Nothing Special   »   [go: up one dir, main page]

You seem to have javascript disabled. Please note that many of the page functionalities won't work as expected without javascript enabled.
 
 
Sign in to use this feature.

Years

Between: -

Subjects

remove_circle_outline
remove_circle_outline
remove_circle_outline
remove_circle_outline
remove_circle_outline
remove_circle_outline
remove_circle_outline
remove_circle_outline
remove_circle_outline

Journals

remove_circle_outline
remove_circle_outline
remove_circle_outline
remove_circle_outline
remove_circle_outline
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

Search Results (1,630)

Search Parameters:
Keywords = ciprofloxacin

Order results
Result details
Results per page
Select all
Export citation of selected articles as:
17 pages, 2168 KiB  
Article
Unveiling the Anticancer Potential of a New Ciprofloxacin-Chalcone Hybrid as an Inhibitor of Topoisomerases I & II and Apoptotic Inducer
by Doaa Mohamed Elroby Ali, Hossameldin A. Aziz, Stefan Bräse, Areej Al Bahir, Abdullah Alkhammash, Gamal El-Din A. Abuo-Rahma, Ali M. Elshamsy, Hamada Hashem and Walid M. Abdelmagid
Molecules 2024, 29(22), 5382; https://doi.org/10.3390/molecules29225382 - 15 Nov 2024
Viewed by 164
Abstract
The current study has yielded promising results in the evaluation of a new ciprofloxacin-chalcone hybrid (CP derivative) for its anticancer activity as potential Topoisomerases (Topo) I and II inhibitors. The in vitro results showed that the CP derivative significantly suppressed the growth of [...] Read more.
The current study has yielded promising results in the evaluation of a new ciprofloxacin-chalcone hybrid (CP derivative) for its anticancer activity as potential Topoisomerases (Topo) I and II inhibitors. The in vitro results showed that the CP derivative significantly suppressed the growth of HCT-116 and LOX IMVI cells, with IC50 values of 5.0 μM and 1.3 μM, respectively, outperforming Staurosporine, which had IC50 values of 8.4 μM and 1.6 μM, respectively. Flow cytometry analysis revealed that the new CP derivative triggered apoptosis and cell cycle arrest at the G2/M phase, associated with the up-regulation of pro-apoptotic genes (Bax and Caspase 9) and downregulation of the anti-apoptotic gene (Bcl-2). Further investigations showed that the CP derivative inhibited Topo I and II enzymes, as expected molecular targets; docking studies further supported its dual inhibitory action on Topo I and II. These findings suggest that the ciprofloxacin-chalcone hybrid could be a promising lead compound for developing new anticancer therapy. Full article
(This article belongs to the Topic Enzymes and Enzyme Inhibitors in Drug Research)
Show Figures

Graphical abstract

Graphical abstract
Full article ">Figure 1
<p>Structure of various N-4 piperazinyl-modified CP compounds with anticancer activity, including target compound <b>V</b> (CP derivative 2).</p>
Full article ">Figure 2
<p>Growth Inhibition Percentages of CP Derivative against Cancer Cell Lines (K-562, SR, HCT-116, LOX IMVI, MCF7, and BT-549) at a dose of 10 μM single screening.</p>
Full article ">Figure 3
<p>Comparative IC<sub>50</sub> (µM) Values of CP Derivative and Staurosporine (Positive Control) against Cancer Cell Lines (HCT-116 and LOX IMVI) and Normal Cell Line (WI-38). Bars display the mean ± SD. Statistical analysis via Two-way ANOVA and the Tukey–Kramer test afterward reveals significant differences (**** <span class="html-italic">p</span> &lt; 0.0001) relative to Staurosporine.</p>
Full article ">Figure 4
<p>The apoptosis and necrosis assay of colon HCT-116 induced by DMSO (control, (<b>A</b>)) and CP derivative (IC<sub>50</sub>, 5 µM, (<b>B</b>)).</p>
Full article ">Figure 5
<p>The percentages of early apoptosis, late apoptosis, total apoptosis, and necrosis induced by the IC<sub>50</sub> concentration of the CP derivative in comparison to the untreated control on HCT-116 colon cancer cells. Bars represent the mean ± SD. Statistical significance was determined using a two-way ANOVA test and the Tukey–Kramer test afterward, with **** <span class="html-italic">p</span> &lt; 0.0001 indicating a significant difference compared to the control group.</p>
Full article ">Figure 6
<p>Flow cytometric analysis illustrating the distribution of cell cycle phases in HCT-116 cells, comparing untreated controls and cells treated with the IC<sub>50</sub> concentration of the CP derivative for 24 h. (<b>A</b>) Dot plots representing the cell cycle phases for untreated cells after PI staining and (<b>B</b>) treated cells under the same conditions. (<b>C</b>) Quantitative comparison of cell proportions in each phase (G0/G1, S, G2/M, and pre-G1) between untreated and treated groups. Data are shown as mean ± SD. A two-way ANOVA and the Tukey–Kramer test afterward were used for statistical analysis, with ** <span class="html-italic">p</span> &lt; 0.01 and **** <span class="html-italic">p</span> &lt; 0.0001 indicating significant differences compared to untreated cells.</p>
Full article ">Figure 7
<p>Quantitative real-time PCR analysis of caspase-9 Bax and Bcl-2 expression levels in HCT-116 cells following 24 h treatment with the IC50 concentration of the drug, normalized to β-actin. Bars indicate mean ± SE. Statistical significance was assessed using an unpaired <span class="html-italic">t</span> test, with **** <span class="html-italic">p</span> &lt; 0.0001 compared to untreated cells (control).</p>
Full article ">Figure 8
<p>Two-dimensional interactions in Topo I active site (PDB: 1K4T); (<b>A</b>) Topotecan binding interactions; (<b>B</b>) CP derivative binding interactions.</p>
Full article ">Figure 9
<p>Two-dimensional interactions in Topo IIβ active site (PDB: 7YQ8); (<b>A</b>) Etoposide binding interactions; (<b>B</b>) CP derivative binding interactions.</p>
Full article ">Figure 10
<p>(<b>A</b>) Rader model for CP derivative; (<b>B</b>) The BOILED-Egg model of CP derivative.</p>
Full article ">Scheme 1
<p>Synthesis of the CP derivative 2. <b>Reagents and conditions:</b> (i) 60% NaOH, ethanol, 0–5 °C stirring overnight.</p>
Full article ">
12 pages, 2210 KiB  
Review
Fluoroquinolone Resistance in Escherichia coli Causing Community-Acquired Urinary Tract Infections: A Systematic Review
by Ana P. Ruiz-Lievano, Fernando Cervantes-Flores, Alessandro Nava-Torres, Paulo J. Carbajal-Morales, Luisa F. Villaseñor-Garcia and Maria G. Zavala-Cerna
Microorganisms 2024, 12(11), 2320; https://doi.org/10.3390/microorganisms12112320 - 15 Nov 2024
Viewed by 219
Abstract
Community-acquired urinary tract infections account for 15% of all outpatient use of antibiotics, and women are primarily affected; the major causative microorganism is uropathogenic Escherichia coli (E. coli). Treatment is indicated for cystitis and pyelonephritis and includes B-lactams (amoxicillin-clavulanic acid or [...] Read more.
Community-acquired urinary tract infections account for 15% of all outpatient use of antibiotics, and women are primarily affected; the major causative microorganism is uropathogenic Escherichia coli (E. coli). Treatment is indicated for cystitis and pyelonephritis and includes B-lactams (amoxicillin-clavulanic acid or third-generation cephalosporins), fluoroquinolones (ciprofloxacin or levofloxacin), nitrofurantoin, fosfomycin, and trimethoprim–sulfamethoxazole. Resistance to antibiotic treatment is of concern; several mechanisms have been associated with the acquisition of genes that confer antimicrobial resistance to fluoroquinolones, which are often associated with other patterns of resistance, especially in extended-spectrum beta-lactamase (ESBL) producers. Several studies have addressed the prevalence of uropathogens producing ESBLs, but only a few have focused on fluoroquinolone resistance, and, to our knowledge, none have addressed the prevalence of phylotypes or genes responsible for antimicrobial resistance to fluoroquinolones. The focus of the present review was to analyze recently published papers that described the E. coli phylotype causing community-acquired UTIs in association with fluoroquinolone resistance. Full article
(This article belongs to the Special Issue Clinical Microbial Infection and Antimicrobial Resistance)
Show Figures

Figure 1

Figure 1
<p>Mechanism of action of fluoroquinolones. <span class="html-italic">E. coli</span> DNA gyrase (which acts by unwinding the DNA double helix) is composed of two A subunits and two B subunits encoded by <span class="html-italic">GyrA</span> and <span class="html-italic">GyrB</span>, respectively. The A subunits carrying the “codon trimming” functions of gyrase are the site of action of fluoroquinolones. The drug inhibits (red circle) gyrase-mediated DNA supercoiling at concentrations that are clearly related to those required to inhibit bacterial proliferation. Created in <a href="http://BioRender.com" target="_blank">BioRender.com</a>.</p>
Full article ">Figure 2
<p>Mechanisms of fluoroquinolone resistance in <span class="html-italic">E. coli</span>. (<b>1</b>) Chromosomal mutations induce changes in drug targets, such as genes that encode <span class="html-italic">GyrA</span>, and <span class="html-italic">ParC</span>; (<b>2</b>) mutations related to reducing the drug concentration in the bacterial cytoplasm through the overexpression of efflux pumps and downregulation of porins; and (<b>3</b>) the expression of genes that encode plasmid-mediated quinolone resistance proteins (PMQR), such as <span class="html-italic">qnr</span> (protects DNA gyrase and topoisomerase IV), <span class="html-italic">aac(6’)-ib-cr</span> (acetylates the fluoroquinolone so that it is inactivated), <span class="html-italic">quepA</span>, and <span class="html-italic">oqxAB</span> (overexpressing multidrug efflux pumps or decreasing the permeability of outer membrane proteins). Created in <a href="http://BioRender.com" target="_blank">BioRender.com</a>.</p>
Full article ">Figure 3
<p>PRISMA flow diagram for systematic reviews.</p>
Full article ">
23 pages, 4154 KiB  
Article
Thermoanalytical and Kinetic Studies for the Thermal Stability of Emerging Pharmaceutical Pollutants Under Different Heating Rates
by Christian Ebere Enyoh, Tochukwu Oluwatosin Maduka, Miho Suzuki, Senlin Lu and Qingyue Wang
J. Xenobiot. 2024, 14(4), 1784-1806; https://doi.org/10.3390/jox14040095 - 14 Nov 2024
Viewed by 334
Abstract
Emerging pharmaceutical pollutants like ciprofloxacin (CIP) and ibuprofen (IBU) are frequently detected in aquatic environments, posing risks to ecosystems and human health. Since pollutants rarely exist alone in the environment, understanding the thermal stability and degradation kinetics of these compounds, especially in mixtures, [...] Read more.
Emerging pharmaceutical pollutants like ciprofloxacin (CIP) and ibuprofen (IBU) are frequently detected in aquatic environments, posing risks to ecosystems and human health. Since pollutants rarely exist alone in the environment, understanding the thermal stability and degradation kinetics of these compounds, especially in mixtures, is crucial for developing effective removal strategies. This study therefore investigates the thermal stability and degradation kinetics of CIP and IBU, under different heating rates. Thermogravimetric analysis (TGA) and differential thermal analysis (DTA) were employed to examine the thermal behavior of these compounds individually and in mixture (CIP + IBU) at heating rates of 10, 20, and 30 °C/min. The kinetics of thermal degradation were analyzed using both model-fitting (Coats–Redfern (CR)) and model-free (Kissinger–Akahira–Sunose (KAS), Flynn–Wall–Ozawa (FWO), and Friedman (FR)) methods. The results showed distinct degradation patterns, with CIP decomposing between 280 and 550 °C and IBU between 152 and 350 °C, while the mixture exhibited multistep decomposition in the 157–500 °C range. The CR model indicated first-order kinetics as a better fit for the degradation (except for IBU). Furthermore, CIP exhibits higher thermal stability and activation energy compared to IBU, with the KAS model yielding activation energies of 58.09 kJ/mol for CIP, 11.37 kJ/mol for IBU, and 41.09 kJ/mol for CIP + IBU mixture. The CIP + IBU mixture generally showed intermediate thermal properties, suggesting synergistic and antagonistic interactions between the compounds. Thermodynamic parameters (ΔH°, ΔG°, ΔS°) were calculated, revealing non-spontaneous, endothermic processes for all samples (except in the FWO method) with a decrease in molecular disorder and positive ΔG° values across all models and heating rates. The study found that higher heating rates led to less thermodynamically favorable conditions for degradation. These findings provide important information concerning the thermal behavior of these pharmaceutical pollutants, which can inform strategies for their removal from the environment and the development of more effective waste-treatment processes. Full article
Show Figures

Graphical abstract

Graphical abstract
Full article ">Figure 1
<p>Structure of (<b>A</b>) ciprofloxacin and (<b>B</b>) ibuprofen. Structure drawn with Chemdraw 8.0.</p>
Full article ">Figure 2
<p>Thermogravimetric analysis of pharmaceutical samples. The study design involved analyzing (<b>A</b>) only ciprofloxacin (CIP), (<b>B</b>) only ibuprofen (IBU), and (<b>C</b>) physical mixture of CIP and IBU at a ratio of 1:1. The left figure is the thermogravimetric analyzer used in this study.</p>
Full article ">Figure 3
<p>Thermogram of TGA and DTA curves for CIP (<b>A</b>,<b>B</b>), IBU (<b>C</b>,<b>D</b>) and CIP + IBU mixture (<b>E</b>,<b>F</b>).</p>
Full article ">Figure 4
<p>Proposed theoretical pathways for the thermal breakdown mechanism for CIP (<b>A</b>) and IBU (<b>B</b>) [<a href="#B29-jox-14-00095" class="html-bibr">29</a>,<a href="#B30-jox-14-00095" class="html-bibr">30</a>,<a href="#B31-jox-14-00095" class="html-bibr">31</a>].</p>
Full article ">Figure 5
<p>(<b>A</b>) Degradation conversion factors and Coats–Redfern kinetic modelling for (<b>B</b>) zero order (ZO) and (<b>C</b>) first order (FO), for the different pharmaceutical samples.</p>
Full article ">Figure 5 Cont.
<p>(<b>A</b>) Degradation conversion factors and Coats–Redfern kinetic modelling for (<b>B</b>) zero order (ZO) and (<b>C</b>) first order (FO), for the different pharmaceutical samples.</p>
Full article ">Figure 6
<p>Various isoconversional kinetic model plots at degree of conversion (<span class="html-italic">α</span> = 0.8) for (<b>A</b>) KAS, (<b>B</b>) FWO, and (<b>C</b>) FR.</p>
Full article ">Figure 6 Cont.
<p>Various isoconversional kinetic model plots at degree of conversion (<span class="html-italic">α</span> = 0.8) for (<b>A</b>) KAS, (<b>B</b>) FWO, and (<b>C</b>) FR.</p>
Full article ">Figure 7
<p>Thermodynamic characteristics showing (<b>A</b>) changes in enthalpy (∆<span class="html-italic">H</span>°, kJ/mol), (<b>B</b>) Gibbs free energy (∆<span class="html-italic">G</span>°, kJ/mol), and (<b>C</b>) entropy (Δ<span class="html-italic">S</span>°, kJ/mol × K) of thermal treatment at different heating rates for different pharmaceutical sample degradations for isoconversional or model-free kinetic models at degree of conversion (<span class="html-italic">α</span> = 0.8). KAS = Kissinger–Akahira–Sunose; FWO = Flynn–Wall–Ozawa; FR (FO) = Friedman First Order; and FR (SO) = Friedman Second Order.</p>
Full article ">Figure 7 Cont.
<p>Thermodynamic characteristics showing (<b>A</b>) changes in enthalpy (∆<span class="html-italic">H</span>°, kJ/mol), (<b>B</b>) Gibbs free energy (∆<span class="html-italic">G</span>°, kJ/mol), and (<b>C</b>) entropy (Δ<span class="html-italic">S</span>°, kJ/mol × K) of thermal treatment at different heating rates for different pharmaceutical sample degradations for isoconversional or model-free kinetic models at degree of conversion (<span class="html-italic">α</span> = 0.8). KAS = Kissinger–Akahira–Sunose; FWO = Flynn–Wall–Ozawa; FR (FO) = Friedman First Order; and FR (SO) = Friedman Second Order.</p>
Full article ">
10 pages, 628 KiB  
Article
Regional Variation in Urinary Escherichia coli Resistance Among Outpatients in Washington State, 2013–2019
by Hannah T. Fenelon, Stephen E. Hawes, Hema Kapoor, Ann E. Salm, Jeff Radcliff and Peter M. Rabinowitz
Microorganisms 2024, 12(11), 2313; https://doi.org/10.3390/microorganisms12112313 - 14 Nov 2024
Viewed by 253
Abstract
Escherichia coli (E. coli) is a predominant pathogen of urinary tract infections (UTIs) in the United States. We analyzed resistance patterns by geographic location in Washington State to assess the need for regional antibiograms. The study included urinary E. coli antibiotic [...] Read more.
Escherichia coli (E. coli) is a predominant pathogen of urinary tract infections (UTIs) in the United States. We analyzed resistance patterns by geographic location in Washington State to assess the need for regional antibiograms. The study included urinary E. coli antibiotic susceptibility tests performed by Quest Diagnostics on Washington outpatient isolates from 2013 to 2019. We conducted logistic regressions with robust standard errors for five antibiotics (ceftriaxone, ciprofloxacin, gentamicin, trimethoprim-sulfamethoxazole), with isolates classified as “susceptible” or “resistant” for each antibiotic tested. Analyses were adjusted for sex, year of isolate collection, and age group (0–18, 19–50, >50). The state’s nine Public Health Emergency Preparedness Regions (PHEPRs) were used as the geographic level for the analysis. The analysis included 40,217 isolates (93% from females, mean age 47 years). Compared to the Central PHEPR (containing Seattle), most other regions had significantly lower adjusted prevalence ratios (aPORs) of antimicrobial resistance (AMR), with aPORs as low as 0.20 (95% CI: 0.06–0.63) for ceftriaxone in the North Central region. Additionally, no regions had significantly higher aPOR of resistance for any antibiotic. Differences in resistance between the Central and other regions varied by antibiotic with the largest difference for ceftriaxone and smallest for ampicillin. The finding of regional variation of E. coli AMR calls for more specific community antibiograms to enable a precise approach to antibiotic prescribing and stewardship. Full article
(This article belongs to the Special Issue Antimicrobial Resistance in Enterobacteriaceae and Enterococci)
Show Figures

Figure 1

Figure 1
<p>Map of Washington State divided into PHEPR and colored by adjusted prevalence odds ratio in relation to the reference region (Central) for the five antibiotics.</p>
Full article ">
19 pages, 4910 KiB  
Article
Investigation of Antibiotic Resistance of E. coli Associated with Farm Animal Feces with Participation of Citizen Scientists
by Anna M. Timofeeva, Maria R. Galyamova, Dmitriy M. Krivosheev, Sergey Yu. Karabanov and Sergey E. Sedykh
Microorganisms 2024, 12(11), 2308; https://doi.org/10.3390/microorganisms12112308 - 13 Nov 2024
Viewed by 419
Abstract
This paper presents the findings of a large-scale study on antibiotic resistance in bacteria found in farm animal feces across Russia. The study included 6578 samples of farm animal manure from 13 regions in Russia, with the help of citizen scientists. Molecular and [...] Read more.
This paper presents the findings of a large-scale study on antibiotic resistance in bacteria found in farm animal feces across Russia. The study included 6578 samples of farm animal manure from 13 regions in Russia, with the help of citizen scientists. Molecular and microbiological methods were used to analyze 1111 samples of E. coli. The microbiological analysis focused on culturing the microorganisms present in the fecal samples on selective media for E. coli and evaluating the sensitivity of the bacteria to different antibiotics, including ampicillin, tetracycline, chloramphenicol, cefotaxime, and ciprofloxacin. The molecular analysis involved isolating the genomic DNA of the bacteria and conducting PCR assays to detect the vanA, vanB, and mcr-1 antibiotic resistance genes. The results demonstrated significant differences in antibiotic sensitivity of the samples that are morphologically identical to E. coli from different regions. For example, 98.0% and 82.5% of E. coli and other fecal bacterial isolates from the Omsk and Vologda regions lacked antibiotic resistance genes, while 97.7% of samples from the Voronezh region possessed three resistance genes simultaneously. The phenotypic antibiotic sensitivity test also revealed regional differences. For instance, 98.1% of fecal bacterial samples from cattle in the Udmurt Republic were sensitive to all five antibiotics tested, whereas 92.8% of samples from the Voronezh region showed resistance to all five antibiotics. The high level of antibiotic resistance observed may be attributed to their use in farming practices. The distinctive feature of our research is that comprehensive geographical coverage was achieved by using a citizen science platform. Citizen scientists, specifically students from colleges and universities, were responsible for the collection and initial analysis of samples. The project attracted 3096 student participants, enabling the collection and analysis of a significant number of samples from various locations in Russia. Full article
(This article belongs to the Special Issue Veterinary Microbiology and Diagnostics)
Show Figures

Figure 1

Figure 1
<p>The general scheme of the study achieved with the use of citizen science tool. A project manager provides the general management of the project and training of mentors. Mentors train students and supervise the collection and analysis of samples. Sampling was conducted at the farm. Students and their mentors analyze the samples using one of the two proposed methods, and professional scientists compile the results in a table for further analysis.</p>
Full article ">Figure 2
<p>Regions of Russia covered by the study. The red circle indicates the region participating in the study, with the circle diameter proportional to the number of samples. The number of collected samples is indicated in brackets. The Voronezh region is indicated by a green circle.</p>
Full article ">Figure 3
<p>Screening of farm animal fecal microorganisms for antibiotic resistance genes: <span class="html-italic">vanA</span> (<b>A</b>), <span class="html-italic">vanB</span> (<b>B</b>), and <span class="html-italic">mcr-1</span> (<b>C</b>) by PCR. Presented here is an example of the analysis of 10 bacterial strains (1–10) derived from cattle fecal samples. The letter M is used to represent markers, while + and − are used to indicate positive and negative controls, respectively.</p>
Full article ">Figure 4
<p>Analysis of antibiotic resistance of microorganisms associated with farm animals by PCR. A represents the data for cattle fecal microorganisms from five regions. n is the number of samples analyzed.</p>
Full article ">Figure 5
<p>Example of microbiological analysis of a bacterial strain for resistance to five antibiotics: ampicillin (1), tetracycline (2), chloramphenicol (3), cefotaxime (4), and ciprofloxacin (5).The results of the analysis of bacterial samples from cattle feces are shown.</p>
Full article ">Figure 6
<p>Analysis of antibiotic resistance of microorganisms associated with farm animals by microbiological methods. (<b>A</b>) The data for cattle fecal microorganisms from six regions of Russia, and (<b>B</b>) the data for different farm animals from the Voronezh region. n—the number of samples analyzed.</p>
Full article ">
13 pages, 4585 KiB  
Article
Analysis of the Prevalence of Bacterial Pathogens and Antimicrobial Resistance Patterns of Edwardsiella piscicida in Largemouth Bass (Micropterus salmoides) from Guangdong, China
by Weimin Huang, Changyi Lin, Caiyi Wen, Biao Jiang and Youlu Su
Pathogens 2024, 13(11), 987; https://doi.org/10.3390/pathogens13110987 - 12 Nov 2024
Viewed by 392
Abstract
To gain insights into the prevalence and antimicrobial resistance patterns of major bacterial pathogens affecting largemouth bass (Micropterus salmoides) in the Pearl River Delta (PRD) region, Guangdong, China, a study was conducted from August 2021 to July 2022. During this period, [...] Read more.
To gain insights into the prevalence and antimicrobial resistance patterns of major bacterial pathogens affecting largemouth bass (Micropterus salmoides) in the Pearl River Delta (PRD) region, Guangdong, China, a study was conducted from August 2021 to July 2022. During this period, bacteria were isolated and identified from the internal organs of diseased largemouth bass within the PRD region. The antimicrobial resistance patterns of 11 antibiotics approved for use in aquaculture in China were analyzed in 80 strains of Edwardsiella piscicida using the microbroth dilution method. The results showed that 151 bacterial isolates were obtained from 532 samples, with E. piscicida (17.29%, 92/532), Aeromonas veronii (4.70%, 25/532), and Nocardia seriolae (2.26%, 12/532) being the main pathogens. Notably, E. piscicida accounted for the highest proportion of all isolated bacteria, reaching 60.92% (92/151), and mainly occurred from November to April, accounting for 68.48% (63/92) of the cases. The symptoms in largemouth bass infected with E. piscicida included ascites, enteritis, and hemorrhaging of tissues and organs. The drug sensitivity results showed that the resistance rates of all E. piscicida strains to ciprofloxacin, all sulfonamides, thiamphenicol, florfenicol, enrofloxacin, doxycycline, flumequine, and neomycin were 96.25%, 60–63%, 56.25%, 43.75%, 40%, 32.5%, 16.25%, and 1.25%, respectively. In addition, 76.25% (61/80) of these strains demonstrated resistance to more than two types of antibiotics. Cluster analysis revealed 23 antibiotic types (A–W) among the 80 isolates, which were clustered into two groups. Therefore, tailored antibiotic treatment based on regional antimicrobial resistance patterns is essential for effective disease management. The findings indicate that in the event of an Edwardsiella infection in largemouth bass, neomycin, doxycycline, and flumequine are viable treatment options. Alternatively, one may choose drugs that are effective as determined by clinical drug sensitivity testing. Full article
(This article belongs to the Special Issue Foodborne Pathogens: The Antimicrobial Resistance from Farm to Fork)
Show Figures

Figure 1

Figure 1
<p>The percentage of pathogenic bacteria isolated from the examined diseased largemouth bass.</p>
Full article ">Figure 2
<p>The monthly analysis results of <span class="html-italic">Edwardsiella piscicida</span>.</p>
Full article ">Figure 3
<p>Clinical signs in largemouth bass infected with <span class="html-italic">Edwardsiella piscicida</span>. (<b>A</b>) Anal red swelling (arrow); (<b>B</b>) abdominal swelling (S), wounds were accompanied by the shedding of fish scales (W) and anal red swelling (arrow); (<b>C</b>) white nodules in the liver (arrow); (<b>D</b>) effusion in the abdominal cavity and swim bladder cavity (arrows) and liver whitening and ischemia (I).</p>
Full article ">Figure 4
<p>The resistance rates of <span class="html-italic">Edwardsiella piscicida</span> to 11 antimicrobial agents. Abbreviations: THI, thiamphenicol; FLO, florfenicol; FLU, flumequine; ENR, enrofloxacin; CIP, ciprofloxacin; DOX, doxycycline; NEO, neomycin; SSM, sodium sulfamonomethoxine; SME, sulfamethoxazole; SDI, sulfadiazine; SMT, sulfamethazine.</p>
Full article ">Figure 5
<p>Antibiogram and cluster analysis of the 80 <span class="html-italic">Edwardsiella piscicida</span> strains. Abbreviations: THI, thiamphenicol; FLO, florfenicol; FLU, flumequine; ENR, enrofloxacin; CIP, ciprofloxacin; DOX, doxycycline; NEO, neomycin; SSM, sodium sulfamonomethoxine; SME, sulfamethoxazole; SDI, sulfadiazine; SMT, sulfamethazine.</p>
Full article ">
14 pages, 3380 KiB  
Article
Optical Properties and Antimicrobial Activity of Si/PVP Hybrid Material Combined with Antibiotics
by Lilia Yordanova, Yoanna Kostova, Elitsa Pavlova, Albena Bachvarova-Nedelcheva, Iliana Ivanova and Elena Nenova
Molecules 2024, 29(22), 5322; https://doi.org/10.3390/molecules29225322 - 12 Nov 2024
Viewed by 343
Abstract
Silica–poly (vinylpyrrolidone) hybrid material was prepared using the sol–gel method. Tetramethyl ortosilane (TMOS) was used as a silica precursor. XRD analysis established that the as-prepared material is amorphous. The morphological structure of the final product was determined by the incorporated PVP. The UV–Vis [...] Read more.
Silica–poly (vinylpyrrolidone) hybrid material was prepared using the sol–gel method. Tetramethyl ortosilane (TMOS) was used as a silica precursor. XRD analysis established that the as-prepared material is amorphous. The morphological structure of the final product was determined by the incorporated PVP. The UV–Vis analysis showed that the obtained hybrid exhibited absorption in the ultraviolet range. The antimicrobial activity of the SiO2/15PVP hybrid material was tested on Staphylococcus epidermidis ATCC 14990, Salmonella typhimurium ATCC BAA-2162, Candida albicans, and Saccharomyces cerevisiae in combination with the following antibiotics: Vancomycin for Gram-positive bacteria, Ciprofloxacin for Gram-negative bacteria, and Nystatin for yeast. The results confirmed a concentration-dependent synergistic effect of the antibiotic in combination with the TM15/PVP hybrid particles, especially at their highest concentration of 100 mg/mL on Gram-positive bacteria and for the Gram-negative Salmonella. On Candida albicans ATCC 18804 and Saccharomyces cerevisiae CCY 21-6-3, the effect was synergistic again, and a fungicidal effect was observed at 6.25 and 1.50 mg/mL for the antibiotic concentration and concentrations of hybrid material at 100 mg/mL. The toxicity on Daphnia magna was also tested. The registered prooxidant activity of SiO2/15PVP shows possible applications at very low concentrations. The obtained results demonstrate the possibility of clinical implementations of the newly synthesized hybrid material. Full article
Show Figures

Figure 1

Figure 1
<p>XRD patterns of SiO<sub>2</sub>/15PVP, TMOS, and PVP.</p>
Full article ">Figure 2
<p>SEM micrographs of a sample with 15% PVP at different magnifications (<b>a</b>,<b>b</b>).</p>
Full article ">Figure 3
<p>UV–Vis spectra for the pure TMOS, PVP, and SiO<sub>2</sub>/15PVP hybrid. The inset shows the spectra in the range of 200–350 nm.</p>
Full article ">Figure 4
<p>AFM images of the investigated SiO<sub>2</sub>/15PVP hybrid: 2D (<b>a</b>) and 3D (<b>b</b>) surface topography and roughness profile (<b>c</b>) of the material.</p>
Full article ">Figure 5
<p>Effect of SiO<sub>2</sub>/15PVP in combination with Vancomycin against <span class="html-italic">Staphylococcus epidermidis</span> ATCC 14990 at 24 h of treatment.</p>
Full article ">Figure 6
<p>SiO<sub>2</sub>/15PVP in combination with Ciprofloxacin against <span class="html-italic">Salmonella typhimurium</span> ATCC BAA-2162 at 24 h of the experiment.</p>
Full article ">Figure 7
<p>Effect of SiO<sub>2</sub>/15PVP in combination with Nystatin against <span class="html-italic">Candida albicans</span> ATCC 18804 at 24 h of treatment.</p>
Full article ">Figure 8
<p>SiO<sub>2</sub>/15PVP in combination with Nystatin against <span class="html-italic">Saccharomyces cerevisiae</span> CCY 21-6-3 at 24 h of the experiment.</p>
Full article ">Figure 9
<p>Chemiluminescence induced by ·OH and ·OOH radicals at pH 7.4/37 °C and the effect of the materials: (<b>a</b>) maximum effects and (<b>b</b>) Fenton’s reaction over time (<span class="html-italic">p</span> ≤ 0.05).</p>
Full article ">Figure 10
<p>TM/15PVP without antibiotic influence on <span class="html-italic">Daphnia magna</span>.</p>
Full article ">
13 pages, 619 KiB  
Article
Antimicrobial Resistance of Escherichia coli for Uncomplicated Cystitis: Korean Antimicrobial Resistance Monitoring System
by Seong Hyeon Yu, Seung Il Jung, Seung-Ju Lee, Mi-Mi Oh, Jin Bong Choi, Chang Il Choi, Yeon Joo Kim, Dong Jin Park, Sangrak Bae, Seung Ki Min and KAUTII Investigators
Antibiotics 2024, 13(11), 1075; https://doi.org/10.3390/antibiotics13111075 - 12 Nov 2024
Viewed by 436
Abstract
Objectives: Uncomplicated cystitis is a leading form of bacterial UTI; the most common causative bacterium worldwide is Escherichia coli. This internet-based, prospective, multicenter, and national observational study aimed to report the antimicrobial resistance of E. coli in patients with uncomplicated cystitis through [...] Read more.
Objectives: Uncomplicated cystitis is a leading form of bacterial UTI; the most common causative bacterium worldwide is Escherichia coli. This internet-based, prospective, multicenter, and national observational study aimed to report the antimicrobial resistance of E. coli in patients with uncomplicated cystitis through the use of the Korean Antimicrobial Resistance Monitoring System (KARMS) in 2023. Results: Data for a total of 654 patients were retrieved from the KARMS database. The mean (standard deviation) patient age was 55.9 (18.3) years. The numbers of postmenopausal women and patients with recurrent cystitis were 381 (59.4%) and 78 (11.9%), respectively. Regarding antimicrobial susceptibility, 96.8% were susceptible to fosfomycin, 98.9% to nitrofurantoin, 50.9% to ciprofloxacin, and 82.4% to cefotaxime. Extended-spectrum beta-lactamase positivity was 14.4% (89/616), and was significantly higher in tertiary hospitals (24.6%, p < 0.001) and recurrent cystitis (27.6%, p < 0.001). Fluoroquinolone resistance was significantly higher in tertiary hospitals (57.8%, p < 0.001), postmenopausal women (54.2%, p < 0.001), and recurrent cystitis (70.3%, p < 0.001). In addition, postmenopausal status (95% confidence interval [CI]: 1.44–3.17, odds ratio [OR] 2.13, p < 0.001), recurrent cystitis (95% CI: 1.40–4.66, OR 2.56, p = 0.002) and tertiary hospitals (95% CI: 1.00–2.93, OR 1.71, p = 0.049) were associated with significantly increased fluoroquinolone resistance. Methods: Any female patient diagnosed with clinical uncomplicated cystitis and microbiologically proven E. coli infection in 2023 was eligible for this study. Patient data were obtained from the web-based KARMS database. The antimicrobial susceptibility of E. coli was analyzed according to clinical factors, including hospital region, hospital type, menopause status, and recurrence status. Conclusions: The antimicrobial resistance of E. coli in patients with uncomplicated cystitis in the Republic of Korea has reached a serious level, especially in fluoroquinolone resistance. Therefore, major efforts should be made to reduce antimicrobial resistance. Full article
Show Figures

Figure 1

Figure 1
<p>ESBL positivity, fluoroquinolone resistance, and third-generation cephalosporin resistance of <span class="html-italic">E. coli</span> according to clinical factors. (<b>a</b>) Hospital type, (<b>b</b>) menopausal status, and (<b>c</b>) recurrence of infection.</p>
Full article ">Figure 1 Cont.
<p>ESBL positivity, fluoroquinolone resistance, and third-generation cephalosporin resistance of <span class="html-italic">E. coli</span> according to clinical factors. (<b>a</b>) Hospital type, (<b>b</b>) menopausal status, and (<b>c</b>) recurrence of infection.</p>
Full article ">
14 pages, 2285 KiB  
Article
Comprehensive Study of Antibiotic Resistance in Enterococcus spp.: Comparison of Influents and Effluents of Wastewater Treatment Plants
by Ji-Hyun Park, Kyung-Seon Bae, Jihyun Kang, Eung-Roh Park and Jeong-Ki Yoon
Antibiotics 2024, 13(11), 1072; https://doi.org/10.3390/antibiotics13111072 - 11 Nov 2024
Viewed by 419
Abstract
Background/Objectives: The spread of antibiotic resistance, particularly through Enterococcus spp., in wastewater treatment plants (WWTPs) poses significant public health risks. Given that research on antibiotic-resistant enterococci and their antibiotic-resistance genes in aquatic environments is limited, we evaluated the role of Enterococcus spp. [...] Read more.
Background/Objectives: The spread of antibiotic resistance, particularly through Enterococcus spp., in wastewater treatment plants (WWTPs) poses significant public health risks. Given that research on antibiotic-resistant enterococci and their antibiotic-resistance genes in aquatic environments is limited, we evaluated the role of Enterococcus spp. in WWTPs by comparing the antibiotic resistance rates, gene prevalence, biofilm formation, and residual antibiotics in the influent and effluent using culture-based methods. Methods: In 2022, influent and effluent samples were collected from 11 WWTPs in South Korea. Overall, 804 Enterococcus strains were isolated, and their resistance to 16 antibiotics was assessed using the microdilution method. Results: High resistance to tetracycline, ciprofloxacin, kanamycin, and erythromycin was observed. However, no significant differences in the overall resistance rates and biofilm formation were observed between the influent and effluent. Rates of resistance to ampicillin, ciprofloxacin, and gentamicin, as well as the prevalence of the tetM and qnrS genes, increased in the effluent, whereas resistance rates to chloramphenicol, florfenicol, erythromycin, and tylosin tartrate, along with the prevalence of the optrA gene, decreased. E. faecium, E. hirae, and E. faecalis were the dominant species, with E. faecalis exhibiting the highest resistance. Conclusions: Our results suggest that WWTPs do not effectively reduce the rates of resistant Enterococcus spp., indicating the need for continuous monitoring and improvement of the treatment process to mitigate the environmental release of antibiotic-resistant bacteria. Full article
(This article belongs to the Special Issue The Spread of Antibiotic Resistance in Natural Environments)
Show Figures

Figure 1

Figure 1
<p>Identification of <span class="html-italic">Enterococcus</span> spp. isolated from the wastewater treatment plants (WWTPs). (<b>a</b>) Species composition in the influent. (<b>b</b>) Species composition in the effluent.</p>
Full article ">Figure 2
<p>Comparison of resistance rates to 16 antibiotics between <span class="html-italic">Enterococcus</span> isolates from influents and effluents of WWTPs. AMP, Ampicillin; CHL, Chloramphenicol; FFC, Florfenicol; CIP, Ciprofloxacin; DAP, Daptomycin; ERY, Erythromycin; TYL, Tylosin tartrate; GEN, Gentamicin; KAN, Kanamycin; STR, Streptomycin; LNZ, Linezolid; SYN, Quinupristin/dalfopristin; TET, Tetracycline; TGC, Tigecycline; VAN, Vancomycin; SAL, Salinomycin. Significant differences between groups are marked with an asterisk (*).</p>
Full article ">Figure 3
<p>Comparison of resistance rates to 16 antibiotics among the three major <span class="html-italic">Enterococcus</span> spp. (<span class="html-italic">E. faecium</span>, <span class="html-italic">E. hirae</span>, and <span class="html-italic">E. faecalis</span>) isolated from the WTTPs. Significant differences between groups are marked with an asterisk (*).</p>
Full article ">Figure 4
<p>Comparison of the prevalence of antibiotic-resistance genes (ARGs) identified in <span class="html-italic">Enterococcus</span> spp. isolated from the influents and effluents of the WTTPs. Significant differences between groups are marked with an asterisk (*).</p>
Full article ">Figure 5
<p>Comparison of the prevalence of ARGs among the three major <span class="html-italic">Enterococcus</span> spp. (<span class="html-italic">E. faecium, E. hirae,</span> and <span class="html-italic">E. faecalis</span>) isolated from the WTTPs.</p>
Full article ">Figure 6
<p>Comparison of residual antibiotic concentrations detected in influents and effluents from the WWTPs.</p>
Full article ">
18 pages, 4538 KiB  
Article
Antimicrobial Profiling of Piper betle L. and Piper nigrum L. Against Methicillin-Resistant Staphylococcus aureus (MRSA): Integrative Analysis of Bioactive Compounds Based on FT-IR, GC-MS, and Molecular Docking Studies
by Budiman Yasir, Suwahyuni Mus, Sitti Rahimah, Rein Mostatian Tandiongan, Kasandra Putri Klara, Nurul Afrida, Nur Rezky Khairun Nisaa, Risna Risna, Agum Wahyudha Jur, Gemini Alam and Abdul Rohman
Separations 2024, 11(11), 322; https://doi.org/10.3390/separations11110322 - 8 Nov 2024
Viewed by 556
Abstract
This study explored the antimicrobial potential of Piper betle L. (PBL) and Piper nigrum L. (PNL) extracts against MRSA. Plant parts including stem, leaf, and fruit were extracted using aquadest, methanol, and hexane, resulting in 18 distinct extracts. FT-IR combined with cluster analysis [...] Read more.
This study explored the antimicrobial potential of Piper betle L. (PBL) and Piper nigrum L. (PNL) extracts against MRSA. Plant parts including stem, leaf, and fruit were extracted using aquadest, methanol, and hexane, resulting in 18 distinct extracts. FT-IR combined with cluster analysis (CA) categorized the extracts, and anti-MRSA activity was assessed through the paper disk diffusion method. The most potent extracts were further analyzed using GC-MS to identify bioactive compounds. Additionally, molecular docking studies were conducted for MRSA protein targets (4DKI, 6H5O, and 4CJN). The hexane extract of PNL and the aqueous extract of PBL fruit showed the strongest inhibitory effects. GC-MS identified piperine (14.22%) and diisooctyl phthalate (14.67%) as major compounds, with piperolein B, piperanine, β-caryophyllene oxide, and α-caryophylladienol as minor compounds in the hexane extract of PNL, while hydroxychavicol (81.89%) and chavibetol (12.01%) were predominant in the aquadest extract of PBL. Molecular docking revealed that piperolein B and piperine had strong binding affinities to MRSA proteins 4DKI, 6H5O, and 4CJN, comparable to ciprofloxacin. In conclusion, this study confirms the potential of PBL and PNL as sources of novel anti-MRSA agents, supporting further research to develop new therapies. Full article
Show Figures

Figure 1

Figure 1
<p>FTIR spectra of <span class="html-italic">Piper</span> extracts (PNL, <span class="html-italic">Piper nigrum</span> L.; PBL, <span class="html-italic">Piper betle</span> L.) extracted using aquadest, methanol, and hexane from stem, leaf, and fruit parts.</p>
Full article ">Figure 2
<p>The dendrogram illustrating the classification of samples through cluster analysis (CA), with samples in clusters 1 to 6 represented by distinct colored lines for each respective cluster.</p>
Full article ">Figure 3
<p>Structure of the MRSA protein target (4DKI) (<b>C</b>) along with the native ligand (ceftobiprole) (<b>B</b>). The re-docking of the native ligand (co-crystal) into the MRSA protein target pocket validates the method, resulting in a root mean square deviation (RMSD) value of 0.991 Å (<b>A</b>). Additionally, the interactions of the compounds in the extract, the native ligand, and the positive control with the MRSA protein target (4DKI) were illustrated, emphasizing their binding affinities and interactions within the target pocket.</p>
Full article ">Figure 4
<p>MRSA protein target (6H5O) structure (<b>F</b>), the native ligand (piperacillin) (<b>E</b>), and re-docking native ligand (co-crystal) result in MRSA protein target pocket for validating the method with RMSD value of 0.991 Å (<b>D</b>). Additionally, the interactions of the compounds in the extract, the native ligand, and the positive control with the MRSA protein target (6H5O) were illustrated, emphasizing their binding affinities and interactions within the target pocket.</p>
Full article ">Figure 5
<p>MRSA protein target (4CJN) Structure (<b>I</b>), the native ligand (E)-3-(2-(4-cyanostyryl)-4-oxoquinazolin-3(4H)-yl)benzoic acid (<b>H</b>) and re-docking native ligand (co-crystal) result in MRSA protein target pocket for validate the method with RMSD value of 0.751 Å (<b>G</b>). Additionally, the interactions of the compounds in the extract, the native ligand, and the positive control with the MRSA protein target (4CJN) were illustrated, emphasizing their binding affinities and interactions within the target pocket.</p>
Full article ">
19 pages, 2936 KiB  
Article
Targeted Screening of Lactiplantibacillus plantarum Strains Isolated from Tomatoes and Its Application in Tomato Fermented Juice
by Nuersiman Tuerhong, Liang Wang, Jie Cui, Dilireba Shataer, Huizhen Yan, Xiaoxiao Dong, Ziqi Gao, Minwei Zhang, Yanan Qin and Jing Lu
Foods 2024, 13(22), 3569; https://doi.org/10.3390/foods13223569 - 8 Nov 2024
Viewed by 393
Abstract
This study explores the functional attributes of Lactiplantibacillus plantarum (L. plantarum) strains isolated from fermented tomato juice, focusing on their physiological, biochemical, and probiotic characteristics. The identified 66 gram-positive strains included 36 L. plantarum ones, which exhibited robust growth in acidic [...] Read more.
This study explores the functional attributes of Lactiplantibacillus plantarum (L. plantarum) strains isolated from fermented tomato juice, focusing on their physiological, biochemical, and probiotic characteristics. The identified 66 gram-positive strains included 36 L. plantarum ones, which exhibited robust growth in acidic environments (pH 2.0–5.0) and utilization of various carbohydrates. Notably, seven strains outperformed a commercial strain in extreme acidic conditions. Antioxidant activity varied, with strain A24 showing the highest hydroxyl radical scavenging ability, while strains with high surface hydrophobicity had lower DPPH scavenging activity, indicating no direct correlation between these properties. Strains also showed strain-specific differences in carbohydrate utilization and antibiotic resistance, with some resistant to gentamicin and ciprofloxacin. Survival rates under simulated gastrointestinal conditions were strain-specific, with some strains demonstrating high survival rates, indicating their potential as probiotics. Furthermore, 13 strains used as fermentation starters in tomato juice significantly enhanced antioxidant activity and reduced pH and total soluble solids, indicating efficient sugar utilization and lactic acid production. These findings suggest that L. plantarum strains are well-suited for functional food fermentation and probiotic applications, with strain-specific traits offering versatility for use in acidic food products and probiotic formulations. Full article
(This article belongs to the Section Food Biotechnology)
Show Figures

Figure 1

Figure 1
<p>Growth tolerance of <span class="html-italic">L. plantarum</span> at various pH levels. Panels (<b>a</b>–<b>d</b>) illustrate the growth capabilities of <span class="html-italic">L. plantarum</span> at pH 2.0, 3.0, 4.0, and 5.0, respectively, under 0, 16, and 24 h. GL represents <span class="html-italic">L. casei</span> CICC6114; ZW represents <span class="html-italic">L. plantarum</span> CICC25155; FJ represents <span class="html-italic">L. fermentum</span> CICC25124.</p>
Full article ">Figure 2
<p>Antioxidant capacity and surface hydrophobicity of <span class="html-italic">L. plantarum strains</span>. (<b>a</b>) Illustrates the hydroxyl radical, antioxidant, and DPPH radical-scavenging activities of <span class="html-italic">L. plantarum strains</span>. (<b>b</b>) Depicts the autoaggregation capacity of <span class="html-italic">L. plantarum strains</span>. GL represents <span class="html-italic">L. casei</span> CICC6114; ZW represents <span class="html-italic">L. plantarum</span> CICC25155; FJ represents <span class="html-italic">L. fermentum</span> CICC25124. Different letters on the top indicate signficant in hyroxyl radical, antioxidant, DPPH and autoaggregation capacity of <span class="html-italic">L. plantarum strains</span> (<span class="html-italic">p</span> &lt; 0.05, ANOVA Significant Difference test. The LSD letter labeling was used to indicate significance).</p>
Full article ">Figure 3
<p>Carbohydrate utilization by <span class="html-italic">L. plantarum</span> strains. Red font and Blue font highlights significant differences in utilization between the two sugars compared. GL represents <span class="html-italic">L. casei</span> CICC6114; ZW represents <span class="html-italic">L. plantarum</span> CICC25155; FJ represents <span class="html-italic">L. fermentum</span> CICC25124. Different letters at the top indicate significant differences in Monosaccharide, Disaccharide, and Polysaccharide levels of <span class="html-italic">L. plantarum</span> strains (<span class="html-italic">p</span> &lt; 0.05, ANOVA Significant Difference test). The LSD letter labeling was used to indicate significance.</p>
Full article ">Figure 4
<p>Antibiotic resistance profile of <span class="html-italic">L. plantarum</span> strains. Panels (<b>a</b>–<b>d</b>) correspond to resistance against ampicillin, gentamicin, kanamycin, and tetracycline, respectively. The dashed line indicates the antibiotic resistance threshold. GL represents <span class="html-italic">L. casei</span> CICC6114; ZW represents <span class="html-italic">L. plantarum</span> CICC25155; FJ represents <span class="html-italic">L. fermentum</span> CICC25124. Different letters at the top indicate significant differences in ampicillin, gentamicin, kanamycin and tetracycline of <span class="html-italic">L. plantarum</span> strains (<span class="html-italic">p</span> &lt; 0.05, ANOVA Significant Difference test). The LSD letter labeling was used to indicate significance.</p>
Full article ">Figure 5
<p>Gastrointestinal tolerance of <span class="html-italic">L. plantarum</span> strains. Panels (<b>a</b>,<b>b</b>) depict the gastrointestinal survival ability of <span class="html-italic">L. plantarum</span> evaluated at inoculation levels of 1 × 10<sup>8</sup> CFU/mL and 1 × 10<sup>9</sup> CFU/mL, respectively. Different letters above the bars indicate significant differences (<span class="html-italic">p</span> &lt; 0.05, one-way ANOVA test). GL represents <span class="html-italic">L. casei</span> CICC6114; ZW represents <span class="html-italic">L. plantarum</span> CICC25155; FJ represents <span class="html-italic">L. fermentum</span> CICC25124.</p>
Full article ">Figure 6
<p>Antioxidant capacity and physiological characteristics of 13 <span class="html-italic">L. plantarum</span> strains in tomato juice fermentation. (<b>a</b>–<b>d</b>) show the SOD activity, ABTS scavenging capacity, TSS characteristics, and pH value of tomato juice fermentation after 22 h fermentation. Different letters above the bars indicate significant differences (<span class="html-italic">p</span> &lt; 0.05, one-way ANOVA test). GL represents <span class="html-italic">L. casei</span> CICC6114; ZW represents <span class="html-italic">L. plantarum</span> CICC25155; FJ represents <span class="html-italic">L. fermentum</span> CICC25124.</p>
Full article ">
12 pages, 2937 KiB  
Article
Isolation and Molecular Characterization of Corynebacterium pseudotuberculosis from Goats in Andaman and Nicobar Islands, India
by Jai Sunder, Arun Kumar De, Tamilvanan Sujatha, Gayatri Chakraborty, Srikoti Chandershekhar Mayuri, Debasis Bhattacharya, Rafeeque Rahman Alyethodi and Eaknath Bhanudasrao Chakurkar
Microbiol. Res. 2024, 15(4), 2274-2285; https://doi.org/10.3390/microbiolres15040152 - 8 Nov 2024
Viewed by 427
Abstract
Caseous lymphadenitis (CLA), caused by the bacteria Corynebacterium pseudotuberculosis, is a highly contagious disease of small ruminants, especially of goats and sheep. Here, we report an outbreak of the disease in goats for the first time from the Andaman and Nicobar Islands [...] Read more.
Caseous lymphadenitis (CLA), caused by the bacteria Corynebacterium pseudotuberculosis, is a highly contagious disease of small ruminants, especially of goats and sheep. Here, we report an outbreak of the disease in goats for the first time from the Andaman and Nicobar Islands along with isolation and molecular characterization of the pathogen. A total of 22 goats were affected, with an attack rate of 12.02%, and six isolates were identified from the clinical samples. Molecular characterization of the pathogen was carried out based on the sequence information of 16S rRNA and RNA polymerase β subunit (rpoB) gene fragments. rpoB-based phylogenetic analysis indicated that the isolates belonged to Corynebacterium pseudotuberculosis biovar ovis. The antimicrobial resistance study revealed that the isolates were 100% resistant against erythromycin and rifampicin. Fifty percent resistance was found against amoxicillin/clavulanic acid, ciprofloxacin, penicillin, and vancomycin. All the isolates were sensitive to tetracycline, chloramphenicol, cotrimoxazole, sulphafurazole, ampicillin/cloxacillin, and oxytetracycline. In conclusion, the present study reports the occurrence of CLA in goats for the first time from an isolated archipelago of India and unveils the molecular signature and antibiotic resistance patten of the pathogen. The findings of this study will be helpful to control or eradicate the disease from the Andaman and Nicobar Islands. Full article
Show Figures

Figure 1

Figure 1
<p>CLA-affected goats.</p>
Full article ">Figure 2
<p>Amplification of PCR product for 16S rRNA (<b>A</b>) and rpoB genes (<b>B</b>).</p>
Full article ">Figure 3
<p>Biovar assignment of <span class="html-italic">C. pseudotuberculosis</span> isolates based on rpoB gene fragments. Phylogenetic tree was constructed by the Neighbor Joining (NJ) method [<a href="#B34-microbiolres-15-00152" class="html-bibr">34</a>] using the Tamura–Nei model [<a href="#B35-microbiolres-15-00152" class="html-bibr">35</a>].</p>
Full article ">Figure 4
<p>Bayesian phylogenetic tree of <span class="html-italic">C. pseudotuberculosis</span> isolates established in BEAST v 1.10.4 [<a href="#B36-microbiolres-15-00152" class="html-bibr">36</a>].</p>
Full article ">Figure 5
<p>Network map of <span class="html-italic">C. pseudotuberculosis</span> isolates. The network was constructed in PopART ver. 1.7 [<a href="#B37-microbiolres-15-00152" class="html-bibr">37</a>]. Numbers in brackets indicate number of mutations.</p>
Full article ">
14 pages, 741 KiB  
Article
Pharmaceutical Residues in Sediments of a Coastal Lagoon in Northwest Mexico—Occurrence and Environmental Risk Assessment
by Oscar Fernando Becerra-Rueda, Griselda Margarita Rodríguez-Figueroa, Ana Judith Marmolejo-Rodríguez, Sergio Aguíñiga-García and Juan Carlos Durán-Álvarez
J. Xenobiot. 2024, 14(4), 1757-1770; https://doi.org/10.3390/jox14040093 - 7 Nov 2024
Viewed by 555
Abstract
Contamination of marine ecosystems by pharmaceutically active compounds (PhACs) deserves more research since their environmental fate differs from that observed in freshwater systems. However, knowledge remains scarce, especially in semi-arid coastal regions of the Global South. This study investigates the occurrence and distribution [...] Read more.
Contamination of marine ecosystems by pharmaceutically active compounds (PhACs) deserves more research since their environmental fate differs from that observed in freshwater systems. However, knowledge remains scarce, especially in semi-arid coastal regions of the Global South. This study investigates the occurrence and distribution of caffeine, carbamazepine, ciprofloxacin, and sulfamethoxazole in sediments from the La Paz lagoon, a coastal system in a semi-arid region of Mexico with inverse estuarine conditions. Samples of superficial sediments (0–5 cm depth) were collected from 18 sampling points distributed through the lagoon, encompassing sites heavily polluted by discharges of municipal sewage and 3 potentially pristine sites far from the urban and peri-urban zones. Also, a 25 cm length sediment core was taken and divided into 1 cm sub-samples to determine the deposition of target PhACs in the sediment bed through time. The extraction of the target PhACs was performed through the accelerated solvent extraction (ASE) technique and quantification was achieved using a validated HPLC-MS/MS analytical method. The concentration of caffeine, carbamazepine, ciprofloxacin, and sulfamethoxazole in superficial sediment oscillated in the range of 1 to 45 ng g−1 (dry weight). The highest mass fraction of target PhACs was detected in sites impacted by wastewater discharges. The caffeine-to-carbamazepine ratio was determined for the first time in marine sediments impacted by wastewater discharges, resulting in values from 4.2 to 9.12. Analysis of the 25 cm length sediment core revealed a high dispersion of caffeine, which was attributed to high water solubility, while antibiotics were predominantly detected in the upper 20 cm of the core. Risk quotients were calculated, observing low risk for caffeine, carbamazepine, and ciprofloxacin, while sulfamethoxazole presented high risk in all the sampling points. PhACs are retained in superficial sediments from a lagoon impacted by wastewater discharges, and the level of impact depends on the properties of the compounds and the TOC content in sediments. Risk assessments should be performed in the future considering the combination of pharmaceuticals and byproducts in marine sediments. This research emphasizes the importance of sewage management in preserving marine ecosystems in semi-arid regions in the Global South. Full article
(This article belongs to the Section Emerging Chemicals)
Show Figures

Graphical abstract

Graphical abstract
Full article ">Figure 1
<p>Location of the sampling points distributed through the La Paz lagoon. Map of La Paz Bay and La Paz Lagoon, Baja California Sur, Mexico. Shows sediment sampling points (red circles) and a sediment core (blue triangle), agricultural fields (green), oxidation ponds (pink), and sewage treatment plants (yellow).</p>
Full article ">
19 pages, 6612 KiB  
Article
From Burst to Sustained Release: The Effect of Antibiotic Structure Incorporated into Chitosan-Based Films
by Nathália F. Sczesny, Helton J. Wiggers, Cecilia Z. Bueno, Pascale Chevallier, Francesco Copes and Diego Mantovani
Antibiotics 2024, 13(11), 1055; https://doi.org/10.3390/antibiotics13111055 - 6 Nov 2024
Viewed by 696
Abstract
Background/Objectives: Medical devices are susceptible to bacterial colonization and biofilm formation, which can result in severe infections, leading to prolonged hospital stays and increased burden on society. Antibacterial films have the potential to assist in preventing biofilm formation, thereby reducing administration of antibiotics [...] Read more.
Background/Objectives: Medical devices are susceptible to bacterial colonization and biofilm formation, which can result in severe infections, leading to prolonged hospital stays and increased burden on society. Antibacterial films have the potential to assist in preventing biofilm formation, thereby reducing administration of antibiotics and the emergence of antibiotic-resistant strains. In a previous study, a chitosan-based matrix crosslinked with tannic acid and loaded with gentamicin was reported. In this study, five different antibiotics (moxifloxacin, ciprofloxacin, trimethoprim, sulfamethoxazole or linezolid) were loaded into these chitosan-based films, and their impact on the release behavior carefully assessed. Methods: The samples were characterized according to their thickness, swelling, and mass loss in phosphate-buffered saline (PBS), as well as by morphology using scanning electron microscopy (SEM) and optical phase contrast microscopy. Antibiotic release over time was quantified in PBS by high-performance liquid chromatography (HPLC). Antibacterial activity was investigated by disk diffusion test and antibiotic release over time. Finally, the cytotoxicity of the samples was assessed with human dermal fibroblasts. Results: The obtained results differed significantly, especially regarding the antibiotic release time and antibacterial activity, which varied from one day to six months, enabling classification of the films from burst/transient to prolonged release. The films also showed antibacterial features against bacteria mostly present in medical devices and displayed to be non-cytotoxic. Conclusions: In conclusion, it was demonstrated that the antibiotics structure significantly alters the release kinetics, and that by carefully selecting the antibiotic, the consequent release can be tuned. This approach yielded films that could be used for potentially-scalable release in antimicrobial coatings specific to medical devices, aiming to reduce biomaterial associated infections (BAIs). Full article
Show Figures

Graphical abstract

Graphical abstract
Full article ">Figure 1
<p>Cumulative antibiotic release from chitosan films crosslinked with tannic acid and iron until 200 min (<b>A</b>) and 200 days (<b>B</b>). (<b>C</b>) represents the antibiotic structure and release time. Studies carried out in PBS buffer at 37 °C, 100 rpm, pH 7.4.</p>
Full article ">Figure 2
<p>Schematic representation of intermolecular interactions between antibiotics and film matrix. Images produced with BioRender.</p>
Full article ">Figure 3
<p>Visual inspection, optical microscopy and scanning electron microscopy analysis of blank film, sulfamethoxazole-loaded film and moxifloxacin-loaded film.</p>
Full article ">Figure 4
<p>Antibiotic concentration released from chitosan-based films in PBS. The samples were made with chitosan, tannic acid, FeSO<sub>4</sub> and the respective antibiotic listed on the label. * BAI: Biomaterial Associated Infection.</p>
Full article ">Figure 5
<p>Indirect antibacterial activity over time of films against <span class="html-italic">E. coli</span> (<b>A</b>) and <span class="html-italic">S. aureus</span> (<b>B</b>). Experiments were performed in triplicates.</p>
Full article ">Figure 6
<p>Indirect cytotoxicity test. Results obtained by treating HDFs with the extracts from the different experimental conditions. Cell viability was measured after 1 day of incubation by means of a resazurin salt solution. * Kruskal-Wallis method with Dunn post-test, <span class="html-italic">p</span> ˂ 0.05 vs. CTRL.</p>
Full article ">
17 pages, 5264 KiB  
Article
Enhanced Antibacterial, Anti-Inflammatory, and Antibiofilm Activities of Tryptophan-Substituted Peptides Derived from Cecropin A-Melittin Hybrid Peptide BP100
by Sukumar Dinesh Kumar, Eun Young Kim, Naveen Kumar Radhakrishnan, Jeong Kyu Bang, Sungtae Yang and Song Yub Shin
Molecules 2024, 29(22), 5231; https://doi.org/10.3390/molecules29225231 - 5 Nov 2024
Viewed by 443
Abstract
The emergence of multidrug-resistant pathogens necessitates the development of novel antimicrobial agents. BP100, a short α-helical antimicrobial peptide (AMP) derived from cecropin A and melittin, has shown promise as a potential therapeutic. To enhance its efficacy, we designed and synthesized 16 tryptophan-substituted BP100 [...] Read more.
The emergence of multidrug-resistant pathogens necessitates the development of novel antimicrobial agents. BP100, a short α-helical antimicrobial peptide (AMP) derived from cecropin A and melittin, has shown promise as a potential therapeutic. To enhance its efficacy, we designed and synthesized 16 tryptophan-substituted BP100 analogs based on helical wheel projections. Among these, BP5, BP6, BP8, BP11, and BP13 exhibited 1.5- to 5.5-fold higher antibacterial activity and improved cell selectivity compared to BP100. These analogs demonstrated superior efficacy in suppressing pro-inflammatory cytokine release in LPS-stimulated RAW 264.7 cells and eradicating preformed biofilms of multidrug-resistant Pseudomonas aeruginosa (MDRPA). Additionally, these analogs showed greater resistance to physiological salts and serum compared to BP100. Mechanistic studies revealed that BP100 and its analogs exert their antibacterial effects through membrane disruption, depolarization, and permeabilization. Notably, these analogs showed synergistic antimicrobial activity with ciprofloxacin against MDRPA. Our findings suggest that these tryptophan-substituted BP100 analogs represent promising candidates for combating multidrug-resistant bacterial infections, offering a multifaceted approach through their antibacterial, anti-inflammatory, and antibiofilm activities. Full article
(This article belongs to the Section Medicinal Chemistry)
Show Figures

Figure 1

Figure 1
<p>The helical wheel projections of BP100 and its analogs are depicted with color-coded residues. Non-polar hydrophobic residues are shown in yellow, and polar basic residues are shown in dark blue. The hydrophobic moment is represented by a black arrow on the helical wheel.</p>
Full article ">Figure 2
<p>Effects of BP100 and its analogs on the release of pro-inflammatory cytokines from lipopolysaccharide (LPS)-stimulated RAW264.7 macrophage cells. (<b>a</b>) Tumor necrosis factor-α (TNF-α) release. (<b>b</b>) Interleukin-6 (IL-6) release. Peptides were administered at a concentration of 2 μM. Data represent the mean ± standard error of the mean (SEM) from at least three independent experiments. Statistical analysis was performed using one-way analysis of variance (ANOVA) followed by Duncan’s test. Compared to control: ** <span class="html-italic">p</span> &lt; 0.01. Compared to LPS: <sup>##</sup> <span class="html-italic">p</span> &lt; 0.01.</p>
Full article ">Figure 3
<p>Antibiofilm activity of BP100 and its analogs against multidrug-resistant <span class="html-italic">Pseudomonas aeruginosa</span> (MDRPA). (<b>a</b>,<b>b</b>) Inhibitory activity against biofilm formation. (<b>c</b>,<b>d</b>) Eradication activity of preformed MDRPA biofilms. Dotted lines indicate 50% inhibition (MBIC<sub>50</sub>), 90% inhibition (MBIC<sub>90</sub>), 50% eradication (MBEC<sub>50</sub>), and 90% eradication (MBEC<sub>90</sub>) concentrations. Values represent the mean ± standard error of the mean (SEM) from three independent experiments.</p>
Full article ">Figure 4
<p>Circular dichroism (CD) spectra of BP100 and its analogs in various environments. (<b>a</b>) With 10 mM of sodium phosphate buffer. (<b>b</b>) With 50% trifluoroethanol (TFE). (<b>c</b>) With 30 mM of sodium dodecyl sulfate (SDS). The mean residue ellipticity was plotted against wavelength. Each spectrum represents the average of three independent scans.</p>
Full article ">Figure 5
<p>Membrane interaction studies of BP100 and its analogs. (<b>a</b>) Time-dependent cytoplasmic membrane depolarization of <span class="html-italic">Staphylococcus aureus</span> (KCTC 1621) treated with peptides at 1× MIC (BP100, BP8, BP12, and BP13: 8 μM, BP5 and BP6: 4 μM, Buforin-2: 32 μM), as assessed by the release of the membrane potential-sensitive dye 3,3′-dipropylthiadicarbocyanine iodide (diSC<sub>3</sub>-5). Control experiments conducted in the absence of bacterial cells showed no direct peptide–dye interactions, confirming that fluorescence changes were specifically due to membrane depolarization. (<b>b</b>) Outer membrane permeabilization of <span class="html-italic">Escherichia coli</span> (KCTC 1682) in the presence of different peptide concentrations, as measured by 1-N-phenylnaphthylamine (NPN) uptake.</p>
Full article ">Figure 6
<p>Assessment of bacterial membrane integrity using flow cytometry. (<b>a</b>) <span class="html-italic">Escherichia coli</span> (KCTC 1682). (<b>b</b>) <span class="html-italic">Staphylococcus aureus</span> (KCTC 1621). Mid-logarithmic phase bacterial cultures were treated with 1× MIC of peptides, and cellular fluorescence was observed using a FACS scan flow cytometer. Membrane integrity damage was assessed by an increase in fluorescent intensity of propidium iodide (PI, 10 μg/mL) after incubation at 37 °C for 1 h. Control samples were processed without peptide treatment. This analysis provides quantitative data on the membrane-disrupting capabilities of the peptides against both Gram-negative and Gram-positive bacteria.</p>
Full article ">
Back to TopTop