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
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

Search Results (15,428)

Search Parameters:
Keywords = reactive oxygen species

Order results
Result details
Results per page
Select all
Export citation of selected articles as:
17 pages, 21053 KiB  
Article
Overexpression of StCDPK13 in Potato Enhances Tolerance to Drought Stress
by Zhenzhen Bi, Simon Dontoro Dekomah, Yihao Wang, Zhuanfang Pu, Xiangdong Wang, Richard Dormatey, Chao Sun, Yuhui Liu, Zhen Liu, Jiangping Bai and Panfeng Yao
Int. J. Mol. Sci. 2024, 25(23), 12620; https://doi.org/10.3390/ijms252312620 (registering DOI) - 24 Nov 2024
Abstract
Calcium-dependent protein kinases (CDPKs), which are activated by transient changes in the Ca2+ concentration in plants, are important for various biological processes, such as growth, development, defense against biotic and abiotic stresses, and others. Mannitol is commonly used as an osmotic regulatory [...] Read more.
Calcium-dependent protein kinases (CDPKs), which are activated by transient changes in the Ca2+ concentration in plants, are important for various biological processes, such as growth, development, defense against biotic and abiotic stresses, and others. Mannitol is commonly used as an osmotic regulatory substance in culture medium or nutrient solutions to create water-deficit conditions. Here, we cloned the potato (Solanum tuberosum L.) StCDPK13 gene and generated stable transgenic StCDPK13-overexpression potato plants. To investigate the potential functions of StCDPK13 in response to drought stress, overexpression-transgenic (OE1, OE2, and OE7) and wild-type (WT) potato seedlings were cultured on MS solid media without or with mannitol, representing the control or drought stress, for 20 days; the elevated mannitol concentrations (150 and 200 mM) were the drought stress conditions. The StCDPK13 gene was consistently expressed in different tissues and was induced by drought stress in both OE and WT plants. The phenotypic traits and an analysis of physiological indicators revealed that the transgenic plants exhibited more tolerance to drought stress than the WT plants. The overexpression lines showed an increased plant height, number of leaves, dry shoot weight, root length, root number, root volume, number of root tips, fresh root weight, and dry root weight under drought stress. In addition, the activities of antioxidant enzymes (CAT, SOD, and POD) and the accumulation of proline and neutral sugars were significantly increased, whereas the levels of malondialdehyde (MDA) and reactive oxygen species (ROS), including hydrogen peroxide (H2O2) and O2•−, were significantly reduced in the OE lines compared to WT plants under drought stress. Moreover, the stomatal aperture of the leaves and the water loss rate in the leaves of the OE lines were significantly reduced under drought stress compared to the WT plants. In addition, the overexpression of StCDPK13 upregulated the expression levels of stress-related genes under drought stress. Collectively, these results indicate that the StCDPK13 gene plays a positive role in drought tolerance by reducing the stomatal aperture, promoting ROS scavenging, and alleviating oxidative damage under drought stress in potatoes. Full article
(This article belongs to the Special Issue Physiology and Molecular Biology of Plant Stress Tolerance)
Show Figures

Figure 1

Figure 1
<p>Sequence analysis of StCDPK13. (<b>A</b>) Amino acid sequence alignment of StCDPK13 with the indicated homologous CDPKs from the indicated plant species (Red line: a variable domain that harbors the ATP binding site, preceding a Ser/Thr protein kinase catalytic domain. Green lines: A CaM-like domain containing four EF hand Ca<sup>2+</sup>-binding motifs). (<b>B</b>) Phylogenetic tree of StCDPK13 and its homologs from different plants species (St: <span class="html-italic">Solanum tuberosum</span>; Le: <span class="html-italic">Lycopersicon esculentum</span>; Os: <span class="html-italic">Oryza sativa</span>; At: <span class="html-italic">Arabidopsis thaliana</span>). The full-length amino acid sequences of CDPKs from four plant species were aligned by DNAMAN and a phylogenetic tree was constructed from the file produced using the Maximum Likelihood method with 1000 bootstrap replicates using MEGA 5.0.</p>
Full article ">Figure 2
<p>Plasmid map and identification of the positive transgenic lines. (<b>A</b>) Schematic representation of the construction of the <span class="html-italic">StCDPK13</span> overexpression vector. The triangular shape indicates the region of the <span class="html-italic">StCDPK13</span> insertion. (<b>B</b>) PCR molecular identification of positive <span class="html-italic">StCDPK13</span> transgenic lines. (<b>C</b>) Expression level of <span class="html-italic">StCDPK13</span> in positive transgenic lines under control condition. All the transgenic lines and WT plants were grown in MS media without mannitol for 20 days, and then the DNA and total RNA were extracted for PCR and qRT-PCR. ** indicate significant differences in the OE line compared with WT, based on Duncan’s Multiple Range Tests (<span class="html-italic">p</span> &lt; 0.01).</p>
Full article ">Figure 3
<p>Expression profile of StCDPK13 in tissues and drought stress in OE lines and WT plants. (<b>A</b>) Tissue-specific expression of StCDPK13 in OE lines and WT plants under control conditions. (<b>B</b>) Relative expression of StCDPK13 in OE lines and WT plants under drought stress (the elevated mannitol concentrations (150 and 200 mM) are the drought stress conditions). All the values represent the means of three independent biological and technical replicates ± SE (standard error). Means with different letters indicate significant differences in OE line compared with WT, based on Duncan’s Multiple Range Tests (<span class="html-italic">p</span> &lt; 0.05).</p>
Full article ">Figure 4
<p>Identification of shoot phenotypic traits in OE lines and WT plants under control and drought conditions (the elevated mannitol concentrations (150 and 200 mM) are the drought stress conditions). (<b>A</b>) Phenotypes of the plants, (<b>B</b>) plant height, (<b>C</b>) number of leaves, (<b>D</b>) number of branches, (<b>E</b>) fresh shoot weight, and (<b>F</b>) dry shoot weight. ** indicate significant differences in the OE line compared with WT, based on Duncan’s Multiple Range Tests (<span class="html-italic">p</span> &lt; 0.01). Bars = 2 cm.</p>
Full article ">Figure 5
<p>Analysis of root phenotypic indicators in OE lines and WT plants under control and drought conditions (the elevated mannitol concentrations (150 and 200 mM) are the drought stress conditions). (<b>A</b>) Morphology of root, (<b>B</b>) root length (cm), (<b>C</b>) number of roots, (<b>D</b>) number of root tips, (<b>E</b>) root volume (cm<sup>3</sup>), (<b>F</b>) fresh root weight, (<b>G</b>) dry root weight. <span class="html-italic">*</span>* indicate significant differences in OE line compared with WT, based on Duncan’s Multiple Range Tests (<span class="html-italic">p</span> &lt; 0.01). Bars = 2.5 cm.</p>
Full article ">Figure 6
<p>Measurement of antioxidant enzyme (CAT, SOD, and POD) activities and the levels of reactive oxygen species (ROS) in the OE lines and WT plants under control and drought conditions (the elevated mannitol concentrations (150 and 200 mM) were the drought stress conditions). (<b>A</b>) CAT activity, (<b>B</b>) SOD activity, (<b>C</b>) POD activity, (<b>D</b>) H<sub>2</sub>O<sub>2</sub> content, (<b>E</b>) DAB staining, and (<b>F</b>) NBT staining. <span class="html-italic">*</span>* indicate significant differences in OE line compared with WT, based on Duncan’s Multiple Range Tests (<span class="html-italic">p</span> &lt; 0.01). Bars = 1 mm.</p>
Full article ">Figure 7
<p>Measurement of malondialdehyde (MDA), proline, and neutral sugars content in the OE lines and WT plants under control and drought conditions (the elevated mannitol concentrations (150 and 200 mM) were the drought stress conditions). (<b>A</b>) MDA content, (<b>B</b>) proline content, and (<b>C</b>) neutral sugars contents. ** indicate significant differences in OE line compared with WT, based on Duncan’s Multiple Range Tests (<span class="html-italic">p</span> &lt; 0.01).</p>
Full article ">Figure 8
<p>Stomatal aperture and water loss in the OE lines and WT plants under control and drought conditions (the elevated mannitol concentrations (150 mM) was the drought stress condition). (<b>A</b>) Observation of stomata in leaves under normal and drought stress conditions, (<b>B</b>) stomatal aperture length, (<b>C</b>) stomatal aperture width, (<b>D</b>) stomatal aperture area, (<b>E</b>) water loss rate of leaves in vitro. ** indicate significant differences in OE line compared with WT, based on Duncan’s Multiple Range Tests (<span class="html-italic">p</span> &lt; 0.01). Bars = 20 μm.</p>
Full article ">Figure 9
<p>Expression levels of stress-related genes in the OE lines and WT plants under control and drought conditions (the elevated mannitol concentrations (150 and 200 mM) were the drought stress conditions). (<b>A</b>) Catalase (PGSC0003DMT400002101), (<b>B</b>) superoxide dismutase (PGSC0003DMT400042937), (<b>C</b>) low molecular weight heat-shock protein 20 (PGSC0003DMT400089913), (<b>D</b>) CAP160 protein (PGSC0003DMT400037083), (<b>E</b>) alcohol dehydrogenase (PGSC0003DMT400 063937), and (<b>F</b>) embryogenesis-abundant protein (PGSC0003DMT400021902). ** indicate significant differences in OE line compared with WT, based on Duncan’s Multiple Range Tests (<span class="html-italic">p</span> &lt; 0.01).</p>
Full article ">
15 pages, 3126 KiB  
Article
Blue Light-Induced Mitochondrial Oxidative Damage Underlay Retinal Pigment Epithelial Cell Apoptosis
by Mohamed Abdouh, Yunxi Chen, Alicia Goyeneche and Miguel N. Burnier
Int. J. Mol. Sci. 2024, 25(23), 12619; https://doi.org/10.3390/ijms252312619 (registering DOI) - 24 Nov 2024
Abstract
Reactive oxygen species (ROS) play a pivotal role in apoptosis. We reported that Blue Light (BL) induced oxidative stress in human retinal pigment epithelial (RPE) cells in vitro and increased drusen deposition and RPE cell apoptosis in human eyes. Here, we investigated the [...] Read more.
Reactive oxygen species (ROS) play a pivotal role in apoptosis. We reported that Blue Light (BL) induced oxidative stress in human retinal pigment epithelial (RPE) cells in vitro and increased drusen deposition and RPE cell apoptosis in human eyes. Here, we investigated the mechanisms underlying BL-induced damage to RPE cells. Cells were exposed to BL with or without the antioxidant N-acetylcysteine. Cells were analyzed for levels of ROS, proliferation, viability, and mitochondria membrane potential (ΔΨM) fluctuation. We performed proteomic analyses to search for differentially expressed proteins. ROS levels increased following RPE cell exposure to BL. While ROS production did not affect RPE cell proliferation, it was accompanied by decreased ΔΨM and increased cell apoptosis due to the caspase cascade activation in a ROS-dependent manner. Proteomic analyses revealed that BL decreased the levels of ROS detoxifying enzymes in exposed cells. We conclude that BL-induced oxidative stress is cytotoxic to RPE cells. These findings bring new insights into the involvement of BL on RPE cell damage and its role in the progression of age-related macular degeneration. The use of antioxidants is an avenue to block or delay BL-mediated RPE cell apoptosis to counteract the disease progression. Full article
(This article belongs to the Special Issue Molecular Mechanisms of Retina Degeneration)
Show Figures

Figure 1

Figure 1
<p>BL-induced oxidative stress in primary RPE cells. Primary human RPE cells were exposed to BL for 30 min. (<b>a</b>) Cells were analyzed for the production of total cellular ROS using the DCF-DA probe. (<b>b</b>) Cells were analyzed for the production of mitochondrial superoxide anion using the MitoSox Red probe. Data are presented as mean ± SD (<span class="html-italic">n</span> = 6 independent experiments each repeated in quadruplicates, ** <span class="html-italic">p</span> ˂ 0.01, *** <span class="html-italic">p</span> ˂ 0.001).</p>
Full article ">Figure 2
<p>BL-induced RPE cells apoptosis in a ROS-dependent manner. Primary human RPE cells were exposed to BL for 30 min. (<b>a</b>) 24 h post-BL exposure, cells were labeled with propidium iodide (PI) and analyzed for their proliferation. Representative cell cycle phase distribution histograms are shown where the first peak (at 50 k) corresponds to cells in the G1 phase, the second peak (at 100 k) corresponds to cells in the G2/M, and area in between the peaks represents cells in the S phase. The graph displays the percentages of cells in these different phases of cell cycle as analyzed using FlowJo software (v10.10). (<b>b</b>) 6 h post-BL exposure, cells were labeled with Annexin V and PI and analyzed for the percentages of apoptotic cells by flow cytometry. Representative Annexin V/PI density plots are shown that display live cells (left-lower quadrants), apoptotic cells (right-lower quadrant), secondary apoptotic cells (right-upper quadrant) and necrotic cells (left-upper quandrant). The numbers represent the percentages of cells in the respective quadrants as analyzed using FlowJo software. The graph displays means of the percentages of primary apoptotic, secondary apoptotic and necrotic cells. Data are presented as mean ± SD (<span class="html-italic">n</span> = 6 independent experiments, * <span class="html-italic">p</span> &lt; 0.05, ** <span class="html-italic">p</span> ˂ 0.01).</p>
Full article ">Figure 3
<p>BL reduced the mitochondrial membrane potential in a ROS-dependent manner. Primary human RPE cells were exposed to BL for 30 min, and cells were stained with JC-1 probe. Fluorescence of J-aggregates and J-monomers were measured. Data are expressed as the ratio between the 2 measures and are presented as mean ± SD (<span class="html-italic">n</span> = 6 independent experiments each repeated in quadruplicates, * <span class="html-italic">p</span> &lt; 0.05, ** <span class="html-italic">p</span> ˂ 0.01.</p>
Full article ">Figure 4
<p>BL increased caspase cascade activation in ROS-dependent manner. Primary human RPE cells were exposed to BL for 30 min. (<b>a</b>) Proteins extracts were analyzed by immunoblot for the activation of Caspase 9. β-actin and red ponceau staining were used as calibrators for proteins loading. The graph shows the levels of caspases activation in the corresponding samples. Data are expressed as the densitometer values relative to the value in control sample set at 1. (<b>b</b>) Cells were loaded with CellEvent Caspase 3/7 Green. Pictures were acquired using a LSM780 confocal microscope. The graph displays the levels of caspases 3 and 7 activation in the corresponding samples. Data are expressed as mean fluorescence intensity (MFI) measured in an Infinite M200 Pro microplate reader relative to the value in control sample set at 1. Data are presented as mean ± SD (<span class="html-italic">n</span> = 3–5 independent RPE cells samples). * <span class="html-italic">p</span> &lt; 0.05, ** <span class="html-italic">p</span> &lt; 0.01.</p>
Full article ">Figure 5
<p>BL decreased the expression levels of ROS detoxifying enzymes. Primary human RPE cells were exposed to BL for 30 min. (<b>a</b>) Venn diagram analyses. Samples datasets were compared for shared proteins between non-exposed cells (CTL), between BL-exposed cells (BL), and between exposed and non-exposed cells. In the insert, CTL and BL-exposed RPE cells shared 2369 proteins, while 288 and 153 proteins were exclusively present in CTL and BL-exposed RPE cells, respectively (see <a href="#app1-ijms-25-12619" class="html-app">Supplementary Table S1</a> for the full list of proteins). (<b>b</b>) Volcano plot representation of proteins significantly and differentially expressed between CTL and BL-exposed RPE cells (See <a href="#app1-ijms-25-12619" class="html-app">Supplementary Tables S2 and S3</a> for the respective protein lists). For statistical analyses, we set the analysis for a T-Test with a significant level at 0.05). ■: Significant, ●: Nonsignificant. (<b>c</b>) Table showing a short list of ROS detoxifying enzymes which expression is decreased in BL-exposed cells. In Orange lines are shown differentially and significantly expressed proteins. In Green are shown proteins with decreased expression levels in BL-exposed cells but not reaching the statistical significance (See <a href="#app1-ijms-25-12619" class="html-app">Supplementary Table S4</a> for the full list of identified proteins). Data are obtained from the analysis of 5 RPE cell lines derived from 5 different eye donors.</p>
Full article ">Figure 6
<p>Gene ontology classification of proteomic data for differentially expressed proteins in primary RPE cells exposed or not to BL. The most enriched categories in biological processes, as analyzed by the DAVID bioinformatics platform, are shown. Data were collected from protein preparations obtained from five patient-derived RPE cells.</p>
Full article ">Figure 7
<p>BL-induced RPE cell apoptosis model. BL-mediated oxidative stress triggered mitochondrial damage and concomitant caspase cascade activation. These effects induced RPE cell apoptosis, that could be reversed by antioxidants (i.e., N-acetyl cysteine; NAC).</p>
Full article ">
17 pages, 2174 KiB  
Article
Lactate Dehydrogenase-B Oxidation and Inhibition by Singlet Oxygen and Hypochlorous Acid
by Lisa M. Landino and Emily E. Lessard
Oxygen 2024, 4(4), 432-448; https://doi.org/10.3390/oxygen4040027 (registering DOI) - 24 Nov 2024
Abstract
Alterations in cellular energy metabolism are a hallmark of cancer and lactate dehydrogenase (LDH) enzymes are overexpressed in many cancers regardless of sufficient oxygen and functional mitochondria. Further, L-lactate plays signaling roles in multiple cell types. We evaluated the effect of singlet oxygen [...] Read more.
Alterations in cellular energy metabolism are a hallmark of cancer and lactate dehydrogenase (LDH) enzymes are overexpressed in many cancers regardless of sufficient oxygen and functional mitochondria. Further, L-lactate plays signaling roles in multiple cell types. We evaluated the effect of singlet oxygen and hypochlorous acid (HOCl) on pig heart LDH-B, which shares 97% homology with human LDH-B. Singlet oxygen was generated photochemically using methylene blue or the chlorophyll metabolites, pheophorbide A and chlorin e6. Singlet oxygen induced protein crosslinks observed by SDS-PAGE under reducing conditions and inhibited LDH-B activity. Ascorbate, hydrocaffeic acid, glutathione and sodium azide were employed as singlet oxygen scavengers and shown to protect LDH-B. Using fluorescein-modified maleimide, no changes in cysteine availability as a result of singlet oxygen damage were observed. This was in contrast to HOCl, which induced the formation of disulfides between LDH-B subunits, thereby decreasing LDH-B labeling with fluorescein. HOCl oxidation inhibited LDH-B activity; however, disulfide reduction did not restore it. LDH-B cysteines were resistant to millimolar H2O2, chloramines and Angeli’s salt. In the absence of pyruvate, LDH-B enhanced NADH oxidation in a chain reaction initiated by singlet oxygen that resulted in H2O2 formation. Once damaged by either singlet oxygen or HOCl, NADH oxidation by LDH-B was impaired. Full article
Show Figures

Figure 1

Figure 1
<p><sup>1</sup>O<sub>2</sub> oxidation of LDH-B. LDH-B (10 µM) samples containing 1 µM MB were irradiated for 0–90 s. (<b>A</b>) Samples were analyzed by SDS-PAGE under reducing conditions on 10% gels and stained with Coomassie blue. (<b>B</b>) LDH-B samples were prepared as in (<b>A</b>). After irradiation, samples were diluted 1:10 with 20 mM Tris pH 8.8. Kinetic assays (200 mL) contained 5 µL 1:10 LDH-B, 4 mM lithium lactate and 2 mM NAD<sup>+</sup> in 20 mM Tris pH 8.8. Data shown is the average of at least three independent experiments performed in duplicate. NAD<sup>+</sup> reduction was monitored at 340 nm in a 96-well plate at 30 °C for 4 min. (<b>C</b>) LDH-B samples prepared as in A were analyzed by native gel electrophoresis on 0.8% agarose gels. Activity stain contained 0.75 mM NAD<sup>+</sup>, 25 mM lithium lactate, 8–10 mg NBT, and 1–2 mg PMS in 100 mM Tris pH 8.6.</p>
Full article ">Figure 2
<p>Cysteine labeling of LDH-B with fluorescein. LDH-B (10 μM) samples containing 1 μM MB were irradiated for 0–120 s. Samples were treated with 500 μM M5F for 30 min at 37 °C and analyzed by SDS-PAGE under reducing conditions on 10% gels. Fluorescent images were captured using a Bio-Rad Chemi-doc XRS imaging system (<b>left</b>). After imaging, gels were stained with Coomassie blue (<b>right</b>).</p>
Full article ">Figure 3
<p>LDH-B protection from <sup>1</sup>O<sub>2</sub> by ascorbate and HCA. (<b>A</b>) LDH-B (10 μM) samples containing 1 μM MB were irradiated for 120 s. Ascorbate (100 or 250 μM) was added prior to irradiation. Samples were analyzed by SDS-PAGE under reducing conditions on 10% gels and stained with Coomassie blue. (<b>B</b>) LDH-B (10 μM) samples containing 1 μM MB or 15 μM pheoA were irradiated for 120 s. Ascorbate (250 or 500 μM) was added prior to irradiation. “D” indicates dark. Samples were analyzed by native gel electrophoresis on 0.8% agarose gels. Activity stain contained 0.75 mM NAD<sup>+</sup>, 25 mM lithium lactate, 8–10 mg NBT, and 1–2 mg PMS in 100 mM Tris pH 8.6. (<b>C</b>) LDH-B samples contained 1 μM MB, 2.5 μM chlorin e6, or 15 μM pheoA. Either 0.5 mM ascorbate or 1.5 mM HCA was added prior to irradiation. LDH-B activity was measured as in <a href="#oxygen-04-00027-f001" class="html-fig">Figure 1</a>B.</p>
Full article ">Figure 4
<p>LDH oxidation and inhibition by HOCl. LDH-B (10 μM) in 10 mM PB pH 7.4 was treated with up to 150 μM HOCl for 30 min at 30 °C. (<b>A</b>) Samples were analyzed by native gel electrophoresis on 0.8% agarose gels. Activity stain contained 0.75 mM NAD<sup>+</sup>, 25 mM lithium lactate, 8–10 mg NBT, and 1–2 mg PMS in 100 mM Tris pH 8.6. (<b>B</b>) Samples were analyzed by SDS-PAGE under reducing and nonreducing conditions (±βME). The gel was stained with Coomassie Blue. (<b>C</b>) Kinetic assays (200 μL) contained 5 μL 1:10 LDH-B, 4 mM lithium lactate, and 2 mM NAD<sup>+</sup> in 20 mM Tris pH 8.8. Data shown is the average of at least three independent experiments performed in duplicate. NAD<sup>+</sup> reduction was monitored at 340 nm in a 96-well plate at 30 °C for 4 min.</p>
Full article ">Figure 5
<p>Cysteine labeling after HOCl oxidation. LDH-B (10 μM) in 10 mM PB pH 7.4 was treated with up to 150 μM HOCl for 30 min at 30 °C. Excess HOCl was quenched with 0.2 mM S-methyl-cys. (<b>A</b>) Samples were treated with 500 μM M5F for 30 min at 37 °C and analyzed by SDS-PAGE under nonreducing conditions on 10% gels. Fluorescent images were captured using a Bio-Rad Chemi-doc XRS imaging system (<b>left</b>). After imaging, gels were stained with Coomassie blue (<b>right</b>). (<b>B</b>) Samples were prepared as in (<b>A</b>) except reactions were 60 μL. M5F-labeled LDH-B was precipitated with 80% ethanol. Fluorescein was quantitated at 495 nm relative to a fluorescein standard curve. These data are the average of two independent trials performed in duplicate.</p>
Full article ">Figure 6
<p>Enhanced NADH oxidation by LDH-A and LDH-B: activation by <sup>1</sup>O<sub>2</sub>. ll reactions (100 μL) contained 200 μM NADH in 50 mM PB pH 7.1 in a 96-well plate. <sup>1</sup>O<sub>2</sub> formation was initiated with 2.5 μM MB and 10 s of red light. Additions included LDH-A or LDH-B (10 μM). Absorbance at 340 nm was measured prior to and immediately after light exposure. These data represent the average of two independent experiments performed in triplicate.</p>
Full article ">
13 pages, 3601 KiB  
Article
Rapid Degradation of Bisphenol F Using Magnetically Separable Bimetallic Biochar Composite Activated by Peroxymonosulfate
by Hui Liang, Ruijuan Li, Tongjin Liu, Rumei Li, Yuxiao Zhu and Feng Fang
Molecules 2024, 29(23), 5545; https://doi.org/10.3390/molecules29235545 (registering DOI) - 24 Nov 2024
Abstract
Peroxymonosulfate (PMS)-based advanced oxidation processes have shown potential for the removal of organic contaminants; however, the preparation of catalysts with high degradation efficiencies and rapid reaction rates remains a challenge. In this study, we have successfully synthesized CoFe bimetallic modified corn cob-derived biochar [...] Read more.
Peroxymonosulfate (PMS)-based advanced oxidation processes have shown potential for the removal of organic contaminants; however, the preparation of catalysts with high degradation efficiencies and rapid reaction rates remains a challenge. In this study, we have successfully synthesized CoFe bimetallic modified corn cob-derived biochar (CoFe/BC) for the activation of PMS, achieving the rapid and efficient degradation of bisphenol F (BPF). The synthesized CoFe/BC catalyst demonstrated excellent catalytic performance, achieving over 99% removal within 3 min and exhibiting a removal rate of 90.0% after five cycles. This could be attributed to the cyclic transformation of Co and Fe, which sustained rapid PMS activation for BPF degradation, and Co0/Fe0 played a significant role in the cyclic transformation. Furthermore, the electron paramagnetic resonance tests confirmed that •SO4 and •OH were the primary reactive oxygen species, while •O2 played a minor role in BPF degradation. This study highlights the high degradation efficiency, rapid reaction rate, excellent magnetic separation properties, and exceptional reusability of CoFe/BC catalysts for BPF removal, providing valuable insights for practical wastewater treatment. Full article
Show Figures

Figure 1

Figure 1
<p>(<b>a</b>–<b>d</b>) SEM images, (<b>e</b>–<b>i</b>) EDS mappings (C, N, O, Co, Fe), and (<b>j</b>–<b>l</b>) HRTEM images of CoFe/BC catalyst.</p>
Full article ">Figure 2
<p>(<b>a</b>) N<sub>2</sub> adsorption–desorption isothermal curve and (<b>b</b>) pore-size distribution curve of CoFe/BC catalyst; (<b>c</b>) XRD patterns of BC, CoFe<sub>2</sub>O<sub>4</sub>/CoFe/BC (Co:Fe = 1:9), CoFe/BC37 (Co:Fe = 3:7), CoFe/BC (Co:Fe = 5:5), CoFe/BC73 (Co:Fe = 7:3), and CoC<sub>x</sub>/Co<sub>7</sub>Fe<sub>3</sub>/BC (Co:Fe = 9:1) catalysts. The proportion of Co and Fe mentioned in this study is the molar ratio, and further information on the catalysts is available in <a href="#sec3dot2-molecules-29-05545" class="html-sec">Section 3.2</a>. Synthesis.</p>
Full article ">Figure 3
<p>(<b>a</b>) XPS survey, (<b>b</b>) C 1s, (<b>c</b>) O 1s, (<b>d</b>) N 1s, (<b>e</b>) Co 2p, and (<b>f</b>) Fe 2p of CoFe/BC catalyst.</p>
Full article ">Figure 4
<p>(<b>a</b>) Degradation efficiencies of BPF degradation on different catalysts. Influence of various conditions on BPF degradation in CoFe/BC/PMS system: (<b>b</b>) catalyst amount, (<b>c</b>) PMS concentration, and (<b>d</b>) reaction temperature. Experimental conditions: Catalyst = 0.05 g, BPF = 10 mg/L, PMS = 100 mg/L, T = 25 °C.</p>
Full article ">Figure 5
<p>EPR spectra of (<b>a</b>) DMPO-X, (<b>b</b>) DMPO-•O<sub>2</sub><sup>−</sup>, and (<b>c</b>) TEMP-<sup>1</sup>O<sub>2</sub>; (<b>d</b>) the proposed mechanism for BPF degradation over the CoFe/BC/PMS system.</p>
Full article ">Figure 6
<p>(<b>a</b>) Recyclability of CoFe/BC for BPF removal and (<b>b</b>) corresponding k value; (<b>c</b>) XRD spectrum of CoFe/BC after reaction; (<b>d</b>) degradation performance of CoFe/BC catalyst on different organic pollutants. Experimental conditions: Catalyst = 0.05 g, BPF = 10 mg/L, BPA = 10 mg/L, phenol = 10 mg/L, MB = 50 mg/L, MG = 50 mg/L, RhB = 50 mg/L, PMS = 100 mg/L, T = 25 °C.</p>
Full article ">
19 pages, 11411 KiB  
Article
Ginsenoside Rg1 Alleviates Blood–Milk Barrier Disruption in Subclinical Bovine Mastitis by Regulating Oxidative Stress-Induced Excessive Autophagy
by Shanshan Yang, Zihao Fang, Hongwei Duan, Weitao Dong and Longfei Xiao
Antioxidants 2024, 13(12), 1446; https://doi.org/10.3390/antiox13121446 (registering DOI) - 24 Nov 2024
Viewed by 72
Abstract
As a critical disease usually infected by Staphylococcus aureus, with a worldwide effect on dairy animals, subclinical mastitis is characterized by persistence and treatment resistance. During mastitis, the blood–milk barrier (BMB)’s integrity is impaired, resulting in pathogen invasion and milk quality decline. [...] Read more.
As a critical disease usually infected by Staphylococcus aureus, with a worldwide effect on dairy animals, subclinical mastitis is characterized by persistence and treatment resistance. During mastitis, the blood–milk barrier (BMB)’s integrity is impaired, resulting in pathogen invasion and milk quality decline. In this study, it was found that ginsenoside Rg1 (Rg1), a natural anti-inflammatory and antioxidant compound derived from ginseng, inhibited the onset of tight junction (TJ) dysfunction and ameliorated lipoteichoic acid (LTA)-induced BMB disruption inside and outside the organisms. According to subsequent mechanistic studies, Rg1 inhibited excessive autophagy and inactivated the NLRP3 inflammasome by blockading ROS generation, thereby alleviating TJ dysfunction. Peroxisome proliferator-activated receptor gamma (PPARγ) was identified as a potential target of Rg1 by means of molecular docking plus network pharmacology analysis. Furthermore, it was demonstrated that Rg1 inhibited the oxidative stress levels by activating PPARγ, and regulating the upstream autophagy-related AMPK/mTOR signaling pathway, thus decreasing excessive in vivo and in vitro autophagy. The ROS/autophagy/NLRP3 inflammasome axis was identified as a promising target for treating subclinical bovine mastitis in this study. In conclusion, Rg1 is proven to alleviate BMB disruption by activating PPARγ to inhibit oxidative stress and subsequent excessive autophagy in the case of subclinical bovine mastitis. Full article
Show Figures

Figure 1

Figure 1
<p>Excessive autophagy and BMB disruption in mammary gland with subclinical bovine mastitis. (<b>A</b>) HE staining for histological changes in mammary glands. Original magnification = 200×, scale bar = 100 μm (n = 3). (<b>B</b>) Ultrastructure of autophagy and TJs in bMECs observed by TEM (n = 3). Original magnification = 10,000×, scale bar = 500 nm. The red arrowheads indicated TJs, the red arrows indicated autophagic structures. (<b>C</b>) Typical immunofluorescence images for localization of LC3 (green) in the mammary glands. Original magnification = 200×, scale bar = 100 μm (n = 3). (<b>D</b>) Typical immunofluorescence images for localization of claudin-1, ZO-1, and occludin (green) in the mammary glands. Original magnification = 200×, scale bar = 100 μm (n = 3). (<b>E</b>) The relative protein expressions of claudin-1, beclin-1, occludin, ZO-1 and LC3 in mammary glands detected by Western blotting (n = 3). The loading control is set as β-actin, and mean ± SD is used for value expression. ** <span class="html-italic">p</span> &lt; 0.01 vs. the healthy group.</p>
Full article ">Figure 2
<p>Rg1 inhibited LTA-induced excessive autophagy to alleviate TJ dysfunction. (<b>A</b>) MAC-T cells processed for 24 h by LTA at 0.1, 1 and 10 μg/mL. Western blotting for proteins associated with TJs (ZO-1, occludin, and claudin-1). (<b>B</b>) MAC-T cells undergoing 24 h LTA (10 μg/mL) and/or Rg1 (at required concentrations) treatment. Western blotting for proteins in relation to autophagy and TJs. (<b>C</b>) MAC-T cells subjected to 24 h of LTA (10 μg/mL) and/or Rg1 (20 μM) processing. After the transfection of MAC-T cells with adenovirus plus autolysosome quantitation via mCherry-GFP-LC3, autophagy was visually observable. Original magnification = 400×. (<b>D</b>) MAC-T cells under 24 h of treatment by LTA (10 μg/mL), Rg1 (20 μM), and/or CQ (50 μM). Western blotting for proteins correlated with autophagy and TJs. The mean ± SD is used for value presentation; compared with the control group, * <span class="html-italic">p</span> &lt; 0.05, ** <span class="html-italic">p</span> &lt; 0.01; compared to the LTA treatment group, ## <span class="html-italic">p</span> &lt; 0.01.</p>
Full article ">Figure 3
<p>Rg1 inhibited LTA-induced TJ dysfunction by blocking the ROS/autophagy/NLRP3 inflammasome axis. (<b>A</b>) SOD, MDA, and CAT content in subclinical bovine mastitis and healthy mammary tissues detected (n = 4). MAC-T cells that had undergone treatment by LTA (10 μg/mL), Rg1 (20 μM), and/or NAC (5 mM) for 24 h. (<b>B</b>) DCFH-DA probe for ROS level evaluation in MAC-T cells (original magnification = 400×, scale bar = 50 μm.). (<b>C</b>) Quantification of MAC-T cells for ROS content (n = 6). (<b>D</b>) Calculated GSH level in MAC-T cells (n = 6). (<b>E</b>) Western blotting for ZO-1, occludin, and claudin-1 (proteins related to TJs). (<b>F</b>) Relative protein expressions concerning ASC, IL-1β, NLRP3, and caspase-1 in mammary glands determined by means of Western blotting. (<b>G</b>) Western blotting for the TJ-related and NLRP3 inflammasome-related proteins. (<b>H</b>) MAC-T cells subjected to LTA (10 μg/mL), Rg1 (20 μM), and/or MCC950 (10 μM) processing for 24 h. Western blotting for the NLRP3 inflammasome-related and TJ-related proteins. (<b>I</b>) MAC-T cells under LTA (10 μg/mL), Rg1 (20 μM), and/or CQ (50 μM) treatment for 24 h. Western blotting for NLRP3 and IL-1β proteins. (<b>J</b>) Twenty-four hour treatment for MAC-T cells using LTA (10 μg/mL), and/or MCC950 (10 μM). Western blotting for beclin-1 and LC3 proteins. The mean ± SD is the expression format for values; compared with the control group, * <span class="html-italic">p</span> &lt; 0.05, ** <span class="html-italic">p</span> &lt; 0.01; compared to the LTA treatment group, ## <span class="html-italic">p</span> &lt; 0.01.</p>
Full article ">Figure 4
<p>Rg1 inhibited LTA-induced excessive autophagy by regulating the ROS/AMPK/mTOR signaling pathway. (<b>A</b>) Relative protein expressions of mTOR, AMPK, p-AMPK, and p-mTOR in mammary glands determined by means of Western blotting (n = 3). (<b>B</b>) Rg1 (20 μM), LTA (10 μg/mL), and/or NAC (5 mM) added to treat MAC-T cells for 24 h. Western blotting for AMPK, mTOR, p-AMPK, and p-mTOR protein. (<b>C</b>) MAC-T cells subjected to 24 h of LTA (10 μg/mL), Rg1 (20 μM), and/or CC (10 μM) treatment. Protein expressions of AMPK, mTOR, beclin-1, p-mTOR, LC3, and p-AMPK detected by Western blotting. The mean ± SD is adopted to describe values; compared to the control group, ** <span class="html-italic">p</span> &lt; 0.01; compared to the LTA treatment group, ## <span class="html-italic">p</span> &lt; 0.01.</p>
Full article ">Figure 5
<p>Repressed oxidative stress plus raised PPARγ are necessary for Rg1 to inhibit LTA-induced excessive autophagy and TJ recovery. (<b>A</b>) Number of intersecting NAFLD, mastitis and R&amp;M target genes. (<b>B</b>) Relative expressions of PPARα, PPARβ/δ, and PPARγ proteins in mammary glands determined by means of Western blotting (n = 3). (<b>C</b>) MAC-T cells subjected to 24 h of LTA (10 μg/mL) and/or Rg1 (20 μM) treatment. Western blotting for PPARα, PPARβ/δ, and PPARγ protein. (<b>D</b>) MAC-T cells undergoing 24 h disposal by LTA (10 μg/mL), Rg1 (20 μM), and/or GW6471 (10 μM). Western blotting for PPARα and TJ-related proteins. (<b>E</b>) MAC-T cells under 24 h of treatment using LTA (10 μg/mL), Rg1 (20 μM), and/or T0070907 (10 μM). Western blotting for PPARγ and TJ-related proteins. (<b>F</b>) Docking of PPARγ with Rg1. Rg1 interacts with 2 amino acids, THR 440 and GLN 437, adjacent to the PPARγ active site, forming crucial binding forces in the periphery of the active site. (<b>G</b>) ROS level quantification in MAC-T cells (n = 6). (<b>H</b>) GSH level in MAC-T cells was calculated (n = 6). (<b>I</b>) Western blotting for autophagy-related plus NLRP3 inflammasome-related proteins like AMPK, mTOR, p-AMPK, and p-mTOR. The mean ± SD is used to express values; compared with the control group, * <span class="html-italic">p</span> &lt; 0.05, ** <span class="html-italic">p</span> &lt; 0.01; compared with the LTA treatment group, ## <span class="html-italic">p</span> &lt; 0.01.</p>
Full article ">Figure 6
<p>Rg1 alleviated BMB disruption through regulating autophagy signaling in organisms. Five groups were set up for the mice: control group (no treatment with LTA), Rg1 (no LTA, 200 mg/kg) group, model group (LTA injection, no treatment), Rg1 (200 mg/kg) + LTA treatment group, and rosiglitazone (10 mg/kg) + LTA treatment group. (<b>A</b>) Hematoxylin-eosin (HE) staining for general examination of mammary glands (original magnification = 200×, scale bar = 100 μm), together with immunofluorescence images for LC3, ZO-1, occludin and claudin-1 (original magnification = 400×, scale bar = 50 μm). (<b>B</b>) Examined variations in SOD, MDA, and CAT content. (<b>C</b>,<b>D</b>) Western blotting for determining the relative protein expressions of AMPK, PPARγ, p-AMPK, mTOR, p-mTOR, autophagy-related, NLRP3 inflammasome-related and TJ-related proteins. The mean ± SD as the value format and β-actin as the loading control. Compared with the control group, ** <span class="html-italic">p</span> &lt; 0.01; Compared with the LTA treatment group, and ## <span class="html-italic">p</span> &lt; 0.01.</p>
Full article ">Figure 7
<p>Proposed mechanism of Rg1 in alleviating BMB disruption in cases of subclinical bovine mastitis by regulating oxidative stress-induced excessive autophagy.</p>
Full article ">
18 pages, 4346 KiB  
Article
Reactive Oxygen Species-Sensitive Nanophotosensitizers Composed of Buthionine Sulfoximine-Conjugated Chitosan Oligosaccharide for Enhanced Photodynamic Treatment of Cancer Cells
by Hyo Young Lee, Ji-Sun Park, Taeyu Grace Kim, Taeyeon Kim, Do Hoon Kim, Jejung Yun and Young-IL Jeong
Int. J. Mol. Sci. 2024, 25(23), 12609; https://doi.org/10.3390/ijms252312609 (registering DOI) - 24 Nov 2024
Viewed by 107
Abstract
The efficacy of photodynamic therapy (PDT) based on traditional photosensitizers is generally limited by the cellular redox homeostasis system due to the reactive oxygen species (ROS) scavenging effect of glutathione (GSH). In this study, buthionine sulfoximine (BSO), a GSH inhibitor, was conjugated with [...] Read more.
The efficacy of photodynamic therapy (PDT) based on traditional photosensitizers is generally limited by the cellular redox homeostasis system due to the reactive oxygen species (ROS) scavenging effect of glutathione (GSH). In this study, buthionine sulfoximine (BSO), a GSH inhibitor, was conjugated with the amine group of chitosan oligosaccharide (COS) using a thioketal linker (COSthBSO) to liberate BSO and chlorine e6 (Ce6) under oxidative stress, and then, Ce6-COSthBSO NP (Ce6-COSthBSO NP), fabricated by a dialysis procedure, showed an accelerated release rate of BSO and Ce6 by the addition of hydrogen peroxide, indicating that nanophotosensitizers have ROS sensitivity. In the in vitro cell culture study using HCT116 colon carcinoma cells, a combination of BSO and Ce6 efficiently suppressed the intracellular GSH and increased ROS production compared to the sole treatment of Ce6. In particular, Ce6-COSthBSO NP showed higher efficacy in the suppression of GSH levels and ROS production compared to the free Ce6 and Ce6/BSO combination. These results were due to the fact that Ce6-COSthBSO NP was efficiently delivered to the intracellular region, suppressed intracellular GSH levels, and elevated ROS levels. The in vivo animal tumor xenograft study demonstrated Ce6-COSthBSO NP being efficiently delivered to the tumor tissue, i.e., the fluorescence intensity in the tumor tissue was higher than those of other organs. The combination of Ce6 and BSO efficiently suppressed tumor growth compared to the sole treatment of Ce6, indicating that BSO might efficiently suppress GSH levels and increase ROS levels in the tumor microenvironment. Specifically, Ce6-COSthBSO NP showed the strongest performance in inhibition of tumor growth than those of Ce6 or the CE6/BSO combination, indicating that they were efficiently delivered to tumor tissue, increased ROS levels, and then efficiently inhibited tumor growth. We suggest that COSthBSO nanophotosensitizers are promising candidates for PDT treatment of cancer cells. Full article
(This article belongs to the Special Issue Photodynamic Therapy and Photodetection, 2nd Edition)
Show Figures

Figure 1

Figure 1
<p>Synthesis scheme (<b>a</b>) and <sup>1</sup>H NMR spectra (<b>b</b>) of COSthBSO conjugates.</p>
Full article ">Figure 2
<p>Particle size distribution (<b>a</b>) and transmission electron microscope image (<b>b</b>) of Ce6-incorporated COSthBSO nanophotosensitizers.</p>
Full article ">Figure 3
<p>Ce6 release from Ce6-COSthBSO NP. The effect of Ce6 contents (<b>a</b>), H<sub>2</sub>O<sub>2</sub> (<b>b</b>), and GSH (<b>c</b>).</p>
Full article ">Figure 4
<p>Intrinsic cytotoxicity of BSO (<b>a</b>), COSthBSO conjugates (<b>b</b>), Ce6 (<b>c</b>), and Ce6-COSthBSO NP (<b>d</b>). Ce6 contents of nanophotosensitizers was 9.2% (<span class="html-italic">w</span>/<span class="html-italic">w</span>).</p>
Full article ">Figure 5
<p>Ce6 uptake of Ce6 and Ce6-COSthBSO NP. (<b>A</b>) Ce6 uptake ratio. (<b>B</b>) Fluorescence microscopic images. (<b>a</b>) Ce6, 1 μg/mL; (<b>b</b>) Ce6, 5 μg/mL; (<b>c</b>) Ce6-COSthBSO NP (NP), 1 μg/mL as a Ce6 concentration; (<b>d</b>) Ce6-COSthBSO NP (NP), 5 μg/mL as a Ce6 concentration. Magnification 200×.</p>
Full article ">Figure 6
<p>(<b>a</b>) ROS generation and (<b>b</b>) phototoxicity of Ce6 and Ce6-COSthBSO NP. For ROS generation and PDT treatment, cells were irradiated with an expanded homogeneous beam at 664 nm with a 2.0 J/cm<sup>2</sup> light dose.</p>
Full article ">Figure 7
<p>The effects of Ce6, BSO, COS, COSthBSO, and Ce6-COSthBSO NP (COSthBSO-Ce6 NP) on intracellular GSH contents (<b>a</b>) and ROS generation (<b>b</b>) in HCT116 cells with or without light irradiation. The light dose was 2.0 J/cm<sup>2</sup> at 664 nm. Ce6 and BSO concentrations were 1 μg/mL and 2 μg/mL, respectively. COS and COSthBSO concentrations were 10 μg/mL and 10 μg/mL. For Ce6-COSthBSO NP (NP), Ce6 concentration was adjusted to 1 μg/mL. For control treatment, phosphate-buffered saline (PBS, pH 7.4, 0.01 M) was treated. Results are expressed as mean ± SD.</p>
Full article ">Figure 8
<p>Fluorescence images of intracellular ROS in HCT116 cells. The light dose was 2.0 J/cm<sup>2</sup> at 664 nm. Ce6 and BSO concentrations were 1 μg/mL and 2 μg/mL, respectively. COSthBSO concentrations were 10 μg/mL. For Ce6-COSthBSO NP, Ce6 concentration was adjusted to 1 μg/mL. For control treatment, phosphate-buffered saline (PBS, pH 7.4, 0.01 M) was treated. (<b>1</b>) Control; (<b>2</b>) free BSO; (<b>3</b>) free Ce6; (<b>4</b>) free Ce6 + free BSO; (<b>5</b>) COSthBSO nanoparticles; and (<b>6</b>) Ce6-incorporated COSthBSO nanophotosensitizers. Magnification, 400×.</p>
Full article ">Figure 9
<p>(<b>a</b>) Animal tumor imaging of HCT116-bearing tumor xenograft model. The BALb/C nude mouse was used. (<b>b</b>) Relative fluorescence intensity at each organ. Results are expressed as average ± SD of three mice. (<b>c</b>) The effects of PBS, BSO, COSthBSO conjugates (as empty nanoparticles), and free Ce6 or Ce6-incorporated COSthBSO nanophotosensitizers on the PDT of HCT116 tumor. (Ce6 dose = 10 mg/kg). *, ***: <span class="html-italic">p</span> &lt; 0.01; **: <span class="html-italic">p</span> &lt; 0.001.</p>
Full article ">Figure 10
<p>Schematic illustration of Ce6-incorporated COSthBSO nanophotosensitizers (Ce6-COSthBSO NP) against cancer cells.</p>
Full article ">
18 pages, 1939 KiB  
Article
Root-Knot Nematode Early Infection Suppresses Immune Response and Elicits the Antioxidant System in Tomato
by Sergio Molinari, Anna Carla Farano and Paola Leonetti
Int. J. Mol. Sci. 2024, 25(23), 12602; https://doi.org/10.3390/ijms252312602 (registering DOI) - 23 Nov 2024
Viewed by 364
Abstract
The immune response in plants is regulated by several phytohormones and involves the overexpression of defense genes, including the pathogenesis-related (PR-) genes. The data reported in this paper indicate that nematodes can suppress the immune response by inhibiting the expression of [...] Read more.
The immune response in plants is regulated by several phytohormones and involves the overexpression of defense genes, including the pathogenesis-related (PR-) genes. The data reported in this paper indicate that nematodes can suppress the immune response by inhibiting the expression of defense genes. Transcripts from nine defense genes were detected by qRT-PCR in the roots of tomato plants at three and seven days post-inoculation (dpi) with living juveniles (J2s) of Meloidogyne incognita (root-knot nematodes, RKNs). All the salicylic acid (SA)-responsive genes tested (PR-1, PR-2, PR-4b, PR-5) were down-regulated in response to nematode infection. On the contrary, the expression of jasmonic acid (JA)-responsive genes, including ACO (encoding the enzyme 1-aminocyclopropane-1-carboxylic acid oxidase, which catalyzes the last step of ethylene (ET) biosynthesis) and JERF3 (Jasmonate Ethylene Response Factor 3), was unaffected by the infection. Conversely, the effect of nematode attack on the activities of the defense enzymes endoglucanase and endochitinase, encoded by PR-2 and PR-3, respectively, changed depending on the tested dpi. At 5 dpi, both enzymes were inhibited in inoculated plants compared to healthy controls. The genes encoding glutathione peroxidase (GPX) and catalase (CAT), both part of the antioxidant plant system, were highly overexpressed. Additionally, the activity of the antioxidant enzymes superoxide dismutase (SOD), CAT, and ascorbate peroxidase (APX) was enhanced in infected roots. Isoelectrofocusing of root extracts revealed novel SOD isoforms in samples from inoculated plants. Furthermore, plants were pre-treated with an array of key compounds, including hormone generators, inhibitors of SA or JA-mediated defense pathways, reactive oxygen species (ROS) scavengers and generators, inhibitors of ROS generation, and compounds that interfere with calcium-mediated metabolism. After treatments, plants were inoculated with RKNs, and nematodes were allowed to complete their life cycle. Factors of plant growth and infection level in treated plants were compared with those from untreated inoculated plants. Generally, compounds that decreased SA and/or ROS levels increased infection severity, while those that reduced JA/ET levels did not affect infection rates. ROS generators induced resistance against the pests. Compounds that silence calcium signaling by preventing its intake augmented infection symptoms. The data shown in this paper indicate that SA-mediated plant immune responses are consistently suppressed during the early stages of nematode infection, and this restriction is associated with the activation of the antioxidant ROS-scavenging system. Full article
(This article belongs to the Special Issue Molecular Interactions between Plants and Pests)
Show Figures

Figure 1

Figure 1
<p>Expression of SA-dependent genes in tomato roots 3 (<b>A</b>) and 7 (<b>B</b>) days after RKN inoculation (nem) and in non-inoculated roots (cntr). Gene expression was detected by quantitative real-time reverse-transcription polymerase chain reaction (qRT-PCR). Gene transcript levels are expressed as 1/ΔC<sub>t</sub>, where ΔC<sub>t</sub>is the difference between the cycle threshold (C<sub>t</sub>) of the tested gene and that of the reference gene (<span class="html-italic">Actin 7</span>). Higher 1/ΔC<sub>t</sub> values indicate higher gene expression. Values are expressed as means (<span class="html-italic">n =</span> 9) ± standard deviations. Means from inoculated roots were compared with their controls using a paired <span class="html-italic">t</span>-test (* <span class="html-italic">p</span> &lt; 0.05; ** <span class="html-italic">p</span> &lt; 0.01).</p>
Full article ">Figure 2
<p>Expression of <span class="html-italic">PR-3</span>, <span class="html-italic">JERF-3</span>, and <span class="html-italic">ACO</span> JA-ET dependent genes in tomato roots 3 (<b>A</b>) and 7 (<b>B</b>) days after RKN inoculation (nem) and non-inoculated roots taken as controls (cntr). Expression of genes was detected by quantitative real-time reverse-transcription polymerase chain reaction (qRT-PCR). Gene transcript levels are expressed as 1/ΔC<sub>t</sub>, where ΔC<sub>t</sub> is the difference between the cycle thresholds (C<sub>t</sub>) of the tested gene and that of the reference gene (<span class="html-italic">Actin 7</span>). Higher 1/ΔC<sub>t</sub> values indicate higher gene expression. Values are expressed as means (<span class="html-italic">n =</span> 9) ± standard deviations. Means from inoculated roots were compared with their controls using a paired <span class="html-italic">t</span>-test (* <span class="html-italic">p</span> &lt; 0.05).</p>
Full article ">Figure 3
<p>Expression of <span class="html-italic">GPX</span> and <span class="html-italic">CAT</span> ROS-scavenging genes in tomato roots 3 (<b>A</b>) and 7 (<b>B</b>) days after RKN inoculation (nem) and non-inoculated roots taken as controls (cntr). Expression of genes was detected by quantitative real-time reverse-transcription polymerase chain reaction (qRT-PCR). Gene transcript levels are expressed as 1/ΔC<sub>t</sub>, where ΔC<sub>t</sub> is the difference between the cycle threshold (C<sub>t</sub>) of the tested gene and that of the reference gene (<span class="html-italic">Actin 7</span>). Higher 1/ΔC<sub>t</sub> values indicate higher gene expression. Values are expressed as means (<span class="html-italic">n =</span> 9) ± standard deviations. Means from inoculated roots were compared with their controls using a paired <span class="html-italic">t</span>-test (** <span class="html-italic">p</span> &lt; 0.01).</p>
Full article ">Figure 4
<p>Isoelectrofocusing of root extracts stained for catalase (CAT) and superoxide dismutase (SOD). Root extracts were obtained from non-inoculated (C) and inoculated (N) plants at 5 days post-inoculation with RKNs. Digital images of mini-gels (3.6 cm separation zone) were adjusted to display white enzyme bands on a dark background. Gels were calibrated using a broad pI calibration kit containing proteins with pI values ranging from 3.5 to 9.3.</p>
Full article ">Figure 5
<p>Scheme of the possible mechanisms by which RKNs suppress basal defense reactions of tomato at the earliest stages of infection. Arrows indicate the processes by which inhibitions/suppressions (X) of biochemical pathways occur.</p>
Full article ">
23 pages, 975 KiB  
Review
MnSOD Mimetics in Therapy: Exploring Their Role in Combating Oxidative Stress-Related Diseases
by Jovan Grujicic and Antiño R. Allen
Antioxidants 2024, 13(12), 1444; https://doi.org/10.3390/antiox13121444 (registering DOI) - 23 Nov 2024
Viewed by 283
Abstract
Reactive oxygen species (ROS) are double-edged swords in biological systems—they are essential for normal cellular functions but can cause damage when accumulated due to oxidative stress. Manganese superoxide dismutase (MnSOD), located in the mitochondrial matrix, is a key enzyme that neutralizes superoxide radicals [...] Read more.
Reactive oxygen species (ROS) are double-edged swords in biological systems—they are essential for normal cellular functions but can cause damage when accumulated due to oxidative stress. Manganese superoxide dismutase (MnSOD), located in the mitochondrial matrix, is a key enzyme that neutralizes superoxide radicals (O2•−), maintaining cellular redox balance and integrity. This review examines the development and therapeutic potential of MnSOD mimetics—synthetic compounds designed to replicate MnSOD’s antioxidant activity. We focus on five main types: Mn porphyrins, Mn salens, MitoQ10, nitroxides, and mangafodipir. These mimetics have shown promise in treating a range of oxidative stress-related conditions, including cardiovascular diseases, neurodegenerative disorders, cancer, and metabolic syndromes. By emulating natural antioxidant defenses, MnSOD mimetics offer innovative strategies to combat diseases linked to mitochondrial dysfunction and ROS accumulation. Future research should aim to optimize these compounds for better stability, bioavailability, and safety, paving the way for their translation into effective clinical therapies. Full article
(This article belongs to the Special Issue Oxidative-Stress in Human Diseases—3rd Edition)
Show Figures

Figure 1

Figure 1
<p>Core structure of Mn Porphyrins.</p>
Full article ">Figure 2
<p>Structure of MitoQ10.</p>
Full article ">Figure 3
<p>Core structure of TEMPO compounds. For TEMPO compounds with longer names, T was used as an abbreviation for TEMPO. DMA-coPFPA is a copolymer made from dimethylacrylamide (DMA) and pentafluorophenyl acrylate (PFPA), often used to enhance targeted delivery.</p>
Full article ">Figure 4
<p>Structure of Mangafodipir and Calmangafodipir [<a href="#B103-antioxidants-13-01444" class="html-bibr">103</a>].</p>
Full article ">
14 pages, 1112 KiB  
Article
Role of Oxidative Stress and Inflammation in Postoperative Complications and Quality of Life After Laryngeal Cancer Surgery
by Andjela Zivkovic, Ana Jotic, Ivan Dozic, Simona Randjelovic, Ivana Cirkovic, Branislava Medic, Jovica Milovanovic, Aleksandar Trivić, Aleksa Korugic, Ivan Vukasinović and Katarina Savic Vujovic
Cells 2024, 13(23), 1951; https://doi.org/10.3390/cells13231951 (registering DOI) - 23 Nov 2024
Viewed by 209
Abstract
(1) Background: Laryngeal surgery due to carcinoma leads to significant tissue disruption, cellular injury, and inflammation. This leads to increased levels of reactive oxygen species (ROS), causing oxidative damage that can influence quality of life (QOL) and recovery and complicate the postoperative course. [...] Read more.
(1) Background: Laryngeal surgery due to carcinoma leads to significant tissue disruption, cellular injury, and inflammation. This leads to increased levels of reactive oxygen species (ROS), causing oxidative damage that can influence quality of life (QOL) and recovery and complicate the postoperative course. The aim of this study was to compare how postoperative quality of life and surgical complication occurrence interacted with the biomarker levels of oxidative stress (malondialdehyde, MDA; superoxide dismutase, SOD; glutathione peroxidase 1, GPX1; and catalase, CAT) and inflammation (interleukin 1, IL-1; interleukin 6, IL-6; C-reactive protein, CRP) in patients treated with conservative and radical laryngeal surgery. (2) Methods: The study included 56 patients who underwent surgical treatment for laryngeal cancer. Blood samples were collected to analyze oxidative stress and inflammation parameters before surgery and on the first and seventh days postoperatively. Serum concentrations of MDA, SOD, GPX, CAT, IL-1, IL-6, and CRP were measured using coated enzyme-linked immunosorbent assay (ELISA) kits. EORTC QLQ-H&H43 questionnaire was used to measure the QOL of patients. (3) Results and Conclusions: T stage, pain intensity, and the extent of the surgical procedure were established as significant predictive factors for QOL in multivariate analysis. There was a significant positive correlation between surgical complication occurrence and preoperative values of GPX and MDA, but significant predictors of surgical complication occurrence on the 7th postoperative day were SOD and MDA values (p < 0.05). Full article
24 pages, 4589 KiB  
Article
Protective Effects of L-Cysteine Against Cisplatin-Induced Oxidative Stress-Mediated Reproductive Damage
by Yi-Fen Chiang, Yi-Tzu Chen, Ko-Chieh Huang, Wei-Lun Hung, Cheng-Pei Chung, Tzong-Ming Shieh, Yun-Ju Huang, Mohamed Ali and Shih-Min Hsia
Antioxidants 2024, 13(12), 1443; https://doi.org/10.3390/antiox13121443 (registering DOI) - 23 Nov 2024
Viewed by 129
Abstract
Cisplatin (CIS) is a widely used chemotherapeutic agent, but its side effects, such as oxidative stress, inflammation, and apoptosis, often lead to male reproductive damage. Oxidative stress, primarily caused by the excessive generation of reactive oxygen species (ROS), plays a critical role in [...] Read more.
Cisplatin (CIS) is a widely used chemotherapeutic agent, but its side effects, such as oxidative stress, inflammation, and apoptosis, often lead to male reproductive damage. Oxidative stress, primarily caused by the excessive generation of reactive oxygen species (ROS), plays a critical role in disrupting testicular homeostasis, resulting in spermatogenic impairment and tissue injury. L-cysteine (CYS), a semi-essential amino acid with potent antioxidant and anti-inflammatory properties, may offer protection against CIS-induced oxidative damage. This study aimed to assess the protective potential of CYS against CIS-induced male reproductive toxicity using in vivo and in vitro models. In vitro, treatment of TM3 (Leydig) and TM4 (Sertoli) cells with CIS led to increased ROS levels, reduced cell viability, and elevated apoptosis and inflammation, all of which were significantly ameliorated by subsequent CYS exposure. In vivo, CIS-treated male rats displayed heightened oxidative stress, impaired spermatogenesis, and histopathological damage in reproductive organs. However, CYS administration for 21 days significantly reduced oxidative stress, improved sperm viability, and protected testicular tissues from damage. These findings suggest that CYS has a protective effect against CIS-induced oxidative stress and male reproductive damage, making it a promising therapeutic agent for mitigating CIS-induced reproductive toxicity. Full article
20 pages, 3793 KiB  
Article
Neuronal Progenitors Suffer Genotoxic Stress in the Drosophila Clock Mutant per0
by Nunzia Colonna Romano, Marcella Marchetti, Anna Marangoni, Laura Leo, Diletta Retrosi, Ezio Rosato and Laura Fanti
Cells 2024, 13(23), 1944; https://doi.org/10.3390/cells13231944 (registering DOI) - 23 Nov 2024
Viewed by 169
Abstract
The physiological role and the molecular architecture of the circadian clock in fully developed organisms are well established. Yet, we have a limited understanding of the function of the clock during ontogenesis. We have used a null mutant (per0) of [...] Read more.
The physiological role and the molecular architecture of the circadian clock in fully developed organisms are well established. Yet, we have a limited understanding of the function of the clock during ontogenesis. We have used a null mutant (per0) of the clock gene period (per) in Drosophila melanogaster to ask whether PER may play a role during normal brain development. In third-instar larvae, we have observed that the absence of functional per results in increased genotoxic stress compared to wild-type controls. We have detected increased double-strand DNA breaks in the central nervous system and chromosome aberrations in dividing neuronal precursor cells. We have demonstrated that reactive oxygen species (ROS) are causal to the genotoxic effect and that expression of PER in glia is necessary and sufficient to suppress such a phenotype. Finally, we have shown that the absence of PER may result in less condensed chromatin, which contributes to DNA damage. Full article
(This article belongs to the Section Cell Nuclei: Function, Transport and Receptors)
18 pages, 4563 KiB  
Article
Exserohilum rostratum-Mediated Synthesis of Silver Nanoparticles: A Case Study on Their Bioherbicidal Activity Against Leptochloa chinensis (L.) Nees
by Ashrit Gulfraz, Yuquan Yuan, Qing Bu, Muhammad Shafiq, Zhiqiu Huang, Mingwei Li, Zhaoxia Dong, Jing An and Yong Chen
Agronomy 2024, 14(12), 2784; https://doi.org/10.3390/agronomy14122784 (registering DOI) - 23 Nov 2024
Viewed by 257
Abstract
The interdisciplinary progress in nanotechnology has yielded environmentally friendly and cost-effective strategies to enhance bioherbicidal efficacy. This study presents the biosynthesis of silver nanoparticles (M-AgNPs) using the fungus Exserohilum rostratum, specifically targeting the Leptochloa chinensis weed in paddy fields. The M-AgNPs were [...] Read more.
The interdisciplinary progress in nanotechnology has yielded environmentally friendly and cost-effective strategies to enhance bioherbicidal efficacy. This study presents the biosynthesis of silver nanoparticles (M-AgNPs) using the fungus Exserohilum rostratum, specifically targeting the Leptochloa chinensis weed in paddy fields. The M-AgNPs were characterized with an aqueous solution size of 107.9 nm and a zeta potential of −24.0 ± 0.20 mV, and their properties were analyzed by UV-Vis spectrophotometry, transmission electron microscopy (TEM), scanning electron microscopy (SEM), X-ray diffraction (XRD), and Fourier-transform infrared spectroscopy (FTIR). The application of M-AgNP suspension at different concentrations of 70 µg∙mL−1, 80 µg∙mL−1, and 100 µg∙mL−1 to L. chinensis at the 3–4 leaf stage resulted in significant herbicidal effects. These nanoparticles induced oxidative stress and significantly reduced the activities of peroxidase, catalase, and superoxide dismutase in the weed seedlings. Meanwhile, M-AgNP treatments significantly increased the activity of cell wall-degrading enzymes, including polygalacturonase and cellulase, in L. chinensis leaves and caused organelle damage in plant leaf cells. Safety assessments showed no significant impact on rice growth after treatment with M-AgNP3 (100 µg∙mL−1) suspension. Our results suggest that M-AgNPs represent a sustainable and eco-friendly approach to weed control that is compatible with rice cultivation, thus supporting the adoption of green agricultural practices. Full article
(This article belongs to the Special Issue Free from Herbicides: Ecological Weed Control)
Show Figures

Figure 1

Figure 1
<p>Characterization of mycosynthesized silver nanoparticles (M-AgNPs). (<b>A</b>) Particle size distribution measured by dynamic light scattering (DLS); (<b>B</b>) zeta potential; (<b>C</b>) UV-Vis absorption spectrum; (<b>D</b>) morphological analysis by transmission electron microscopy (TEM); (<b>E</b>) functional group identification by Fourier-transform infrared spectroscopy (FTIR); (<b>F</b>) surface morphology examination by scanning electron microscopy (SEM); (<b>G</b>) X-ray diffraction (XRD) pattern simulation; (<b>H</b>) energy-dispersive X-ray (EDX) spectrum with the area of analysis indicated; (<b>I</b>) elemental composition analysis of M-AgNPs.</p>
Full article ">Figure 2
<p>Herbicidal effects of M-AgNPs on <span class="html-italic">L. chinensis</span> at the 4-leaf stage. Effects of different concentrations of M-AgNPs on the (<b>A</b>) growth of seedlings; (<b>B</b>) detached leaf; (<b>C</b>) fresh weight; (<b>D</b>) fresh weight control efficiency; and (<b>E</b>) SPAD value of leaf. CK: healthy plant; +ive CK: <span class="html-italic">Exserohilum rostratum</span> filtrate; M-AgNP<sub>1</sub>, M-AgNP<sub>2</sub>, and M-AgNP<sub>3</sub>: treatment groups with 70 µg∙mL<sup>−1</sup>, 80 µg∙mL<sup>−1</sup>, and 100 µg∙mL<sup>−1</sup> of M-AgNPs, respectively. Statistical analysis was performed using the one-way analysis of variance (ANOVA) based on Duncan’s test. Different lowercase letters denote significant differences among treatments (<span class="html-italic">p</span> &lt; 0.05). Error bars represent the standard error of the mean.</p>
Full article ">Figure 3
<p>M-AgNPs caused oxidative stress and decreased antioxidant enzyme activities in <span class="html-italic">L. chinensis</span> leaves. Effect of M-AgNPs on (<b>A</b>) malondialdehyde (MDA); (<b>B</b>) hydrogen peroxide (H<sub>2</sub>O<sub>2</sub>) content; (<b>C</b>) electrolyte leakage (EL); (<b>D</b>) superoxide dismutase (SOD); (<b>E</b>) catalase (CAT); and (<b>F</b>) peroxidase (POD) activity. CK: healthy plant; +ive CK: <span class="html-italic">Exserohilum rostratum</span> filtrate; M-AgNP<sub>1</sub>, M-AgNP<sub>2</sub>, and M-AgNP<sub>3</sub>: treatment groups with 70 µg∙mL<sup>−1</sup>, 80 µg∙mL<sup>−1</sup>, and 100 µg∙mL<sup>−1</sup> of M-AgNPs, respectively. Statistical analysis was performed using the one-way analysis of variance (ANOVA) based on Duncan’s test. Different lowercase letters denote significant differences among treatments (<span class="html-italic">p</span> &lt; 0.05). Error bars represent the standard error of the mean.</p>
Full article ">Figure 4
<p>M-AgNPs modulated the activities of cell wall-degrading enzymes in the leaves of <span class="html-italic">L. chinensis</span>. (<b>A</b>) Polygalacturonase (PG); (<b>B</b>) cellulase (CX); (<b>C</b>) polymethyl-galacturonase (PMG); and (<b>D</b>) β-glucuronidase (βG). +ive CK: <span class="html-italic">Exserohilum rostratum</span> filtrate; M-AgNP<sub>1</sub>, M-AgNP<sub>2</sub>, and M-AgNP<sub>3</sub>: treatment groups with 70 µg∙mL<sup>−1</sup>, 80 µg∙mL<sup>−1</sup>, and 100 µg∙mL<sup>−1</sup> of M-AgNPs, respectively. Statistical analysis was performed using the one-way analysis of variance (ANOVA) based on Duncan’s test. Different lowercase letters denote significant differences among treatments (<span class="html-italic">p</span> &lt; 0.05). Error bars represent the standard error of the mean.</p>
Full article ">Figure 5
<p>M-AgNPs damaged the subcellular structure of leaves in <span class="html-italic">L. chinensis</span>. (<b>A</b>–<b>C</b>) Cell wall and vacuole; (<b>D</b>–<b>F</b>) chloroplast; and (<b>G</b>–<b>I</b>) mitochondria of CK, <span class="html-italic">E. rostratum</span> (100 µg∙mL<sup>−1</sup>), and M-AgNP<sub>3</sub> (100 µg∙mL<sup>−1</sup>), respectively. The red arrow indicates the abnormal organelles. The circle denotes the damage in the cell wall (CW), mitochondria (M), vacuole (V), and chloroplast (Ch).</p>
Full article ">Figure 6
<p>Safety evaluation of M-AgNPs on rice plants. The effects of different concentrations of M-AgNPs on the (<b>A</b>) growth of seedlings; (<b>B</b>) plant height; (<b>C</b>) SPAD value of leaf; (<b>D</b>) shoot fresh weight; and (<b>E</b>) root fresh weight. CK: healthy plant; +ive CK: <span class="html-italic">Exserohilum rostratum</span> filtrate; M-AgNP<sub>1</sub>, M-AgNP<sub>2</sub>, and M-AgNP<sub>3</sub>: 70 µg∙mL<sup>−1</sup>, 80 µg∙mL<sup>−1</sup>, and 100 µg∙mL<sup>−1</sup> of M-AgNPs, respectively. The data were analyzed using the one-way analysis of variance (ANOVA) based on Duncan’s test. Data with different lowercase letters indicate significant differences (<span class="html-italic">p</span> &lt; 0.05). Error bars indicate standard error.</p>
Full article ">Figure 7
<p>Mechanism of mycosynthesis of silver nanoparticle by <span class="html-italic">E. rostratum</span> against <span class="html-italic">L. chinensis</span>. M-AgNPs induced excessive accumulation of ROS by suppressing the activity of antioxidant enzymes and disrupting plant cell structure by enhancing the activities of cell wall-degrading enzymes, polygalacturonase (PG) and cellulase (CX), resulting in leaf necrosis of <span class="html-italic">L. chinensis</span>. An upward arrow indicates an increase, while a downward arrow signifies a decrease.</p>
Full article ">
15 pages, 2385 KiB  
Article
Eleutherin and Isoeleutherin Activity against Staphylococcus aureus and Escherichia coli Strain’s: Molecular Docking and Antibacterial Evaluation
by Mírian Letícia Carmo Bastos, Houéfa Egidia Fallon Adido, Ananda Karolyne Martins de Brito, Cristian Kallahan Silva Chagas, Ana Laura Gadelha Castro, Gleison Gonçalves Ferreira, Pedro Henrique Costa Nascimento, Walice Rans da Silva Padilha, Rosana Moura Sarmento, Viviane Vasconcelos Garcia, Andrey Moacir do Rosario Marinho, Patrícia Santana Barbosa Marinho, Johnatt Allan Rocha de Oliveira, Valdicley Vieira Vale, Sandro Percário and Maria Fâni Dolabela
Int. J. Mol. Sci. 2024, 25(23), 12583; https://doi.org/10.3390/ijms252312583 (registering DOI) - 23 Nov 2024
Viewed by 391
Abstract
Naphthoquinones eleutherin and isoeleutherin have demonstrated promising antibacterial activity, probably due to their quinone structure, which can generate reactive oxygen species. The study examines the activities of pathogens, such as Staphylococcus aureus and Escherichia coli, associated with antimicrobial resistance and explores their [...] Read more.
Naphthoquinones eleutherin and isoeleutherin have demonstrated promising antibacterial activity, probably due to their quinone structure, which can generate reactive oxygen species. The study examines the activities of pathogens, such as Staphylococcus aureus and Escherichia coli, associated with antimicrobial resistance and explores their potential mechanisms of action. The MIC, IC50, and MBC were determined. PharmMapper 2017 server and GOLD 2020.1 software were utilized for molecular docking to identify protein targets and interaction mechanisms. The docking predictions were verified by redocking, focusing on structures with RMSD below 2 Å. The molecular docking revealed a significant affinity of eleutherin for the peptide, transcriptional regulator QacR, and regulatory protein BlaR1 with better interactions with BlaR1 than the crystallographic ligand (benzylpenicillin). Isoeleutherin demonstrated specific interactions with methionine aminopeptidase, indicating specificity and affinity. In summary, the difference in naphthoquinones activities may be related to structural differences. Eleutherin exhibits potential as a therapeutic adjuvant to reverse bacterial resistance in S. aureus, suggesting this molecule interferes with the antibiotic resistance mechanism. The absence of homologous proteins or variations in the structure of the target proteins could be the cause of the inactivity against E. coli. Full article
Show Figures

Figure 1

Figure 1
<p>Bacterial inhibition of <span class="html-italic">Staphylococcus aureus</span> by eleutherin, isoeleutherin, and controls. Solvent—methanol.</p>
Full article ">Figure 2
<p>Bacterial inhibition of <span class="html-italic">Escherichia coli</span> by eleutherin, isoeleutherin, and controls. Solvent—methanol.</p>
Full article ">Figure 3
<p>Molecular interactions of eleutherin and actinonin with peptide deformylase–PDF. (<b>A</b>) Interactions of eleutherin; (<b>B</b>) interactions with the crystallographic ligand actinonin; and RMSD—Root Mean Square Deviation, value in angstrom.</p>
Full article ">Figure 4
<p>Molecular interactions of eleutherin and pentamidine with the regulator QacR. (<b>A</b>) Interactions of eleutherin; (<b>B</b>) interactions with the crystallographic ligand pentamidine; and RMSD—Root Mean Square Deviation, value in angstrom.</p>
Full article ">Figure 5
<p>Molecular interactions of eleutherin and benzylpenicillin with the transcriptional regulator BlaR1. (<b>A</b>) Interactions of eleutherin; (<b>B</b>) interactions with the crystallographic ligand benzylpenicillin; and RMSD—Root Mean Square Deviation, value in angstrom.</p>
Full article ">Figure 6
<p>Molecular interactions of isoeleutherin and ketoheterocycle 618 with methionine aminopeptidase—MetAP. (<b>A</b>) Interactions of isoeleutherin; (<b>B</b>) interactions with the crystallographic ligand ketoheterocycle 618; and RMSD—Root Mean Square Deviation, value in angstrom.</p>
Full article ">
21 pages, 6998 KiB  
Article
Effect of Dihydroquercetin During Long-Last Growth of Yarrowia lipolytica Yeast: Anti-Aging Potential and Hormetic Properties
by Maxim S. Pusev, Olga I. Klein, Natalya N. Gessler, Galina P. Bachurina, Svetlana Yu. Filippovich, Elena P. Isakova and Yulia I. Deryabina
Int. J. Mol. Sci. 2024, 25(23), 12574; https://doi.org/10.3390/ijms252312574 - 22 Nov 2024
Viewed by 333
Abstract
Polyphenols are powerful natural antioxidants with numerous biological activities. They change cell membrane permeability, interact with receptors, intracellular enzymes, and cell membrane transporters, and quench reactive oxygen species (ROS). Yarrowia lipolytica yeast, being similar to mammalian cells, can be used as a model [...] Read more.
Polyphenols are powerful natural antioxidants with numerous biological activities. They change cell membrane permeability, interact with receptors, intracellular enzymes, and cell membrane transporters, and quench reactive oxygen species (ROS). Yarrowia lipolytica yeast, being similar to mammalian cells, can be used as a model to study their survival ability upon long-lasting cultivation, assaying the effect of dihydroquercetin polyphenol (DHQ). The complex assessment of the physiological features of the population assaying cell respiration, survival, ROS detection, and flow cytometry was used. Y. lipolytica showed signs of chronological aging by eight weeks of growth, namely a decrease in the cell number, and size, increased ROS generation, a decrease in colony-forming unit (CFU) and metabolic activity, and decreased respiratory rate and membrane potential. An amount of 150 µM DHQ decreased ROS generation at the 6-week growth stage upon adding an oxidant of 2,2′-azobis (2-amidinopropane) dihydrochloride (AAPH). Moreover, it decreased CFU at 1–4 weeks of cultivation, inhibited cell metabolic activity of the 24-h-old culture and stimulated that on 14–56 days of growth, induced the cell respiration rate in the 24-h-old culture, and blocked alternative mitochondrial oxidase at growth late stages. DHQ serves as a mild pro-oxidant on the first day of age-stimulating anti-stress protection. In the deep stationary stage, it can act as a powerful antioxidant, stabilizing cell redox status and reducing free radical oxidation in mitochondria. It provides a stable state of population. The hormetic effects of DHQ using lower eukaryotes of Y. lipolytica have been previously discussed, which can be used as a model organism for screening geroprotective compounds of natural origin. Full article
(This article belongs to the Special Issue Stress Response Research: Yeast as Models: 2nd Edition)
Show Figures

Figure 1

Figure 1
<p>Dynamics of the total number of the cells (<b>a</b>) and the share of budding cells (<b>b</b>) upon the prolonged cultivation upon addition of DHQ. (<b>c</b>,<b>d</b>)—Micro images of the cells upon long-lasting cultivation. The photos were taken with an AxioCam MRc camera (magnification 100×). (<b>c</b>)—24 h of cultivation; (<b>d</b>)—8 weeks of cultivation. The white arrows show vacuoles. *—Statistically significant difference between samples, <span class="html-italic">p</span> ≤ 0.05. (<b>b</b>) Does not show statistically significant difference. Full statistical analysis is presented in the <a href="#app1-ijms-25-12574" class="html-app">Supplementary Materials</a> section.</p>
Full article ">Figure 2
<p>The medians of forward (FSC-A) (<b>a</b>) and side (SSC-A) (<b>b</b>) scatter of 10 × 10<sup>3</sup> cells for the control and experimental samples. (<b>c</b>–<b>e</b>)—Histogram superimposition of forward and side scatter signal upon long-lasting cultivation; (<b>c</b>)—24 h of growth; (<b>d</b>)—2 weeks of cultivation; (<b>e</b>)—8 weeks of cultivation. *—Statistically significant difference between samples, <span class="html-italic">p</span> ≤ 0.05. (<b>a</b>) Does not show statistically significant difference. Full statistical analysis is presented in the <a href="#app1-ijms-25-12574" class="html-app">Supplementary Materials</a> section.</p>
Full article ">Figure 2 Cont.
<p>The medians of forward (FSC-A) (<b>a</b>) and side (SSC-A) (<b>b</b>) scatter of 10 × 10<sup>3</sup> cells for the control and experimental samples. (<b>c</b>–<b>e</b>)—Histogram superimposition of forward and side scatter signal upon long-lasting cultivation; (<b>c</b>)—24 h of growth; (<b>d</b>)—2 weeks of cultivation; (<b>e</b>)—8 weeks of cultivation. *—Statistically significant difference between samples, <span class="html-italic">p</span> ≤ 0.05. (<b>a</b>) Does not show statistically significant difference. Full statistical analysis is presented in the <a href="#app1-ijms-25-12574" class="html-app">Supplementary Materials</a> section.</p>
Full article ">Figure 3
<p>Dynamics of the ROS generation in the control and experimental samples upon treatment with H<sub>2</sub>DCFDA (<b>a</b>) and upon the AAPH influence of (<b>b</b>), presented as the fluorescence ratio of the samples with DHQ to the control samples of the same age. *—Statistically significant difference between samples, <span class="html-italic">p</span> ≤ 0.05. Full statistical analysis is presented in the <a href="#app1-ijms-25-12574" class="html-app">Supplementary Materials</a> section.</p>
Full article ">Figure 4
<p>Dynamics of the CFU number in the control and experimental samples upon long-lasting cultivation. *—Statistically significant difference between samples, <span class="html-italic">p</span> ≤ 0.05. Full statistical analysis is presented in the <a href="#app1-ijms-25-12574" class="html-app">Supplementary Materials</a> section.</p>
Full article ">Figure 5
<p>The metabolic activity (MTT test) of the control and experimental samples upon long-lasting cultivation. (<b>a</b>) A<sub>590</sub>; (<b>b</b>) A<sub>590</sub> of the experimental samples to the A<sub>590</sub> of the control samples of the same age. *—Statistically significant difference between samples, <span class="html-italic">p</span> ≤ 0.05. Full statistical analysis is presented in the <a href="#app1-ijms-25-12574" class="html-app">Supplementary Materials</a> section.</p>
Full article ">Figure 6
<p>Dynamics of the respiratory rate in the control and experimental samples upon long-lasting cultivation. *—Statistically significant difference between samples, <span class="html-italic">p</span> ≤ 0.05. Full statistical analysis is presented in the <a href="#app1-ijms-25-12574" class="html-app">Supplementary Materials</a> section.</p>
Full article ">Figure 7
<p>The ratio of the medians of the red fluorescence signal dye to the green one using JC-1 (<b>a</b>), superimposition of population fluorescence histograms cultured on the seventh and fifty-sixth days (<b>b</b>), the ratio of the mCherry signal to the FITC signal regarding 1 day control + DHQ (<b>c</b>), fluorescence microscopy of control (<b>d</b>), and experimental (<b>e</b>) samples on the seventh day in <span class="html-italic">Y. lipolytica</span>. Cells were incubated with 0.5 µM JC-1 for 20 min. The incubation medium contained 0.01 M phosphate-buffered saline (PBS), 1% glycerol, pH 7.4. The areas of high mitochondrial polarization are indicated by bright-red fluorescence due to the concentrated dye. To examine the JC-1-stained preparations, filters 02 and 15 (Zeiss) were used (magnification 100×). Photos were taken using an AxioCam MRc camera. *—Statistically significant difference between samples, <span class="html-italic">p</span> ≤ 0.05. Full statistical analysis is presented in the <a href="#app1-ijms-25-12574" class="html-app">Supplementary Materials</a> section.</p>
Full article ">Figure 7 Cont.
<p>The ratio of the medians of the red fluorescence signal dye to the green one using JC-1 (<b>a</b>), superimposition of population fluorescence histograms cultured on the seventh and fifty-sixth days (<b>b</b>), the ratio of the mCherry signal to the FITC signal regarding 1 day control + DHQ (<b>c</b>), fluorescence microscopy of control (<b>d</b>), and experimental (<b>e</b>) samples on the seventh day in <span class="html-italic">Y. lipolytica</span>. Cells were incubated with 0.5 µM JC-1 for 20 min. The incubation medium contained 0.01 M phosphate-buffered saline (PBS), 1% glycerol, pH 7.4. The areas of high mitochondrial polarization are indicated by bright-red fluorescence due to the concentrated dye. To examine the JC-1-stained preparations, filters 02 and 15 (Zeiss) were used (magnification 100×). Photos were taken using an AxioCam MRc camera. *—Statistically significant difference between samples, <span class="html-italic">p</span> ≤ 0.05. Full statistical analysis is presented in the <a href="#app1-ijms-25-12574" class="html-app">Supplementary Materials</a> section.</p>
Full article ">Figure 8
<p>The median of the mitochondrial volume (number of free thiol groups, MitoTracker Green) (<b>a</b>) and mitochondrial potential (MitoTracker Red) (<b>b</b>) upon long-lasting cultivation. (<b>c</b>–<b>e</b>) Superimposition of forward and side scatter histograms of the populations cultured for 24 h, 4 weeks, and 8 weeks. (<b>f</b>–<b>h</b>) Microimages of the cells upon long-lasting cultivation. The photos were taken with an AxioCam MRc camera (magnification 100×); 24 h of cultivation (<b>f</b>); 4 weeks of cultivation (<b>g</b>); and 8 weeks of cultivation (<b>h</b>). *—Statistically significant difference between samples, <span class="html-italic">p</span> ≤ 0.05. (<b>b</b>) Does not show statistically significant difference.</p>
Full article ">Figure 8 Cont.
<p>The median of the mitochondrial volume (number of free thiol groups, MitoTracker Green) (<b>a</b>) and mitochondrial potential (MitoTracker Red) (<b>b</b>) upon long-lasting cultivation. (<b>c</b>–<b>e</b>) Superimposition of forward and side scatter histograms of the populations cultured for 24 h, 4 weeks, and 8 weeks. (<b>f</b>–<b>h</b>) Microimages of the cells upon long-lasting cultivation. The photos were taken with an AxioCam MRc camera (magnification 100×); 24 h of cultivation (<b>f</b>); 4 weeks of cultivation (<b>g</b>); and 8 weeks of cultivation (<b>h</b>). *—Statistically significant difference between samples, <span class="html-italic">p</span> ≤ 0.05. (<b>b</b>) Does not show statistically significant difference.</p>
Full article ">Figure 9
<p>Distribution of the cells into subpopulations with low (P2) and high (P1) mitochondrial activity in the control (<b>a</b>) and experimental (<b>b</b>) samples. *—Statistically significant difference between samples. Full statistical analysis is presented in the <a href="#app1-ijms-25-12574" class="html-app">Supplementary Materials</a> section.</p>
Full article ">Figure 10
<p>Superimposition of the mitochondrial volume (number of free thiol groups, MitoTracker Green) (<b>a</b>) and mitochondrial potential (MitoTracker Red) (<b>b</b>) in the control and experimental samples. Full statistical analysis is presented in the <a href="#app1-ijms-25-12574" class="html-app">Supplementary Materials</a> section.</p>
Full article ">
31 pages, 913 KiB  
Review
Versatile Porphyrin Arrangements for Photodynamic Therapy—A Review
by Arleta Glowacka-Sobotta, Beata Czarczynska-Goslinska, Daniel Ziental, Marcin Wysocki, Maciej Michalak, Emre Güzel and Lukasz Sobotta
Nanomaterials 2024, 14(23), 1879; https://doi.org/10.3390/nano14231879 - 22 Nov 2024
Viewed by 179
Abstract
Nanotechnology is an emerging field that involves the development of nanoscale particles, their fabrication methods, and potential applications. From nanosized inorganic particles to biopolymers, the variety of nanoparticles is unstoppably growing, offering huge opportunities for drug delivery. Various nanoformulations, such as nanoparticles, nanocomposites, [...] Read more.
Nanotechnology is an emerging field that involves the development of nanoscale particles, their fabrication methods, and potential applications. From nanosized inorganic particles to biopolymers, the variety of nanoparticles is unstoppably growing, offering huge opportunities for drug delivery. Various nanoformulations, such as nanoparticles, nanocomposites, and nanoemulsions, have been developed to enhance drug stability, solubility, and tissue penetration. Moreover, nanocarriers can be specifically engineered to target diseased cells or release the drug in a controllable manner, minimizing damage to surrounding healthy tissues and reducing side effects. This review focuses on the combinations between porphyrin derivatives and nanocarriers applied in photodynamic therapy (PDT). PDT has emerged as a significant advance in medicine, offering a low-invasive method for managing infections, the treatment of tumors, and various dermatoses. The therapy relies on the activation of a photosensitizer by light, which results in the generation of reactive oxygen species. Despite their favorable properties, porphyrins reveal non-specific distribution within the body. Nanotechnology has the capability to enhance the PS delivery and its activation. This review explores the potential improvements that are provided by the use of nanotechnology in the PDT field. Full article
Back to TopTop