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

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

Search Results (1,421)

Search Parameters:
Keywords = EPR

Order results
Result details
Results per page
Select all
Export citation of selected articles as:
16 pages, 20305 KiB  
Article
Preparation and Antibacterial Performance Study of CeO2/g-C3N4 Nanocomposite Materials
by Jingtao Zhang, Ruichun Nan, Tianzhu Liang, Yuheng Zhao, Xinxin Zhang, Mengzhen Zhu, Ruoyu Li, Xiaodong Sun, Yisong Chen and Bingkun Liu
Molecules 2024, 29(23), 5557; https://doi.org/10.3390/molecules29235557 - 25 Nov 2024
Viewed by 96
Abstract
In response to the challenges of food spoilage and water pollution caused by pathogenic microorganisms, CeO2/g-C3N4 nanocomposites were synthesized via one-step calcination using thiourea and urea as precursors. Steady-state photoluminescence (PL) spectroscopy analysis demonstrated that 8 wt% CeO [...] Read more.
In response to the challenges of food spoilage and water pollution caused by pathogenic microorganisms, CeO2/g-C3N4 nanocomposites were synthesized via one-step calcination using thiourea and urea as precursors. Steady-state photoluminescence (PL) spectroscopy analysis demonstrated that 8 wt% CeO2/g-C3N4 exhibited superior electron–hole separation efficiency. Quantitative antimicrobial assays demonstrated that the nanocomposites displayed enhanced bactericidal activity against Escherichia coli, Ralstonia solanacearum, and Staphylococcus aureus. Electron paramagnetic resonance (EPR) spectroscopy analysis verified the generation of hydroxyl radicals (·OH) and superoxide radicals (·O2) during the photo-Fenton process utilizing CeO2/g-C3N4 nanocomposites. Additionally, 8 wt% CeO2/g-C3N4 nanocomposites demonstrated enhanced photocatalytic degradation of rhodamine B (RhB) and tetracycline hydrochloride (TC) under photo-Fenton conditions. Full article
Show Figures

Graphical abstract

Graphical abstract
Full article ">Figure 1
<p>PL spectra of samples with different proportions.</p>
Full article ">Figure 2
<p>XRD patterns of different samples.</p>
Full article ">Figure 3
<p>UV-vis spectra of different samples.</p>
Full article ">Figure 4
<p>TEM (<b>a</b>) and HRTEM (<b>b</b>) spectrum of the 8 wt% CeO<sub>2</sub>/g-C<sub>3</sub>N<sub>4</sub> heterojunction.</p>
Full article ">Figure 5
<p>FTIR spectra of different samples.</p>
Full article ">Figure 6
<p>XPS spectra of 8 wt% CeO<sub>2</sub>/g-C<sub>3</sub>N<sub>4</sub> nanocomposites. (<b>a</b>) Survey, (<b>b</b>) N 1s, (<b>c</b>) Ce 3d, and (<b>d</b>) O 1s.</p>
Full article ">Figure 7
<p>Different samples of (<b>a</b>) I-T and (<b>b</b>) EIS curves.</p>
Full article ">Figure 8
<p>(<b>a</b>) Fenton-like sterilization curves of different samples. (<b>b</b>) Photocatalytic sterilization curve; (species of bacteria: <span class="html-italic">E. coli</span>; H<sub>2</sub>O<sub>2</sub>: 5 mM; light intensity: 30 mW/cm<sup>2</sup>).</p>
Full article ">Figure 9
<p>Photo-Fenton (<b>a</b>) <span class="html-italic">E. coli</span> curve; (<b>b</b>) <span class="html-italic">R. solanacearum;</span> and (<b>c</b>) <span class="html-italic">S. aureus</span> (H<sub>2</sub>O<sub>2</sub>: 5 mM; light intensity: 30 mW/cm<sup>2</sup>).</p>
Full article ">Figure 10
<p>The 8 wt%CeO<sub>2</sub>/g-C<sub>3</sub>N<sub>4</sub> nanomaterials of the (<b>a</b>) photo-Fenton reaction cycle experiment; (<b>b</b>) comparison of XRD spectra before and after (experimental conditions: <span class="html-italic">R. solanacearum</span> as model bacteria, [H<sub>2</sub>O<sub>2</sub>] = 5 mM, light intensity = 30 mW/cm<sup>2</sup>).</p>
Full article ">Figure 11
<p>Photo-Fenton degradation of 8 wt% CeO<sub>2</sub>/g-C<sub>3</sub>N<sub>4</sub> samples; (<b>a</b>) RhB and (<b>b</b>) kinetic curves; (<b>c</b>) TC and (<b>d</b>) kinetic curves (amount of catalyst: 20 mg; H<sub>2</sub>O<sub>2</sub>: 5 mM; light intensity: 100 mW/cm<sup>2</sup>).</p>
Full article ">Figure 12
<p>Photo-Fenton degradation of RhB by 8 wt% CeO<sub>2</sub>/g-C<sub>3</sub>N<sub>4</sub> samples under different experimental conditions; (<b>a</b>) catalyst dosage; (<b>b</b>) pH value; (<b>c</b>) H<sub>2</sub>O<sub>2</sub> concentration (pollutant concentration: 30 mg/L; light intensity: 100 mW/cm<sup>2</sup>).</p>
Full article ">Figure 13
<p>Scanning electron microscopy images of cells before (<b>a</b>) and after (<b>b</b>) photo-Fenton treatment.</p>
Full article ">Figure 14
<p>(<b>a</b>,<b>d</b>) Phase contrast microscope images of the corresponding fluorescent images (<b>b</b>,<b>c</b>,<b>e</b>,<b>f</b>) showing the cell viability of <span class="html-italic">E. coli</span> post 40 min illumination, (<b>a</b>–<b>c</b>) without photo-Fenton treatment or (<b>d</b>–<b>f</b>) with photo-Fenton of 8 wt% CeO<sub>2</sub>/g-C<sub>3</sub>N<sub>4</sub>. Cells were stained with CFDA, AM/Propidium Iodide: the live cells fluoresce green; the dead cells fluoresce (red, dead bacteria; green, live bacteria).</p>
Full article ">Figure 15
<p>The (<b>a</b>) ·OH and (<b>b</b>) ·O<sub>2</sub><sup>−</sup> signals of different samples.</p>
Full article ">Figure 16
<p>Effect of scavenger on photo-Fenton bacteriostatic rate. (IPA: ·OH; TEMPOL: ·O<sub>2</sub><sup>−</sup> ).</p>
Full article ">Figure 17
<p>Toxicity of 8 wt% CeO<sub>2</sub>/g-C<sub>3</sub>N<sub>4</sub> to HEK293 cells.</p>
Full article ">
21 pages, 3674 KiB  
Article
Preclinical Photodynamic Therapy Targeting Blood Vessels with AGuIX® Theranostic Nanoparticles
by Ewa Kowolik, Dariusz Szczygieł, Małgorzata Szczygieł, Agnieszka Drzał, Kalyani Vemuri, Anna-Karin Olsson, Arjan W. Griffioen, Patrycja Nowak-Sliwinska, Agnieszka Wolnicka-Glubisz and Martyna Elas
Cancers 2024, 16(23), 3924; https://doi.org/10.3390/cancers16233924 - 23 Nov 2024
Viewed by 341
Abstract
Background: Glioblastoma multiforme (GBM) is the most common highly aggressive, primary malignant brain tumor in adults. Current experimental strategies include photodynamic therapy (PDT) and new drug delivery technologies such as nanoparticles, which could play a key role in the treatment, diagnosis, and imaging [...] Read more.
Background: Glioblastoma multiforme (GBM) is the most common highly aggressive, primary malignant brain tumor in adults. Current experimental strategies include photodynamic therapy (PDT) and new drug delivery technologies such as nanoparticles, which could play a key role in the treatment, diagnosis, and imaging of brain tumors. Objectives: The purpose of this study was to test the efficacy of PDT using AGuIX-TPP, a polysiloxane-based nanoparticle (AGuIX) that contains TPP (5,10,15,20-tetraphenyl-21H,23H-porphine), in biological models of glioblastoma multiforme and to investigate the vascular mechanisms of action at multiple complexity levels. Methods: PDT effects were studied in monolayer and spheroid cell culture, as well as tumors in chicken chorioallantoic membranes (CAMs) and in mice were studied. Results: Treatment was effective in both endothelial ECRF and glioma U87 cells, as well as in the inhibition of growth of the glioma spheroids. PDT using AGuIX-TPP inhibited U87 tumors growing in CAM and destroyed their vascularization. The U87 tumors were also grown in nude mice. Their vascular network, as well as oxygen partial pressure, were assessed using ultrasound and EPR oximetry. The treatment damaged tumor vessels and slightly decreased oxygen levels. Conclusions: PDT with AGuIX-TPP was effective against glioma cells, spheroids, and tumors; however, in mice, its efficacy appeared to be strongly related to the presence of blood vessels in the tumor before the treatment. Full article
(This article belongs to the Collection Combination Therapies in Cancers)
Show Figures

Figure 1

Figure 1
<p>Scheme of AGuIX-TPP nanoparticles targeting vascular endothelial cells. AGuIX contains a KDKPPR peptide that binds to NRP-1 (created in BiorendeR).</p>
Full article ">Figure 2
<p>AGuIX-TPP uptake by ECRF and U87 cancer cells. The cells were incubated with 5 μM of AGuIX-TPP for 6 h up to 24 h. The fluorescence intensity of cellular extracts from (<b>a</b>) ECRF cells and (<b>b</b>) U87 glioma cells; <span class="html-italic">n</span> = 2 biological repetitions, in triplicate. <span class="html-italic">** p</span> &lt; 0.01 and *** &lt; <span class="html-italic">p</span> 0.001 vs. 6 h.</p>
Full article ">Figure 3
<p>AGuIX-TPP, combined with irradiation, reduced the metabolic activity of ECRF vascular endothelial cells and U87 glioma cells 24 h after treatment. (<b>a</b>) Cells incubated with AGuIX-TPP (0–10 μM) and kept in the dark (gray bars) or irradiated with 10 J/cm<sup>2</sup> (white bars); <span class="html-italic">n</span> = 9. (<b>b</b>) Cells incubated in the absence of (white bars; 0) or with 1 μM of AGuIX-TPP (gray bars; 1) and irradiated with light doses of 5, 10, and 20 J/cm<sup>2</sup>; <span class="html-italic">n</span> = 9. Statistically significant difference, <span class="html-italic">** p</span> &lt; 0.01 and *** <span class="html-italic">p</span> &lt; 0.001; vs. control (dark control) in (<b>a</b>) and vs. control (light alone) (<b>b</b>). <span class="html-italic">n</span> = 3 biological repetitions, in triplicate.</p>
Full article ">Figure 4
<p>Photodynamic effect with AGuIX-TPP in a 3D culture of U87 cells 24 h after treatment with 1, 5, and 10 μM of AGuIX-TPP and exposure to 650 nm light with a dose of 5–20 J/cm<sup>2</sup>. (<b>a</b>) Metabolic activity; <span class="html-italic">n</span> = 3–9; ** <span class="html-italic">p</span> &lt; 0.01 and *** <span class="html-italic">p</span> &lt; 0.001 vs. the lowest concentration. (<b>b</b>) Representative fluorescent images of U87 spheroids after PDE; cells stained with calcein AM (green) and EtHD (red). The scalebar represents 200 μm.</p>
Full article ">Figure 5
<p>AGuIX-TPP photodynamic therapy inhibits U87 tumor growth in CAMs. (<b>a</b>) U87 tumor growth after PDT or dark control (only an injection of 1.75 μM of AGuIX-TPP/kg b.w.) in relation to untreated control tumor growing on the CAM. This graph represents the mean of two independent experiments with SEM. Statistically significant differences between the growth of control and PDT-treated tumors: <span class="html-italic">* p</span> &lt; 0.05 and ** <span class="html-italic">p</span> &lt; 0.01. (<b>b</b>) Photograph of the U87 tumor on the CAM 24 h after PDT. Scale bar represents 1 mm. (<b>c</b>) Fluorescence angiograms taken before (<b>A</b>,<b>C</b>) and immediately after PDT (<b>B</b>,<b>D</b>). The vasculature is visualized by FITC-dextran fluorescence angiography (25 mg/kg b.w., 20 kDa, λ<sub>ex</sub> = 470 nm, λ<sub>em</sub> &gt; 520 nm). Red arrows show the changes after PDT in the blood vessels and blood stasis. Scale bar represents (<b>b</b>) 100 μm and (<b>c</b>) 0.1 mm.</p>
Full article ">Figure 6
<p>AGuIX-TPP photodynamic therapy inhibits U87 tumor growth. Mice with U87 tumors (approximately 100 mm<sup>3</sup>) were injected with 1.75 µM of AGuIX-TPP/kg b.w. and exposed for 10–17 min to 650 nm of light (130–151 mW/cm<sup>2</sup>). (<b>a</b>) Growth curves of tumors after PDT (<span class="html-italic">n</span> = 7) or control treatment (light alone); <span class="html-italic">n</span> = 2, shown as a mean with SEM, * <span class="html-italic">p</span> &lt; 0.05. (<b>b</b>) Representative photographs of tumors in the leg after the indicated time.</p>
Full article ">Figure 7
<p>Vascular disruption after PDT with AGuIX-TPP in human glioblastoma tumors in an ectopic mouse model. (<b>a</b>) The graphs show the percentage of vessels in the tumor volume, as determined by the US, after PDT (<span class="html-italic">n</span> = 14) and light control (<span class="html-italic">n</span> = 5), i.e., <span class="html-italic">n</span> = 7 for the PDT responders (2 cured, 5 partially responded), <span class="html-italic">n</span> = 7 for the PDT non-responders, and <span class="html-italic">n</span> = 4–5 for the light control. A statistically significant difference was noted in responding tumors with respect to % vessels in tumor volume before and 24 h after therapy; * <span class="html-italic">p</span> &lt; 0.05, ** <span class="html-italic">p</span> &lt; 0.01. (<b>b</b>) Representative single-slice images of ectopic U87 tumors, before and after PDT therapy, were acquired with Power Doppler, showing the vasculature with blood flow (red) and the tumor area (yellow line). The original scale bar on the left side of each image shows that the tumors were between 2 and 7 mm deep. Yellow dot marks a 5 mm distance from the transducer.</p>
Full article ">Figure 8
<p>Expression of VEGFR2, VEGFA, ANG1, and ANG2 mRNA in the ectopic U87 tumors. Real-time qPCR of VEGFR2 (<b>a</b>), VEGFA (<b>b</b>), ANG1 (<b>c</b>), and ANG2 (<b>d</b>), showing mRNA levels in the ectopic U87 tumors (<span class="html-italic">n</span> = 3–4 mice/group). Values represent fold changes relative to HPRT mRNA levels.</p>
Full article ">Figure 9
<p>Changes in oxygenation in the ectopic U87 tumors after PDT with AGuIX-TPP. The graphs show the mean pO<sub>2</sub> in tumors at the indicated time points, before and after PDT; <span class="html-italic">n</span> = 3 for the group responding to PDT, <span class="html-italic">n</span> = 4 for non-responders to PDT, and <span class="html-italic">n</span> = 2 for the light control group.</p>
Full article ">
21 pages, 29750 KiB  
Article
Effect of Ultrasonic Nanocrystalline Surface Modification (UNSM) on Stress Corrosion Cracking of 304L Stainless Steel
by Hyunhak Cho, Young-Ran Yoo and Young-Sik Kim
Metals 2024, 14(12), 1315; https://doi.org/10.3390/met14121315 - 21 Nov 2024
Viewed by 288
Abstract
The nuclear industry uses 304L stainless steel to construct canisters for storing spent nuclear fuel. The spent nuclear fuel canisters require the lifetime prediction and robustness of their corrosion behavior over periods ranging from thousands to hundreds of thousands of years. Since nuclear [...] Read more.
The nuclear industry uses 304L stainless steel to construct canisters for storing spent nuclear fuel. The spent nuclear fuel canisters require the lifetime prediction and robustness of their corrosion behavior over periods ranging from thousands to hundreds of thousands of years. Since nuclear power plants are predominantly located in coastal areas, where storage conditions are highly vulnerable to chloride environments, extensive research has been conducted to enhance the canisters’ stress corrosion cracking (SCC) resistance. The welded canisters inherently possess residual tensile stress, prompting the application of plastic deformation-based techniques to boost their SCC resistance, with peening being the most prevalent method. It is reported that UNSM increases the SCC resistance by plastic deformation through surface treatment. In this study, the commercial 304L stainless steel was subjected to UNSM treatment on base metal, heat affected zone (HAZ), and weld metal U-bend test specimens to induce compressive residual stresses up to a depth of 1 mm. The impact of peening treatment on SCC properties was explored through microstructural analysis, corrosion properties analysis, and compressive residual stress assessments. The U-bend specimens underwent SCC testing (in 42% MgCl2 at 155 °C), microstructure examination using an optical microscope (OM) and a scanning electron microscope (SEM), electron backscatter diffraction (EBSD) analysis, and compressive residual stress measurements via the hole-drilling method. Corrosion behavior on the surface and cross-section was evaluated using anodic polarization tests, electrochemical impedance spectroscopy (EIS) measurements, double loop-electrochemical potentiokinetic reactivation (DL-EPR) tests, and ASTM A262 Pr. C tests. The compressive residual stress imparted by UNSM refined the outermost grains of the cross-section and enhanced the corrosion resistance of 304L stainless steel. Furthermore, it led to a longer crack initiation time, a reduced crack propagation rate, and improved SCC properties. Full article
(This article belongs to the Section Corrosion and Protection)
Show Figures

Figure 1

Figure 1
<p>Impact of UNSM on crack times in U-bended 304LB subjected to SCC testing at 155 °C and 42% MgCl<sub>2</sub>; (<b>a</b>) crack time and (<b>b</b>) crack propagation rates.</p>
Full article ">Figure 2
<p>Image of the cross-section of U-bended 304LB after SCC test (OM, ×50, 42% MgCl<sub>2</sub> at 155 °C) (A: top of the crack, B: middle of the crack, C: bottom of the crack); (<b>a</b>) 304LB (Non-peened) and (<b>b</b>) 304LB-UNSM.</p>
Full article ">Figure 3
<p>EBSD results of the cross-section of 304LB (Non-peened) after SCC test (EBSD: step size 0.3 µm); (<b>a</b>) Band contrast, (<b>b</b>) Phase color, and (<b>c</b>) IPF color.</p>
Full article ">Figure 4
<p>EBSD results of the cross-section of 304LB-UNSM after SCC test (EBSD: step size 0.3 µm); (<b>a</b>) Band contrast, (<b>b</b>) Phase color, and (<b>c</b>) IPF color.</p>
Full article ">Figure 5
<p>Effect of UNSM on the crack times of U-bended 304LW by SCC test at 155 °C, 42% MgCl<sub>2</sub>; (<b>a</b>) crack time and (<b>b</b>) crack propagation rates.</p>
Full article ">Figure 6
<p>Image of the cross-section of U-bended 304LW following the SCC test (OM, ×50, 42% MgCl<sub>2</sub> at 155 °C) (A: top of the crack, B: middle of the crack, C: bottom of the crack); (<b>a</b>) 304LW (Non-peened) and (<b>b</b>) 304LW-UNSM.</p>
Full article ">Figure 7
<p>EBSD results of the cross-section of 304LW (Non-peened) after SCC test (EBSD: step size 0.3 µm); (<b>a</b>) Band contrast, (<b>b</b>) Phase color, and (<b>c</b>) IPF color.</p>
Full article ">Figure 8
<p>EBSD results of the cross-section of 304LW-UNSM after SCC test (EBSD: step size 0.3 µm); (<b>a</b>) Band contrast, (<b>b</b>) Phase color, and (<b>c</b>) IPF color.</p>
Full article ">Figure 8 Cont.
<p>EBSD results of the cross-section of 304LW-UNSM after SCC test (EBSD: step size 0.3 µm); (<b>a</b>) Band contrast, (<b>b</b>) Phase color, and (<b>c</b>) IPF color.</p>
Full article ">Figure 9
<p>Crack mode of 304LB and 304LW after the U-bend test (SEM, ×400; 42% MgCl<sub>2</sub> at 155 °C; Areas A, B, and C); (<b>a</b>) 304LB (Non-peened), (<b>b</b>) 304LB-UNSM, (<b>c</b>) 304LW (Non-peened), and (<b>d</b>) 304LW-UNSM.</p>
Full article ">Figure 9 Cont.
<p>Crack mode of 304LB and 304LW after the U-bend test (SEM, ×400; 42% MgCl<sub>2</sub> at 155 °C; Areas A, B, and C); (<b>a</b>) 304LB (Non-peened), (<b>b</b>) 304LB-UNSM, (<b>c</b>) 304LW (Non-peened), and (<b>d</b>) 304LW-UNSM.</p>
Full article ">Figure 10
<p>Effect of UNSM on the polarization behavior of 304L stainless steel surface in de-aerated 1% NaCl at 30 °C at a scan rate of 0.33 mV/s; (<b>a</b>) 304LB (Base metal) and (<b>b</b>) 304LW (Welded metal).</p>
Full article ">Figure 11
<p>Effect of UNSM on the EIS of the 304L stainless steel surface in de-aerated 1% NaCl at 30 °C; (<b>a</b>) Nyquist plot, (<b>b</b>) Bode plot, and (<b>c</b>) Polarization resistance.</p>
Full article ">Figure 12
<p>Optical microstructure of 304L stainless steel surface post-UNSM (OM, ×200, 10% oxalic acid); (<b>a</b>) 304LB (base metal), (<b>b</b>) 304LW-H (HAZ), and (<b>c</b>) 304LW-W (welded metal).</p>
Full article ">Figure 13
<p>Effect of UNSM on the polarization behavior of the cross-section of 304L stainless steel in de-aerated 1% NaCl at 30 °C at a scan rate of 0.33 mV/s; (<b>a</b>) 304LB (Base metal), (<b>b</b>) 304LW-H (HAZ), and (<b>c</b>) 304LW-W (Welded metal).</p>
Full article ">Figure 14
<p>Effect of UNSM on the EIS of the 304L stainless steel cross-section in de-aerated 1% NaCl at 30 °C; (<b>a</b>) Nyquist plot, (<b>b</b>) Bode plot, and (<b>c</b>) Polarization resistance.</p>
Full article ">Figure 14 Cont.
<p>Effect of UNSM on the EIS of the 304L stainless steel cross-section in de-aerated 1% NaCl at 30 °C; (<b>a</b>) Nyquist plot, (<b>b</b>) Bode plot, and (<b>c</b>) Polarization resistance.</p>
Full article ">Figure 15
<p>Optical microstructure of the cross-section of 304LW after UNSM (OM, ×200, 10% oxalic acid); (<b>a</b>) 304LB (Base metal), (<b>b</b>) 304LW-H (HAZ), and (<b>c</b>) 304LW-W (Welded metal).</p>
Full article ">Figure 16
<p>Residual stress in welds of 304L stainless steel post-UNSM; (<b>a</b>) <span class="html-italic">X</span>-axis and (<b>b</b>) <span class="html-italic">Y</span>-axis.</p>
Full article ">Figure 17
<p>Impact of UNSM on Residual Stress and Crack Time in 304L Stainless Steel; (<b>a</b>) 304LB (Base metal) and (<b>b</b>) 304LW (Welded metal).</p>
Full article ">Figure 18
<p>(<b>a</b>) Typical DL-EPR test graph and (<b>b</b>) IGC results.</p>
Full article ">Figure 19
<p>Relationship between the crack propagation and residual stress; (<b>a</b>) Total crack propagation rate and (<b>b</b>) Net crack propagation rate.</p>
Full article ">Figure 20
<p>Proposed model of crack initiation and propagation of 304L stainless steel upon UNSM treatment; (<b>a</b>) Non-peened, (<b>b</b>) UNSM-treated and crack propagation of the refined area.</p>
Full article ">Figure 20 Cont.
<p>Proposed model of crack initiation and propagation of 304L stainless steel upon UNSM treatment; (<b>a</b>) Non-peened, (<b>b</b>) UNSM-treated and crack propagation of the refined area.</p>
Full article ">
13 pages, 2731 KiB  
Article
EPR Spectroscopy Coupled with Spin Trapping as an Alternative Tool to Assess and Compare the Oxidative Stability of Vegetable Oils for Cosmetics
by Giulia Di Prima, Viviana De Caro, Cinzia Cardamone, Giuseppa Oliveri and Maria Cristina D’Oca
Appl. Sci. 2024, 14(22), 10766; https://doi.org/10.3390/app142210766 - 20 Nov 2024
Viewed by 323
Abstract
Antioxidants are the most popular active ingredients in anti-aging cosmetics as they can restore the physiological radical balance and counteract the photoaging process. Instead of adding pure compounds into the formulations, some “precious” vegetable oils could be used due to their content of [...] Read more.
Antioxidants are the most popular active ingredients in anti-aging cosmetics as they can restore the physiological radical balance and counteract the photoaging process. Instead of adding pure compounds into the formulations, some “precious” vegetable oils could be used due to their content of tocopherols, phenols, vitamins, etc., constituting a powerful antioxidant unsaponifiable fraction. Here, electron paramagnetic resonance (EPR) spectroscopy coupled with spin trapping was proven to provide a valid method for evaluating the antioxidant properties and the oxidative resistance of vegetable oils which, following UV irradiation, produce highly reactive radical species although hardly detectable. Extra virgin olive oil, sweet almond oil, apricot kernel oil, and jojoba oil were then evaluated by using N-t-butyl-α-phenylnitrone as a spin trapper and testing different UV irradiation times followed by incubation for 5 to 180 min at 70 °C. The EPR spectra were manipulated to obtain quantitative information useful for comparing the different tested samples. As a result, the knowledge acquired via the EPR analyses demonstrated jojoba oil as the best of the four considered oils in terms of both starting antioxidant ability and oxidative stability overtime. The obtained results confirmed the usefulness of the EPR spin trapping technique for the main proposed purpose. Full article
Show Figures

Figure 1

Figure 1
<p>EPR spectra of the PBN spin adducts in EVOO after 30 min of incubation at 70 °C with the not-irradiated (black line) sample and UV-irradiated ones (30 min: green line; 60 min: blue line; 120 min: red line). Results are reported as means (n = 6).</p>
Full article ">Figure 2
<p>Determination of the induction time (IT) for EVOO (sample not irradiated). Results are reported as means (n = 6).</p>
Full article ">Figure 3
<p>H<sub>p/p</sub> intensity against incubation time at 70 °C for EVOO samples. Means (n = 6) ± SD.</p>
Full article ">Figure 4
<p>H<sub>p/p</sub> intensity against incubation time at 70 °C for SAO samples. Means (n = 6) ± SD.</p>
Full article ">Figure 5
<p>H<sub>p/p</sub> intensity against incubation time at 70 °C for AKO samples. Means (n = 6) ± SD.</p>
Full article ">Figure 6
<p>H<sub>p/p</sub> intensity against incubation time at 70 °C for JO samples. Means (n = 6) ± SD.</p>
Full article ">Figure 7
<p>Slope of the linear portion of the H<sub>p/p</sub> intensity against incubation time at 70 °C graphs as a function of UV irradiation time for EVOO, SAO, AKO, and JO samples. Means (n = 6).</p>
Full article ">
26 pages, 1813 KiB  
Review
Towards Zero Waste: An In-Depth Analysis of National Policies, Strategies, and Case Studies in Waste Minimisation
by Mohammed Almansour and Mohammad Akrami
Sustainability 2024, 16(22), 10105; https://doi.org/10.3390/su162210105 - 19 Nov 2024
Viewed by 483
Abstract
This review provides a detailed analysis of zero waste (ZW) initiatives, focusing on national policies, strategies, and case studies aimed at minimising municipal solid waste (MSW). It evaluates the environmental, social, and economic impacts of waste and explores the transition from conventional landfill [...] Read more.
This review provides a detailed analysis of zero waste (ZW) initiatives, focusing on national policies, strategies, and case studies aimed at minimising municipal solid waste (MSW). It evaluates the environmental, social, and economic impacts of waste and explores the transition from conventional landfill reliance to sustainable waste management practices. Key ZW approaches, including circular economy frameworks and extended producer responsibility (EPR), are examined through case studies from countries such as China, Germany, and the United States. The review highlights advancements in waste-to-energy (WTE) technologies, the development of zero waste cities, and the critical role of policies in achieving significant MSW reduction. Additionally, it identifies key challenges such as infrastructure gaps and regulatory weaknesses and offers practical solutions to overcome these barriers. This study serves as a valuable resource for policymakers aiming to implement effective waste reduction strategies and enhance sustainable waste management systems globally. Full article
Show Figures

Figure 1

Figure 1
<p>The phases of a waste economy in South Africa [<a href="#B31-sustainability-16-10105" class="html-bibr">31</a>].</p>
Full article ">Figure 2
<p>Goals for managing MSW by 2025 and 2030 by Nottingham City Council.</p>
Full article ">
8 pages, 1239 KiB  
Proceeding Paper
Sustainable Supply Chain Management for Plastic Manufacturing in Small- and Medium-Sized Enterprises Using MCDA Method
by Hansraj Tundiya, Mohammad Ali Palsaniya, Nishi Panchal, Pranav Topre, Yash Bhavsar and Sharfuddin Ahmed Khan
Eng. Proc. 2024, 76(1), 84; https://doi.org/10.3390/engproc2024076084 - 19 Nov 2024
Viewed by 264
Abstract
This study analyzed the operational environment of supply chains involved in the production of eco-friendly plastics, with a particular emphasis on small- and medium-sized businesses worldwide. Qualitative research methods were used to highlight the significance of extended producer responsibility (EPR) regulations and strong [...] Read more.
This study analyzed the operational environment of supply chains involved in the production of eco-friendly plastics, with a particular emphasis on small- and medium-sized businesses worldwide. Qualitative research methods were used to highlight the significance of extended producer responsibility (EPR) regulations and strong recycling programs for the sustainability of small- and medium-sized enterprises (SMEs). Also, this study looked at international regulations affecting the implementation of circular economy strategies and the difficulties in creating sustainable supply chains. It concluded that waste reduction, effective supply chain management, and sustainable practices are crucial aspects of a more effective and sustainable global plastic-manufacturing sector. This research highlighted the significance of government policies in SME revival and used Multiple-Criteria Decision Analysis (MCDA) to help SMEs adopt sustainable supply chain practices that act as a catalyst for industry reform. Full article
Show Figures

Figure 1

Figure 1
<p>PRISMA methodology.</p>
Full article ">Figure 2
<p>Fuzzy ELECTRE and AHP MCDA technique framework.</p>
Full article ">Figure 3
<p>Areas to work on to attend sustainability.</p>
Full article ">Figure 4
<p>Outranking graph.</p>
Full article ">
24 pages, 3148 KiB  
Article
Nitroxyl Hybrids with Curcumin and Stilbene Scaffolds Display Potent Antioxidant Activity, Remodel the Amyloid Beta Oligomer, and Reverse Amyloid Beta-Induced Cytotoxicity
by Madhu S. Budamagunta, Hidetoshi Mori, Joshua Silk, Ryan R. Slez, Balázs Bognár, Ulises Ruiz Mendiola, Tamás Kálai, Izumi Maezawa and John C. Voss
Antioxidants 2024, 13(11), 1411; https://doi.org/10.3390/antiox13111411 - 18 Nov 2024
Viewed by 354
Abstract
The disorder and heterogeneity of low-molecular-weight amyloid-beta oligomers (AβOs) underlie their participation in multiple modes of cellular dysfunction associated with the etiology of Alzheimer’s disease (AD). The lack of specified conformational states in these species complicates efforts to select or design small molecules [...] Read more.
The disorder and heterogeneity of low-molecular-weight amyloid-beta oligomers (AβOs) underlie their participation in multiple modes of cellular dysfunction associated with the etiology of Alzheimer’s disease (AD). The lack of specified conformational states in these species complicates efforts to select or design small molecules to targeting discrete pathogenic states. Furthermore, targeting AβOs alone may be therapeutically insufficient, as AD progresses as a multifactorial, self-amplifying cascade. To address these challenges, we have screened the activity of seven new candidates that serve as Paramagnetic Amyloid Ligand (PAL) candidates. PALs are bifunctional small molecules that both remodel the AβO structure and localize a potent antioxidant that mimics the activity of SOD within live cells. The candidates are built from either a stilbene or curcumin scaffold with nitroxyl moiety to serve as catalytic antioxidants. Measurements of PAL AβO binding and remolding along with assessments of bioactivity allow for the extraction of useful SAR information from screening data. One candidate (HO-4450; PMT-307), with a six-membered nitroxyl ring attached to a stilbene ring, displays the highest potency in protecting against cell-derived Aβ. A preliminary low-dose evaluation in AD model mice provides evidence of modest treatment effects by HO-4450. The results for the curcumin PALs demonstrate that the retention of the native curcumin phenolic groups is advantageous to the design of the hybrid PAL candidates. Finally, the PAL remodeling of AβO secondary structures shows a reasonable correlation between a candidate’s bioactivity and its ability to reduce the fraction of antiparallel β-strand. Full article
Show Figures

Figure 1

Figure 1
<p>The potent radical scavenging of the nitroxide in cells. The N-oxyl cycles through alternative redox states. In the presence of the principal intracellular antioxidants glutathione (GSH) and ascorbate (ASC), the nitroxide undergoes bioreduction, enabling multiple rounds of PAL antioxidant activity.</p>
Full article ">Figure 2
<p>The chemical structure of PAL candidates. For convenience, PAL family codes are designated as PMT. The original HO- designations, as previously published, are also provided.</p>
Full article ">Figure 3
<p>PAL neuronal protection activity. Induced (TC–) C99 expression in the MC65 model results in amyloid beta cytotoxicity that can be eliminated with PAL titration. Error bars are the SEM from the assay of three separate measurements.</p>
Full article ">Figure 4
<p>CD spectra of AβOs treated with a stoichiometric amount of PAL agent. CD spectra of AβOs treated for 1 h with 40 μM PAL. The control (black) was treated with an equal volume of vehicle. ∆ε represents the molar circular dichroism.</p>
Full article ">Figure 5
<p>PALs’ effect on AβO-associated dyes. (<b>A</b>) The ThT fluorescence intensity of Aβ+PAL samples normalized to the sample of Aβ alone. (<b>B</b>) The NR fluorescence intensity of Aβ+PAL samples normalized to the sample of Aβ alone. Values represent the average intensities of triplicate samples measured at 20 h.</p>
Full article ">Figure 6
<p>Decreases in the A11 antibody recognition of AβOs following PAL treatment. Amounts of captured A11 are reported from the oxidized luminol signal of the HRP-secondary antibody. Samples were tested in quadruplicate in each assay. The results are the average of three independent assays with errors reported as SEM.</p>
Full article ">Figure 7
<p>Relative antioxidant capacities of PAL candidates. The oxidation of DCFH generates a strong fluorescence reflective of ROS levels. (<b>A</b>) shows the DCF intensities following peroxidase-dependent DCFH oxidation (by O<sub>2</sub>•−), in the presence of PALs, relative to the ACN control. The effect of PALs on DCFH oxidation (relative to the ACN control) in the presence of •OH radical generation is shown in (<b>B</b>). A PAL concentration of 20 μM was used to compare the superoxide oxidation of 20 μM DCFH. A PAL concentration of 40 μM was used to compare the hydroxyl radical oxidation of 50 μM DCFH. All assays were performed in quadruplicate.</p>
Full article ">Figure 8
<p>Effect of ROS on PAL nitroxide signal as measured by EPR spectroscopy. Shown are center line amplitudes of PAL nitroxide (160 μM) following incubation with either O<sub>2</sub>•− or •OH radical relative to PAL in buffer. Reduction in EPR amplitude is indicative of oxidation of N-oxyl moiety to diamagnetic state. We included 4 mM GSH in •OH radical samples to diminish differential production of Cu(II)-catalyzed •OH radical generation in stilbene PALs.</p>
Full article ">Scheme 1
<p>Synthesis of HO-4897.</p>
Full article ">
16 pages, 2254 KiB  
Article
The Potential of Agaricus bisporus in Mitigating Pesticide-Induced Oxidative Stress in Honey Bees Infected with Nosema ceranae
by Stefan Jelisić, Zoran Stanimirović, Marko Ristanić, Đura Nakarada, Miloš Mojović, Dušan Bošnjaković and Uroš Glavinić
Life 2024, 14(11), 1498; https://doi.org/10.3390/life14111498 - 17 Nov 2024
Viewed by 649
Abstract
Global climate change, environmental pollution, and frequent pesticide use severely reduce bee populations, greatly challenging beekeeping. Pesticides such as deltamethrin, a pyrethroid insecticide commonly used to control mosquitoes, can kill individual bees and entire colonies, depending on the exposure. Due to mosquito resistance [...] Read more.
Global climate change, environmental pollution, and frequent pesticide use severely reduce bee populations, greatly challenging beekeeping. Pesticides such as deltamethrin, a pyrethroid insecticide commonly used to control mosquitoes, can kill individual bees and entire colonies, depending on the exposure. Due to mosquito resistance to pyrethroid insecticides, components that enhance their effect are commonly used. This study explores the potential of Agaricus bisporus mushroom extract in mitigating oxidative stress in bees triggered by pesticides and Nosema ceranae infection. Our findings indicate that A. bisporus extract significantly reduced mortality rates of bees and spore counts of N. ceranae. Furthermore, the extract demonstrated antioxidant properties that lower enzyme activity related to oxidative stress (CAT, SOD, and GST) and MDA concentration, which is linked to lipid peroxidation. These results indicate that natural extracts like A. bisporus can aid bee health by mitigating the effects of pesticides and pathogens on honey bees, thus improving biodiversity. Full article
(This article belongs to the Section Animal Science)
Show Figures

Figure 1

Figure 1
<p>Survival rate of bees across control and various treated groups over a 14-day experimental period. The Kaplan–Meier survival curve demonstrates significant differences in mortality rates, with the highest mortality observed in the N-DPb group, followed by the N-D group. The Ab group, treated solely with <span class="html-italic">Agaricus bisporus</span> extract, shows the lowest mortality rate. Significance values are included to compare differences between groups. The group labels are described in <a href="#life-14-01498-t001" class="html-table">Table 1</a>.</p>
Full article ">Figure 2
<p>Mean spore counts of <span class="html-italic">Nosema ceranae</span> in bees from different experimental groups on days 7 and 14. The treatment groups are described in <a href="#life-14-01498-t001" class="html-table">Table 1</a>.</p>
Full article ">Figure 3
<p>Percentage of DPPH radical scavenging activity of the aqueous <span class="html-italic">Agaricus bisporus</span> extract (Ab), deltamethrin (D), and the combination of deltamethrin and piperonyl butoxide (DPb).</p>
Full article ">Figure 4
<p>Percentage of hydroxyl radical (•OH) scavenging activity of aqueous extract of <span class="html-italic">Agaricus bisporus</span> (Ab), deltamethrin (D), and deltamethrin with piperonyl butoxide (DPb).</p>
Full article ">Figure 5
<p>Electron Paramagnetic Resonance (EPR) signal intensity indicating •OH radical generation in reaction with test solutions: <span class="html-italic">Agaricus bisporus</span> extract (Ab), deltamethrin (D), and deltamethrin with piperonyl butoxide (DPb).</p>
Full article ">Figure 6
<p>CAT activity. Catalase (CAT) activity in the experimental groups on days 7 and 14. Groups and group labels are listed in <a href="#life-14-01498-t001" class="html-table">Table 1</a>. <span class="html-italic">p</span>-values of pairwise comparison are presented in <a href="#app1-life-14-01498" class="html-app">Supplementary Tables S1 (for day 7) and S2 (for day 14)</a>.</p>
Full article ">Figure 7
<p>SOD activity. Superoxide dismutase (SOD) activity in the experimental groups on days 7 and 14. Groups and group labels are listed in <a href="#life-14-01498-t001" class="html-table">Table 1</a>. <span class="html-italic">p</span>-values of pairwise comparison are presented in <a href="#app1-life-14-01498" class="html-app">Supplementary Tables S3 (for day 7) and S4 (for day 14)</a>.</p>
Full article ">Figure 8
<p>GST activity. Glutathione S-transferase (GST) activity in the experimental groups on days 7 and 14. Groups and group labels are listed in <a href="#life-14-01498-t001" class="html-table">Table 1</a>. <span class="html-italic">p</span>-values of pairwise comparison are presented in <a href="#app1-life-14-01498" class="html-app">Supplementary Tables S5 (for day 7) and S6 (for day 14)</a>.</p>
Full article ">Figure 9
<p>MDA concentrations. Malondialdehyde (MDA) concentrations (nmol/mg of total proteins) in the experimental groups. Groups and group labels are listed in <a href="#life-14-01498-t001" class="html-table">Table 1</a>. <span class="html-italic">p</span>-values of pairwise comparison are presented in <a href="#app1-life-14-01498" class="html-app">Supplementary Tables S7 (for day 7) and S8 (for day 14)</a>.</p>
Full article ">Figure 10
<p>Antioxidant enzyme (CAT, SOD, GST) activity and MDA concentration. Heat map of median values for the activity of catalase (CAT), superoxide dismutase (SOD), glutathione S-transferase (GST), and concentration of malondialdehyde (MDA) in honey bees from experimental groups. Groups and group labels are listed in <a href="#life-14-01498-t001" class="html-table">Table 1</a>. <span class="html-italic">p</span>-values of pairwise comparison are presented in <a href="#app1-life-14-01498" class="html-app">Supplementary Tables S1–S8</a>.</p>
Full article ">
14 pages, 3347 KiB  
Article
Efficient Degradation of Tetracycline by Peroxymonosulfate Activated with Ni-Co Bimetallic Oxide Derived from Bimetallic Oxalate
by Qi Zhang, Mingling Yu, Hang Liu, Jin Tang, Xiaolong Yu, Haochuan Wu, Ling Jin and Jianteng Sun
Toxics 2024, 12(11), 816; https://doi.org/10.3390/toxics12110816 - 14 Nov 2024
Viewed by 382
Abstract
In this work, NiCo2O4 was synthesized from bimetallic oxalate and utilized as a heterogeneous catalyst to active peroxymonosulfate (PMS) for the degradation of tetracycline (TC). The degradation efficiency of TC (30 mg/L) in the NiCo2O4 + PMS [...] Read more.
In this work, NiCo2O4 was synthesized from bimetallic oxalate and utilized as a heterogeneous catalyst to active peroxymonosulfate (PMS) for the degradation of tetracycline (TC). The degradation efficiency of TC (30 mg/L) in the NiCo2O4 + PMS system reached 92.4%, with NiCo2O4 exhibiting satisfactory reusability, stability, and applicability. Radical trapping test and electron paramagnetic resonance (EPR) results indicated that SO4•−, •OH, O2•−, and 1O2 were the dominating reactive oxygen species (ROS) for TC degradation in the NiCo2O4 + PMS system. Seven intermediates were identified, and their degradation pathways were proposed. Toxicity assessment using T.E.S.T software (its version is 5.1.1.0) revealed that the identified intermediates had lower toxicity compared to intact TC. A rice seed germination test further confirmed that the NiCo2O4 + PMS system effectively degraded TC into low-toxicity or non-toxic products. In conclusion, NiCo2O4 shows promise as a safe and efficient catalyst in advanced oxidation processes (AOPs) for the degradation of organic pollutants. Full article
Show Figures

Graphical abstract

Graphical abstract
Full article ">Figure 1
<p>The XRD pattern (<b>a</b>), FTIR spectrum (<b>b</b>), SEM (<b>c</b>), and EDS (<b>d</b>) of the obtained NiCo<sub>2</sub>O<sub>4</sub>.</p>
Full article ">Figure 2
<p>The comparison of different catalysts for TC degradation (<b>a</b>) and degradation kinetics (<b>b</b>). Conditions (unless otherwise specified in the figures): TC = 30 mg/L; pH = 6.8; temperature = 30 ± 2 °C; PMS = 0.75 mM.</p>
Full article ">Figure 3
<p>The degradation efficiency of TC in the presence of different scavengers: TBA (<b>a</b>); EtOH (<b>b</b>); <span class="html-italic">p</span>-BQ (<b>c</b>); and FFA (<b>d</b>). Conditions (unless otherwise specified in the figures): TC = 30 mg/L; pH = 6.8; temperature = 30 ± 2 °C; PMS = 0.75 mM.</p>
Full article ">Figure 4
<p>EPR spectrum of the NiCo<sub>2</sub>O<sub>4</sub> + PMS system with the addition of DMPO or TEMP: (<b>a</b>) SO<sub>4</sub><sup>−</sup> and OH; (<b>b</b>) O<sub>2</sub><sup>−</sup>; (<b>c</b>) <sup>1</sup>O<sub>2</sub>.</p>
Full article ">Figure 5
<p>The reusability, stability, and applicability of NiCo<sub>2</sub>O<sub>4</sub> (<b>a</b>). Conditions (unless otherwise specified in the figures): TC = 30 mg/L; pH = 6.8; temperature = 30 ± 2 °C; PMS = 0.75 mM. The XRD of the fresh and used of NiCo<sub>2</sub>O<sub>4</sub> (<b>b</b>). The degradation efficiency of other organic contaminants in the NiCo<sub>2</sub>O<sub>4</sub> + PMS system (<b>c</b>).</p>
Full article ">Figure 6
<p>The effect of experimental factors on TC degradation: pH (<b>a</b>); PMS concentration (<b>b</b>); NiCo<sub>2</sub>O<sub>4</sub> dosage (<b>c</b>); and TC concentration (<b>d</b>). Conditions (unless otherwise specified in the figures): TC = 30 mg/L; pH = 6.8; temperature = 30 ± 2 °C; PMS = 0.75 mM.</p>
Full article ">Figure 7
<p>The germination of rice seedlings in before reaction solution and after reaction solution.</p>
Full article ">
25 pages, 8370 KiB  
Article
The Analysis of the ZnO/Por-Si Hierarchical Surface by Studying Fractal Properties with High Accuracy and the Behavior of the EPR Spectra Components in the Ordering of Structure
by Tatyana Seredavina, Rashid Zhapakov, Danatbek Murzalinov, Yulia Spivak, Nurzhan Ussipov, Tatyana Chepushtanova, Aslan Bolysbay, Kulzira Mamyrbayeva, Yerik Merkibayev, Vyacheslav Moshnikov, Aliya Altmyshbayeva and Azamat Tulegenov
Processes 2024, 12(11), 2541; https://doi.org/10.3390/pr12112541 - 14 Nov 2024
Viewed by 399
Abstract
A hierarchical surface that includes objects with different sizes, as a result of creating local fields, initiates a large number of effects. Micropores in the composition of macropores, as well as nanoclusters of the substance, were detected by scanning electron and atomic force [...] Read more.
A hierarchical surface that includes objects with different sizes, as a result of creating local fields, initiates a large number of effects. Micropores in the composition of macropores, as well as nanoclusters of the substance, were detected by scanning electron and atomic force microscopies on the surface of ZnO/Por-Si samples. An identical fractal dimension for all levels of the hierarchy was determined for these structures, which is associated with the same response to external excitation. Photoluminescence studies have shown the presence of localized levels in the band gap, with the probability of capturing both electrons and holes, which ensures charge transitions between energy bands. Decomposition of the electron paramagnetic resonance (EPR) signal into components made it possible to determine the manifestations of various types of interaction between paramagnetic particles, including the hyperfine structure of the spectrum. The ordering of the structure of the substance as a result of sequential annealing in the range from 300 to 500 °C was revealed in the EPR spectrum. This fact, as well as photo- and gas sensitivity for all types of samples studied, confirms the prospects of using these structures as sensors. Full article
(This article belongs to the Section Materials Processes)
Show Figures

Figure 1

Figure 1
<p>Images of the sample surface without depositing ZnO layers: (<b>a</b>) SEM image taken at an angle of 12° to the horizontal axis, at magnification ×550, (<b>b</b>) SEM image taken at an angle of 12° to the horizontal axis, at magnification ×1200, (<b>c</b>) Optical microscope image of the surface of ground side of a silicon wafer.</p>
Full article ">Figure 2
<p>SEM image of a sample without depositing ZnO, taken vertically to the surface.</p>
Full article ">Figure 3
<p>Schematic representation of the structure of the porous layer of the sample without depositing ZnO layers.</p>
Full article ">Figure 4
<p>(<b>a</b>) SEM image of a macropore of a sample with 20 layers of ZnO; (<b>b</b>) The height distribution of structures at the boundary of macropores of a sample with 20 layers of ZnO, obtained by the Gwyddion program v2.64.</p>
Full article ">Figure 5
<p>(<b>a</b>) SEM image of a macropore of a sample with 25 layers of ZnO; (<b>b</b>) The height distribution of structures at the boundary of macropores of a sample with 25 layers of ZnO, obtained by the Gwyddion program v2.64.</p>
Full article ">Figure 6
<p>Microscopy images of the sample with 25 ZnO layers: (<b>a</b>) SEM image of the macroporous structure of the sample; (<b>b</b>) SEM image of the surface inside the macro pores; (<b>c</b>) AFM image of the microporous structure of the sample; (<b>d</b>) AFM image of the structure of nanocrystals formed between micropores, transformed in the Gwyddion program v2.64.</p>
Full article ">Figure 6 Cont.
<p>Microscopy images of the sample with 25 ZnO layers: (<b>a</b>) SEM image of the macroporous structure of the sample; (<b>b</b>) SEM image of the surface inside the macro pores; (<b>c</b>) AFM image of the microporous structure of the sample; (<b>d</b>) AFM image of the structure of nanocrystals formed between micropores, transformed in the Gwyddion program v2.64.</p>
Full article ">Figure 7
<p>AFM images of the microporous structure of the sample with 25 layers of ZnO: (<b>a</b>) 10 × 10 µm, (<b>b</b>) 200 nm resolution.</p>
Full article ">Figure 8
<p>Dependence of the logarithm of the number of pores N(δ) on the scale δ for (<b>a</b>) macroporous level of surface hierarchy, (<b>b</b>) microporous level of surface hierarchy.</p>
Full article ">Figure 9
<p>A percentage error matrix for determining the number of pores of porous silicon using YOLOv8 neural network.</p>
Full article ">Figure 10
<p>(<b>a</b>) AFM image of nanoclusters located between micropores; (<b>b</b>) Dependence of the logarithm of the number of pores N(δ) on the scale δ for nanoscale level of surface hierarchy.</p>
Full article ">Figure 11
<p>The photoluminescence spectrum, decomposed by Gaussian for a sample of porous silicon without ZnO.</p>
Full article ">Figure 12
<p>Photoluminescence spectrum decomposed into Gaussians for a sample: (<b>a</b>) with 20 layers of ZnO, (<b>b</b>) with 25 layers of ZnO.</p>
Full article ">Figure 13
<p>The comparison of photoluminescence peak intensities for samples with 20 and 25 ZnO layers.</p>
Full article ">Figure 14
<p>The dependence of the resistance of the samples on the number of deposited layers of ZnO.</p>
Full article ">Figure 15
<p>EPR spectrum of the sample with 25 ZnO layers before annealing.</p>
Full article ">Figure 16
<p>Comparison of EPR spectra for extreme points at signal saturation (P<sub>1</sub>= 1 mW, P<sub>2</sub> = 7.4 mW).</p>
Full article ">Figure 17
<p>Changing the signal parameters for the components of the right doublet at microwave powers from 3.4 to 6.6 mW: (<b>a</b>) changing the signal intensity from microwave power; (<b>b</b>) changing the signal width from microwave power.</p>
Full article ">Figure 18
<p>The EPR spectrum for the sample without ZnO deposition, obtained by subtracting the spectra at 5.8 mW and 5.4 mW: 1—signal in the middle of the magnetic field sweep, 2—left doublet signal, 3—right doublet signal.</p>
Full article ">Figure 19
<p>The dependence of the intensity of the fourth component of the spectrum on the sequential increase in microwave power for a sample with 25 ZnO layers.</p>
Full article ">Figure 20
<p>Comparison of the dependences of the intensity of the fourth component of the spectrum on the sequential increase in microwave power for samples: 1—25 layers of ZnO, 2—without deposition ZnO.</p>
Full article ">Figure 21
<p>EPR spectrum decomposed into components for a sample with 25 ZnO layers: (<b>a</b>) after annealing at 300 °C, (<b>b</b>) after annealing at 400 °C, (<b>c</b>) after annealing at 500 °C.</p>
Full article ">
12 pages, 2690 KiB  
Article
Perborate Activated Peroxymonosulfate Process for Improving the Coagulation Efficiency of Microcystis aeruginosa by Polymeric Aluminum Chloride
by Fan Chen, Lu Li, Shunfan Qiu, Shiyang Chen, Lingfang Yang, Lin Deng and Zhou Shi
Molecules 2024, 29(22), 5352; https://doi.org/10.3390/molecules29225352 - 14 Nov 2024
Viewed by 327
Abstract
In this study, the sodium perborate (SP)-activated peroxymonosulfate (PMS) process was used to enhance the coagulation efficiency of cyanobacteria with polymeric aluminum chloride (PAC), aiming to efficiently mitigate the impact of algal blooms on the safety of drinking water production. The optimal concentrations [...] Read more.
In this study, the sodium perborate (SP)-activated peroxymonosulfate (PMS) process was used to enhance the coagulation efficiency of cyanobacteria with polymeric aluminum chloride (PAC), aiming to efficiently mitigate the impact of algal blooms on the safety of drinking water production. The optimal concentrations of SP, PMS, and PAC were determined by evaluating the removal rate of OD680 and zeta potential of the algae. Experimental results demonstrated that the proposed ternary PMS/SP/PAC process achieved a remarkable OD680 removal efficiency of 95.2%, significantly surpassing those obtained from individual treatments with PMS (19.5%), SP (5.2%), and PAC (42.1%), as well as combined treatments with PMS/PAC (55.7%) and PMS/SP (28%). The synergistic effect of PMS/SP/PAC led to the enhanced aggregation of cyanobacteria cells due to a substantial reduction in their zeta potential. Flow cytometry was performed to investigate cell integrity before and after treatment with PMS/SP/PAC. Disinfection by-products (DBPs) (sodium hypochlorite disinfection) of the algae-laden water subsequent to PMS/SP/PAC treatment declined by 57.1%. Moreover, microcystin-LR was completely degraded by PMS/SP/PAC. Electron paramagnetic resonance (EPR) analysis evidenced the continuous production of SO4, •OH, 1O2, and O2, contributing to both cell destruction and organic matter degradation. This study highlighted the significant potential offered by the PMS/SP/PAC process for treating algae-laden water. Full article
Show Figures

Figure 1

Figure 1
<p>Comparison of different processes on cyanobacteria removal (initial algal cell density: 5.0 × 10<sup>6</sup> cells/mL, SP dosage: 1 mM, PMS dosage: 3 mM, PAC dosage: 10 mg/L).</p>
Full article ">Figure 2
<p>Effect of SP and PMS dosage on cyanobacteria removal using PMS/SP/PAC treatment (initial algal cell density: 5.0 × 10<sup>6</sup> cells/mL, PAC dosage: 10 mg/L).</p>
Full article ">Figure 3
<p>Effect of PAC dosage on cyanobacteria removal using PMS/SP/PAC treatment (initial algal cell density: 5.0 × 10<sup>6</sup> cells/mL, SP dosage: 1 mM, PMS dosage: 3 mM).</p>
Full article ">Figure 4
<p>Cell integrity of algae treated with (<b>a</b>) control, (<b>b</b>) PAC, (<b>c</b>) PMS/PAC, and (<b>d</b>) PMS/SP/PAC (initial algal cell density: 5.0 × 10<sup>6</sup> cells/mL, SP dosage: 1 mM, PMS dosage: 3 mM, PAC dosage: 10 mg/L).</p>
Full article ">Figure 5
<p>DBP yield from different treatments of cyanobacteria under chlorination in DI water (initial algal cell density: 5.0 × 10<sup>6</sup> cells/mL, SP dosage: 1 mM, PMS dosage: 3 mM, PAC dosage: 10 mg/L).</p>
Full article ">Figure 6
<p>Microcystin-LR elimination using different treatment processes in the removal of cyanobacteria (initial algal cell density: 5.0 × 10<sup>6</sup> cells/mL, SP dosage: 1 mM, PMS dosage: 3 mM, PAC dosage: 10 mg/L).</p>
Full article ">Figure 7
<p>The EPR spectra of (<b>a</b>) DMPO−OH and DMPO−SO<sub>4</sub> adducts, (<b>b</b>) DMPO−<math display="inline"><semantics> <mrow> <msubsup> <mrow> <mi mathvariant="normal">O</mi> </mrow> <mrow> <mn>2</mn> </mrow> <mrow> <mo>•</mo> <mo>−</mo> </mrow> </msubsup> </mrow> </semantics></math> adducts, and (<b>c</b>) TEMP−<sup>1</sup>O<sub>2</sub> adducts (initial algal cell density: 5.0 × 10<sup>6</sup> cells/mL, SP dosage: 1 mM, PMS dosage: 3 mM, PAC dosage: 10 mg/L, [DMPO] = [TEMP] = 100 mM).</p>
Full article ">
29 pages, 7114 KiB  
Review
Photocatalytic Reactions over TiO2-Based Interfacial Charge Transfer Complexes
by Vesna Lazić and Jovan M. Nedeljković
Catalysts 2024, 14(11), 810; https://doi.org/10.3390/catal14110810 - 11 Nov 2024
Viewed by 499
Abstract
The present review is related to the novel approach for improvement of the optical properties of wide bandgap metal oxides, in particular TiO2, based on the formation of the inorganic–organic hybrids that display absorption in the visible spectral range due to [...] Read more.
The present review is related to the novel approach for improvement of the optical properties of wide bandgap metal oxides, in particular TiO2, based on the formation of the inorganic–organic hybrids that display absorption in the visible spectral range due to the formation of interfacial charge transfer (ICT) complexes. We outlined the property requirements of TiO2-based ICT complexes for efficient photo-induced catalytic reactions, emphasizing the simplicity of the synthetic procedure, the possibility of the fine-tuning of the optical properties supported by the density functional theory (DFT) calculations, and the formation of a covalent linkage between the inorganic and organic components of hybrids, i.e., the nature of the interface. In addition, this study provides a comprehensive insight into the potential applications of TiO2-based ICT complexes in photo-driven catalytic reactions (water splitting and degradation of organic molecules), including the identification of the reactive species that participate in photocatalytic reactions by the spin-trapping electron paramagnetic resonance (EPR) technique. Considering the practically limitless number of combinations between the inorganic and organic components capable of forming oxide-based ICT complexes and with the knowledge that this research area is unexplored, we are confident it is worth studying, and we emphasized some further perspectives. Full article
(This article belongs to the Special Issue TiO2 Photocatalysts—Towards Sustainable Chemistry)
Show Figures

Figure 1

Figure 1
<p>The pH–dependent energy level position of VB<sub>max</sub> and CB<sub>min</sub> in anatase toward the vacuum level and the normal hydrogen electrode [<a href="#B4-catalysts-14-00810" class="html-bibr">4</a>].</p>
Full article ">Figure 2
<p>Absorption spectra and photo images of colloidal solutions of 45-Å TiO<sub>2</sub> nanoparticles surface-modified with different ligands: (A) bare TiO<sub>2</sub>, (B) 2-hydroxybenzoic acid, (C) 2,5-dihydroxybenzoic acid, (D) 2,3-dihydroxybenzoic acid, (E) 3,4-dihydroxybenzoic acid, and (F) catechol [<a href="#B29-catalysts-14-00810" class="html-bibr">29</a>].</p>
Full article ">Figure 3
<p>(<b>A</b>) Reaction mechanism for the in situ formation of surface-modified TiO<sub>2</sub> NPs attached to polymer support, and (<b>B</b>) TEM image of obtained composite. (<b>C</b>) Kubelka–Munk transformations of diffuse reflection data for surface-modified TiO<sub>2</sub> NPs supported by polymer, including photo image [<a href="#B24-catalysts-14-00810" class="html-bibr">24</a>].</p>
Full article ">Figure 4
<p>(<b>A</b>) Kubelka–Munk function spectra of TiO<sub>2</sub> (blank) and the TiO<sub>2</sub>–phenol sample together with photographs of these samples, and (<b>B</b>) FT-IR spectra of phenol (above) and the TiO<sub>2</sub>–phenol sample (below) together with that of TiO<sub>2</sub> (blank; dashed curve) [<a href="#B37-catalysts-14-00810" class="html-bibr">37</a>].</p>
Full article ">Figure 5
<p>Job’s curves for ligand–T<sub>surf</sub> complexes (ligands are CAT and SA; [Ti<sub>surf</sub>] + [L] = 2.0 mM) [<a href="#B29-catalysts-14-00810" class="html-bibr">29</a>].</p>
Full article ">Figure 6
<p>(<b>A</b>) IR spectra of free and adsorbed CAT to TiO<sub>2</sub> nanoparticles and calculated IR spectra for bridging and chelating coordination (top–down) [<a href="#B43-catalysts-14-00810" class="html-bibr">43</a>], (<b>B</b>) Optimized structures and (<b>C</b>) calculated vibrational spectra of [Ti<sub>8</sub>O<sub>14</sub>(OH)<sub>3</sub>–o-AP] (green), [Ti<sub>8</sub>O<sub>14</sub>(OH)<sub>3</sub>–m-AP] (orange), and [Ti<sub>8</sub>O<sub>14</sub>(OH)<sub>3</sub>–p-AP] (red). Large white: titanium; small white: hydrogen; gray: carbon; blue: nitrogen; red: oxygen atom [<a href="#B38-catalysts-14-00810" class="html-bibr">38</a>].</p>
Full article ">Figure 7
<p>Kubelka–Munk spectra (solid curves) of TiO<sub>2</sub> and TiO<sub>2</sub>–O–Ph–R (R: H, C(CH<sub>3</sub>)<sub>3</sub>, OCH<sub>3</sub>, and F<sub>5,</sub> together with absorption spectra (dashed curves) of HO–Ph–R in CH<sub>3</sub>CN solution [<a href="#B53-catalysts-14-00810" class="html-bibr">53</a>].</p>
Full article ">Figure 8
<p>(<b>A</b>) Absorption spectra of pristine 50-Å TiO<sub>2</sub> colloid (a) and surface-modified TiO<sub>2</sub> colloids with SA (b) and 5-ASA (c) and corresponding photoimages. (<b>B</b>) Photoelectron spectra of TiO<sub>2</sub>, TiO<sub>2</sub>–SA, and TiO<sub>2</sub>–5-ASA recorded at 11 eV photon energy [<a href="#B49-catalysts-14-00810" class="html-bibr">49</a>].</p>
Full article ">Figure 9
<p>(<b>A</b>) Kubelka–Munk spectra of pristine titanate nanotubes, surface-modified titanate nanotubes with 5-ASA, and surface-modified titanate nanotubes with 5-ASA decorated with Ag nanoparticles. (<b>B</b>) TEM image of surface-modified titanate nanotubes with 5-ASA decorated with Ag nanoparticles [<a href="#B76-catalysts-14-00810" class="html-bibr">76</a>].</p>
Full article ">Figure 10
<p>Effects of electron acceptor on visible-light-induced photocatalytic degradation of 4-CP [<a href="#B117-catalysts-14-00810" class="html-bibr">117</a>].</p>
Full article ">Figure 11
<p>Adsorption in dark and photocatalytic degradation of crystal violet (CV) over TiO<sub>2</sub>-based ICT complex with dopamine supported by macroporous polymer upon illumination with visible light (hν &lt; 2.75 eV) followed by absorption spectroscopy; inset: adsorption in dark and photocatalytic degradation kinetic curves of the CV using for excitation light with or without UV part of the spectrum [<a href="#B24-catalysts-14-00810" class="html-bibr">24</a>].</p>
Full article ">Figure 12
<p>Dependence of rate of H<sub>2</sub> evolution by 1,1′-binaphtalene-2,2′-diol–modified TiO<sub>2</sub> loaded with 0.5 wt.-% of Pt on the wavelength of photoirradiated light of (A) &gt;430, (B) &gt;490, (C) &gt;540, and (D) &gt;580 nm [<a href="#B145-catalysts-14-00810" class="html-bibr">145</a>].</p>
Full article ">Figure 13
<p>(<b>A</b>) Photocatalytic hydrogen evolution from aqueous TEOA (10 vol.-%) over surface-modified TiO<sub>2</sub> with (a) 4-tert-butylcatechol, (b) 3-methoxycatechol, (c) catechol, (d) 2,3-dihydroxybenzoic acid, (e) 3,4-dihydroxybenzonitrile, and (f) tiron. (<b>B</b>) Photocatalytic activity of surface-modified TiO<sub>2</sub> with catechol derivatives as a function of the corresponding oxidative modifier potentials (E<sub>A</sub>). The arrow represents the E<sub>A</sub> of TEOA [<a href="#B21-catalysts-14-00810" class="html-bibr">21</a>].</p>
Full article ">Figure 14
<p>Rates of photocatalytic generation of hydrogen as a function of illumination time (medium-pressure Hg lamp) over TiO<sub>2</sub> nanoparticles coordinated over dopamine to polymer support (a) and TiO<sub>2</sub> Degussa P25 powder (b) [<a href="#B24-catalysts-14-00810" class="html-bibr">24</a>].</p>
Full article ">Figure 15
<p>(<b>A</b>) EPR spectra of pristine TiO<sub>2</sub> and (<b>B</b>) surface-modified TiO<sub>2</sub> with 4-CP in the dark and upon UV or Vis excitation at 100 K [<a href="#B40-catalysts-14-00810" class="html-bibr">40</a>].</p>
Full article ">Figure 16
<p>Time dependence of ABTS<sup>•+</sup> relative concentration evaluated from double-integrated EPR spectra monitored in the aqueous aerated suspensions of TiO<sub>2</sub>, surface-modified TiO<sub>2</sub> with squaric acid, and ABTS<sup>•+</sup> reference (photocatalyst-free) solution upon excitation with (<b>A</b>) UVA (λ = 365 nm) and (<b>B</b>) Vis light (λ <span class="html-italic">&gt;</span> 420 nm) [<a href="#B67-catalysts-14-00810" class="html-bibr">67</a>].</p>
Full article ">Scheme 1
<p>A schematic presentation for the formation of the TiO<sub>2</sub>-based ICT complexes.</p>
Full article ">Scheme 2
<p>Energy diagram of organic-to-inorganic ICT transition (<b>A</b>) and photoexcitation of dye-sensitized semiconductor (<b>B</b>).</p>
Full article ">Scheme 3
<p>(<b>A</b>) Chelating and (<b>B</b>) bridging coordination of CAT to TiO<sub>2</sub> surface.</p>
Full article ">Scheme 4
<p>The influence of the EDG/EWG on the alignment of energy levels in the ICT complexes.</p>
Full article ">
15 pages, 315 KiB  
Article
Bell vs. Bell: A Ding-Dong Battle over Quantum Incompleteness
by Michael J. W. Hall
Foundations 2024, 4(4), 658-672; https://doi.org/10.3390/foundations4040041 - 8 Nov 2024
Viewed by 324
Abstract
Does determinism (or even the incompleteness of quantum mechanics) follow from locality and perfect correlations? In a 1964 paper, John Bell gave the first demonstration that quantum mechanics is incompatible with local hidden variables. Since then, a vigorous debate has rung out over [...] Read more.
Does determinism (or even the incompleteness of quantum mechanics) follow from locality and perfect correlations? In a 1964 paper, John Bell gave the first demonstration that quantum mechanics is incompatible with local hidden variables. Since then, a vigorous debate has rung out over whether he relied on an assumption of determinism or instead, as he later claimed in a 1981 paper, derived determinism from assumptions of locality and perfect correlation. This paper aims to bring clarity to the debate via simple examples and rigorous results. It is first recalled, via quantum and classical counterexamples, that the weakest statistical form of locality consistent with Bell’s 1964 paper (parameter independence) is insufficient for the derivation of determinism. Attention is then turned to critically assess Bell’s appeal to the Einstein–Rosen–Podolsky (EPR) incompleteness argument to support his claim. It is shown that this argument is itself incomplete, via counterexamples that expose two logical gaps. Closing these gaps via a strong “counterfactual” reality criterion enables a rigorous derivation of both determinism and parameter independence, and in this sense justifies Bell’s claim. Conversely, however, it is noted that whereas the EPR argument requires a weaker “measurement choice” assumption than Bell’s demonstration, it nevertheless leads to a similar incompatibility with quantum predictions rather than quantum incompleteness. Full article
(This article belongs to the Section Physical Sciences)
12 pages, 2875 KiB  
Article
Two New 2p–3d Metal Complexes with a Nitronyl-Nitroxide Ligand Derived from o-Vanillin: Synthesis, Crystals Structures and Magnetic Properties
by Cristian Andrei Spinu, Daniel O. T. A. Martins, Teodora Mocanu, Mihaela Hillebrand, Jean-Pascal Sutter, Floriana Tuna and Marius Andruh
Magnetochemistry 2024, 10(11), 86; https://doi.org/10.3390/magnetochemistry10110086 - 1 Nov 2024
Viewed by 783
Abstract
Two new 2p–3d complexes, (Et3NH)[ML(hfac)2], have been obtained using the nitronyl-nitroxide radical (HL) derived from 2-hydroxy-3-methoxy-5-nitrobenzaldehyde (M = Mn 1; Co 2). The two compounds are isomorphous and their structures consist of anionic mononuclear species, [M(hfac)2 [...] Read more.
Two new 2p–3d complexes, (Et3NH)[ML(hfac)2], have been obtained using the nitronyl-nitroxide radical (HL) derived from 2-hydroxy-3-methoxy-5-nitrobenzaldehyde (M = Mn 1; Co 2). The two compounds are isomorphous and their structures consist of anionic mononuclear species, [M(hfac)2L], M = Mn 1; Co 2, and triethylammonium cations, Et3NH+. The metal ions adopt an octahedral geometry, being coordinated by phenoxido and aminoxyl oxygen atoms from the ligand and four oxygen atoms from the hexafluoroacetylacetonato (hfac) ligand. The cryomagnetic behaviors of the two compounds reveal relatively strong antiferromagnetic M(II)-Rad interactions (JMnRad = −191 cm−1, JCoRad = −166 cm−1 with H = −JSMSRad). The EPR spectra (X- and Q-band) of compound 1 below 70 K show the characteristical features of a S = 2 spin system with zero field splitting terms of D = 0.26 cm−1 and E = 0.031 cm−1. Full article
Show Figures

Figure 1

Figure 1
<p>The X-ray structure of the complex anion in <b>1</b>; the hydrogen and fluorine atoms have been omitted for clarity.</p>
Full article ">Figure 2
<p>(<b>a</b>) Experimental (○) and calculated (<b>―</b>) temperature dependence of <span class="html-italic">χ</span><sub>M</sub><span class="html-italic">T</span> and (<b>b</b>) field dependence of the magnetization at 2 K for compound <b>1</b>.</p>
Full article ">Figure 3
<p>The spin density surfaces (isodensity = 0.04) for the BS (<b>a</b>) and HS (<b>b</b>) spin states at the uB3LYP/TZVP level.</p>
Full article ">Figure 4
<p>Experimental (□) and best fits (full red lines) for (<b>a</b>) <span class="html-italic">χ</span><sub>M</sub><span class="html-italic">T</span> = f(<span class="html-italic">T</span>) and (<b>b</b>) M = f(<span class="html-italic">H/T</span>) behaviors for compound <b>2</b>.</p>
Full article ">Figure 5
<p>X-band (9.4 GHz) CW-EPR spectra of powder samples of <b>1</b> at temperatures between 5.4 and 293 K.</p>
Full article ">Figure 6
<p>Q-band (34 GHz) CW-EPR spectra of powdered <b>1</b> at temperatures between 7 and 293 K.</p>
Full article ">Figure 7
<p>Experimental (black) and simulated (red) Q-band (34 GHz) CW-EPR spectra of powdered compound <b>1</b> at 9 K (g = 2, <span class="html-italic">D</span> = 0.26 cm<sup>−1</sup>, <span class="html-italic">E</span> = 0.031 cm<sup>−1</sup>, and H-Strain (<span class="html-italic">x</span>, <span class="html-italic">y</span>, <span class="html-italic">z</span>) = 1300, 2100, 2000 MHz).</p>
Full article ">Scheme 1
<p>Structure of HL ligand.</p>
Full article ">
8 pages, 562 KiB  
Article
Use of the Micro-Agar Larval Development Test to Differentiate Resistant and Susceptible Cooperia spp. Isolates in Cattle Within the Context of Parasite Population Replacement
by Mariana Elisabet Fuentes, Mercedes Lloberas, Gisele Bernat, Eliana Riva, Milagros Junco and Silvina Fernández
Pathogens 2024, 13(11), 952; https://doi.org/10.3390/pathogens13110952 - 31 Oct 2024
Viewed by 499
Abstract
Gastrointestinal nematode infections are a global concern in grazing cattle production systems, even more so due to the widespread problem of anthelmintic resistance. In response, early anthelmintic resistance detection methods, such as the micro-agar larval development test (MALDT), and parasite management strategies, such [...] Read more.
Gastrointestinal nematode infections are a global concern in grazing cattle production systems, even more so due to the widespread problem of anthelmintic resistance. In response, early anthelmintic resistance detection methods, such as the micro-agar larval development test (MALDT), and parasite management strategies, such as the replacement of resistant parasite populations with susceptible ones, have been developed. This study aimed to characterize ivermectin-susceptible and -resistant isolates of Cooperia spp. using MALDT in the context of a parasite population replacement strategy. Three Cooperia spp. field isolates were evaluated: a susceptible one (Coop-S), a resistant one (Coop-R), and a post-replacement one (Coop-PR). The MALDT was performed in 96-well plates with 12 known concentrations of eprinomectin (EPR) on an agar base. Each test was performed in quadruplicate. Data analysis included nonlinear regression to determine EC50, EC90, and EC99 values, resistance ratios (RRs), and R2. The results showed clear differentiation between the isolates, with RR values of 5.78 and 1.28 for Coop-R and Coop-PR, respectively, compared to Coop-S. The MALDT proved to be a reliable tool for differentiating ivermectin-susceptible from ivermectin-resistant isolates of Cooperia spp., and future evaluations of this test in mixed nematode populations are recommended for routine diagnosis of anthelmintic resistance. Full article
(This article belongs to the Special Issue Pathogenesis, Epidemiology, and Drug Resistance in Nematode Parasites)
Show Figures

Graphical abstract

Graphical abstract
Full article ">Figure 1
<p>Dose–response for <span class="html-italic">Coop-S</span>, <span class="html-italic">Coop-R</span>, and <span class="html-italic">Coop-PR</span> isolates obtained from the MALDT (CI: 95%). EPR: Eprinomectin.</p>
Full article ">
Back to TopTop