−5</sup>, (6) 6.0 × 10<sup>−5</sup>, (7) 9.6 × 10<sup>−5</sup>, (8) 1.2 × 10<sup>−4</sup>, (9) 1.44 × 10<sup>−4</sup>, (10) 1.68 × 10<sup>−4</sup> M.</p>
Full article ">Figure 19
<p><b>A</b> Stern–Volmer plot for the quenching of the fluorescence of EB-DNA caused by the Pd(II) compound.</p>
Full article ">Figure 20
<p>Plot of −log (<span class="html-italic">F<sub>o</sub></span> − <span class="html-italic">F</span>)/<span class="html-italic">F</span> versus log[<span class="html-italic">Q</span>].</p>
Full article ">Figure 21
<p>Effect of Pd(II) complex on viscosity of DNA solution. Pd(II) complex was titrated into 2.4 × 10<sup>−5</sup> M DNA solution at 298 k.</p>
Full article ">Figure 22
<p>B3LYP/6–31g(d)-optimized structures of L2H2 and L2<sup>2−</sup>.</p>
Full article ">Figure 23
<p>HOMO of L2H2 (<b>left top</b>) and L2<sup>2−</sup> (<b>right top</b>). Electrostatic potential maps of L2H2 (<b>left bottom</b>) and L2<sup>2−</sup> (<b>right bottom</b>). Electrostatic potential maps are superimposed over total electronic density. Colors demonstrate the following: red—negative, blue—positive, and green—zero electrostatic potential.</p>
Full article ">Figure 24
<p>B3LYP/6–31g(d)-optimized structures of L1.</p>
Full article ">Figure 25
<p>HOMO of L1 (<b>top</b>). Electrostatic potential maps of L1 (<b>bottom</b>). Electrostatic potential maps are superimposed over total electronic density. Colors demonstrate the following: red—negative, blue—positive, and green—zero electrostatic potential.</p>
Full article ">Figure 26
<p>B3LYP-optimized structures of (<b>a</b>) Pd(L2)(<sup>N</sup>L1), (<b>b</b>) Pd(L2)(<sup>NH2</sup>L1), (<b>c</b>) Pd(L2)(<sup>S</sup>L1).</p>
Full article ">Figure 27
<p>Molecular electrostatic potential superimposed over electronic isodensity surface for Pd(L2)(<sup>N</sup>L1).</p>
Full article ">Figure 28
<p>Frontier orbitals and their energies for (<b>A</b>) Pd(L2)(<sup>N</sup>L1), (<b>B</b>) L1, and (<b>C</b>) L2H2. (Energy in a.u.).</p>
Full article ">
11 pages, 655 KiB  
Article
Insights into Aspergillus fumigatus Colonization in Cystic Fibrosis and Cross-Transmission between Patients and Hospital Environments
by Laís Pontes, Ana Luisa Perini Leme Giordano, Franqueline Reichert-Lima, Caio Augusto Gualtieri Beraquet, Guilherme Leite Pigolli, Teppei Arai, José Dirceu Ribeiro, Aline Cristina Gonçalves, Akira Watanabe, Gustavo Henrique Goldman, Maria Luiza Moretti and Angélica Zaninelli Schreiber
J. Fungi 2024, 10(7), 461; https://doi.org/10.3390/jof10070461 - 29 Jun 2024
Viewed by 815
Abstract
Background: Approximately 60% of individuals with cystic fibrosis (CF) are affected by Aspergillus fumigatus infection. This condition is correlated with a decline in lung function and is identified as an independent risk factor contributing to hospital admissions among CF patients. This study investigates [...] Read more.
Background: Approximately 60% of individuals with cystic fibrosis (CF) are affected by Aspergillus fumigatus infection. This condition is correlated with a decline in lung function and is identified as an independent risk factor contributing to hospital admissions among CF patients. This study investigates the dynamic interplay of A. fumigatus within the context of CF patients, tracing its evolution over time, with a specific emphasis on colonization dynamics. Methods: An analysis was conducted on 83 sequential A. fumigatus isolates derived from sputum samples of six patients receiving care at a renowned CF hospital in Brazil. Employing microsatellite genotyping techniques, alongside an investigation into cyp51A gene mutations, this research sheds light on the genetic variations, colonization, and resistance of A. fumigatus within the CF respiratory environment. Results: Our research findings indicate that CF patients can harbor A. fumigatus strains from the same clonal complexes for prolonged periods. Additionally, we identified that clinical isolates have the potential to spread among patients in the same healthcare facility, evidencing hospital contamination. Two patients who underwent long-term Itraconazole treatment did not show phenotypic resistance. However, one of these patients exhibited mutations in the cyp51A gene, indicating the need to monitor resistance to azoles in these patients colonized for long periods by A. fumigatus. We also observed co-colonization or co-infection involving multiple genotypes in all patients over time. Conclusion: This comprehensive examination offers valuable insights into the pathogenesis of A. fumigatus infections in CF patients, potentially shaping future therapeutic strategies and management approaches. This enhanced understanding contributes to our knowledge of A. fumigatus impact on disease progression in individuals with cystic fibrosis. Additionally, the study provides evidence of cross-contamination among patients undergoing treatment at the same hospital. Full article
(This article belongs to the Special Issue Young Investigators of Human Pathogenic Fungi)
Show Figures

Figure 1

Figure 1
<p>Genotypic relationship between 80 isolates of <span class="html-italic">Aspergillus fumigatus</span> isolates. The dendrogram is based on a categorical analysis of nine microsatellite markers in combination with an unweighted pair group method with arithmetic mean (UPGMA) clustering using BioNumerics V7.6 software (Applied Math Inc., Austin, TX, USA).</p>
Full article ">
20 pages, 4350 KiB  
Article
Synthesis and Biological Evaluation of New Compounds with Nitroimidazole Moiety
by Katarzyna Dziduch, Sara Janowska, Sylwia Andrzejczuk, Paulina Strzyga-Łach, Marta Struga, Marcin Feldo, Oleg Demchuk and Monika Wujec
Molecules 2024, 29(13), 3023; https://doi.org/10.3390/molecules29133023 - 26 Jun 2024
Viewed by 1304
Abstract
Heterocyclic compounds, particularly those containing azole rings, have shown extensive biological activity, including anticancer, antibacterial, and antifungal properties. Among these, the imidazole ring stands out due to its diverse therapeutic potential. In the presented study, we designed and synthesized a series of imidazole [...] Read more.
Heterocyclic compounds, particularly those containing azole rings, have shown extensive biological activity, including anticancer, antibacterial, and antifungal properties. Among these, the imidazole ring stands out due to its diverse therapeutic potential. In the presented study, we designed and synthesized a series of imidazole derivatives to identify compounds with high biological potential. We focused on two groups: thiosemicarbazide derivatives and hydrazone derivatives. We synthesized these compounds using conventional methods and confirmed their structures via nuclear magnetic resonance spectroscopy (NMR), MS, and elemental analysis, and then assessed their antibacterial and antifungal activities in vitro using the broth microdilution method against Gram-positive and Gram-negative bacteria, as well as Candida spp. strains. Our results showed that thiosemicarbazide derivatives exhibited varied activity against Gram-positive bacteria, with MIC values ranging from 31.25 to 1000 µg/mL. The hydrazone derivatives, however, did not display significant antibacterial activity. These findings suggest that structural modifications can significantly influence the antimicrobial efficacy of imidazole derivatives, highlighting the potential of thiosemicarbazide derivatives as promising candidates for further development in antibacterial therapies. Additionally, the cytotoxic activity against four cancer cell lines was evaluated. Two derivatives of hydrazide-hydrazone showed moderate anticancer activity. Full article
Show Figures

Scheme 1

Scheme 1
<p>The synthetic route of thiosemicarbazide derivatives (<b>2</b>–<b>18</b>) and hydrazide-hydrazone derivatives (<b>19</b>–<b>33</b>).</p>
Full article ">
22 pages, 3615 KiB  
Article
Bioprospecting, Synergistic Antifungal and Toxicological Aspects of the Hydroxychalcones and Their Association with Azole Derivates against Candida spp. for Treating Vulvovaginal Candidiasis
by Lígia de Souza Fernandes, Letícia Sayuri Ogasawara, Kaila Petronila Medina-Alarcón, Kelvin Sousa dos Santos, Samanta de Matos Silva, Letícia Ribeiro de Assis, Luís Octavio Regasini, Anselmo Gomes de Oliveira, Maria José Soares Mendes Giannini, Maria Virginia Scarpa and Ana Marisa Fusco Almeida
Pharmaceutics 2024, 16(7), 843; https://doi.org/10.3390/pharmaceutics16070843 - 21 Jun 2024
Viewed by 663
Abstract
Vulvovaginal candidiasis (VVC) remains a prevalent fungal disease, characterized by challenges, such as increased fungal resistance, side effects of current treatments, and the rising prevalence of non-albicans Candida spp. naturally more resistant. This study aimed to propose a novel therapeutic approach by [...] Read more.
Vulvovaginal candidiasis (VVC) remains a prevalent fungal disease, characterized by challenges, such as increased fungal resistance, side effects of current treatments, and the rising prevalence of non-albicans Candida spp. naturally more resistant. This study aimed to propose a novel therapeutic approach by investigating the antifungal properties and toxicity of 2-hydroxychalcone (2-HC) and 3′-hydroxychalcone (3′-HC), both alone and in combination with fluconazole (FCZ) and clotrimazole (CTZ). A lipid carrier (LC) was also developed to deliver these molecules. The study evaluated in vitro anti-Candida activity against five Candida species and assessed cytotoxicity in the C33-A cell line. The safety and therapeutic efficacy of in vivo were tested using an alternative animal model, Galleria mellonella. The results showed antifungal activity of 2-HC and 3′-HC, ranging from 7.8 to 31.2 as fungistatic and 15.6 to 125.0 mg/L as fungicide effect, with cell viability above 80% from a concentration of 9.3 mg/L (2-HC). Synergistic and partially synergistic interactions of these chalcones with FCZ and CTZ demonstrated significant improvement in antifungal activity, with MIC values ranging from 0.06 to 62.5 mg/L. Some combinations reduced cytotoxicity, achieving 100% cell viability in many interactions. Additionally, two LCs with suitable properties for intravaginal application were developed. These formulations demonstrated promising therapeutic efficacy and low toxicity in Galleria mellonella assays. These results suggest the potential of this approach in developing new therapies for VVC. Full article
Show Figures

Figure 1

Figure 1
<p>Structural formulas of 2-hydroxychalcone (2-HC) and 3′-hydroxychalcone (3′-HC).</p>
Full article ">Figure 2
<p>Percentage of cell viability after exposure to different concentrations of 2-hydroxychalcone and 3′-hydroxychalcone in cell line: C33A. Positive control: DMSO 40% (cell death). Negative control: 100% living cells. The results are expressed as mean ± standard deviation one-way ANOVA followed by Bonferroni (<span class="html-italic">p</span> &lt; 0.001). *** Statistically significant difference, the assays were performed from three replicates of three independent experiments.</p>
Full article ">Figure 3
<p>Visual classification of the systems: (<b>A</b>)—low viscosity and translucent system; (<b>B</b>)—low viscosity and opaque system; (<b>C</b>)—medium viscosity and translucent system; (<b>D</b>)—medium viscosity and opaque system; (<b>E</b>)—high viscosity and translucent system; (<b>F</b>)—high viscosity and opaque system; (<b>G</b>)—phase separation.</p>
Full article ">Figure 4
<p>Pseudo-ternary phase diagram of hydrogenated castor oil and soy phosphatidylcholine (75:25) (surfactant/co-surfactant), capric/caprylic triglyceride (oily phase), and water. LC (1) and LC (2) were the selected formulations for mechanical characterization and encapsulation efficiency. LVTS: low viscosity and translucent system, LVOS: low viscosity and opaque system, MVTS: medium viscosity and translucent system, MVOS: medium viscosity and opaque system, HVTS: high viscosity and translucent system, HVOS: high viscosity and opaque system, PS: phase separation.</p>
Full article ">Figure 5
<p>Evaluation of the mechanical properties of empty lipid carriers (LC1 and LC2), lipid carriers containing 1.0% of 2-HC, and clotrimazole 1% (commercial vaginal cream). (<b>A</b>): hardness (N); (<b>B</b>): compressibility (N.sec); (<b>C</b>): adhesiveness (N.sec); (<b>D</b>): cohesion; (<b>E</b>): mucoadhesive strength (N.s). Mean ± SD. One-way ANOVA followed by Bonferroni’s post hoc comparison tests were used in all statistical analyses (<span class="html-italic">p</span> &lt; 0.05). ns; Not significant. LC (1): Lipid carrier composed of 30% surfactant/co-surfactant (75:25, <span class="html-italic">w</span>/<span class="html-italic">w</span>), 35% oil phase, and 35% water. LC (2): lipid carrier composed of 30% surfactant/co-surfactant (75:25, <span class="html-italic">w</span>/<span class="html-italic">w</span>), 22% oil phase, and 48% water. 2-HC: 2-hydroxychalcone. Tests performed on independent quadruplicates. *** <span class="html-italic">p</span> &lt; 0.001. ** 0.001 &lt; <span class="html-italic">p</span> &lt; 0.01. * 0.01 &lt; <span class="html-italic">p</span> &lt; 0.05.</p>
Full article ">Figure 6
<p>Kaplan–Meier survival curves and health index after application of different concentrations (mg/L) of the combination of 3′-HC + 2′-HC (<b>A</b>,<b>B</b>), 2-HC + FCZ (<b>C</b>,<b>D</b>), and the MIC values obtained and LCs solutions (<b>E</b>,<b>F</b>) in the <span class="html-italic">G. mellonella.</span> Statistical analyses to survival curves: log-rank (Mantel–Cox) Test (<span class="html-italic">p</span> &lt; 0.001). * Survival curve statistically different compared to the control group (infection). Health index: results expressed as mean ± standard deviation. One-way ANOVA followed by Bonferroni. *** <span class="html-italic">p</span> &lt; 0.001. ** 0.001 &lt; <span class="html-italic">p</span> &lt; 0.01. Experiments were performed twice.</p>
Full article ">Figure 7
<p>Kaplan–Meier survival curves and health index after infection (5 × 10<sup>6</sup> cells/larva—<span class="html-italic">C. albicans</span> ATCC 90028) and treatment with different concentrations (mg/L) of the combination of 3′-HC + 2-HC (<b>A</b>,<b>B</b>), 2-HC + FCZ (<b>C</b>,<b>D</b>) and of the substances at the respective MIC values obtained and of the LCs (<b>E</b>,<b>F</b>). Statistical analyses to survival curves: log-rank (Mantel–Cox) test (<span class="html-italic">p</span> &lt; 0.001). Health index: results are expressed as mean ± standard deviation. One-way ANOVA followed by Bonferroni. ** 0.001 &lt; <span class="html-italic">p</span> &lt; 0.01. * 0.01 &lt; <span class="html-italic">p</span> &lt; 0.05. Experiments were performed twice. (<b>F</b>): * Health index = 0.</p>
Full article ">
Back to TopTop