[go: up one dir, main page]

Next Issue
Volume 12, July
Previous Issue
Volume 12, May
 
 

Antioxidants, Volume 12, Issue 6 (June 2023) – 175 articles

Cover Story (view full-size image): Thrombosis-related diseases are some of the leading causes of illness and death in the general population and, despite significant improvements in long-term survival rates due to remarkable advancements in pharmacologic therapy, they continue to represent a tremendous burden for healthcare systems. Translational research offers interesting pathophysiological insight, gradually shedding light on the pivotal role of oxidative stress in the genesis of thrombosis and cardiovascular diseases. The aim of the present review is to describe the role of oxidative stress in thrombosis-related vascular disease and to summarize the main studies exploring the effects of oral antithrombotic drugs commonly used in patients with atherosclerotic disease and atrial fibrillation on the oxidative stress pathways. View this paper
  • Issues are regarded as officially published after their release is announced to the table of contents alert mailing list.
  • You may sign up for e-mail alerts to receive table of contents of newly released issues.
  • PDF is the official format for papers published in both, html and pdf forms. To view the papers in pdf format, click on the "PDF Full-text" link, and use the free Adobe Reader to open them.
Order results
Result details
Section
Select all
Export citation of selected articles as:
24 pages, 6300 KiB  
Article
Nrf2 as a Therapeutic Target in the Resistance to Targeted Therapies in Melanoma
by Marie Angèle Cucci, Margherita Grattarola, Chiara Monge, Antonella Roetto, Giuseppina Barrera, Emilia Caputo, Chiara Dianzani and Stefania Pizzimenti
Antioxidants 2023, 12(6), 1313; https://doi.org/10.3390/antiox12061313 - 20 Jun 2023
Cited by 2 | Viewed by 3299
Abstract
The use of specific inhibitors towards mutant BRAF (BRAFi) and MEK (MEKi) in BRAF-mutated patients has significantly improved progression-free and overall survival of metastatic melanoma patients. Nevertheless, half of the patients still develop resistance within the first year of therapy. Therefore, understanding the [...] Read more.
The use of specific inhibitors towards mutant BRAF (BRAFi) and MEK (MEKi) in BRAF-mutated patients has significantly improved progression-free and overall survival of metastatic melanoma patients. Nevertheless, half of the patients still develop resistance within the first year of therapy. Therefore, understanding the mechanisms of BRAFi/MEKi-acquired resistance has become a priority for researchers. Among others, oxidative stress-related mechanisms have emerged as a major force. The aim of this study was to evaluate the contribution of Nrf2, the master regulator of the cytoprotective and antioxidant response, in the BRAFi/MEKi acquired resistance of melanoma. Moreover, we investigated the mechanisms of its activity regulation and the possible cooperation with the oncogene YAP, which is also involved in chemoresistance. Taking advantage of established in vitro melanoma models resistant to BRAFi, MEKi, or dual resistance to BRAFi/MEKi, we demonstrated that Nrf2 was upregulated in melanoma cells resistant to targeted therapy at the post-translational level and that the deubiquitinase DUB3 participated in the control of the Nrf2 protein stability. Furthermore, we found that Nrf2 controlled the expression of YAP. Importantly, the inhibition of Nrf2, directly or through inhibition of DUB3, reverted the resistance to targeted therapies. Full article
(This article belongs to the Special Issue Oxidative Stress and NRF2 in Health and Disease)
Show Figures

Graphical abstract

Graphical abstract
Full article ">Figure 1
<p>Viability (MTT assay) in D4M_SENS, D4M_DMSO, D4M_DABres, D4M_TRAres, and D4M_(D+T)res untreated or exposed to dabrafenib (DAB) (<b>A</b>), or trametinib (TRA) (<b>B</b>) at the indicated concentrations 72 h after the treatment. Results are expressed as a percent of respective untreated control (C) and are the mean ± SD of six separate experiments. a: <span class="html-italic">p</span> &lt; 0.01 vs. D4M_SENS; b: <span class="html-italic">p</span> &lt; 0.01 vs. D4M_DMSO. (<b>C</b>) Viability (MTT assay) in D4M_SENS, D4M_DMSO, D4M_DABres, D4M_TRAres, and D4M_(D+T)res untreated or exposed to DMSO (dilution 1/1000), 1.5 μM DAB, 36 nM TRA, and 1.5 μM DAB plus 36 nM TRA combined treatments for 72 h. Results are expressed as a percent of respective untreated control (C) and are the mean ± SD of six separate experiments. ** <span class="html-italic">p</span> &lt; 0.01 vs. D4M_SENS; §§ <span class="html-italic">p</span> &lt; 0.01 vs. D4M_DMSO.</p>
Full article ">Figure 2
<p>Anchorage-independent cell growth in untreated D4M_SENS, D4M_DMSO, D4M_DABres, D4M_TRAres, and D4M_(D+T)res. (<b>A</b>) Representative images of cell morphology obtained with a phase-contrast microscope in the sphere formation Assay. (<b>B</b>) Sphere diameter was expressed as the percent of spheres obtained in the sensitive clone D4M_SENS; sphere numbers greater than 50 μm were expressed as an arbitrary unit, normalized to the value obtained in the sensitive clone D4M_SENS. (<b>C</b>) Soft agar assay in untreated D4M_SENS, D4M_DMSO, D4M_DABres, D4M_TRAres, and D4M_(D+T)res. Colonies &gt; 0.5 mm were counted using ImageJ software. Results were presented as the mean ± SD of triplicate samples from representative data of three independent experiments. ** <span class="html-italic">p</span> &lt; 0.01 vs. D4M_SENS; §§ <span class="html-italic">p</span> &lt; 0.01 vs. D4M_DMSO.</p>
Full article ">Figure 3
<p>Apoptosis in D4M_SENS, D4M_DMSO, D4M_DABres, D4M_TRAres, and D4M_(D+T)res treated with 1.5 μM DAB, 36 nM TRA or combination (DAB + TRA). (<b>A</b>) The flow cytometry profiles of a representative experiment in Annexin V/IP-stained cells at 24 h are shown. Q1-LL = live (Annexin V−/PI−), Q1-LR = early stage of apoptosis (Annexin V+/PI−), Q1-UR = late stage of apoptosis (Annexin V+/PI+), and Q1-UL = necrosis (Annexin V−/PI+). (<b>B</b>) Histograms reporting cytofluorimetric analysis of Annexin V/PI staining in D4M treated sublines. Results of early and late apoptosis and necrosis were expressed as means ± SD of three independent experiments. ** <span class="html-italic">p</span> &lt; 0.01 vs. D4M_SENS; §§ <span class="html-italic">p</span> &lt; 0.01 vs. D4M_DMSO.</p>
Full article ">Figure 4
<p>(<b>A</b>) Boyden chamber assay at 6 h in D4M_SENS, D4M_DMSO, D4M_DABres, D4M_TRAres, and D4M_(D+T)res sublines treated with 1.5 μM DAB, 36 nM TRA or combination (DAB + TRA). The results are expressed as a percentage of invasion inhibition, as the mean ± SD of five independent experiments. ** <span class="html-italic">p</span> &lt; 0.01 vs. D4M_SENS; §§ <span class="html-italic">p</span> &lt; 0.01 vs. D4M_DMSO. (<b>B</b>) Representative images of the tube formation assay on HUVECs after exposure to the conditioned media from untreated D4M_SENS, D4M_DMSO, D4M_DABres, D4M_TRAres, and D4M_(D+T)res subclones. Tube formation was photographed after 6 h incubation with these conditioned media and evaluated by counting the total number of tubes in three wells; three different experiments were performed. The results are illustrated in the histogram below. The data are the mean ± SD of three independent experiments ** <span class="html-italic">p</span> &lt; 0.01 vs. D4M_SENS; §§ <span class="html-italic">p</span> &lt; 0.01 vs. D4M_DMSO.</p>
Full article ">Figure 5
<p>Nrf2 expression in D4M cell lines. (<b>A</b>) Intracellular oxidative stress levels in D4M_SENS, D4M_DMSO, D4M_DABres, D4M_TRAres, and D4M_(D+T)res untreated cells, measured by incubating cells with dichlorodihydrofluorescein diacetate (DCF-DA). The amount of fluorescent product (2,7-dichlorodihydrofluorescein, DCF) was measured by FACScan cytometer (Becton Dickinson Accuri). Bar graph showing median fluorescence intensity (MFI) values, expressed as means ± SD. ** <span class="html-italic">p</span> &lt; 0.01 vs. D4M_SENS; §§ <span class="html-italic">p</span> &lt; 0.01 vs. D4M_DMSO. (<b>B</b>) GSH level was evaluated in D4M_SENS, D4M_DMSO, D4M_DABres, D4M_TRAres, and D4M_(D+T)res untreated cells. Values are the mean ± SD of three separate evaluations. ** <span class="html-italic">p</span> &lt; 0.01 vs. D4M_SENS; §§ <span class="html-italic">p</span> &lt; 0.01 vs. D4M_DMSO. (<b>C</b>) Western blot analysis of Nrf2, and its target gene HO-1 in D4M_DMSO, D4M_SENS, D4M_(D+T)res, D4M_TRAres, and D4M_DABres untreated cells. (<b>D</b>) Densitometric analysis of the protein expression, normalized using the β-actin signal. Data are the mean ± SD of three independent experiments. * <span class="html-italic">p</span> &lt; 0.05 and ** <span class="html-italic">p</span> &lt; 0.01 vs. D4M_SENS; § <span class="html-italic">p</span> &lt; 0.05 and §§ <span class="html-italic">p</span> &lt; 0.01 vs. D4M_DMSO.</p>
Full article ">Figure 6
<p>YAP expression and its regulation by Nrf2. (<b>A</b>) Western blot analysis of YAP and its target genes Survivin (the arrow is indicated the right band) and FoxM1 basal expression in D4M_DMSO, D4M_SENS, D4M_(D+T)res, D4M_TRAres, and D4M_DABres untreated cells. On the right densitometric analysis of the protein expressions, normalized using the β-actin signal. Data are the mean ± SD of three independent experiments. ** <span class="html-italic">p</span> &lt; 0.01 vs. D4M_SENS; §§ <span class="html-italic">p</span> &lt; 0.01 vs. D4M_DMSO; # <span class="html-italic">p</span> ≤ 0.05. (<b>B</b>) Western blot analysis of Nrf2 and YAP expressions in D4M-resistant cells in untreated control cells (C) or 24 h after the treatment with siRNA targeting Nfr2 (siNrf2). On the right is a densitometric analysis of protein expressions. Data were normalized using the β-actin signal and are indicated in the percentage of control values as the mean ± SD of three independent experiments. ** <span class="html-italic">p</span> ≤ 0.01 vs. C.</p>
Full article ">Figure 7
<p>Nrf2 expression regulation. (<b>A</b>) Nrf2 mRNA expression in D4M_DMSO, D4M_SENS, D4M_(D+T)res, D4M_TRAres, and D4M_DABres untreated cells. mRNA expression was evaluated by qRT-PCR in triplicate. Abelson (Abl) gene was utilized as a housekeeping control. Results showing a discrepancy greater than one cycle threshold in one of the wells were excluded. The results were analyzed using the ΔΔCt method. (<b>B</b>) Western blot analysis of KEAP1 in D4M_DMSO, D4M_SENS, D4M_(D+T)res, D4M_TRAres, and D4M_DABres untreated cells. Below is a densitometric analysis of the protein expression, normalized using the β-actin signal. Data are the mean ± SD of three independent experiments. (<b>C</b>) Western blot analysis of DUB3 in D4M_DMSO, D4M_SENS, D4M_(D+T)res, D4M_TRAres, and D4M_DABres untreated cells. Below is a densitometric analysis of the protein expression, normalized using the β-actin signal. Data are the mean ± SD of three independent experiments. ** <span class="html-italic">p</span> &lt; 0.01 vs. D4M_SENS; §§ <span class="html-italic">p</span> &lt; 0.01 vs. D4M_DMSO. ## <span class="html-italic">p</span> &lt; 0.01. (<b>D</b>) Western blot analysis of DUB3, Nrf2, and YAP expressions in D4M resistant cells in untreated control cells (C) or after 24 h from the treatment with siRNA targeting DUB3 (siDUB3). On the right is a densitometric analysis of protein expressions. Data were normalized using the β-actin signal and are indicated in the percentage of control values as the mean ± SD of three independent experiments. ** <span class="html-italic">p</span> ≤ 0.01 vs. C.</p>
Full article ">Figure 8
<p>Viability (MTT assay) in D4M_SENS or resistant subclones treated with specific siRNAs targeting Nrf2 (siNrf2) or DUB3 (siDUB3). (<b>A</b>) Viability in untreated D4M_SENS (Control, C) or treated with DMSO or 1.5 μM DAB; viability in untreated D4M_DABres cells (Control, C) or treated with DMSO (DMSO), 1.5 μM DAB (DAB), siNrf2, siNrf2 plus 1.5 μM DAB (siNrf2+DAB), siDUB3, siDUB3 plus 1.5 μM DAB (siDUB3+DAB), siNeg, siNeg plus 1.5 μM DAB (siNeg+DAB). Results are expressed as a percent of control and are the mean ± SD of three separate experiments. ** <span class="html-italic">p</span> &lt; 0.01 vs. respective Control untreated cells; §§ <span class="html-italic">p</span> &lt; 0.01 vs. respective DMSO treated cells; <span>$</span><span>$</span> <span class="html-italic">p</span> &lt; 0.01 vs. D4M_DABres or D4M_SENS cells treated with DAB; ∫∫ <span class="html-italic">p</span> &lt; 0.01. (<b>B</b>) Viability in untreated D4M_SENS (Control, C) or treated with DMSO or 36 nM TRA; viability in untreated D4M_TRAres cells (Control, C) or treated with DMSO (DMSO), 36 nM TRA (TRA), siNrf2, siNrf2 plus 36 nM TRA (siNrf2+tra), siDUB3, siDUB3 plus 36 nM TRA (siDUB3+TRA), siNeg, siNeg plus 36 nM TRA (siNeg+TRA). Results are expressed as a percent of control and are the mean ± SD of three separate experiments. ** <span class="html-italic">p</span> &lt; 0.01 vs. respective Control untreated cells; §§ <span class="html-italic">p</span> &lt; 0.01 vs. respective DMSO treated cells; <span>$</span><span>$</span> <span class="html-italic">p</span> &lt; 0.01 vs. D4M_TRAres or D4M_SENS cells treated with 36 nM TRA; ∫∫ <span class="html-italic">p</span> &lt; 0.01. (<b>C</b>) Viability in untreated D4M_SENS (Control, C) or treated with DMSO or 1.5 μM DAB plus 36 nM TRA combined treatments (D+T); viability in untreated D4M_(D+T)res cells (Control, C) or treated with DMSO (DMSO), 1.5 μM DAB plus 36 nM TRA combined treatments (D+T), siNrf2, siNrf2 plus combined treatment (siNrf2+ (D+T), siDUB3, siDUB3 plus plus combined treatment (siNrf2+ (D+T), siNneg, siNeg plus combined treatment (siNrf2+ (D+T). Results are expressed as a percent of control and are the mean ± SD of three separate experiments. ** <span class="html-italic">p</span> &lt; 0.01 vs. respective Control untreated cells; §§ <span class="html-italic">p</span> &lt; 0.01 vs. respective DMSO treated cells; <span>$</span><span>$</span> <span class="html-italic">p</span> &lt; 0.01 vs. D4M_(D+T)res cells or D4M_SENS cells treated with 1.5 μM DAB plus 36 nM TRA combined treatments (D+T); ∫∫ <span class="html-italic">p</span> &lt; 0.01.</p>
Full article ">Figure 9
<p>Analysis in A375_sens and A375_DABres human melanoma cell lines. (<b>A</b>) Viability (MTT assay) in A375_sens and A375_DABres untreated (control, C) or treated with dabrafenib 200 nM (DAB). Results are the mean ± SD of three separate experiments. ** <span class="html-italic">p</span> &lt; 0.01 vs. A375_sens. (<b>B</b>) Western blot analysis of Nrf2 and its target gene HO-1 in A375_sens and A375_DABres untreated cells. On the right is a densitometric analysis of the protein expressions, normalized using the β-actin signal. Data are the mean ± SD of three independent experiments. ** <span class="html-italic">p</span> &lt; 0.01 vs. A375_sens. (<b>C</b>) Western blot analysis of YAP and its target gene Survivin in A375_sens and A375_DABres untreated cells. On the right is a densitometric analysis of the protein expressions, normalized using the β-actin signal. Data are the mean ± SD of three independent experiments. ** <span class="html-italic">p</span> &lt; 0.01 vs. A375_sens. (<b>D</b>) Western blot analysis of Nrf2 and YAP expressions in untreated A375_DABres untreated control cells (C) or after 24 h from the treatment with siRNA targeting Nfr2 (siNrf2). On the right is a densitometric analysis of protein expressions. Data were normalized using the β-actin signal and are indicated in the percentage of control values as the mean ± SD of three independent experiments. ** <span class="html-italic">p</span> ≤ 0.01 vs. C.</p>
Full article ">Figure 10
<p>Nrf2 expression regulation in A375 cells. (<b>A</b>) Nrf2 mRNA expression in A375_sens and A375_DABres untreated cells. mRNA expression was evaluated by qRT-PCR in triplicate. Abelson (Abl) gene was utilized as a housekeeping control. Results showing a discrepancy greater than one cycle threshold in one of the wells were excluded. The results were analyzed using the ΔΔCt method. (<b>B</b>) Western blot analysis of KEAP1 in A375_sens and A375_DABres untreated cells. Below is a densitometric analysis of the protein expression, normalized using the β-actin signal. Data are the mean ± SD of three independent experiments. (<b>C</b>) Western blot analysis of DUB3 in A375_sens and A375_DABres untreated cells. Below is a densitometric analysis of the protein expression, normalized using the β-actin signal. Data are the mean ± SD of three independent experiments. ** <span class="html-italic">p</span> &lt; 0.01 vs. A375_sens. (<b>D</b>) Western blot analysis of DUB3, Nrf2, and YAP expressions in A375_DABres untreated control cells (C) or after 24 h from the treatment with siRNA targeting DUB3 (siDUB3). On the right is a densitometric analysis of protein expressions. Data were normalized using the β-actin signal and are indicated in the percentage of control values as the mean ± SD of three independent experiments. ** <span class="html-italic">p</span> ≤ 0.01 vs. C. (<b>E</b>) Viability in untreated A375_sens (Control, C) or treated with 200 nM DAB; viability in A375_DABres cell untreated (control, C) or treated with siNrf2, siNrf2 plus 200 nM DAB, siDUB3, siDUB3 plus 200 nM DAB, siNneg, siNeg plus 200 nM DAB. Results are expressed as percent of control and are the mean ± SD of three separate experiments. ** <span class="html-italic">p</span> &lt; 0.01 vs. respective Control untreated cells; §§ <span class="html-italic">p</span> &lt; 0.01 vs. respective 200 nM DAB treated cells; ∫ <span class="html-italic">p</span> &lt; 0.05.</p>
Full article ">
16 pages, 1700 KiB  
Article
Impact of Modern Oven Treatments on Lipid Oxidation and Vitamin E Content of Fillets from Sardine (Sardina pilchardus) at Different Reproductive Cycle Phases
by Ancuta Nartea, Lama Ismaiel, Emanuela Frapiccini, Pasquale Massimiliano Falcone, Deborah Pacetti, Natale Giuseppe Frega, Paolo Lucci and Sabrina Colella
Antioxidants 2023, 12(6), 1312; https://doi.org/10.3390/antiox12061312 - 20 Jun 2023
Viewed by 1742
Abstract
The beneficial effects of sardine consumption can be related to the presence of bioactive compounds, such as vitamin E and ω3 polyunsaturated fatty acids. In any case, the levels of these compounds in sardine fillet depend on different factors mainly related to the [...] Read more.
The beneficial effects of sardine consumption can be related to the presence of bioactive compounds, such as vitamin E and ω3 polyunsaturated fatty acids. In any case, the levels of these compounds in sardine fillet depend on different factors mainly related to the diet and reproductive cycle phase of the fish as well as the technological treatments carried out to cook the fillets. The aim of the present study is two-fold: first, to evaluate changes in the total fatty acid profile, lipid oxidation, and vitamin E content of raw fillets from sardine (Sardina pilchardus) at different reproductive cycle phases (pre-spawning, spawning, and post-spawning); and second, to highlight how these nutritional profiles are affected by three oven treatments (conventional, steam, and sous-vide). For this purpose, raw fish was grouped into pre-spawning, spawning, and post-spawning phases according to the mesenteric fat frequency and the gonadosomatic index evaluation, and submitted to conventional (CO), steam (SO), and sous-vide (SV) baking. The ratio of EPA/DHA and vitamin E increased from post-spawning to pre-spawning, to spawning. Considering the reproductive phases, baking affected the oxidative degree differently: a CO > SO ≥ SV impact was found in the worst scenario (post-spawning), mitigated by vitamin E, to CO ≥ SO > SV in the best scenario (spawning). SV was the best treatment with high values of vitamin E in pre-spawning individuals (110.1 mg/kg). This study shows how vitamin E is correlated to the combined effect of endogenous and exogenous factors. Full article
(This article belongs to the Special Issue Impact of Processing on Antioxidant Rich Foods - 2nd Edition)
Show Figures

Figure 1

Figure 1
<p>Evolution of SFA, MUFA, ω6 PUFA, EPA, and DHA contents (weight % of total fatty acids), and ω3/ω6 PUFA ratio during the reproductive cycle is reported as three phases: pre-spawning, spawning, and post-spawning. Data are expressed as mean value ± standard deviation (<span class="html-italic">n</span> = 3). Different letters per variable mean statistical difference (<span class="html-italic">p</span> &lt; 0.01).</p>
Full article ">Figure 2
<p>Evolution of PV, TBARS, and vitamin E during the reproductive cycle is reported as three phases: pre-spawning, spawning, and post-spawning. Data are expressed as mean value ± standard deviation. Different letters per variable means statistical difference (<span class="html-italic">p</span> &lt; 0.01). PV = meqO<sub>2</sub>/kg oil; TBARS = µmol/g oil; vitamin E = mg/kg oil.</p>
Full article ">Figure 3
<p>Partial least squares-discriminant analysis (PLS-DA) of fatty acid composition (weight % of total fatty acids) based on (<b>a</b>) reproductive cycle stage (pre-spawning, Pre-Sp; spawning; and post-spawning, Post-SP) and on (<b>b</b>) oven treatment (CO, SO, and SV). CO = conventional oven; SO = steam oven; SV = sous-vide oven.</p>
Full article ">Figure 4
<p>Evolution of (<b>a</b>) vitamin E (mg/kg oil), (<b>b</b>) PV (meqO<sub>2</sub>/kg oil), and (<b>c</b>) TBARS (µmol/g oil), in pre-spawning, spawning, and post-spawning sardines, raw and baked with different techniques: conventional oven (CO), steam oven (SO), and sous-vide (SV) oven. Values are means of three replicates ± standard deviation. Bars having different lowercase letters are significantly different at <span class="html-italic">p</span> &lt; 0.05 within the same reproductive phase.</p>
Full article ">Figure 5
<p>Box plots of normalized concentrations of PV and vitamin E of baked (CO, SO, and SV) fillets in prespawning (Pre Sp), spawning, and post-spawning (Post Sp) groups (<b>a</b>). CO = conventional oven; SO = steam oven; SV = sous-vide oven. Pearson’s correlation of vitamin E/PV, vitamin E/TBARs of all raw and baked samples (<b>b</b>).</p>
Full article ">
18 pages, 5859 KiB  
Article
Whey Improves In Vitro Endothelial Mitochondrial Function and Metabolic Redox Status in Diabetic State
by Elisa Martino, Amalia Luce, Anna Balestrieri, Luigi Mele, Camilla Anastasio, Nunzia D’Onofrio, Maria Luisa Balestrieri and Giuseppe Campanile
Antioxidants 2023, 12(6), 1311; https://doi.org/10.3390/antiox12061311 - 20 Jun 2023
Cited by 4 | Viewed by 1670
Abstract
Endothelial dysfunction plays a critical role in the progression of type 2 diabetes mellitus (T2DM), leading to cardiovascular complications. Current preventive antioxidant strategies to reduce oxidative stress and improve mitochondrial function in T2DM highlight dietary interventions as a promising approach, stimulating the deepening [...] Read more.
Endothelial dysfunction plays a critical role in the progression of type 2 diabetes mellitus (T2DM), leading to cardiovascular complications. Current preventive antioxidant strategies to reduce oxidative stress and improve mitochondrial function in T2DM highlight dietary interventions as a promising approach, stimulating the deepening of knowledge of food sources rich in bioactive components. Whey (WH), a dairy by-product with a considerable content of bioactive compounds (betaines and acylcarnitines), modulates cancer cell metabolism by acting on mitochondrial energy metabolism. Here, we aimed at covering the lack of knowledge on the possible effect of WH on the mitochondrial function in T2DM. The results showed that WH improved human endothelial cell (TeloHAEC) function during the in vitro diabetic condition mimicked by treating cells with palmitic acid (PA) (0.1 mM) and high glucose (HG) (30 mM). Of note, WH protected endothelial cells from PA+HG-induced cytotoxicity (p < 0.01) and prevented cell cycle arrest, apoptotic cell death, redox imbalance, and metabolic alteration (p < 0.01). Moreover, WH counteracted mitochondrial injury and restored SIRT3 levels (p < 0.01). The SiRNA-mediated suppression of SIRT3 abolished the protective effects exerted by WH on the mitochondrial and metabolic impairment caused by PA+HG. These in vitro results reveal the efficacy of whey as a redox and metabolic modulator in the diabetic state and pave the way for future studies to consider whey as the source of dietary bioactive molecules with health benefits in preventive strategies against chronic diseases. Full article
Show Figures

Graphical abstract

Graphical abstract
Full article ">Figure 1
<p>Whey effects on PA+HG-perturbated cell cycle. EC viability following exposure for different times and doses to (<b>A</b>) glucose (HG), (<b>B</b>) palmitic acid (PA), and (<b>C</b>) whey (WH). (<b>D</b>) Viability and (<b>E</b>) cytotoxicity evaluated on EC after 48 h exposure to PA+HG (0.1 mM and 30 mM, respectively) with or without WH supplementation (20% <span class="html-italic">v</span>/<span class="html-italic">v</span>). (<b>F</b>,<b>G</b>) Representative cell cycle detection by cytometric analysis and immunoblotting analysis with cropped blots of (<b>H</b>) cyclin D1 and (<b>I</b>) cyclin E1. Data are reported as mean ± SD of n = 3 independent experiments. M = molecular weight markers, lane 1 = Ctr, lane 2 = WH, Lane 3 = PA+HG, lane 4 = WH+PA+HG. Western blotting results are expressed as arbitrary units (AU). * <span class="html-italic">p</span> &lt; 0.05 vs. 0 or Ctr; † <span class="html-italic">p</span> &lt; 0.01 vs. 0 or Ctr; ‡ <span class="html-italic">p</span> &lt; 0.001 vs. 0 or Ctr; • <span class="html-italic">p</span> &lt; 0.05 vs. PA+HG; § <span class="html-italic">p</span> &lt; 0.01 vs. PA+HG.</p>
Full article ">Figure 2
<p>Whey reduced the PA+HG-mediated apoptosis. (<b>A</b>,<b>B</b>) Representative FACS analysis of annexin V-FITC and PI-staining and immunoblotting analysis with cropped blots of (<b>C</b>) procaspase-9, (<b>D</b>) caspase-3, (<b>E</b>) PARP, (<b>F</b>) Bax, (<b>G</b>) Bcl-2, and (<b>H</b>) Bax/Bcl-2 ratio. Q1: necrotic cells; Q2: late apoptotic cells; Q3: early apoptotic cells; Q4: viable cells. Data are reported as mean ± SD of n = 3 independent experiments. M = molecular weight markers, lane 1 = Ctr, lane 2 = WH, Lane 3 = PA+HG, lane 4 = WH+PA+HG. Western blotting results are expressed as arbitrary units (AU). * <span class="html-italic">p</span> &lt; 0.05 vs. Ctr; † <span class="html-italic">p</span> &lt; 0.01 vs. Ctr; ‡ <span class="html-italic">p</span> &lt; 0.001 vs. Ctr; • <span class="html-italic">p</span> &lt; 0.05 vs. PA+HG; § <span class="html-italic">p</span> &lt; 0.01 vs. PA+HG.</p>
Full article ">Figure 3
<p>Whey prevented the PA+HG-related ROS accrual. (<b>A</b>) MDA content, (<b>B</b>) extracellular ROS levels and representative fluorescent images and FACS analysis of (<b>C</b>,<b>D</b>) intracellular and (<b>E</b>,<b>F</b>) mitochondrial ROS. Scale bars = 100 μm. (<b>G</b>) Immunoblotting analysis with cropped blots of COX-IV protein levels. Data are reported as mean ± SD of n = 3 independent experiments. M = molecular weight markers, lane 1 = Ctr, lane 2 = WH, Lane 3 = PA+HG, lane 4 = WH+PA+HG. Western blotting results are expressed as arbitrary units (AU). † <span class="html-italic">p</span> &lt; 0.01 vs. Ctr; ‡ <span class="html-italic">p</span> &lt; 0.001 vs. Ctr; • <span class="html-italic">p</span> &lt; 0.05 vs. PA+HG; § <span class="html-italic">p</span> &lt; 0.01 vs. PA+HG.</p>
Full article ">Figure 4
<p>Whey ameliorated the PA+HG-induced mitochondrial damage. (<b>A</b>–<b>C</b>) Representative fluorescent images and FACS analysis of mitochondrial membrane potential and (<b>D</b>) HDAC3 activity. SIRT3 levels evaluated by (<b>E</b>) ELISA kit and (<b>F</b>) immunoblotting analysis. Scale bars = 100 μm. Data are reported as mean ± SD of <span class="html-italic">n</span> = 3 independent experiments. M = molecular weight markers, lane 1 = Ctr, lane 2 = WH, Lane 3 = PA+HG, lane 4 = WH+PA+HG. Western blotting results are expressed as arbitrary units (AU). * <span class="html-italic">p</span> &lt; 0.05 vs. Ctr; † <span class="html-italic">p</span> &lt; 0.01 vs. Ctr; ‡ <span class="html-italic">p</span> &lt; 0.001 vs. Ctr; • <span class="html-italic">p</span> &lt; 0.05 vs. PA+HG.</p>
Full article ">Figure 5
<p>Whey beneficial effects on mitochondrial and metabolic homeostasis. (<b>A</b>) ATP production coupled respiration, (<b>B</b>) maximal and (<b>C</b>) basal respiration, and (<b>D</b>) coupling efficiency assessed by Seahorse analyzer. (<b>E</b>) NAD<sup>+</sup>/NADH and (<b>F</b>) GSH/GSSG ratios, (<b>G</b>) ATP levels, (<b>H</b>) lactate content and immunoblotting analysis with cropped blots of (<b>I</b>) LDH, (<b>J</b>) PPAR-α, (<b>K</b>) PPAR-γ, and (<b>L</b>) SREBP1. Data are reported as mean ± SD of n = 3 independent experiments. M = molecular weight markers, lane 1 = Ctr, lane 2 = WH, Lane 3 = PA+HG, lane 4 = WH+PA+HG. Western blotting results are expressed as arbitrary units (AU). * <span class="html-italic">p</span> &lt; 0.05 vs. Ctr; † <span class="html-italic">p</span> &lt; 0.01 vs. Ctr; ‡ <span class="html-italic">p</span> &lt; 0.001 vs. Ctr; • <span class="html-italic">p</span> &lt; 0.05 vs. PA+HG; § <span class="html-italic">p</span> &lt; 0.01 vs. PA+HG.</p>
Full article ">Figure 6
<p>SIRT3 depletion abolished the protective effects of whey. (<b>A</b>) Representative immunoblotting analysis with cropped blot of SIRT3 protein levels in EC treated with empty transfection reagent (Vehicle) or transfected with nontargeting siRNA control (NT) or SIRT3 siRNA (SIRT3<sup>−</sup>). M = weight markers, lane 1 = Ctr, lane 2 = Vehicle, lane 3 = NT, lane 4 = SIRT3<sup>-</sup>. Western blotting data are expressed as arbitrary units (AU). (<b>B</b>–<b>D</b>) Representative fluorescent images and FACS analysis of mitochondrial membrane potential. Scale bars = 100 μm. Data are reported as mean ± SD of n = 3 independent experiments. ‡ <span class="html-italic">p</span> &lt; 0.001 vs. NT; • <span class="html-italic">p</span> &lt; 0.05 vs. NT+PA+HG.</p>
Full article ">Figure 7
<p>SIRT3 silencing suppressed the whey beneficial effects on metabolic pathways. (<b>A</b>–<b>C</b>) Representative fluorescent images and FACS analysis of mitochondrial ROS. Scale bars = 100 μm. Detection of (<b>D</b>) NAD<sup>+</sup>/NADH and (<b>E</b>) GSH/GSSG ratios, (<b>F</b>) ATP levels, and (<b>G</b>) lactate content. Data are reported as mean ± SD of <span class="html-italic">n</span> = 3 independent experiments. * <span class="html-italic">p</span> &lt; 0.05 vs. NT; † <span class="html-italic">p</span> &lt; 0.01 vs. NT; ‡ <span class="html-italic">p</span> &lt; 0.001 vs. NT; • <span class="html-italic">p</span> &lt; 0.05 vs. NT+PA+HG; § <span class="html-italic">p</span> &lt; 0.01 vs. NT+PA+HG.</p>
Full article ">
23 pages, 2542 KiB  
Article
Tyrosine Nitroxidation Does Not Affect the Ability of α-Synuclein to Bind Anionic Micelles, but It Diminishes Its Ability to Bind and Assemble Synaptic-like Vesicles
by Ana Belén Uceda, Juan Frau, Bartolomé Vilanova and Miquel Adrover
Antioxidants 2023, 12(6), 1310; https://doi.org/10.3390/antiox12061310 - 20 Jun 2023
Cited by 1 | Viewed by 1322
Abstract
Parkinson’s disease (PD) is characterized by dopaminergic neuron degeneration and the accumulation of neuronal inclusions known as Lewy bodies, which are formed by aggregated and post-translationally modified α-synuclein (αS). Oxidative modifications such as the formation of 3-nitrotyrosine (3-NT) or di-tyrosine are found in [...] Read more.
Parkinson’s disease (PD) is characterized by dopaminergic neuron degeneration and the accumulation of neuronal inclusions known as Lewy bodies, which are formed by aggregated and post-translationally modified α-synuclein (αS). Oxidative modifications such as the formation of 3-nitrotyrosine (3-NT) or di-tyrosine are found in αS deposits, and they could be promoted by the oxidative stress typical of PD brains. Many studies have tried to elucidate the molecular mechanism correlating nitroxidation, αS aggregation, and PD. However, it is unclear how nitroxidation affects the physiological function of αS. To clarify this matter, we synthetized an αS with its Tyr residues replaced by 3-NT. Its study revealed that Tyr nitroxidation had no effect on either the affinity of αS towards anionic micelles nor the overall structure of the micelle-bound αS, which retained its α-helical folding. Nevertheless, we observed that nitroxidation of Y39 lengthened the disordered stretch bridging the two consecutive α-helices. Conversely, the affinity of αS towards synaptic-like vesicles diminished as a result of Tyr nitroxidation. Additionally, we also proved that nitroxidation precluded αS from performing its physiological function as a catalyst of the clustering and the fusion of synaptic vesicles. Our findings represent a step forward towards the completion of the puzzle that must explain the molecular mechanism behind the link between αS-nitroxidation and PD. Full article
(This article belongs to the Section Aberrant Oxidation of Biomolecules)
Show Figures

Graphical abstract

Graphical abstract
Full article ">Figure 1
<p>Features of αS primary sequence and of its nitroxidation pathways in vivo. (<b>A</b>) Primary sequence of αS in which the Tyr residues are circled to highlight them and the regions corresponding to the distinct domains are in squares in different colors (i.e., the N-terminal domain in purple; the NAC domain in green; the C-terminal domain in orange). (<b>B</b>) Scheme of the free radical pathways of in vivo peroxynitrite-mediated Tyr nitroxidation and cross-linking that lead to the formation of 3-nitrotyrosine and dityrosine, respectively. Peroxynitrite and nitrogen dioxide radicals have been coloured in purple in order to highlight their participation in the process. The nitro group of 3-NT and the new bond resulting from dityrosine formation have been also coloured in purple to highlight their formation.</p>
Full article ">Figure 2
<p>Impact of Tyr nitroxidation on the αS-bound α-helical folding. Panels (<b>A</b>,<b>C</b>–<b>E</b>) show the overlapping of the CD spectra of 20 µM αS alone or in the presence of 10 mM SDS (<b>A</b>), 5 mM DOPC-SUVs (<b>C</b>), 5 mM DOPS-SUVs (<b>D</b>), or 5 mM ESC-SUVs (<b>E</b>) on that corresponding to αS-NO<sub>2</sub> in the presence of those different lipids. All the CD spectra were acquired in 20 mM phosphate buffer (pH 7.4) enriched with 150 mM NaCl and at 25 °C. Panel (<b>B</b>) shows the percentage of α-helicity achieved for αS (black) and αS-NO<sub>2</sub> (purple) in the absence or presence of SDS micelles and distinct SUVs. By using Equation (2), the percentages were determined from the CD data.</p>
Full article ">Figure 3
<p>CD study of the effect of temperature on the α-helicities of αS and αS-NO<sub>2</sub>. (<b>A</b>,<b>B</b>) Overlapping of the CD spectra of solutions containing αS (20 μM) (<b>A</b>) or aS-NO<sub>2</sub> (20 μM) (<b>B</b>) in the presence of SDS (10 mM) collected at different temperatures (10–50 °C). (<b>C</b>,<b>D</b>) Plots of the values of [<span class="html-italic">θ</span>]<sub>222nm</sub> (<b>C</b>) or [<span class="html-italic">θ</span>]<sub>200nm</sub> (<b>D</b>) collected at different temperatures for solutions containing αS (20 μM) and SDS (10 mM) (black) or αS-NO<sub>2</sub> (20 μM) and SDS (10 mM) (purple). In Panels (<b>C</b>,<b>D</b>), the experimental data are shown as dots, whereas their fits to linear functions are shown as lines.</p>
Full article ">Figure 4
<p>Study of the impact of Tyr nitroxidation on the affinity of αS-ESC-SUV. (<b>A</b>) Overlapping of the <sup>1</sup>H,<sup>15</sup>N-HSQC spectra of 135 µM αS-NO<sub>2</sub> before (black) and after (red) the addition of 1.3 mM ESC-SUVs. Both spectra were acquired in 20 mM phosphate buffer (pH 6.5) at 37 °C. (<b>B</b>) Fractional signal attenuation of the <sup>1</sup>H,<sup>15</sup>N-HSQC signals relative to the lipid-free spectrum as a function of the residue number for αS-NO<sub>2</sub> (135 µM) in the presence of ESC-SUVs at 250 µM (black), 610 µM (grey), and 1.3 mM (red) concentrations. The experiments were acquired at 12.5 °C. (<b>C</b>) Lipid-bound fraction of αS-NO<sub>2</sub> at increasing ESC-SUV concentrations. The data were obtained from the <sup>1</sup>H,<sup>15</sup>N-HSQC spectra αS-NO<sub>2</sub> in the presence of different concentrations of ESC-SUVs (0–30 mM). The spectra were recorded in 20 mM phosphate buffer (pH 6.5) at 12.5 °C. The experimental data were fit to the bimolecular binding curve (Equation (6)) by using the software Sigma Plot (version 10), which allowed obtaining the dissociation constant (<span class="html-italic">K<sub>d</sub></span>).</p>
Full article ">Figure 5
<p>Impact of Tyr nitroxidation on the chemical shifts of αS bound to SDS micelles. (<b>A</b>) Overlapping of the <sup>1</sup>H,<sup>15</sup>N-HSQC spectra of 130 µM αS-NO<sub>2</sub> (purple) and 100 µM αS (black) obtained in the presence of 40 mM d<sub>25</sub>-SDS micelles. The spectra were collected in 20 mM phosphate buffer (pH 6.5) at 37 °C. Residues whose signals were shifted as a result of Tyr nitroxidation are labelled in red. (<b>B</b>) Amide chemical shift perturbations (Δ<span class="html-italic">δ</span>) of the H<sub>N</sub> and N backbone resonances of SDS bound-αS as a result of Tyr nitroxidation. For each residue, <math display="inline"><semantics> <mrow> <mo>∆</mo> <mi>δ</mi> <mo>=</mo> <msqrt> <msup> <mrow> <mfenced separators="|"> <mrow> <msub> <mrow> <mo>∆</mo> <mi>δ</mi> </mrow> <mrow> <mi>H</mi> <mi>N</mi> </mrow> </msub> </mrow> </mfenced> </mrow> <mrow> <mn>2</mn> </mrow> </msup> <mo>+</mo> <msup> <mrow> <mi>x</mi> <mo>·</mo> <mfenced separators="|"> <mrow> <msub> <mrow> <mo>∆</mo> <mi>δ</mi> </mrow> <mrow> <mi>N</mi> </mrow> </msub> </mrow> </mfenced> </mrow> <mrow> <mn>2</mn> </mrow> </msup> </msqrt> </mrow> </semantics></math>, where x is 0.2 for Gly and 0.14 for the other residues. Δ<span class="html-italic">δ<sub>HN</sub></span> and Δ<span class="html-italic">δ<sub>N</sub></span> are the amide proton and the amide nitrogen chemical shift differences between αS and αS-NO<sub>2</sub> in the presence of SDS (Δ<span class="html-italic">δ</span><sub><span class="html-italic">x</span></sub> = <span class="html-italic">δ</span><sub><span class="html-italic">x</span>,αSNO2</sub> − <span class="html-italic">δ</span><sub><span class="html-italic">x</span>,αS</sub>). The chemical shift assignments of H<sub>N</sub> and N resonances of the SDS bound αS were achieved in a previous work of our group [<a href="#B11-antioxidants-12-01310" class="html-bibr">11</a>]. Data corresponding to Tyr residues are colored in green. (<b>C</b>) Residue-specific ncSPC α-helical scores (<a href="https://st-protein02.chem.au.dk/ncSPC/" target="_blank">https://st-protein02.chem.au.dk/ncSPC/</a>) (accessed on 7 July 2022) obtained for αS (black and grey) and for αS-NO<sub>2</sub> (purple) in the absence (grey) or in the presence (black and purple) of SDS calculated from the H<sub>N</sub>, H<sub>α</sub>, C<sub>α</sub>, C<sub>β</sub>, and CO chemical shifts. Here, “+1” denotes the highest propensity to form a completely formed α-helix, “0” denotes disorder, and “−1” denotes a fully formed β-sheet.</p>
Full article ">Figure 6
<p>SDS micelle-bound αS-NO<sub>2</sub>: structure and dynamics. (<b>A</b>) NMR bundles of the 10 lowest-energy structures of αS-NO<sub>2</sub> (<b>left</b>). Purple sticks represent the backbone. Average structure of αS-NO<sub>2</sub> (<b>right</b>) obtained from the ensemble (<b>left</b>) using MOLMOL. For visualization purposes, the disordered C-terminal domain was deleted in both representations. (<b>B</b>) Alignment of the D2-G41 (H1; (<b>bottom</b>)) and E46-L100 (H2; (<b>top</b>)) regions of the average structures of αS (grey) [<a href="#B11-antioxidants-12-01310" class="html-bibr">11</a>] and αS-NO<sub>2</sub> (purple). The Pymol software (version 2.5.3) was used to carry out the alignment. The side chains of Y39 in αS and αS-NO<sub>2</sub> are shown as sticks. (<b>C</b>) Plot of the <span class="html-italic">R</span><sub>2</sub> (s<sup>−1</sup>) relaxation data collected for αS (black) and αS-NO<sub>2</sub> (purple) in the presence of SDS micelles. Relaxation values of the different Tyr residues are colored in green. The relaxation measurements were performed at 37 °C in 20 mM phosphate buffer (pH 6.5).</p>
Full article ">Figure 7
<p>Studying the impact of Tyr nitroxidation on the ability of αS to modulate the ordering and fusion of SUVs mimicking SVs. (<b>A</b>,<b>B</b>) Lipid order parameters (<span class="html-italic">S</span>) of 130 µM DOPS-SUVs labelled with the TMA-DPH ((<b>A</b>), 2 µM) or DPH ((<b>B</b>), 1 µM) probes in the absence (grey) or in the presence of αS (black) or αS-NO<sub>2</sub> (purple). In Panels (<b>A</b>,<b>B</b>), empty and full bars represent the <span class="html-italic">S</span> values of the DOPS-SUVs before and after the addition of 13 µM αS or αS-NO<sub>2</sub>, respectively. (<b>C</b>,<b>D</b>) DLS size distributions of 130 µM DOPS-SUVs before (red) and after (green) 96 h of incubation with αS (13 µM) (<b>C</b>) or αS-NO<sub>2</sub> (13 µM) (<b>D</b>). All the measurements were performed in Buffer B1 and at 25 °C.</p>
Full article ">
23 pages, 1667 KiB  
Review
Benefits of Natural Antioxidants on Oral Health
by Giuseppina Malcangi, Assunta Patano, Anna Maria Ciocia, Anna Netti, Fabio Viapiano, Irene Palumbo, Irma Trilli, Mariafrancesca Guglielmo, Alessio Danilo Inchingolo, Gianna Dipalma, Francesco Inchingolo, Elio Minetti and Angelo Michele Inchingolo
Antioxidants 2023, 12(6), 1309; https://doi.org/10.3390/antiox12061309 - 20 Jun 2023
Cited by 10 | Viewed by 3658
Abstract
In recent years, special attention has been paid to the correlation between oxidation–reduction mechanisms and human health. The free radicals produced via physiological cellular biochemical processes are major contributors to oxidation phenomena. Their instability is the major cause of cellular damage. Free radical [...] Read more.
In recent years, special attention has been paid to the correlation between oxidation–reduction mechanisms and human health. The free radicals produced via physiological cellular biochemical processes are major contributors to oxidation phenomena. Their instability is the major cause of cellular damage. Free radical reactive oxygen species containing oxygen are the best-known ones. The body neutralises the harmful effects of free radicals via the production of endogenous antioxidants (superoxide dismutase, catalase, glutathione, and melatonin). The field of study of nutraucetics has found antioxidant capacity in substances such as vitamins A, B, C, E, coenzyme Q-10, selenium, flavonoids, lipoic acid, carotenoids, and lycopene contained in some foods. There are several areas of investigation that aim to research the interaction between reactive oxygen species, exogenous antioxidants, and the microbiota to promote increased protection via the peroxidation of macromolecules (proteins, and lipids) by maintaining a dynamic balance among the species that make up the microbiota. In this scoping review, we aim to map the scientific literature on oxidative stress related to the oral microbiota, and the use of natural antioxidants to counteract it, to assess the volume, nature, characteristics, and type of studies available to date, and to suggest the possible gaps that will emerge from the analysis. Full article
(This article belongs to the Special Issue The Role of Oxidative Stress and Antioxidant Systems in Oral Health)
Show Figures

Figure 1

Figure 1
<p>Factors inducing ROS formation.</p>
Full article ">Figure 2
<p>PRISMA-ScR Flow diagram.</p>
Full article ">Figure 3
<p>Oral diseases affected by antioxidant intake.</p>
Full article ">Figure 4
<p>Redox mechanism in pathological and physiological cellular metabolism.</p>
Full article ">
19 pages, 2520 KiB  
Article
In Vitro Characterization of Antioxidant, Antibacterial and Antimutagenic Activities of the Green Microalga Ettlia pseudoalveolaris
by Andrea Vornoli, Teresa Grande, Valter Lubrano, Francesco Vizzarri, Chiara Gorelli, Andrea Raffaelli, Clara Maria Della Croce, Santiago Zarate Baca, Carla Sandoval, Vincenzo Longo, Luisa Pozzo and Cristina Echeverria
Antioxidants 2023, 12(6), 1308; https://doi.org/10.3390/antiox12061308 - 20 Jun 2023
Cited by 3 | Viewed by 1792
Abstract
Recently, green microalgae have gained importance due to their nutritional and bioactive compounds, which makes them some of the most promising and innovative functional foods. The aim of this study was to evaluate the chemical profile and the in vitro antioxidant, antimicrobial and [...] Read more.
Recently, green microalgae have gained importance due to their nutritional and bioactive compounds, which makes them some of the most promising and innovative functional foods. The aim of this study was to evaluate the chemical profile and the in vitro antioxidant, antimicrobial and antimutagenic activity of an aqueous extract of the green microalga Ettlia pseudoalveolaris, obtained from the freshwater lakes of the Ecuadorian Highlands. Human microvascular endothelial cells (HMEC-1) were used to determine the ability of the microalga to reduce the endothelial damage caused by hydrogen peroxide-induced oxidative stress. Furthermore, the eukaryotic system Saccharomyces cerevisiae was used to evaluate the possible cytotoxic, mutagenic and antimutagenic effect of E. pseudoalveolaris. The extract showed a notable antioxidant capacity and a moderate antibacterial activity mostly due to the high content in polyphenolic compounds. It is likely that the antioxidant compounds present in the extract were also responsible for the observed reduction in endothelial damage of HMEC-1 cells. An antimutagenic effect through a direct antioxidant mechanism was also found. Based on the results of in vitro assays, E. pseudoalveolaris proved to be a good source of bioactive compounds and antioxidant, antibacterial and antimutagenic capacities making it a potential functional food. Full article
Show Figures

Figure 1

Figure 1
<p>LOX-1 protein levels in HMEC-1 cells by immunoenzymatic assay, expressed as pg of LOX-1 per 100,000 cells, measured in CTR, H<sub>2</sub>O<sub>2</sub>, H<sub>2</sub>O<sub>2</sub> + E (<span class="html-italic">E. pseudoalveolaris</span>) and E cells. Data represent the mean ± SD (bars). a, b: values significantly different by one way ANOVA-test, <span class="html-italic">p</span> = 0.0359.</p>
Full article ">Figure 2
<p>IL-6 protein levels in HMEC-1 cells by immunoenzymatic assay, expressed as pg of IL-6 per 100,000 cells, measured in CTR, H<sub>2</sub>O<sub>2</sub>, H<sub>2</sub>O<sub>2</sub> + E (<span class="html-italic">E. pseudoalveolaris</span>) and E cells. Data represent the mean ± SD (bars). a, b: values significantly different by one way ANOVA-test, <span class="html-italic">p</span> = 0.0478.</p>
Full article ">Figure 3
<p>Nitrites and nitrates levels in HMEC-1 cells based on Griess reaction, expressed as µM NO<sub>2</sub><sup>−</sup> + NO<sub>3</sub><sup>−</sup>/100,000 cells, measured in CTR, H<sub>2</sub>O<sub>2</sub>, H<sub>2</sub>O<sub>2</sub> + E (<span class="html-italic">E. pseudoalveolaris</span>) and E cells. Data represent the mean ± SD (bars). a, b: values significantly different by one way ANOVA-test, <span class="html-italic">p</span> &lt; 0.0001.</p>
Full article ">Figure 4
<p>Growth percentage of Gram-negative (<span class="html-italic">E. coli</span> and <span class="html-italic">S. typhimurium</span>) and Gram-positive (<span class="html-italic">S. auresus</span> and <span class="html-italic">E. faecalis</span>) bacterial strains at different concentrations (0, 0.0031, 0.31, 6.25 and 50 mg/mL) of <span class="html-italic">E. pseudoalveolaris</span> (E) extract. Data represent the mean ± SD (bars). a, b: values significantly different by one way ANOVA-test, <span class="html-italic">p</span> = 0.0003 (<span class="html-italic">E. coli</span>); <span class="html-italic">p</span> = 0.0007 (<span class="html-italic">S. typhimurium</span>); <span class="html-italic">p</span> = 0.0006 (<span class="html-italic">S. aureus</span>); <span class="html-italic">p</span> &lt; 0.0001 (<span class="html-italic">E. faecalis</span>).</p>
Full article ">Figure 5
<p>Mitotic gene conversion (GC) frequency on the D7 strain of yeast <span class="html-italic">S. cerevisiae</span> incubated with different concentrations (0, 50, 100, 200 and 400 μg/mL) of <span class="html-italic">E. pseudoalveolaris</span> (E) extract, added during cell growth (<span class="html-italic">growth assay</span>, empty bar) or after cell growth (<span class="html-italic">incubation assay</span>, striped bar). Results are reported as convertants/10<sup>5</sup> survivors. Data represent the mean ± SD (bars). Capital letters are used for experimental mode (1), A, a: values significantly different by one way ANOVA-test.</p>
Full article ">Figure 6
<p>Point reverse mutation (PM) frequency on the D7 strain of yeast <span class="html-italic">S. cerevisiae</span> incubated with different concentrations (0, 50, 100, 200 and 400 μg/mL) of <span class="html-italic">E. pseudoalveolaris</span> (E) extract, added during cell growth (<span class="html-italic">growth assay</span>, empty bar) or after cell growth (<span class="html-italic">incubation assay</span>, striped bar). Results are reported as revertants/10<sup>6</sup> survivors. Data represent the mean ± SD (bars). Capital letters are used for experimental mode (1), A, a: values significantly different by one way ANOVA-test.</p>
Full article ">Figure 7
<p>Mitotic gene conversion (GC) frequency on the D7 strain of yeast <span class="html-italic">S. cerevisiae</span> incubated with 4 mM H<sub>2</sub>O<sub>2</sub> and different concentrations (0, 50, 100, 200 and 400 μg/mL) of <span class="html-italic">E. pseudoalveolaris</span> (E) extract, added during cell growth (<span class="html-italic">growth assay</span>, empty bar) or after cell growth (<span class="html-italic">incubation assay,</span> striped bar), compared to untreated cells (CTR). Results are reported as convertants/10<sup>5</sup> survivors. Data represent the mean ± SD (bars). Capital letters are used for experimental mode (1), A, B, C; a, b, c: values significantly different by one way ANOVA-test, <span class="html-italic">p</span> = 0.001.</p>
Full article ">Figure 8
<p>Point reverse mutation (PM) frequency on the D7 strain of yeast <span class="html-italic">S. cerevisiae</span> incubated with 4 mM H<sub>2</sub>O<sub>2</sub> and different concentrations (0, 50, 100, 200 and 400 μg/mL) of <span class="html-italic">E. pseudoalveolaris</span> (E) extract, added during cell growth (<span class="html-italic">growth assay</span>, empty bar) or after cell growth (<span class="html-italic">incubation assay</span>, striped bar), compared to untreated cells (CTR). Results are reported as revertants/10<sup>6</sup> survivors. Data represent the mean ± SD (bars). Capital letters are used for experimental mode (1), A, B, a, b, c: values significantly different by one way ANOVA-test, <span class="html-italic">p</span> = 0.0235.</p>
Full article ">
16 pages, 5419 KiB  
Article
3-Bromo-4,5-dihydroxybenzaldehyde Protects Keratinocytes from Particulate Matter 2.5-Induced Damages
by Ao-Xuan Zhen, Mei-Jing Piao, Kyoung-Ah Kang, Pincha-Devage-Sameera-Madushan Fernando, Herath-Mudiyanselage-Udari-Lakmini Herath, Suk-Ju Cho and Jin-Won Hyun
Antioxidants 2023, 12(6), 1307; https://doi.org/10.3390/antiox12061307 - 20 Jun 2023
Cited by 3 | Viewed by 1473
Abstract
Cellular senescence can be activated by several stimuli, including ultraviolet radiation and air pollutants. This study aimed to evaluate the protective effect of marine algae compound 3-bromo-4,5-dihydroxybenzaldehyde (3-BDB) on particulate matter 2.5 (PM2.5)-induced skin cell damage in vitro and in vivo. [...] Read more.
Cellular senescence can be activated by several stimuli, including ultraviolet radiation and air pollutants. This study aimed to evaluate the protective effect of marine algae compound 3-bromo-4,5-dihydroxybenzaldehyde (3-BDB) on particulate matter 2.5 (PM2.5)-induced skin cell damage in vitro and in vivo. The human HaCaT keratinocyte was pre-treated with 3-BDB and then with PM2.5. PM2.5-induced reactive oxygen species (ROS) generation, lipid peroxidation, mitochondrial dysfunction, DNA damage, cell cycle arrest, apoptotic protein expression, and cellular senescence were measured using confocal microscopy, flow cytometry, and Western blot. The present study exhibited PM2.5-generated ROS, DNA damage, inflammation, and senescence. However, 3-BDB ameliorated PM2.5-induced ROS generation, mitochondria dysfunction, and DNA damage. Furthermore, 3-BDB reversed the PM2.5-induced cell cycle arrest and apoptosis, reduced cellular inflammation, and mitigated cellular senescence in vitro and in vivo. Moreover, the mitogen-activated protein kinase signaling pathway and activator protein 1 activated by PM2.5 were inhibited by 3-BDB. Thus, 3-BDB suppressed skin damage induced by PM2.5. Full article
(This article belongs to the Special Issue Oxidative Stress Induced by Air Pollution)
Show Figures

Figure 1

Figure 1
<p>Inhibition of PM<sub>2.5</sub>-induced ROS generation and lipid peroxidation were performed by 3-BDB in keratinocytes. (<b>a</b>) Cells were added to 10, 20, and 30 μΜ of 3-BDB or 1 mM of N-acetyl cysteine (NAC) for 1 h and then exposed to 50 μg/mL of PM<sub>2.5</sub> for 30 min. ROS were measured by a flow cytometer after H<sub>2</sub>DCFDA staining. (<b>b</b>) Depletion of PM<sub>2.5</sub>-induced ROS by 30 μM of 3-BDB was visualized by a confocal microscope after H<sub>2</sub>DCFDA staining. (<b>c</b>) Prevention of PM<sub>2.5</sub>-induced lipid peroxidation analysis by 3-BDB was performed using a confocal microscope after DPPP staining. (<b>a</b>–<b>c</b>) * <span class="html-italic">p</span> &lt; 0.05 and <sup>#</sup> <span class="html-italic">p</span> &lt; 0.05 compared to control cells and PM<sub>2.5</sub>-exposed cells, respectively.</p>
Full article ">Figure 2
<p>Prevention of PM<sub>2.5</sub>-induced mitochondrial dysfunction was performed by 3-BDB in keratinocytes. Cells were treated with 30 μΜ of 3-BDB for 1 h and then exposed to 50 μg/mL of PM<sub>2.5</sub> for 24 h. (<b>a</b>) Rhod-2 AM was used to detect the mitochondrial calcium. (<b>b</b>,<b>c</b>) The mitochondrial membrane potential was obtained by (<b>b</b>) flow cytometry and (<b>c</b>) confocal microscopy by JC-1 staining. (<b>a</b>–<b>c</b>) * <span class="html-italic">p</span> &lt; 0.05 and <sup>#</sup> <span class="html-italic">p</span> &lt; 0.05 compared to control cells and PM<sub>2.5</sub>-exposed cells, respectively.</p>
Full article ">Figure 3
<p>Reversibility of PM<sub>2.5</sub>-induced DNA damage and cell cycle arrest was performed by 3-BDB. Cells were treated with 30 μΜ of 3-BDB for 1 h and then exposed to 50 μg/mL of PM<sub>2.5</sub> for 24 h. (<b>a</b>) Avidin-TRITC conjugate was used to detect the 8-oxoG. (<b>b</b>) A Comet assay was performed to analyze DNA damage. (<b>c</b>,<b>d</b>) The proteins were obtained from both (<b>c</b>) cells and (<b>d</b>) tissues, and phospho-H2A.X, phospho-p53, and p53 were examined by Western blot. (<b>e</b>) The checkpoint of the G<sub>1</sub> phase was measured by flow cytometry. (<b>a</b>–<b>e</b>) * <span class="html-italic">p</span> &lt; 0.05 and <sup>#</sup> <span class="html-italic">p</span> &lt; 0.05 compared to control groups and PM<sub>2.5</sub>-exposed groups, respectively.</p>
Full article ">Figure 3 Cont.
<p>Reversibility of PM<sub>2.5</sub>-induced DNA damage and cell cycle arrest was performed by 3-BDB. Cells were treated with 30 μΜ of 3-BDB for 1 h and then exposed to 50 μg/mL of PM<sub>2.5</sub> for 24 h. (<b>a</b>) Avidin-TRITC conjugate was used to detect the 8-oxoG. (<b>b</b>) A Comet assay was performed to analyze DNA damage. (<b>c</b>,<b>d</b>) The proteins were obtained from both (<b>c</b>) cells and (<b>d</b>) tissues, and phospho-H2A.X, phospho-p53, and p53 were examined by Western blot. (<b>e</b>) The checkpoint of the G<sub>1</sub> phase was measured by flow cytometry. (<b>a</b>–<b>e</b>) * <span class="html-italic">p</span> &lt; 0.05 and <sup>#</sup> <span class="html-italic">p</span> &lt; 0.05 compared to control groups and PM<sub>2.5</sub>-exposed groups, respectively.</p>
Full article ">Figure 4
<p>Reduction in PM<sub>2.5</sub>-induced cell apoptosis was by 3-BDB in vitro and in vivo. Cells were treated with 30 μΜ of 3-BDB for 1 h and then exposed to 50 μg/mL of PM<sub>2.5</sub> for 24 h. Mice skin was treated with 3-BDB and PM<sub>2.5</sub> according to the animal experiment in Materials and Methods. (<b>a</b>–<b>d</b>) The proteins were isolated from (<b>a</b>,<b>c</b>) cells, (<b>b</b>,<b>d</b>) tissue, and Bcl-2, Bax, and cleaved caspase-9, and cleaved caspase-3 were detected by Western blot. (<b>e</b>) The apoptotic bodies were counted by using Hoechst 33342 staining. The arrows indicate the apoptotic bodies. (<b>a</b>–<b>e</b>) * <span class="html-italic">p</span> &lt; 0.05 and <sup>#</sup> <span class="html-italic">p</span> &lt; 0.05 compared to control groups and PM<sub>2.5</sub>-exposed groups, respectively.</p>
Full article ">Figure 4 Cont.
<p>Reduction in PM<sub>2.5</sub>-induced cell apoptosis was by 3-BDB in vitro and in vivo. Cells were treated with 30 μΜ of 3-BDB for 1 h and then exposed to 50 μg/mL of PM<sub>2.5</sub> for 24 h. Mice skin was treated with 3-BDB and PM<sub>2.5</sub> according to the animal experiment in Materials and Methods. (<b>a</b>–<b>d</b>) The proteins were isolated from (<b>a</b>,<b>c</b>) cells, (<b>b</b>,<b>d</b>) tissue, and Bcl-2, Bax, and cleaved caspase-9, and cleaved caspase-3 were detected by Western blot. (<b>e</b>) The apoptotic bodies were counted by using Hoechst 33342 staining. The arrows indicate the apoptotic bodies. (<b>a</b>–<b>e</b>) * <span class="html-italic">p</span> &lt; 0.05 and <sup>#</sup> <span class="html-italic">p</span> &lt; 0.05 compared to control groups and PM<sub>2.5</sub>-exposed groups, respectively.</p>
Full article ">Figure 5
<p>Inactivation of PM<sub>2.5</sub>-induced MAPK signaling pathway, the transcription factor AP-1 was performed by 3-BDB. (<b>a</b>,<b>b</b>) Cells were treated with 30 μΜ of 3-BDB for 1 h, and then the cells were stimulated by PM<sub>2.5</sub> for 24 h, the proteins were separated from cells, and (<b>a</b>) phospho-MEK1/2, MEK1/2, phospho-ERK1/2, ERK2, phospho-SEK1, SEK1, phospho-JNK, JNK, phospho-p38, p38, as well as (<b>b</b>) c-Fos, phospho-c-Jun, and c-Jun expressions were detected by Western blot. (<b>a</b>,<b>b</b>) * <span class="html-italic">p</span> &lt; 0.05 and <sup>#</sup> <span class="html-italic">p</span> &lt; 0.05 compared to control cells and PM<sub>2.5</sub>-exposed cells, respectively.</p>
Full article ">Figure 6
<p>Inhibition of PM<sub>2.5</sub>-induced pro-inflammatory cytokines and matrix metalloproteinases was performed by 3-BDB in vitro and in vivo. Cells were treated with 30 μΜ of 3-BDB for 1 h and then were exposed to 50 μg/mL of PM<sub>2.5</sub> for 24 h. Mice skin was treated with 3-BDB and PM<sub>2.5</sub> according to the animal experiment in Materials and Methods. (<b>a</b>,<b>b</b>) IL-1β and IL-6 concentrations in HaCaT cells were assessed using a human IL-1β and IL-6 ELISA kits, respectively. (<b>c</b>–<b>f</b>) From the proteins of (<b>c</b>,<b>e</b>) cells and (<b>d</b>,<b>f</b>) tissues, IL-1β, IL-6, MMP-1, MMP-2, and MMP-9 were examined by Western blot. (<b>g</b>) Senescence cells were available to visualize under a confocal microscope. (<b>a</b>–<b>g</b>) * <span class="html-italic">p</span> &lt; 0.05 and <sup>#</sup> <span class="html-italic">p</span> &lt; 0.05 compared to control groups and PM<sub>2.5</sub>-exposed groups, respectively.</p>
Full article ">Figure 6 Cont.
<p>Inhibition of PM<sub>2.5</sub>-induced pro-inflammatory cytokines and matrix metalloproteinases was performed by 3-BDB in vitro and in vivo. Cells were treated with 30 μΜ of 3-BDB for 1 h and then were exposed to 50 μg/mL of PM<sub>2.5</sub> for 24 h. Mice skin was treated with 3-BDB and PM<sub>2.5</sub> according to the animal experiment in Materials and Methods. (<b>a</b>,<b>b</b>) IL-1β and IL-6 concentrations in HaCaT cells were assessed using a human IL-1β and IL-6 ELISA kits, respectively. (<b>c</b>–<b>f</b>) From the proteins of (<b>c</b>,<b>e</b>) cells and (<b>d</b>,<b>f</b>) tissues, IL-1β, IL-6, MMP-1, MMP-2, and MMP-9 were examined by Western blot. (<b>g</b>) Senescence cells were available to visualize under a confocal microscope. (<b>a</b>–<b>g</b>) * <span class="html-italic">p</span> &lt; 0.05 and <sup>#</sup> <span class="html-italic">p</span> &lt; 0.05 compared to control groups and PM<sub>2.5</sub>-exposed groups, respectively.</p>
Full article ">Figure 7
<p>The schematic diagram for the protective effect of 3-BDB induced by PM<sub>2.5</sub> was exhibited. 3-BDB inhibited ROS generation induced by PM<sub>2.5</sub>, which caused macromolecular damage, cell cycle arrest, and apoptosis. In addition, 3-BDB inhibited inflammatory cytokines release and MMPs expression through the MAPK signaling pathway, thus alleviating cell senescence through a complex intracellular mechanism.</p>
Full article ">
20 pages, 2855 KiB  
Article
Comparative Analysis of Hot and Cold Brews from Single-Estate Teas (Camellia sinensis) Grown across Europe: An Emerging Specialty Product
by Patricia Carloni, Alfonso Albacete, Purificación A. Martínez-Melgarejo, Federico Girolametti, Cristina Truzzi and Elisabetta Damiani
Antioxidants 2023, 12(6), 1306; https://doi.org/10.3390/antiox12061306 - 20 Jun 2023
Cited by 6 | Viewed by 3627
Abstract
Tea is grown around the world under extremely diverse geographic and climatic conditions, namely, in China, India, the Far East and Africa. However, recently, growing tea also appears to be feasible in many regions of Europe, from where high-quality, chemical-free, organic, single-estate teas [...] Read more.
Tea is grown around the world under extremely diverse geographic and climatic conditions, namely, in China, India, the Far East and Africa. However, recently, growing tea also appears to be feasible in many regions of Europe, from where high-quality, chemical-free, organic, single-estate teas have been obtained. Hence, the aim of this study was to characterize the health-promoting properties in terms of the antioxidant capacity of traditional hot brews as well as cold brews of black, green and white teas produced across the European territory using a panel of antioxidant assays. Total polyphenol/flavonoid contents and metal chelating activity were also determined. For differentiating the characteristics of the different tea brews, ultraviolet-visible (UV-Vis) spectroscopy and ultra-high performance liquid chromatography coupled with high-resolution mass spectrometry were employed. Overall, our findings demonstrate for the first time that teas grown in Europe are good quality teas that are endowed with levels of health-promoting polyphenols and flavonoids and that have an antioxidant capacity similar to those grown in other parts of the world. This research is a vital contribution to the characterization of European teas, providing essential and important information for both European tea growers and consumers, and could be of guidance and support for the selection of teas grown in the old continent, along with having the best brewing conditions for maximizing the health benefits of tea. Full article
Show Figures

Figure 1

Figure 1
<p>Graphical distribution of tea micro-plantations across Europe (teacups). The fuchsia teacups indicate the origin of the teas that were studied in this work.</p>
Full article ">Figure 2
<p>Total polyphenol content (TPC) of the tea brews measured using Folin-Ciocalteu’s reagent. Bars are colored according to the type of tea (brown = black tea; green = green tea; grey = white tea) and the type of infusion (light shade = cold brew; dark shade = hot brew). The letters above the bars indicate the homogeneous sub-classes resulting from Tukey’s post hoc multiple comparison tests (<span class="html-italic">p</span> &lt; 0.05). Neth. = The Netherlands; Switz. = Switzerland.</p>
Full article ">Figure 3
<p>Antioxidant activity of the tea brews measured with the ORAC assay. Bars are colored according to the type of tea (brown = black tea; green = green tea; grey = white tea) and the type of brew (light shade = cold brew; dark shade = hot brew). The letters above the bars indicate the homogeneous sub-classes resulting from Tukey’s post hoc multiple comparison tests (<span class="html-italic">p</span> &lt; 0.05). Neth. = The Netherlands; Switz. = Switzerland.</p>
Full article ">Figure 4
<p>The metal chelating activity of the tea brews measured with the ferrozine assay. Bars are colored according to the type of tea (brown = black tea; green = green tea; grey = white tea) and the type of brew (light shade = cold brew; dark shade = hot brew). The. letters above the bars indicate the homogeneous sub-classes resulting from Tukey’s post hoc multiple comparison tests (<span class="html-italic">p</span> &lt; 0.05). Neth. = The Netherlands; Switz. = Switzerland.</p>
Full article ">Figure 5
<p>Averaged UV-Vis spectra (320–500 nm, Δλ = 5 nm) of the various cold and hot tea brews. Spectra are colored according to the type of tea (red = black tea; green = green tea; grey = white tea) and the type of infusion (light shade = cold brew; dark shade = hot brew).</p>
Full article ">Figure 6
<p>Score plots for the first two components resulting from the principal component analysis performed using the UV-Vis data (320–500 nm, Δλ = 5 nm) of the cold and hot brews. Dots are colored according to the type of tea (brown = black tea; green = green tea; grey = white tea) and the type of infusion (light shade = cold brew; dark shade = hot brew).</p>
Full article ">Figure 7
<p>Score plots by multigroup analysis and loading plot for the first two components resulting from the principal component analysis performed using the pairwise jobs of the metabolomic data of the green tea cold brews.</p>
Full article ">
15 pages, 4257 KiB  
Article
Hepatic Anti-Oxidative Genes CAT and GPX4 Are Epigenetically Modulated by RORγ/NRF2 in Alphacoronavirus-Exposed Piglets
by Haotian Gu, Yaya Liu, Yahui Zhao, Huan Qu, Yanhua Li, Abdelkareem A. Ahmed, Hao-Yu Liu, Ping Hu and Demin Cai
Antioxidants 2023, 12(6), 1305; https://doi.org/10.3390/antiox12061305 - 19 Jun 2023
Cited by 2 | Viewed by 2236
Abstract
As a member of alpha-coronaviruses, PEDV could lead to severe diarrhea and dehydration in newborn piglets. Given that lipid peroxides in the liver are key mediators of cell proliferation and death, the role and regulation of endogenous lipid peroxide metabolism in response to [...] Read more.
As a member of alpha-coronaviruses, PEDV could lead to severe diarrhea and dehydration in newborn piglets. Given that lipid peroxides in the liver are key mediators of cell proliferation and death, the role and regulation of endogenous lipid peroxide metabolism in response to coronavirus infection need to be illuminated. The enzymatic activities of SOD, CAT, mitochondrial complex-I, complex-III, and complex-V, along with the glutathione and ATP contents, were significantly decreased in the liver of PEDV piglets. In contrast, the lipid peroxidation biomarkers, malondialdehyde, and ROS were markedly elevated. Moreover, we found that the peroxisome metabolism was inhibited by the PEDV infection using transcriptome analysis. These down-regulated anti-oxidative genes, including GPX4, CAT, SOD1, SOD2, GCLC, and SLC7A11, were further validated by qRT-PCR and immunoblotting. Because the nuclear receptor RORγ-driven MVA pathway is critical for LPO, we provided new evidence that RORγ also controlled the genes CAT and GPX4 involved in peroxisome metabolism in the PEDV piglets. We found that RORγ directly binds to these two genes using ChIP-seq and ChIP-qPCR analysis, where PEDV strongly repressed the binding enrichments. The occupancies of histone active marks such as H3K9/27ac and H3K4me1/2, together with active co-factor p300 and polymerase II at the locus of CAT and GPX4, were significantly decreased. Importantly, PEDV infection disrupted the physical association between RORγ and NRF2, facilitating the down-regulation of the CAT and GPX4 genes at the transcriptional levels. RORγ is a potential factor in modulating the CAT and GPX4 gene expressions in the liver of PEDV piglets by interacting with NRF2 and histone modifications. Full article
Show Figures

Figure 1

Figure 1
<p>Oxidative stress was observed after being infected by PEDV. (<b>A</b>,<b>B</b>) The activities of anti–oxidant enzymes SOD and CAT were analyzed and normalized to tissue weight. (<b>C</b>,<b>D</b>) The contents of GSH and MDA were analyzed and normalized to tissue weight. (<b>E</b>) The relative ROS in the PEDV–infected piglets were analyzed. (<b>F</b>–<b>I</b>) The relative parameters of mitochondria ATP, enzyme complex–I, complex–III, and complex–V were analyzed. (*) <span class="html-italic">p</span> &lt; 0.05, (**) <span class="html-italic">p</span> &lt; 0.01 and (***) <span class="html-italic">p</span> &lt; 0.001 compared with the uninfected sample. The circles represent distribution of results for different samples.</p>
Full article ">Figure 2
<p>The transcriptional profiling showed that the anti–oxidative pathway was modulated by PEDV infection. (<b>A</b>) Gene ontology (GO) analysis showed that the anti–oxidative metabolic process is one of the most enriched pathways. (<b>B</b>) Volcano plot for the hepatic transcriptome of anti–oxidation measured by RNA–seq analysis. (<b>C</b>) The GSEA plot of the differentially expressed genes in the anti–oxidative pathway of the two groups. FDR, false–discovery rate.</p>
Full article ">Figure 3
<p>The expressions of genes included in anti–oxidation and fatty acid oxidation are inhibited by PEDV. (<b>A</b>) Heatmap of mRNA expression changes of the anti–oxidant metabolism genes from the data of RNA–seq. (<b>B</b>) The qRT–PCR analysis confirmed the mRNA expression changes of the anti–oxidant metabolism genes after being infected by PEDV. (<b>C</b>) Heatmap of mRNA expression changes of the fatty acid oxidation genes from the data of RNA–seq. (*) <span class="html-italic">p</span> &lt; 0.05 compared with the uninfected sample. The circles represent distribution of results for different samples.</p>
Full article ">Figure 4
<p>The expressions of RORγ and NRF2 in livers were dysregulated after being infected by PEDV. (<b>A</b>,<b>B</b>) Fragments Per Kilobase of transcript per Million mapped reads of RORC (encoding RORγ) and NRF2 were measured by RNA–seq and the mRNA expressions were measured by qRT–PCR. (<b>C</b>,<b>D</b>) The expression of RORγ and NRF2 at the protein level was evaluated via Western blotting. The intensity of these two proteins was normalized to GAPDH protein contents. (*) <span class="html-italic">p</span> &lt; 0.05 and (***) <span class="html-italic">p</span> &lt; 0.001 compared with the uninfected sample. The circles represent distribution of results for different samples.</p>
Full article ">Figure 5
<p>RORγ binding occupancies were reduced in the PEDV piglets. (<b>A</b>) ChIP–seq profiles (top) and heatmaps of ChIP–seq signal intensity (bottom) of RORγ within ±10 kb windows around the center of peak regions. (<b>B</b>) ChIP–seq signal visualization of RORγ at representative anti–oxidant metabolism genes <span class="html-italic">CAT</span> and <span class="html-italic">GPX4</span>, the numbers coupled with the group names of Uninfected or PEDV represent the maximum range of data. (<b>C</b>) Schematic diagram depicting the putative RORE sequence on the enhancers of <span class="html-italic">CAT</span> and <span class="html-italic">GPX4</span>.</p>
Full article ">Figure 6
<p>PEDV infection decreased the enrichment of RORγ and NRF2 at target loci of <span class="html-italic">CAT</span> and <span class="html-italic">GPX4</span> genes and their physical interaction. (<b>A</b>–<b>D</b>) ChIP–qPCR analyses of RORγ and NRF2 occupancies at the locus of <span class="html-italic">CAT</span>, <span class="html-italic">GPX4</span>, <span class="html-italic">SOD2</span>, and <span class="html-italic">GSS</span>, normalized to IgG. (<b>E</b>,<b>F</b>) ChIP–re–ChIP analysis of the combined binding of RORγ and NRF2 at the locus of <span class="html-italic">CAT</span> and <span class="html-italic">GPX4</span>. (*) <span class="html-italic">p</span> &lt; 0.05 and (***) <span class="html-italic">p</span> &lt; 0.001 compared with the uninfected sample. The circles represent distribution of results for different samples.</p>
Full article ">Figure 7
<p>PEDV infection modifies histone modification at the locus of <span class="html-italic">CAT</span> and <span class="html-italic">GPX4</span>. (<b>A</b>–<b>F</b>) The relative enrichment of histone marks’ (H3K4me1/2/3, H3K9ac, H3K18ac, H3K27ac) occupancy was analyzed by ChIP–qPCR. (<b>G</b>,<b>H</b>) The relative enrichment of co–activator p300 and RNA polymerase II at the locus of <span class="html-italic">CAT</span> and <span class="html-italic">GPX4</span>. (*) <span class="html-italic">p</span> &lt; 0.05 and (***) <span class="html-italic">p</span> &lt; 0.001 compared with the uninfected sample. The circles represent distribution of results for different samples.</p>
Full article ">
16 pages, 4519 KiB  
Article
Glycolysis Aids in Human Lens Epithelial Cells’ Adaptation to Hypoxia
by Yuxin Huang, Xiyuan Ping, Yilei Cui, Hao Yang, Jing Bao, Qichuan Yin, Hailaiti Ailifeire and Xingchao Shentu
Antioxidants 2023, 12(6), 1304; https://doi.org/10.3390/antiox12061304 - 19 Jun 2023
Cited by 2 | Viewed by 1650
Abstract
Hypoxic environments are known to trigger pathological damage in multiple cellular subtypes. Interestingly, the lens is a naturally hypoxic tissue, with glycolysis serving as its main source of energy. Hypoxia is essential for maintaining the long-term transparency of the lens in addition to [...] Read more.
Hypoxic environments are known to trigger pathological damage in multiple cellular subtypes. Interestingly, the lens is a naturally hypoxic tissue, with glycolysis serving as its main source of energy. Hypoxia is essential for maintaining the long-term transparency of the lens in addition to avoiding nuclear cataracts. Herein, we explore the complex mechanisms by which lens epithelial cells adapt to hypoxic conditions while maintaining their normal growth and metabolic activity. Our data show that the glycolysis pathway is significantly upregulated during human lens epithelial (HLE) cells exposure to hypoxia. The inhibition of glycolysis under hypoxic conditions incited endoplasmic reticulum (ER) stress and reactive oxygen species (ROS) production in HLE cells, leading to cellular apoptosis. After ATP was replenished, the damage to the cells was not completely recovered, and ER stress, ROS production, and cell apoptosis still occurred. These results suggest that glycolysis not only performs energy metabolism in the process of HLE cells adapting to hypoxia, but also helps them continuously resist cell apoptosis caused by ER stress and ROS production. Furthermore, our proteomic atlas provides possible rescue mechanisms for cellular damage caused by hypoxia. Full article
Show Figures

Figure 1

Figure 1
<p><b>HLE cells adapt to a hypoxic environment.</b> (<b>A</b>) Light photomicrographs of HLECs under hypoxic and normoxic conditions. Scale bars: 500 μm. (<b>B</b>) CCK8 assay was used to detect the cell viability, and the optical density (OD) value of the hypoxia and normoxia conditions showed no statistically significant change. Cell viability is expressed as the mean ± standard deviation of three independent experiments (ns: <span class="html-italic">p</span> &gt; 0.05). (<b>C</b>) Schematic diagram of the proteomics process.</p>
Full article ">Figure 2
<p><b>Proteomic profiles of HLE cells under hypoxia and normoxia conditions.</b> TMT-based proteomic analysis on HLECs subjected to either hypoxia (HYP) or normoxia (CON) conditions. (<b>A</b>) Volcano analysis representing the significantly differentially expressed proteins (DEPs). (<b>B</b>) The heatmap demonstrates that expression patterns were altered under hypoxia. Each column represents a sample, and each row represents the expression levels of a single proteomic in various samples. The colour scale of the heatmap ranges from blue (low expression) to red (high expression). Top 10 enrichment analysis of GO terms for the differentially expressed proteins. The upregulated protein terms (<b>C</b>) in the biological process (BP), cellular component (CC), and molecular function (MF) categories are depicted. (<b>D</b>) KEGG based on upregulated differentially expressed proteins (DEPs) annotated based on proteomic analysis.</p>
Full article ">Figure 3
<p><b>Validation of glycolysis pathway in HLE cells.</b> Images represent protein levels of glycolysis-related proteins in HLECs under hypoxia and normoxia conditions. The protein levels of HK1, HK2, PGK1, ENO2, and LDHA were analysed via Western blot (<b>A</b>). (<b>B</b>) Bar graph shows quantification. Results are combined data from four experiments with different cell preparations, and each value represents mean ± SEM; *: <span class="html-italic">p</span> &lt; 0.05, **: <span class="html-italic">p</span> &lt; 0.01, ns: <span class="html-italic">p</span> &gt; 0.05. (<b>C</b>) Light photomicrographs of HLECs cultured with or without glycolysis inhibitor 2DG under hypoxia conditions. Scale bars: 500 μm. (<b>D</b>) CCK8 assay was used to detect the cell viability and the optical density (OD) values of HLECs; those cultured with glycolysis inhibitor 2DG have significantly lower values under hypoxia conditions. Images represent protein levels of glycolysis-related proteins in HLECs cultured with or without glycolysis inhibitor 2DG under hypoxia conditions (***: <span class="html-italic">p &lt;</span> 0.001). The protein levels of HK1, HK2, PGK1, ENO2, and LDHA were analysed via Western blot (<b>E</b>). (<b>F</b>) Bar graph shows quantification. Results are combined data from four experiments with different cell preparations, and each value represents mean ± SEM; *: <span class="html-italic">p</span> &lt; 0.05, **: <span class="html-italic">p</span> &lt; 0.01, ns: <span class="html-italic">p</span> &gt; 0.05.</p>
Full article ">Figure 4
<p><b>Inhibition of glycolysis induced the upregulation of ROS-, ER stress-, and apoptosis-related proteins.</b> The expression of ER stress markers (GRP94, GRP78, CHOP, XBP1), ROS markers (NRF2, NOX4, SOD1, CAT), and apoptosis markers (BCL2, BAX, CYCS, CASP1, CASP3, CASP4, CASP7, CASP8, CASP9) in HLECs was measured after culturing with or without glycolysis inhibitor 2-deoxyglucose (2DG) under hypoxia conditions by qRT-PCR (<b>A</b>–<b>C</b>). Images represent protein levels of ER stress markers, apoptosis markers, and ROS markers in HLECs cultured with or without glycolysis inhibitor 2-deoxyglucose (2DG) under hypoxia conditions (<b>D</b>,<b>F</b>,<b>H</b>). The protein levels of GRP78, CHOP, XBP1, NRF2, NOX4, CASP3, CASP4, CASP9, and CYCS were analysed via Western blot; bar graph shows quantification (<b>E</b>,<b>G</b>,<b>I</b>). Results are combined data from four experiments with different cell preparations, and each value represents mean ± SEM; *: <span class="html-italic">p</span> &lt; 0.05, **: <span class="html-italic">p</span> &lt; 0.01.</p>
Full article ">Figure 5
<p><b>Inhibition of glycolysis induced ROS, ER stress, and apoptosis.</b> (<b>A</b>) Representative immunofluorescence images of GRP78 (red), CHOP (red), XBP1 (green), and DAPI (blue) in HLE cells treated or untreated with 2DG under a hypoxia environment (magnification 40×). Scale bars: 50 μm. (<b>B</b>) The average fluorescence intensities of GRP78, CHOP, and XBP1 in immunofluorescence experiments (<span class="html-italic">n</span> = 3, &gt;30 cells per experiment, *: <span class="html-italic">p</span> &lt; 0.05, **: <span class="html-italic">p</span> &lt; 0.01). (<b>C</b>,<b>D</b>) Representative fluorescence images depicting the intracellular Annexin V-FITC and ROS levels of HLE cells treated or untreated with 2DG under a hypoxia environment (magnification 40×). Scale bars: 100 μm. (<b>E</b>) The average fluorescence intensities of Annexin V-FITC and ROS in HLE cells (<span class="html-italic">n</span> = 3, &gt;30 cells per experiment, *: <span class="html-italic">p</span> &lt; 0.05).</p>
Full article ">Figure 6
<p><b>Replenishing ATP did not rescue damage caused by glycolytic inhibition in HLE cells.</b> (<b>A</b>) Representative immunofluorescence images of GRP78 (green), CHOP (green), XBP1 (red), and DAPI (blue) in HLE cells treated with 2DG/2DG + ATP or untreated under a hypoxia environment (magnification 40×). Scale bars: 50 μm. (<b>B</b>) The average fluorescence intensities of GRP78, CHOP, and XBP1 in immunofluorescence experiments (<span class="html-italic">n</span> = 3, &gt;30 cells per experiment, *: <span class="html-italic">p</span> &lt; 0.05, **: <span class="html-italic">p</span> &lt; 0.01). (<b>C</b>,<b>D</b>) Representative fluorescence images depicting the intracellular Annexin V-FITC and ROS levels of HLE cells treated with 2DG/2DG + ATP or untreated with 2DG under a hypoxia environment (magnification 40×). Scale bars: 100 μm. (<b>E</b>) The average fluorescence intensities of Annexin V-FITC and ROS in HLE cells (<span class="html-italic">n</span> = 3, &gt;30 cells per experiment, *: <span class="html-italic">p</span> &lt; 0.05, **: <span class="html-italic">p</span> &lt; 0.01).</p>
Full article ">
23 pages, 4966 KiB  
Article
Effects of Dietary Oleacein Treatment on Endothelial Dysfunction and Lupus Nephritis in Balb/C Pristane-Induced Mice
by Rocío Muñoz-García, Marina Sánchez-Hidalgo, Manuel Alcarranza, María Victoria Vazquéz-Román, María Alvarez de Sotomayor, María Luisa González-Rodríguez, María C. de Andrés and Catalina Alarcón-de-la-Lastra
Antioxidants 2023, 12(6), 1303; https://doi.org/10.3390/antiox12061303 - 19 Jun 2023
Cited by 1 | Viewed by 1808
Abstract
Systemic lupus erythematosus (SLE) is a chronic immune-inflammatory disease characterized by multiorgan affectation and lowered self-tolerance. Additionally, epigenetic changes have been described as playing a pivotal role in SLE. This work aims to assess the effects of oleacein (OLA), one of the main [...] Read more.
Systemic lupus erythematosus (SLE) is a chronic immune-inflammatory disease characterized by multiorgan affectation and lowered self-tolerance. Additionally, epigenetic changes have been described as playing a pivotal role in SLE. This work aims to assess the effects of oleacein (OLA), one of the main extra virgin olive oil secoiridoids, when used to supplement the diet of a murine pristane-induced SLE model. In the study, 12-week-old female BALB/c mice were injected with pristane and fed with an OLA-enriched diet (0.01 % (w/w)) for 24 weeks. The presence of immune complexes was evaluated by immunohistochemistry and immunofluorescence. Endothelial dysfunction was studied in thoracic aortas. Signaling pathways and oxidative-inflammatory-related mediators were evaluated by Western blotting. Moreover, we studied epigenetic changes such as DNA methyltransferase (DNMT-1) and micro(mi)RNAs expression in renal tissue. Nutritional treatment with OLA reduced the deposition of immune complexes, ameliorating kidney damage. These protective effects could be related to the modulation of mitogen-activated protein kinases, the Janus kinase/signal transducer and transcription activator of transcription, nuclear factor kappa, nuclear-factor-erythroid-2-related factor 2, inflammasome signaling pathways, and the regulation of miRNAs (miRNA-126, miRNA-146a, miRNA-24-3p, and miRNA-123) and DNMT-1 expression. Moreover, the OLA-enriched diet normalized endothelial nitric oxide synthase and nicotinamide adenine dinucleotide phosphate (NADPH) oxidase-1 overexpression. These preliminary results suggest that an OLA-supplemented diet could constitute a new alternative nutraceutical therapy in the management of SLE, supporting this compound as a novel epigenetic modulator of the immunoinflammatory response. Full article
(This article belongs to the Special Issue Olive Tree Products and Antioxidants)
Show Figures

Graphical abstract

Graphical abstract
Full article ">Figure 1
<p>Immunomodulatory effect of an OLA-supplemented diet on the weight of the spleen and thymus organs 6 months after pristane induction. Data are expressed as the mean ± SEM (<span class="html-italic">n</span> = 10). One-way ANOVA followed by Tukey’s post hoc test results: ## <span class="html-italic">p</span> &lt; 0.01 vs. naïve control group; * <span class="html-italic">p</span> &lt; 0.05 vs. SD-pristane group.</p>
Full article ">Figure 2
<p>Histological effects of OLA dietary treatment on pristane-induced SLE mice. Representative images of (<b>A</b>–<b>C</b>) H&amp;E staining; (<b>D</b>–<b>F</b>) PAS, and (<b>G</b>–<b>I</b>) MT stains from renal tissues from mice. Normal histology architecture in the cortex and medullary renal naïve mice with low stroma between renal tubules (<b>A</b>,<b>G</b>) and normal glomeruli in the cortex parenchyma (<b>D</b>, arrows). Compressed glomeruli in SD-Pristane mice (<b>B</b>, arrows), probably due to Bowman space proliferation with the partial sclerosis of glomerular tufts. Inflammatory cells occupying the medullary interstitium (<b>E</b>, arrows). Fibrotic tissue characterized by an abundance of collagen along with hemorrhagic signs identified in the medullary stroma (<b>H</b>). OLA-treated mice showed no histopathological changes either in the medullary or renal cortex. Normal distribution of renal tubules in medullary parenchyma (<b>C</b>), two glomeruli showing regular size and shape (<b>F</b>, arrows), and a renal cortex with no fibrosis or pathologic changes (<b>I</b>). Scale bar, 25 µm (<b>A</b>–<b>I</b>). Scale bar, 100 µm (<b>H</b> black-dotted squared).</p>
Full article ">Figure 3
<p>An OLA-supplemented diet reduced IgG and IgM deposits in the kidney tissue of pristane-induced SLE mice. Representative figures of (<b>A</b>–<b>F</b>) IHC and (<b>G</b>–<b>L</b>) IF staining of IgG or IgM in kidneys from pristane-induced SLE mice. (<b>A</b>,<b>D</b>,<b>G</b>,<b>J</b>) Absence of deposits of IgG and IgM in the naïve group. (<b>B</b>,<b>E</b>,<b>H</b>,<b>K</b>) Marked IgG and IgM deposits in the renal glomeruli of SD-pristane mice. (<b>B</b>,<b>E</b>, black arrows) Deposits in the capillary loops and mesangial cells and into the tubular interstices (extraglomerular nephritis). Deposits in capillary loops (<b>H</b>, white arrow) and mesangial deposits (<b>H</b>, yellow arrow) (mesangio—capillary glomerulonephritis). They were also accompanied by extraglomerular deposits. Mesangial glomerulonephritis (<b>K</b>, yellow arrow) also appeared with intracapillary deposits (<b>K</b>, white arrow) and tubular involvement. Scale bar, 25 µm.</p>
Full article ">Figure 4
<p>Dietary OLA administration reduced COX-2, iNOS, and mPGES-1 protein overexpression. Protein expression was measured in total kidney homogenates from mice. Densitometry was performed after normalization to the control (β-actin housekeeping gene). One-way ANOVA followed by Tukey’s post hoc test results: ## <span class="html-italic">p</span> &lt; 0.01, ### <span class="html-italic">p</span> &lt; 0.001 vs. naïve group; ** <span class="html-italic">p</span> &lt; 0.01, *** <span class="html-italic">p</span> &lt; 0.001 vs. SD-pristane group. Data were represented as the mean ± SEM (<span class="html-italic">n</span> = 6).</p>
Full article ">Figure 5
<p>Role of the OLA dietary treatment in the regulation of SLE-related intracellular signaling pathways. Protein expressions of (<b>A</b>) phosphorylated JAK3/STAT-3; (<b>B</b>) the Nrf-2/HO-1 axis; (<b>C</b>) the IκB-α and p65/p50 nuclear subunits, and (<b>D</b>) phosphorylated P38, JNK, and ERK MAPKs were evaluated in renal homogenates from mice. Densitometry was performed after normalization to the control (JNK, ERK, p38, or β-actin). Data are represented as the means ± SEM (<span class="html-italic">n</span> = 6). One-way ANOVA followed by Tukey’s post hoc test results: # <span class="html-italic">p</span> &lt; 0.05, ## <span class="html-italic">p</span> &lt; 0.01, ### <span class="html-italic">p</span> &lt; 0.001 vs. naïve group; * <span class="html-italic">p</span> &lt; 0.05, ** <span class="html-italic">p</span> &lt; 0.01, *** <span class="html-italic">p</span> &lt; 0.001 significant difference versus SD-pristane group.</p>
Full article ">Figure 6
<p>The inflammasome signaling pathway was downregulated by an OLA-supplemented diet. (<b>A</b>) NLRP3/ASC, (<b>B</b>) Caspase 11, (<b>C</b>) IL-18, and (<b>D</b>) Caspase 1. Protein expressions were analyzed by immunoblots in kidney lysates. Densitometry was performed following normalization to the control (β-actin housekeeping gene). Data are represented as the means ± SEM (<span class="html-italic">n</span> = 6). One-way analysis of variance (ANOVA), using Tukey–Kramer multiple comparisons test as post hoc test: # <span class="html-italic">p</span> &lt; 0.05, ## <span class="html-italic">p</span> &lt; 0.01, ### <span class="html-italic">p</span> &lt; 0.001 vs. naïve group; * <span class="html-italic">p</span> &lt; 0.05 ** <span class="html-italic">p</span> &lt; 0.01; significant difference versus SD-pristane group.</p>
Full article ">Figure 7
<p>The OLA-supplemented diet enhanced redox-sensitive effectors and endothelial function in pristane-induced mice. (<b>B</b>) NOX-1 and (<b>C</b>) penos Thr495 protein expression quantification in homogenates of aortic rings from mice. Data are represented as the mean ± SEM (<span class="html-italic">n</span> = 6). β-Actin was used as a loading control. One-way ANOVA followed by Tukey’s post hoc test results: # <span class="html-italic">p</span> &lt; 0.05, ### <span class="html-italic">p</span> &lt; 0.001 vs. naïve group; * <span class="html-italic">p</span> &lt; 0.05, vs. SD-pristane group. (<b>A</b>) Endothelial relaxation was induced by ACh (0.001–10 mM) in U46619 (0.003 mM) precontracted intact aorta rings from mice. The Ach-induced relaxant responses are expressed as a percentage of precontraction induced by U46619 (<span class="html-italic">n</span> = 7 per group). One-way ANOVA followed by Fisher’s LSD test results: * <span class="html-italic">p</span> &lt; 0.05; ** <span class="html-italic">p</span> &lt; 0.01 significant difference versus naïve control group (U46619), thromboxane receptor agonist 9,11-didesoxi-11α,9α-epoximetanoprostaglandina F2α.</p>
Full article ">Figure 8
<p>OLA dietary treatment modulated miRNAs and dnmt1 expression in pristane-induced nephritis. The relative renal expression of miRNAs: (<b>A</b>) miRNA-23b, (<b>B</b>) miRNA-146a, (<b>C</b>) miRNA-24-3p, and (<b>D</b>) miRNA-126, as well as (<b>E</b>) DNMT-1 mRNA from mice of the naïve, SD-pristane, and OLA-enriched diet groups, evaluated by RT-qPCR. Expression was normalized with MammU6 or β-actin, (<span class="html-italic">n</span> = 8). One-way ANOVA followed by Tukey’s post hoc test results: # <span class="html-italic">p</span> &lt; 0.05, ## <span class="html-italic">p</span> &lt; 0.01; ### <span class="html-italic">p</span> &lt; 0.001 vs. naïve group; * <span class="html-italic">p</span> &lt; 0,05, ** <span class="html-italic">p</span> &lt; 0.01; *** <span class="html-italic">p</span> &lt; 0.001, significant difference versus SD-pristane group.</p>
Full article ">
13 pages, 1236 KiB  
Review
Redox Mechanisms Underlying the Cytostatic Effects of Boric Acid on Cancer Cells—An Issue Still Open
by Giulia Paties Montagner, Silvia Dominici, Simona Piaggi, Alfonso Pompella and Alessandro Corti
Antioxidants 2023, 12(6), 1302; https://doi.org/10.3390/antiox12061302 - 19 Jun 2023
Cited by 4 | Viewed by 1821
Abstract
Boric acid (BA) is the dominant form of boron in plasma, playing a role in different physiological mechanisms such as cell replication. Toxic effects have been reported, both for high doses of boron and its deficiency. Contrasting results were, however, reported about the [...] Read more.
Boric acid (BA) is the dominant form of boron in plasma, playing a role in different physiological mechanisms such as cell replication. Toxic effects have been reported, both for high doses of boron and its deficiency. Contrasting results were, however, reported about the cytotoxicity of pharmacological BA concentrations on cancer cells. The aim of this review is to briefly summarize the main findings in the field ranging from the proposed mechanisms of BA uptake and actions to its effects on cancer cells. Full article
(This article belongs to the Section Health Outcomes of Antioxidants and Oxidative Stress)
Show Figures

Figure 1

Figure 1
<p>Overview of the proposed mechanisms of action of BA. The mechanisms underlying the apparent dichotomous functions of BA and—more importantly—the possibility of reaching suitably high concentrations of BA in the tumor microenvironment need to be deeply investigated.</p>
Full article ">Figure 2
<p>Overview of the proposed effects of boron on NAD<sup>+</sup> and possible consequences on cellular antioxidant defenses. B, boron; CAT, catalase; Cys, cysteine; CysGly, cysteinylglycine; DP, dipeptidase; GCL, glutamate cysteine ligase; GGT, gamma-glutamyl trasnferase; GS, glutathione synthetase; Glu, glutamate; Gly, glycine; GPx, glutathione peroxidase; GR, glutathione reductase; GSH, reduced glutathione; GSSG, glutathione disulfide; G6P, glucose 6−phosphate; G6PGL, glucono 1,5−lactone 6−phosphate; G6PD, glucose−6−phosphate dehydrogenase; Lipid−OOH, lipid peroxide; Lipid−OH, lipid alcohol; NAD<sup>+</sup>, nicotinamide adenine dinucleotide (oxidized form); NADH, nicotinamide adenine dinucleotide (reduced form); NADK, nicotinamide adenine dinucleotide kinase; NADP<sup>+</sup>, nicotinamide adenine dinucleotide phosphate (oxidized form); NADPH, nicotinamide adenine dinucleotide phosphate (reduced form); NAM, nicotinamide; Nrf2, nuclear factor erythroid 2−related factor 2; PARPs, poly (adenosine diphosphate-ribose) polymerases; PRx, peroxiredoxin; SIRTs, sirtuins; Trx, thioredoxin; TrxR, thioredoxin reductase; Trx(SH)2, thioredoxin reduced form.</p>
Full article ">
18 pages, 7027 KiB  
Article
Phaeanthus vietnamensis Ban Ameliorates Lower Airway Inflammation in Experimental Asthmatic Mouse Model via Nrf2/HO-1 and MAPK Signaling Pathway
by Thi Van Nguyen, Chau Tuan Vo, Van Minh Vo, Cong Thuy Tram Nguyen, Thi My Pham, Chun Hua Piao, Yan Jing Fan, Ok Hee Chai and Thi Tho Bui
Antioxidants 2023, 12(6), 1301; https://doi.org/10.3390/antiox12061301 - 19 Jun 2023
Cited by 1 | Viewed by 1740
Abstract
Asthma is a chronic airway inflammatory disease listed as one of the top global health problems. Phaeanthus vietnamensis BÂN is a well-known medicinal plant in Vietnam with its anti-oxidant, anti-microbial, anti-inflammatory potential, and gastro-protective properties. However, there is no study about P. vietnamensis [...] Read more.
Asthma is a chronic airway inflammatory disease listed as one of the top global health problems. Phaeanthus vietnamensis BÂN is a well-known medicinal plant in Vietnam with its anti-oxidant, anti-microbial, anti-inflammatory potential, and gastro-protective properties. However, there is no study about P. vietnamensis extract (PVE) on asthma disease. Here, an OVA-induced asthma mouse model was established to evaluate the anti-inflammatory and anti-asthmatic effects and possible mechanisms of PVE. BALB/c mice were sensitized by injecting 50 μg OVA into the peritoneal and challenged by nebulization with 5% OVA. Mice were orally administered various doses of PVE once daily (50, 100, 200 mg/kg) or dexamethasone (Dex; 2.5 mg/kg) or Saline 1 h before the OVA challenge. The cell infiltrated in the bronchoalveolar lavage fluid (BALF) was analyzed; levels of OVA-specific immunoglobulins in serum, cytokines, and transcription factors in the BALF were measured, and lung histopathology was evaluated. PVE, especially PVE 200mg/kg dose, could improve asthma exacerbation by balancing the Th1/Th2 ratio, reducing inflammatory cells in BALF, depressing serum anti-specific OVA IgE, anti-specific OVA IgG1, histamine levels, and retrieving lung histology. Moreover, the PVE treatment group significantly increased the expressions of antioxidant enzymes Nrf2 and HO-1 in the lung tissue and the level of those antioxidant enzymes in the BALF, decreasing the oxidative stress marker MDA level in the BALF, leading to the relieving the activation of MAPK signaling in asthmatic condition. The present study demonstrated that Phaeanthus vietnamensis BÂN, traditionally used in Vietnam as a medicinal plant, may be used as an efficacious agent for treating asthmatic disease. Full article
(This article belongs to the Special Issue Oxidative Stress and Lung Inflammation)
Show Figures

Graphical abstract

Graphical abstract
Full article ">Figure 1
<p>Animal experimental protocol. The male 5-week-old BALB/c mice were divided into 6 groups: group (1) Naïve; (2) OVA; (3) PVE 50; (4) PVE 100; (5) PVE 200; and (6) Dex. The asthma mouse model was established by sensitization OVA (i.p) on day 1 and day 15; then, the OVA challenge (nebulization) was performed from day 27 to day 29. The asthma mice in groups (3), (4), and (5) were orally administered with corresponding concentrations of PVE 50, 100, 200 mg/kg. The mice in the Dex group were treated with Dex 2.5 mg/kg. The mice in the OVA group were given Saline. All the mice were sacrificed on day 30. PVE = <span class="html-italic">Phaeanthus vietnamensis</span> extract; Dex = Dexamethasone; OVA = Ovalbumin.</p>
Full article ">Figure 2
<p>Analyzed compounds from Phaeanthus vietnamensis 70% ethanol extract with UPLC-Q-TOF-MS/MS. The main important biomolecules identified were spathulenol, neophytadiene, octadecanoic acid ethyl ester, n-hexadecanoic acid, hexadecanoic acid ethyl ester, oleic acid, linoleic acid ethyl ester, stigmasterol.</p>
Full article ">Figure 2 Cont.
<p>Analyzed compounds from Phaeanthus vietnamensis 70% ethanol extract with UPLC-Q-TOF-MS/MS. The main important biomolecules identified were spathulenol, neophytadiene, octadecanoic acid ethyl ester, n-hexadecanoic acid, hexadecanoic acid ethyl ester, oleic acid, linoleic acid ethyl ester, stigmasterol.</p>
Full article ">Figure 3
<p>PVE prevented the degranulation of RPMCs by compound 48/80. (<b>A</b>) MTT assay. (<b>B</b>) PVE protected RPMCs degranulation from C48/80. (<b>C</b>) The level of histamine released from RPMCs. (<b>D</b>) Inverted light microscopy of RPMCs. RPMCs (2 × 10<sup>5</sup> cells/well) were incubated with different concentrations of PVE (0.01, 0.1, and 1 mg/mL) at 37 °C for 3 h; then, absorbance was measured at 570 nm with a spectrophotometer. RPMCs were pretreated with PVE (10, 1, 0.1 mg/mL) or saline for 10 min at 37 °C and then incubated with C48/80 (5 µg/mL) or saline for 15 min. PVE dose-dependently inhibited the C48/80-induced RPMCs degranulation. The red arrow indicates degranulated mast cells. * compared to OVA; ###, *** <span class="html-italic">p</span> &lt; 0.00, * <span class="html-italic">p</span> &lt; 0.05. Scale bar = 25 µm. PVE = <span class="html-italic">Phaeanthus vietnamensis</span> extract; RPMC = Rat peritoneal mast cell; C48/80 = Compound 48/80.</p>
Full article ">Figure 4
<p>PVE reduced the recruitment of inflammatory cells in BALF of the OVA-induced asthmatic mouse model. (<b>A</b>) Diff-Quick stain. (<b>B</b>) The differential cells in BALF. The number of epithelial cells and inflammatory cells (macrophage, eosinophil, neutrophil) was significantly increased in BALF of the OVA group compared with the Naive group, and those cell numbers were significantly decreased in PVE 50, 100, 200 mg/kg, and Dex 2.5 mg/kg groups. The red arrows indicated eosinophils. # Compared to Naive; * compared to OVA. ###, *** <span class="html-italic">p</span> &lt; 0.001; ** <span class="html-italic">p</span> &lt; 0.01; #, * <span class="html-italic">p</span> &lt; 0.05. BALF = Bronchoalveolar lavage fluid; Dex = Dexamethasone; PVE = Phaeanthus vietnamensis extract; OVA = Ovalbumin. Scale bar = 50 µm.</p>
Full article ">Figure 5
<p>PVE prevented histopathological changes in the lung tissue of the OVA-induced asthmatic mouse model. (<b>A</b>) Histology of lung tissue. (<b>B)</b> Lung inflammatory score. (<b>C</b>) The density of PAS (+) area (%). (<b>D</b>) The density of collagen (+) area. (<b>E</b>) The density of α-SMA (+) area. All pictures were at a magnification of ×200. Via H&amp;E staining, typical inflammation features were observed in lung tissue of asthma mice. Via PAS staining, asthma mice showed an increase in goblet cells resulting in oversecreted mucus into the lumen of the bronchia (which was stained with purple color and indicated by purple arrows). Via Trichrome staining, the collagen fiber (which was stained with blue color and indicated by blue arrows) was abundantly expressed surrounding the bronchi and vessels in the lung tissue of asthma mice. IHC stain with α-SMA again confirmed the majored fibrosis state in the OVA group. However, the inflammation feature, goblet cell hyperplasia, and collagen fiber deposition in asthma mice were considerably attenuated with PVE. * compared to OVA. ###, *** <span class="html-italic">p</span> &lt; 0.001; ** <span class="html-italic">p</span> &lt; 0.01; * <span class="html-italic">p</span> &lt; 0.05. Dex = Dexamethasone; H&amp;E = hematoxylin and eosin; IHC = Immunohistochemistry; PAS = Periodic acid–Schiff; PVE = <span class="html-italic">Phaeanthus vietnamensis</span> extract; OVA = Ovalbumin.</p>
Full article ">Figure 6
<p>PVE reduced the levels of OVA-specific antibodies and histamine in the serum of the OVA-induced asthmatic mouse model. (<b>A</b>) OVA-specific IgE. (<b>B</b>) OVA-specific IgG<sub>2a</sub>. (<b>C</b>) OVA-specific IgG<sub>1</sub>. (<b>D</b>) Histamine in serum. Oral administration of PVE 200 mg/kg significantly down-regulated the levels of OVA-specific IgE, IgG<sub>1</sub>, and histamine in the serum. PVE also strongly up-regulated the level of OVA-specific IgG<sub>2a</sub> in the serum of asthma mice. * compared to OVA. ###, *** <span class="html-italic">p</span> &lt; 0.001; ** <span class="html-italic">p</span> &lt; 0.01; * <span class="html-italic">p</span> &lt; 0.05. BALF = Bronchoalveolar lavage fluid; Dex = Dexamethasone; Ig = Immunoglobulin; PVE = Phaeanthus vietnamensis extract; OVA = Ovalbumin.</p>
Full article ">Figure 7
<p>PVE notably increased Th1-related cytokines while decreasing Th2-related cytokines in the BALF of the OVA-induced asthmatic mouse model. The levels of Th2-related cytokines (<b>A</b>) IL-4, (<b>B</b>) IL-5, (<b>C</b>) IL-13, (<b>D</b>) eotaxin, and Th1-related cytokines (<b>E</b>) IL-12, (<b>F</b>) IFN-γ in BALF. The levels of IL-12 in BALF were significantly improved by PVE 200 mg/kg. However, PVE did not increase the level of IFN-γ. The levels of Th2 cytokines IL-4, IL-5, IL-13, and eotaxin in BALF were significantly suppressed by PVE 200 mg/kg or Dex 2.5 mg/kg. # compared to Naive; * compared to OVA. ### <span class="html-italic">p</span> &lt; 0.001; ##, ** <span class="html-italic">p</span> &lt; 0.01; * <span class="html-italic">p</span> &lt; 0.05. BALF = Bronchoalveolar lavage fluid; Dex = Dexamethasone; IFN-γ = Interferon-γ; IL = Interleukin; IL = Interleukin; PVE = Phaeanthus vietnamensis extract; OVA = Ovalbumin; Th = T helper.</p>
Full article ">Figure 8
<p>PVE enhanced the expression of Th1-transcription factor T-bet and decreased the Th2-transcription factor GATA-3 in BALF of the OVA-induced asthmatic mouse model. (<b>A</b>) Western blot data. The relative expression of (<b>B</b>) T-bet and (<b>C</b>) GATA-3 in the lung tissue. (<b>D</b>) The T-bet/GATA-3 expression ratio. The treatment of Dex or PVE at high dose of 200 mg/kg increased the expression of T-bet and suppressed the expression of GATA-3 in the lung tissue of asthmatic mice. It led to enhancing the T-bet/GATA-3 ratio, which is represented by Th1/Th2 ratio. * compared to OVA. ###, *** <span class="html-italic">p</span> &lt; 0.001; ##, ** <span class="html-italic">p</span> &lt; 0.01; * <span class="html-italic">p</span> &lt; 0.05. BALF = Bronchoalveolar lavage fluid; Dex = Dexamethasone; GATA-3 = GATA Binding Protein 3; IL = Interleukin; PVE = Phaeanthus vietnamensis extract; OVA = Ovalbumin; Th = T helper; T-bet = T-box expressed in T cells.</p>
Full article ">Figure 9
<p>PVE effectively regulated oxidative stress in the OVA-induced asthmatic mouse model. The levels of (<b>A</b>) MDA, (<b>B</b>) Nrf-2, and (<b>C</b>) HO-1 in BALF. (<b>D</b>) The expression of HO-1 in the lung tissue. (<b>E</b>) The relative level of HO-1 in the lung tissue. # Compared to Naive; * compared to OVA. ###, *** <span class="html-italic">p</span> &lt; 0.001; ** <span class="html-italic">p</span> &lt; 0.01; #, * <span class="html-italic">p</span> &lt; 0.05. BALF = Bronchoalveolar lavage fluid; Dex = Dexamethasone; HO-1 = heme oxygenase-1; MDA = Malondialdehyde; Nrf2 = nuclear factor erythroid 2–related factor 2; PVE = <span class="html-italic">Phaeanthus vietnamensis</span> extract; OVA = Ovalbumin.</p>
Full article ">Figure 10
<p>PVE suppressed the MAPK signaling in the OVA-induced asthmatic mouse model. (<b>A</b>) Western blot results of MAPKs related protein. The relative levels of (<b>B</b>) p-ERK/ERK. (<b>C</b>) p-JNK/JNK, and (<b>D</b>) p-P38/P-38. * compared to OVA. ### <span class="html-italic">p</span> &lt; 0.001; ** <span class="html-italic">p</span> &lt; 0.01; * <span class="html-italic">p</span> &lt; 0.05. BALF = Bronchoalveolar lavage fluid; Dex = Dexamethasone; MAPK = mitogen-activated protein kinases; PVE = <span class="html-italic">Phaeanthus vietnamensis</span> extract; OVA = Ovalbumin.</p>
Full article ">
12 pages, 1004 KiB  
Review
Effect of 8-Hydroxyguanine DNA Glycosylase 1 on the Function of Immune Cells
by Weiran Zhang, Ranwei Zhong, Xiangping Qu, Yang Xiang and Ming Ji
Antioxidants 2023, 12(6), 1300; https://doi.org/10.3390/antiox12061300 - 19 Jun 2023
Cited by 1 | Viewed by 1834
Abstract
Excess reactive oxygen species (ROS) can cause an imbalance between oxidation and anti-oxidation, leading to the occurrence of oxidative stress in the body. The most common product of ROS-induced base damage is 8-hydroxyguanine (8-oxoG). Failure to promptly remove 8-oxoG often causes mutations during [...] Read more.
Excess reactive oxygen species (ROS) can cause an imbalance between oxidation and anti-oxidation, leading to the occurrence of oxidative stress in the body. The most common product of ROS-induced base damage is 8-hydroxyguanine (8-oxoG). Failure to promptly remove 8-oxoG often causes mutations during DNA replication. 8-oxoG is cleared from cells by the 8-oxoG DNA glycosylase 1 (OGG1)-mediated oxidative damage base excision repair pathway so as to prevent cells from suffering dysfunction due to oxidative stress. Physiological immune homeostasis and, in particular, immune cell function are vulnerable to oxidative stress. Evidence suggests that inflammation, aging, cancer, and other diseases are related to an imbalance in immune homeostasis caused by oxidative stress. However, the role of the OGG1-mediated oxidative damage repair pathway in the activation and maintenance of immune cell function is unknown. This review summarizes the current understanding of the effect of OGG1 on immune cell function. Full article
(This article belongs to the Special Issue Cellular ROS and Antioxidants: Physiological and Pathological Role)
Show Figures

Figure 1

Figure 1
<p>The function and mechanism of OGG1. (<b>a</b>) Oxidative stress can lead to base mismatch. (<b>b</b>) OGG1-BER pathway diagram. (<b>c</b>) TH10785 can activate the new lyase function of OGG1. (<b>d</b>) The function and mechanism of OGG1 in innate immunity. (<b>e</b>) OGG1 can act as a signaling molecule to regulate DNA demethylation. Abbreviations: APE1, AP nucleic acid endonuclease 1; PNKP1: polynucleotide kinase phosphatase; GEF: Guanine-Nucleotide Exchange Factor; GDP, Guanosine diphosphate; GTP, Guanosine triphosphate; MAPK, Mitogen-activated protein kinases; PI3K, Phosphoinositide 3-Kinase; NF-κB, nuclear factor kappa-B pathwaypathway; IL-6, Interleukin-6; TNF-α, Tumor Necrosis Factor alpha; TET1, Tet Methylcytosine Dioxygenase 1; 5mC, 5-methylcytosine; 5hmC, 5-hydroxymethylcytosine (Created with Figdraw).</p>
Full article ">Figure 2
<p>The effect of OGG1-BER pathway on immune cells. OGG1 can initiate the BER pathway to repair oxidatively damaged DNA, a process that affects the function of immune cells, including macrophages, dendritic cells, granulocytes, and lymphocytes. Immune cells are regulated by participating in a variety of signaling pathways (Created with Figdraw).</p>
Full article ">
11 pages, 2739 KiB  
Communication
Relationships between Serum Biomarkers of Oxidative Stress and Tobacco Smoke Exposure in Patients with Mental Disorders
by Ana-Maria Vlasceanu, Daniela Gradinaru, Miriana Stan, Viorela G. Nitescu and Daniela Luiza Baconi
Antioxidants 2023, 12(6), 1299; https://doi.org/10.3390/antiox12061299 - 19 Jun 2023
Viewed by 2947
Abstract
The role of cigarette smoking as an aggravating factor of systemic oxidative stress in patients with mental disorders has not been extensively investigated, although significantly higher rates of smoking are recorded in these subjects in comparison with the general population. In the present [...] Read more.
The role of cigarette smoking as an aggravating factor of systemic oxidative stress in patients with mental disorders has not been extensively investigated, although significantly higher rates of smoking are recorded in these subjects in comparison with the general population. In the present study, we tested the hypothesis that smoking might be an exacerbator of systemic oxidative stress, being directly correlated with the degree of exposure to tobacco smoke. We analyzed, in 76 adult subjects from a public health care unit, the relationships between serum cotinine levels as a marker of tobacco smoke exposure, and three biomarkers of oxidative stress: the serum glutathione (GSH), the advanced oxidation protein products (AOPPs), and the total serum antioxidant status (FRAP). The results indicate that the degree of tobacco smoke exposure was inversely associated with GSH levels in both passive and active smokers, suggesting that smoke particulate components’ toxicity is associated with a systemic GSH depletion. Paradoxically, the lowest AOPP levels which were positively associated with GSH, were recorded in active smoking patients whereas in passive smokers individual values of AOPPs decreased along with the increase in GSH levels. Our data suggest that an enhanced inhalation of particulate constituents of cigarette smoke could induce critical changes in systemic redox homeostasis and GSH can no longer exert its antioxidant role. Full article
(This article belongs to the Special Issue Cigarette Smoking: Associated Oxidative Stress and Health Hazards)
Show Figures

Figure 1

Figure 1
<p>The distribution by serum cotinine levels of patients with mental disorders enrolled in the study.</p>
Full article ">Figure 2
<p>Pearson’s correlation (r) between serum glutatione (GSH), ferric reducing ability of plasma (FRAP), advanced oxidation protein products (AOPP), and smoking exposure measured as serum cotinine levels, in patients with serum cotinine levels &lt;4 ng/mL serum (illustrating the lowest exposure to tobacco smoke, tertile-1, n = 25) (<b>a</b>) versus patients with cotinine levels &gt;70 ng/mL serum (illustrating the highest exposure to tobacco smoke, tertile-3, n = 25) (<b>b</b>). NS: non-significant.</p>
Full article ">Figure 2 Cont.
<p>Pearson’s correlation (r) between serum glutatione (GSH), ferric reducing ability of plasma (FRAP), advanced oxidation protein products (AOPP), and smoking exposure measured as serum cotinine levels, in patients with serum cotinine levels &lt;4 ng/mL serum (illustrating the lowest exposure to tobacco smoke, tertile-1, n = 25) (<b>a</b>) versus patients with cotinine levels &gt;70 ng/mL serum (illustrating the highest exposure to tobacco smoke, tertile-3, n = 25) (<b>b</b>). NS: non-significant.</p>
Full article ">Figure 3
<p>Pearson’s correlation (r) between ferric reducing ability of plasma (FRAP) and advanced oxidation protein products (AOPPs) with serum glutatione (GSH) levels in patients with the lowest exposure to tobacco smoke (tertile-1, n = 25) (<b>a</b>) versus patients with the highest exposure to tobacco smoke (tertile-3, n = 25) (<b>b</b>). NS: non-significant.</p>
Full article ">Figure 3 Cont.
<p>Pearson’s correlation (r) between ferric reducing ability of plasma (FRAP) and advanced oxidation protein products (AOPPs) with serum glutatione (GSH) levels in patients with the lowest exposure to tobacco smoke (tertile-1, n = 25) (<b>a</b>) versus patients with the highest exposure to tobacco smoke (tertile-3, n = 25) (<b>b</b>). NS: non-significant.</p>
Full article ">
27 pages, 5786 KiB  
Article
Antioxidant Activities of Photoinduced Phycogenic Silver Nanoparticles and Their Potential Applications
by Vijayakumar Maduraimuthu, Jayappriyan Kothilmozhian Ranishree, Raja Mohan Gopalakrishnan, Brabakaran Ayyadurai, Rathinam Raja and Klaus Heese
Antioxidants 2023, 12(6), 1298; https://doi.org/10.3390/antiox12061298 - 18 Jun 2023
Cited by 9 | Viewed by 2676
Abstract
While various methods exist for synthesizing silver nanoparticles (AgNPs), green synthesis has emerged as a promising approach due to its affordability, sustainability, and suitability for biomedical purposes. However, green synthesis is time-consuming, necessitating the development of efficient and cost-effective techniques to minimize reaction [...] Read more.
While various methods exist for synthesizing silver nanoparticles (AgNPs), green synthesis has emerged as a promising approach due to its affordability, sustainability, and suitability for biomedical purposes. However, green synthesis is time-consuming, necessitating the development of efficient and cost-effective techniques to minimize reaction time. Consequently, researchers have turned their attention to photo-driven processes. In this study, we present the photoinduced bioreduction of silver nitrate (AgNO3) to AgNPs using an aqueous extract of Ulva lactuca, an edible green seaweed. The phytochemicals found in the seaweed functioned as both reducing and capping agents, while light served as a catalyst for biosynthesis. We explored the effects of different light intensities and wavelengths, the initial pH of the reaction mixture, and the exposure time on the biosynthesis of AgNPs. Confirmation of AgNP formation was achieved through the observation of a surface plasmon resonance band at 428 nm using an ultraviolet-visible (UV-vis) spectrophotometer. Fourier transform infrared spectroscopy (FTIR) revealed the presence of algae-derived phytochemicals bound to the outer surface of the synthesized AgNPs. Additionally, high-resolution transmission electron microscopy (HRTEM) and atomic force microscopy (AFM) images demonstrated that the NPs possessed a nearly spherical shape, ranging in size from 5 nm to 40 nm. The crystalline nature of the NPs was confirmed by selected area electron diffraction (SAED) and X-ray diffraction (XRD), with Bragg’s diffraction pattern revealing peaks at 2θ = 38°, 44°, 64°, and 77°, corresponding to the planes of silver 111, 200, 220, and 311 in the face-centered cubic crystal lattice of metallic silver. Energy-dispersive X-ray spectroscopy (EDX) results exhibited a prominent peak at 3 keV, indicating an Ag elemental configuration. The highly negative zeta potential values provided further confirmation of the stability of AgNPs. Moreover, the reduction kinetics observed via UV-vis spectrophotometry demonstrated superior photocatalytic activity in the degradation of hazardous pollutant dyes, such as rhodamine B, methylene orange, Congo red, acridine orange, and Coomassie brilliant blue G-250. Consequently, our biosynthesized AgNPs hold great potential for various biomedical redox reaction applications. Full article
Show Figures

Figure 1

Figure 1
<p>AgNPs were synthesized from the aqueous extract of <span class="html-italic">U. lactuca</span> via photoinduction (LU-AgNPs). (<b>A</b>) <span class="html-italic">U. lactuca</span> used for biosynthesis of AgNPs, (<b>B</b>) AgNO<sub>3</sub> solution, (<b>C</b>) aqueous extract from <span class="html-italic">U. lactuca</span>, (<b>D</b>–<b>F</b>) <span class="html-italic">U. lactuca</span> extract treated with AgNO<sub>3</sub> solution in the (<b>D</b>) dark, (<b>E</b>) sunlight, and (<b>F</b>) normal LED white light, respectively.</p>
Full article ">Figure 2
<p>UV-vis spectra of biosynthesized LU-AgNPs. (<b>A</b>) LU-AgNP synthesis under sunlight, normal white light (LED), and under complete dark. AgNO<sub>3</sub> solution (red) and aqueous extract from <span class="html-italic">U. lactuca</span> (black) served as references. (<b>B</b>–<b>E</b>) LU-AgNP synthesis under different sunlight intensities (<b>B</b>), at different time intervals under sunlight (<b>C</b>), at different wavelengths of sunlight (<b>D</b>), and at different pH values kept under sunlight and dark conditions (<b>E</b>).</p>
Full article ">Figure 3
<p>Simultaneous evaluation of photocatalytic LU-AgNP biosynthesis under opaque/light-exposed conditions. <span class="html-italic">U. lactuca</span> extract, along with AgNO<sub>3</sub> (9:1), was simultaneously exposed to (<b>A</b>) 3/4 opaque conditions and (<b>B</b>) 1/4 light exposure. (<b>C</b>) LU-AgNP biosynthesis occurred only in the light-exposed portion of the test tube, and (<b>D</b>) dispersion of formed LU-AgNPs to the opaque portion occurred after manual disturbance of the test tube.</p>
Full article ">Figure 4
<p>(<b>A</b>) Transmission electron microscopy (TEM) image of photocatalytic biosynthesized LU-AgNPs at 50 nm resolution, (<b>B</b>) SAED pattern of LU-AgNPs, (<b>C</b>) High-resolution transmission electron microscopy (HRTEM) image of a single spherical NP, and (<b>D</b>) histogram showing different size distributions of LU-AgNPs.</p>
Full article ">Figure 5
<p>Energy-dispersive X-ray (EDX) spectrum of photocatalytic biosynthesized LU-AgNPs confirming nanoparticle synthesis.</p>
Full article ">Figure 6
<p>2D and 3D atomic force microscopy (AFM) images displaying topographical features (such as homogeneity, spherical shape, and dense packing) of photocatalytic biosynthesized LU-AgNPs.</p>
Full article ">Figure 7
<p>The Fourier transform infrared spectroscopy (FTIR) spectrum of <span class="html-italic">U. lactuca</span> extract shows major peaks representing hydroxyl, primary amine (O–H and N–H bonds, peak at 3420 cm<sup>−1</sup>), alkyne (C=C bonds, peak at 2280 cm<sup>−1</sup>), and vibrational stretching carbonyl (C≡C bonds, peak at 1760 cm<sup>−1</sup>) groups. FTIR spectrum of LU-AgNPs shows peaks that represent hydroxyl, primary amine (O–H and N–H bonds, peak at 3300 cm<sup>−1</sup>), terminal alkyne (C=C bonds, peak at 2100 cm<sup>−1</sup>), and primary amide (N–H stretching and C=O bending vibration at 1628 cm<sup>−1</sup>) groups.</p>
Full article ">Figure 8
<p>X-ray diffraction (XRD) spectra analysis of LU-AgNPs and Bragg’s reflection of the face-centered cubic (fcc) elemental silver (JCPDS No. 98-002-1923).</p>
Full article ">Figure 9
<p>Catalytic redox potential of photocatalytically biosynthesized LU-AgNPs-dye degradation studies. (<b>A</b>) UV-vis spectra for photocatalytic degradation of RB using LU-AgNPs, (<b>B</b>) ln(A/A<sub>0</sub>) vs. time for the reduction of RB using LU-AgNPs, (<b>C</b>) UV-vis spectra for photocatalytic degradation of MO using LU-AgNPs, (<b>D</b>) ln(A/A<sub>0</sub>) vs. time for the reduction of MO using LU-AgNPs, (<b>E</b>) UV-vis spectra for photocatalytic degradation of AO using LU-AgNPs, (<b>F</b>) ln(A/A<sub>0</sub>) vs. time for the reduction of AO using LU-AgNPs, (<b>G</b>) UV-vis spectra for photocatalytic degradation of CR using LU-AgNPs, (<b>H</b>) ln(A/A<sub>0</sub>) vs. time for the reduction of CR using LU-AgNPs, (<b>I</b>) UV-vis spectra for photocatalytic degradation of CBB G-250 using LU-AgNPs, and (<b>J</b>) ln(A/A<sub>0</sub>) vs. time for the reduction of CBB G-250 using LU-AgNPs. AO: acridine orange, CBB G-250: Coomassie brilliant blue G-250, CR: Congo red, MO: methylene orange, RB: rhodamine B.</p>
Full article ">
15 pages, 1348 KiB  
Article
Anti-Inflammatory, Anti-Oxidative and Anti-Apoptotic Effects of Thymol and 24-Epibrassinolide in Zebrafish Larvae
by Germano A. B. Lanzarin, Luís M. Félix, Sandra M. Monteiro, Jorge M. Ferreira, Paula A. Oliveira and Carlos Venâncio
Antioxidants 2023, 12(6), 1297; https://doi.org/10.3390/antiox12061297 - 18 Jun 2023
Cited by 3 | Viewed by 1952
Abstract
Thymol (THY) and 24-epibrassinolide (24-EPI) are two examples of plant-based products with promising therapeutic effects. In this study, we investigated the anti-inflammatory, antioxidant and anti-apoptotic effects of the THY and 24-EPI. We used zebrafish (Danio rerio) larvae transgenic line (Tg(mpx [...] Read more.
Thymol (THY) and 24-epibrassinolide (24-EPI) are two examples of plant-based products with promising therapeutic effects. In this study, we investigated the anti-inflammatory, antioxidant and anti-apoptotic effects of the THY and 24-EPI. We used zebrafish (Danio rerio) larvae transgenic line (Tg(mpxGFP)i114) to evaluate the recruitment of neutrophils as an inflammatory marker to the site of injury after tail fin amputation. In another experiment, wild-type AB larvae were exposed to a well known pro-inflammatory substance, copper (CuSO4), and then exposed for 4 h to THY, 24-EPI or diclofenac (DIC), a known anti-inflammatory drug. In this model, the antioxidant (levels of reactive oxygen species—ROS) and anti-apoptotic (cell death) effects were evaluated in vivo, as well as biochemical parameters such as the activity of antioxidant enzymes (superoxide dismutase, catalase and glutathione peroxidase), the biotransformation activity of glutathione-S-transferase, the levels of glutathione reduced and oxidated, lipid peroxidation, acetylcholinesterase activity, lactate dehydrogenase activity, and levels of nitric acid (NO). Both compounds decreased the recruitment of neutrophils in Tg(mpxGFP)i114, as well as showed in vivo antioxidant effects by reducing ROS production and anti-apoptotic effects in addition to a decrease in NO compared to CuSO4. The observed data substantiate the potential of the natural compounds THY and 24-EPI as anti-inflammatory and antioxidant agents in this species. These results support the need for further research to understand the molecular pathways involved, particularly their effect on NO. Full article
(This article belongs to the Special Issue Antioxidant and Anti-inflammatory Compounds from Natural Products)
Show Figures

Figure 1

Figure 1
<p>Treatment effects on neutrophil migration to the lesion site in zebrafish larvae (Tg(<span class="html-italic">mpx</span>GFP)<sup>i114</sup>) produced by tail transection. (<b>A</b>) Diagram of the experimental design for the neutrophil migration (<b>B</b>) Image of a normal Tg(<span class="html-italic">mpx</span>GFP)<sup>i114</sup> zebrafish larva (control), and a larva with tail transection (Control cut) and detailed photos of the transection site after 4 h of treatment. The scale bar represents 125 μm. (<b>C</b>) Graph showing the normalized number of neutrophils that migrated to the tail after 4 hpi (Mean of control: 30.7 ± 9.2). Data are expressed as mean ± SD from at least five independent samples from five random animals each. Different letters represent statistical differences among treatment groups (<span class="html-italic">p</span> &lt; 0.05).</p>
Full article ">Figure 2
<p>Effect of different treatments on cell death and ROS production in zebrafish WT larvae with 72 hpf exposed for 4 h. Data from at least five independent samples from twenty-five random animals each. (<b>A</b>) Schematic diagram showing the experimental exposure protocol for the study of cell death and ROS production. (<b>B</b>) Illustrative images from larvae exposed to the DCF probe. (<b>C</b>) Illustrative images from larvae exposed to the AO probe. (<b>D</b>) Result of the DCF fluorescence intensities in homogenized larvae (Mean of control: 37.9 ± 8.8). (<b>E</b>) Result of AO fluorescence intensities in homogenized larvae (Mean of control: 48.4 ± 7.5). Data are expressed as mean ± SD and normalized according to the control group. Different letters represent statistical differences among treatment groups (<span class="html-italic">p</span> &lt; 0.05).</p>
Full article ">Figure 3
<p>Biochemical indicators were examined in 72-hour-old zebrafish WT larvae subjected to various treatments for 4 h. (<b>A</b>) Schematic diagram of the experimental methodology for biochemical marker exposure and analysis. (<b>B</b>) Graphs of biochemical indicators that differed significantly following exposure to various treatments. Data were obtained from at least five independent samples from fifty random animals each and values normalized to the control group. Data are expressed as median (interquartile range) for non-parametric data (Median of control: SOD = 7.2 (6.3–7.9) U/mg.protein) or mean ± SD for parametric data distribution (Mean of control: CAT = 2.9 ± 0.9 U/mg.protein; GPx = 2.24 ± 0.3 nmol NADPH/min.mg protein; OSI = 0.18 ± 0.05; AChE = 10.43 ± 1.96 μmol TNB/min.mg protein; LDH = 65.45 ± 5.15 nmol NADH/min.mg protein; NO = 23.4 ± 11.3 nmol NO/mg protein). Different letters represent statistical differences among treatment groups (<span class="html-italic">p</span> &lt; 0.05).</p>
Full article ">
16 pages, 1796 KiB  
Article
Effect of Exercise Repetitions on Arylesterase Activity of PON1 in Plasma of Average-Trained Men—The Dissociation between Activity and Concentration
by Aneta Otocka-Kmiecik, Monika Orłowska-Majdak, Robert Stawski, Urszula Szkudlarek, Gianluca Padula, Szymon Gałczyński and Dariusz Nowak
Antioxidants 2023, 12(6), 1296; https://doi.org/10.3390/antiox12061296 - 17 Jun 2023
Viewed by 1550
Abstract
Exercise may increase the antioxidant capacity of plasma by stimulating antioxidant enzymes. The study aimed to measure the effect of three repetitions of acute exercise on arylesterase (ARE) activity of the paraoxonase 1 (PON1) enzyme. Eleven average-trained men (age 34.0 ± 5.2 years) [...] Read more.
Exercise may increase the antioxidant capacity of plasma by stimulating antioxidant enzymes. The study aimed to measure the effect of three repetitions of acute exercise on arylesterase (ARE) activity of the paraoxonase 1 (PON1) enzyme. Eleven average-trained men (age 34.0 ± 5.2 years) completed three treadmill runs. ARE activity in plasma was evaluated spectrophotometrically and compared with PON1 concentration (PON1c), paraoxonase (PON) activity, and high-density lipoprotein cholesterol (HDL-C) at rest and after exercise. In all repetitions of the exercise, ARE activity remained stable, and ARE activity standardized for PON1c (ARE/PON1c) was lower post- than pre-exercise. The ARE/PON1c ratio changes returned to baseline levels during rest after each exercise session. Pre-exercise ARE activity correlated negatively with post-exercise C-reactive protein (CRP) (ρ = −0.35, p = 0.049), white blood cell count (WBC) (ρ = −0.35, p = 0.048), polymorphonuclear leukocytes (PMN) (ρ = −0.37, p = 0.037), and creatine kinase (CK) (ρ = −0.37, p = 0.036). ARE activity may be depleted under conditions of oxidative stress, as increases in PON1c during acute exercise did not result in parallel increases in ARE activity. No adaptation of the response of ARE activity to exercise was detected in subsequent exercise sessions. Individuals with lower pre-exercise ARE activity may develop a higher inflammatory response to strenuous exercise. Full article
(This article belongs to the Special Issue Exercise-Induced Oxidative Stress in Health and Disease)
Show Figures

Figure 1

Figure 1
<p>Plausible mechanisms of reversible increase in PON1 concentration with concomitant decrease in ARE activity in plasma during acute exercise. <span class="html-fig-inline" id="antioxidants-12-01296-i001"><img alt="Antioxidants 12 01296 i001" src="/antioxidants/antioxidants-12-01296/article_deploy/html/images/antioxidants-12-01296-i001.png"/></span>-increase, <span class="html-fig-inline" id="antioxidants-12-01296-i002"><img alt="Antioxidants 12 01296 i002" src="/antioxidants/antioxidants-12-01296/article_deploy/html/images/antioxidants-12-01296-i002.png"/></span>- decrease, PON1—paraoxonase 1, HDL—high-density lipoprotein, ROS—reactive oxygen species, Ca<sup>2+</sup>—catalytic calcium ion in the active site tunnel of PON1, His—Histidine, Cys—Cysteine, -SH groups—sulfhydryl groups.</p>
Full article ">Figure 2
<p>Flowchart of the study. BP—blood pressure measurement, ECG—electrocardiography, VO<sub>2</sub>max—maximal oxygen consumption.</p>
Full article ">Figure 3
<p>No significant effect of repeated exercise on arylesterase activity (ARE) of paraoxonase 1 (PON1) in the plasma of average-trained men (mean ± SD).</p>
Full article ">Figure 4
<p>Effect of repeated exercise on arylesterase to paraoxonase 1 concentration ratio (ARE/PON1c) in the plasma of average-trained men (mean ± SD). * vs. corresponding pre-exercise value, <span class="html-italic">p</span> &lt; 0.05.</p>
Full article ">Figure 5
<p>Scatterplot of variables x and y. Correlation of pre-exercise values of arylesterase (ARE) activity and post-exercise C-reactive protein (CRP) in the plasma of average-trained men. Pooled individual data from three exercise bouts (<span class="html-italic">n</span> = 33; <span class="html-italic">ρ</span> = −0.35, <span class="html-italic">p</span> = 0.049).</p>
Full article ">
19 pages, 1245 KiB  
Review
The Influence of Antioxidants on Oxidative Stress-Induced Vascular Aging in Obesity
by Hiva Sharebiani, Shayan Keramat, Abdolali Chavoshan, Bahar Fazeli and Agata Stanek
Antioxidants 2023, 12(6), 1295; https://doi.org/10.3390/antiox12061295 - 17 Jun 2023
Cited by 5 | Viewed by 1912
Abstract
Obesity is a worldwide trend that is growing in incidence very fast. Adipose tissue dysfunction caused by obesity is associated with the generation of oxidative stress. Obesity-induced oxidative stress and inflammation play a key role in the pathogenesis of vascular diseases. Vascular aging [...] Read more.
Obesity is a worldwide trend that is growing in incidence very fast. Adipose tissue dysfunction caused by obesity is associated with the generation of oxidative stress. Obesity-induced oxidative stress and inflammation play a key role in the pathogenesis of vascular diseases. Vascular aging is one of the main pathogenesis mechanisms. The aim of this study is to review the effect of antioxidants on vascular aging caused by oxidative stress in obesity. In order to achieve this aim, this paper is designed to review obesity-caused adipose tissue remodeling, vascular aging generated by high levels of oxidative stress, and the effects of antioxidants on obesity, redox balance, and vascular aging. It seems that vascular diseases in obese individuals are complex networks of pathological mechanisms. In order to develop a proper therapeutic tool, first, there is a need for a better understanding of interactions between obesity, oxidative stress, and aging. Based on these interactions, this review suggests different lines of strategies that include change in lifestyle to prevent and control obesity, strategies for adipose tissue remodelling, oxidant–antioxidant balance, inflammation suppression, and strategies against vascular aging. Some antioxidants support different lines of these strategies, making them appropriate for complex conditions such as oxidative stress-induced vascular diseases in obese individuals. Full article
Show Figures

Figure 1

Figure 1
<p>Causes of obesity, remodelling of adipose tissue and management strategies. PAI-1, plasminogen activator inhibitor-1; TG, triglyceride; WAT, white adipose tissue; BAT, brown adipose tissue; LDL, low-density lipoprotein; ↓ decrease; ↑ increase.</p>
Full article ">Figure 2
<p>Generation and consequences of oxidative stress and management strategies. SOD, superoxide dismutase; CAT, catalase; GSH, glutathione; MDA, malondialdehyde; H<sub>2</sub>O<sub>2</sub> hydrogen peroxide; eNOS, endothelial nitric oxide synthase; ROS, reactive oxygen species; MAPK, mitogen-activated protein kinase; PI3K, phosphoinositide 3-kinases; NF-<span class="html-small-caps">k</span>B, nuclear factor-kappa B; PKA, protein kinase A; PKB, protein kinase B; ↓ decrease; ↑ increase.</p>
Full article ">Figure 3
<p>Endothelial cell aging mechanisms, vascular aging consequences and their management strategies. ICAM-1, intercellular adhesion molecule 1; iNOS, inducible nitric oxide synthase; NO, nitric oxide; eNOS, endothelial nitric oxide synthase; ↓ decrease; ↑ increase.</p>
Full article ">
13 pages, 1522 KiB  
Review
Revised Aspects into the Molecular Bases of Hydroxycinnamic Acid Metabolism in Lactobacilli
by Félix López de Felipe
Antioxidants 2023, 12(6), 1294; https://doi.org/10.3390/antiox12061294 - 17 Jun 2023
Cited by 3 | Viewed by 1754
Abstract
Hydroxycinnamic acids (HCAs) are phenolic compounds produced by the secondary metabolism of edible plants and are the most abundant phenolic acids in our diet. The antimicrobial capacity of HCAs is an important function attributed to these phenolic acids in the defense of plants [...] Read more.
Hydroxycinnamic acids (HCAs) are phenolic compounds produced by the secondary metabolism of edible plants and are the most abundant phenolic acids in our diet. The antimicrobial capacity of HCAs is an important function attributed to these phenolic acids in the defense of plants against microbiological threats, and bacteria have developed diverse mechanisms to counter the antimicrobial stress imposed by these compounds, including their metabolism into different microbial derivatives. The metabolism of HCAs has been intensively studied in Lactobacillus spp., as the metabolic transformation of HCAs by these bacteria contributes to the biological activity of these acids in plant and human habitats or to improve the nutritional quality of fermented foods. The main mechanisms known to date used by Lactobacillus spp. to metabolize HCAs are enzymatic decarboxylation and/or reduction. Here, recent advances in the knowledge regarding the enzymes that contribute to these two enzymatic conversions, the genes involved, their regulation and the physiological significance to lactobacilli are reviewed and critically discussed. Full article
(This article belongs to the Section Health Outcomes of Antioxidants and Oxidative Stress)
Show Figures

Figure 1

Figure 1
<p>Graphic representation of metabolic decarboxylation of caffeic, <span class="html-italic">p</span>-coumaric and ferulic acid into their corresponding vinyl derivatives by <span class="html-italic">Lactobacillus</span> spp. The predicted p<span class="html-italic">K</span>a of the hydroxycinnamic acids and their vinyl derivatives are shown. Each vinyl derivative shares the same R group as the hydroxycinnamic acid from which it is derived; PAD, phenolic acid decarboxylase.</p>
Full article ">Figure 2
<p>Genetic organization of the <span class="html-italic">L. plantarum</span> WCFS1 chromosomal regions containing clusters harboring genetically and biochemically validated genes determining hydroxycinnamate metabolic conversions: hydroxycinnamate decarboxylation <span class="html-italic">padA</span> or <span class="html-italic">lp_3665</span> (red arrow) (<b>A</b>) [NCBI accession NC_004567; <span class="html-italic">lp_3663</span> (or <span class="html-italic">usp1</span>) through <span class="html-italic">lp_3665</span> (or <span class="html-italic">padA</span>)]; hydroxycinnamate reduction <span class="html-italic">hcrB</span> or <span class="html-italic">lp_1425</span> (green arrow) (<b>B</b>) [NCBI accession NC_004567; <span class="html-italic">lp_1422</span> (or <span class="html-italic">hcrR</span>) through <span class="html-italic">lp_1426</span> (or <span class="html-italic">hcrC</span>)] or vinylphenol reduction <span class="html-italic">vprA</span> or <span class="html-italic">lp_3125</span> (blue arrow) (<b>C</b>) [NCBI accession NC_004567; <span class="html-italic">lp_3124</span> (or <span class="html-italic">vprR</span>) through <span class="html-italic">lp_3125</span> (or <span class="html-italic">vprA</span>)]. Lys-R type regulators that control decarboxylation or reduction of hydroxycinnamates as well as the reduction of vinylphenols, are indicated by lined arrows.</p>
Full article ">Figure 3
<p>Graphic representation of metabolic reduction of vinyl derivatives of caffeic (vinylcatechol), <span class="html-italic">p</span>-coumaric (vinylphenol) and ferulic (vinylguaicol) acids by <span class="html-italic">Lactobacillus spp</span>. Each ethyl derivative shares the same R group as the vinyl derivative from which it is derived. VprA, vinylphenol reductase A.</p>
Full article ">Figure 4
<p>Graphic representation of metabolic reduction of caffeic, <span class="html-italic">p</span>-coumaric and ferulic acid into phenylpropionic derivatives by <span class="html-italic">Lactobacillus</span> spp. HcrB, hydroxycinnamate reductase B. Each phenylpropionic acid shares the same R group as the hydroxycinnamic acid from which it is derived.</p>
Full article ">
20 pages, 2434 KiB  
Article
Improvement of Fresh Ovine “Tuma” Cheese Quality Characteristics by Application of Oregano Essential Oils
by Giuliana Garofalo, Marialetizia Ponte, Carlo Greco, Marcella Barbera, Michele Massimo Mammano, Giancarlo Fascella, Giuseppe Greco, Giulia Salsi, Santo Orlando, Antonio Alfonzo, Antonino Di Grigoli, Daniela Piazzese, Adriana Bonanno, Luca Settanni and Raimondo Gaglio
Antioxidants 2023, 12(6), 1293; https://doi.org/10.3390/antiox12061293 - 17 Jun 2023
Cited by 5 | Viewed by 1570
Abstract
In the present work, oregano essential oils (OEOs) were applied to process the fresh ovine cheese “Tuma” obtained by pressed cheese technology. Cheese making trials were performed under industrial conditions using ewe’s pasteurized milk and two strains of Lactococcus lactis (NT1 and NT4) [...] Read more.
In the present work, oregano essential oils (OEOs) were applied to process the fresh ovine cheese “Tuma” obtained by pressed cheese technology. Cheese making trials were performed under industrial conditions using ewe’s pasteurized milk and two strains of Lactococcus lactis (NT1 and NT4) as fermenting agents. Two experimental cheese products (ECP) were obtained through the addition of 100 (ECP100) and 200 (ECP200) µL/L of OEO to milk, while the control cheese product (CCP) was OEO-free. Both Lc. lactis strains showed in vitro and in vivo ability to grow in the presence of OEOs and to dominate over indigenous milk lactic acid bacteria (LAB) resistant to pasteurization. In the presence of OEOs, the most abundant compound found in cheese was carvacrol, constituting more than 65% of the volatile fraction in both experimental products. The addition of OEOs did not influence ash, fat, or protein content, but it increased by 43% the antioxidant capacity of the experimental cheeses. ECP100 cheeses showed the best appreciation scores by the sensory panel. In order to investigate the ability OEOs to be used as a natural preservative, a test of artificial contamination was carried out, and the results showed a significant reduction of the main dairy pathogens in OEO-added cheeses. Full article
(This article belongs to the Section Extraction and Industrial Applications of Antioxidants)
Show Figures

Figure 1

Figure 1
<p>Scheme of dairy plant. (<b>a</b>) diesel burner; (<b>b</b>) furnace (first and second stage of smoke circuits); (<b>c</b>) third stage of smoke circuits; (<b>d</b>) steam accumulator with dispensing valve; (<b>e</b>) water inlet; (<b>f</b>) multipurpose coagulation vats; (<b>g</b>) gearmotor for mechanically cutting curd; (<b>h</b>) perforated steel table.</p>
Full article ">Figure 2
<p>Flow diagram of Tuma cheese production. Abbreviation: OEOs, oregano essential oils.</p>
Full article ">Figure 3
<p>Volatile organic compounds emitted from oregano essential oil. Results indicate mean percentage values ± standard deviation (S.D.) of three measurements and are expressed as relative peak areas (peak area of each compound/total area of the significant peaks to all samples) × l00.</p>
Full article ">Figure 4
<p>Growth of starter cultures during Tuma cheese productions. (<b>a</b>) total mesophilic microorganisms; (<b>b</b>) <span class="html-italic">Lactococcus lactis</span>. Units are Log CFU/mL for milk and whey samples and Log CFU/g for curd and cheese samples. Abbreviations: CCP, control cheese product inoculated with the milk starter cultures (MSC); ECPO100, experimental cheese product inoculated with MSC + 100 μL/L of oregano essential oils (OEOs); ECPO200, experimental cheese product inoculated with MSC + 200 μL/L of OEOs.</p>
Full article ">Figure 5
<p>Dendrogram obtained from combined RAPD-PCR patterns of LAB strains isolated from pasteurized ewe’s milk to Tuma cheeses. Abbreviations: CCP, control cheese product inoculated with the milk starter cultures (MSC); ECPO100, experimental cheese product inoculated with MSC + 100 μL/L of oregano essential oils (OEOs); ECPO200, experimental cheese product inoculated with MSC + 200 μL/L of OEOs; PM, pasteurized milk; IM, inoculated milk; W, whey; C, curd; Ch, cheese; <span class="html-italic">Lc.</span>, <span class="html-italic">Lactococcus</span>; <span class="html-italic">Str.</span>, <span class="html-italic">Streptococcus</span>.</p>
Full article ">Figure 6
<p>Spider diagram of descriptive sensory analysis of Tuma cheeses. Abbreviations: CCP, control cheese product inoculated with the milk starter cultures (MSC); ECPO100, experimental cheese product inoculated with MSC + 100 μL/L of oregano essential oils (OEOs); ECPO200, experimental cheese product inoculated with MSC + 200 μL/L of OEOs. ** <span class="html-italic">p</span> &lt; 0.01; *** <span class="html-italic">p</span> &lt; 0.001; n.s., not significant.</p>
Full article ">
23 pages, 6215 KiB  
Article
A Deadly Liaison between Oxidative Injury and p53 Drives Methyl-Gallate-Induced Autophagy and Apoptosis in HCT116 Colon Cancer Cells
by Antonietta Notaro, Marianna Lauricella, Diana Di Liberto, Sonia Emanuele, Michela Giuliano, Alessandro Attanzio, Luisa Tesoriere, Daniela Carlisi, Mario Allegra, Anna De Blasio, Giuseppe Calvaruso and Antonella D’Anneo
Antioxidants 2023, 12(6), 1292; https://doi.org/10.3390/antiox12061292 - 16 Jun 2023
Cited by 6 | Viewed by 1784
Abstract
Methyl gallate (MG), which is a gallotannin widely found in plants, is a polyphenol used in traditional Chinese phytotherapy to alleviate several cancer symptoms. Our studies provided evidence that MG is capable of reducing the viability of HCT116 colon cancer cells, while it [...] Read more.
Methyl gallate (MG), which is a gallotannin widely found in plants, is a polyphenol used in traditional Chinese phytotherapy to alleviate several cancer symptoms. Our studies provided evidence that MG is capable of reducing the viability of HCT116 colon cancer cells, while it was found to be ineffective on differentiated Caco-2 cells, which is a model of polarized colon cells. In the first phase of treatment, MG promoted both early ROS generation and endoplasmic reticulum (ER) stress, sustained by elevated PERK, Grp78 and CHOP expression levels, as well as an upregulation in intracellular calcium content. Such events were accompanied by an autophagic process (16–24 h), where prolonging the time (48 h) of MG exposure led to cellular homeostasis collapse and apoptotic cell death with DNA fragmentation and p53 and γH2Ax activation. Our data demonstrated that a crucial role in the MG-induced mechanism is played by p53. Its level, which increased precociously (4 h) in MG-treated cells, was tightly intertwined with oxidative injury. Indeed, the addition of N-acetylcysteine (NAC), which is a ROS scavenger, counteracted the p53 increase, as well as the MG effect on cell viability. Moreover, MG promoted p53 accumulation into the nucleus and its inhibition by pifithrin-α (PFT-α), which is a negative modulator of p53 transcriptional activity, enhanced autophagy, increased the LC3-II level and inhibited apoptotic cell death. These findings provide new clues to the potential action of MG as a possible anti-tumor phytomolecule for colon cancer treatment. Full article
Show Figures

Graphical abstract

Graphical abstract
Full article ">Figure 1
<p>IC<sub>50</sub> determination and colony formation assay of MG-treated colon cancer cells. (<b>A</b>) MTT assay of colon cancer cells (HCT116 and Caco-2) and differentiated Caco-2 cells incubated with incremental doses of MG. Cell viability was determined at 48 h as reported in the Materials and Methods section. IC<sub>50</sub> values were assessed using GraphPad Prism 7 software. (<b>B</b>) LDH cytotoxicity test in MG-treated HCT116 cells. After incubation in the presence or absence of the phytocompound, cells were centrifuged and supernatants were used to assess the LDH content using a commercial kit. Data are reported as a percentage of the total LDH released from cells using as a positive control of cells incubated with 0.1% Triton 100×. (<b>C</b>) Clonogenic assay in MG-treated HCT116 cells. The colony formation inhibition was assessed by crystal violet staining after 10 days of exposure to MG. Representative images of colony formation are reported in the upper panel. The quantitative analysis of colonies (lower panel) was performed as reported in the Materials and Methods section. All experiments were performed in triplicate. (*) <span class="html-italic">p</span> &lt; 0.05 and (**) <span class="html-italic">p</span> &lt; 0.01 compared with the untreated sample.</p>
Full article ">Figure 2
<p>Oxidative stress is required for the cytotoxic efficacy of MG in colon cancer cells. (<b>A</b>) Effect of NAC pre-incubation on HCT116 cell viability of MG-treated cells for 48 h. Each value reported in the histogram represents the mean of three independent experiments ± SD. (*) <span class="html-italic">p</span> &lt; 0.05 and (**) <span class="html-italic">p</span> &lt; 0.01 compared with the untreated sample. (#) <span class="html-italic">p</span> &lt; 0.05 compared with the MG-treated sample. (<b>B</b>) Phase-contrast micrographs of morphological changes of HCT116 cells treated for 48 h with MG and the protective effect of NAC pre-incubation (original magnification 200×). (<b>C</b>) ROS generation induced by MG treatment in HCT116 cells. The ROS level was measured using H2-DCFDA, which is a redox-sensitive fluorescent probe, as reported in the Materials and Methods section. Original pictures were taken using a Leica fluorescence microscope equipped with a CCD camera and FITC filter (original magnification 200×).</p>
Full article ">Figure 3
<p>Upregulation of the antioxidant enzymatic systems in HCT116 cells treated with MG. After treatment with the indicated doses of the phytocompound in the presence or absence of NAC, the cell lysates were prepared and the level of stress-associated proteins (MnSOD and catalase) was detected using Western blotting. Representative blots from three independent experiments were considered and a densitometry analysis histogram was normalized to γ-tubulin, which was used as a loading control. (**) <span class="html-italic">p</span> &lt; 0.01 and (***) <span class="html-italic">p</span> &lt; 0.001 compared with the untreated sample. (#) <span class="html-italic">p</span> &lt; 0.05, and (##) <span class="html-italic">p</span> &lt; 0.01 compared with the MG-treated condition.</p>
Full article ">Figure 4
<p>MG exposure activated the ER-stress-associated protein levels. HCT116 cells were treated in the presence of 30 and 90 μg/mL MG for 24 h and 48 h. Western blot analysis was performed to evaluate the protein expression of the ER stress markers PERK, phospho-PERK, eiF2α, Grp78 and CHOP. The amount of analyzed proteins was assessed using γ-tubulin as the loading control protein and for band density normalization. The data are presented as the mean ± SD; (*) <span class="html-italic">p</span> &lt; 0.05 and (**) <span class="html-italic">p</span> &lt; 0.01 compared with the untreated sample.</p>
Full article ">Figure 5
<p>MG exposure increased the intracellular calcium level. Changes in the content of intracellular calcium analyzed at the indicated times via flow cytometry using Fluo 3-AM fluorochrome as reported in the Materials and Methods section.</p>
Full article ">Figure 6
<p>MG exposure provoked an autophagic process in the early phases of treatment. (<b>A</b>) Monodansylcadaverine (MDC) staining that enabled the visualization of autophagic vacuoles as dot-like structures was performed in MG-treated cells. HCT116 cells were incubated with MG for the indicated time periods and autophagic vacuoles were highlighted via fluorescence microscopy using a Leica microscope equipped with a DAPI filter. Representative fluorescence microscopy images were taken at a magnification of 400×, as reported in <a href="#sec2-antioxidants-12-01292" class="html-sec">Section 2</a>. (<b>B</b>) Immunoblots of autophagic markers performed in MG-treated HCT116 cells. Proteins were detected using different antibodies directed against the LC3-I and LC3-II forms, p62, Beclin 1, Atg1/Ulk1 and Atg7. γ-tubulin was used as the loading control. All graphs show the mean ± SD of three independent experiments. (*) <span class="html-italic">p</span> &lt; 0.05 and (**) <span class="html-italic">p</span> &lt; 0.01 compared with the untreated cells.</p>
Full article ">Figure 7
<p>MG-induced DNA damage. (<b>A</b>) Morphological analysis of HCT116 cells after vital staining with Hoechst 33342. Following Hoechst staining, cells were treated for 48 h with different doses of MG in the presence or absence of NAC. Chromatin fragmentation and condensation were observed under a fluorescence microscope. The images (original magnification at 200×) were acquired with a DAPI filter using an inverted fluorescence microscope and processed with Leica Q Fluoro Software. (<b>B</b>) Analysis of DNA damage markers: γH2AX and p53. After treatment with MG in the presence or absence of NAC for 48 h, cells were lysed and proteins were analyzed using Western blotting. The γ-tubulin blot was reported as a loading control. The blots and histograms of densitometric analyses reported are representative of three independent experiments. (*) <span class="html-italic">p</span> &lt; 0.05 and (**) <span class="html-italic">p</span> &lt; 0.01 compared with untreated cells. (#) <span class="html-italic">p</span> &lt; 0.05 compared with the MG-treated condition.</p>
Full article ">Figure 8
<p>MG-induced apoptosis in colon cancer cells. (<b>A</b>) Annexin V/PI staining to evaluate apoptosis. HCT116 cells were treated with MG and compared with untreated cells. The rate of apoptosis was assessed via flow cytometry using the Annexin V/PI double staining assay. The data represent one of three independent experiments. (<b>B</b>) MG treatment evoked an increase in apoptotic markers. Caspase-3 activation and PARP1 fragmentation were analyzed using Western blotting. The relative quantification was assessed after densitometric analysis of bands and normalization to γ-tubulin used as a loading control. Histograms of densitometric analyses report the average values of three independent experiments. (*) <span class="html-italic">p</span> &lt; 0.05, (***) <span class="html-italic">p</span> &lt; 0.001 and (****) <span class="html-italic">p</span>&lt; 0.0001 compared with the untreated sample. ns, not significant.</p>
Full article ">Figure 9
<p>Time course analysis of the MG action on DNA damage markers and p53 nuclear accumulation. (<b>A</b>) MG treatment provoked an early upregulation of DNA damage markers p53 and γH2Ax in HCT116 cells. After treatment with MG for various periods, p53 and γH2Ax were detected using Western blotting analyses. Data were normalized to γ-tubulin, which was used as the loading control. The blots and histograms of densitometric analyses reported are representative of three independent experiments. (*) <span class="html-italic">p</span> &lt; 0.05, (**) <span class="html-italic">p</span> &lt; 0.01 and (***) <span class="html-italic">p</span> &lt; 0.001 compared with the untreated sample. (<b>B</b>) Subcellular fractionation for cytosolic and nuclear protein extract displayed nuclear enrichment in the p53 content after the MG incubation. The relative quantification was assessed after densitometric analysis of the bands and normalization to the correspondent loading control. Lamin B or GADPH was used to assess the possible changes in the loaded protein amount for the cytosolic and nuclear fractions, respectively. (*) <span class="html-italic">p</span> &lt; 0.05 and (**) <span class="html-italic">p</span> &lt; 0.01 compared with the untreated sample.</p>
Full article ">Figure 10
<p>Effects of p53 inhibition on MG-induced autophagy. (<b>A</b>) The inhibition of p53 transcriptional activity by PFT-α enhanced the MG-induced autophagy in HCT116 cells. Cells were pre-incubated with NAC (5 mM), BafA1 (100 nM) or PFT-α (20 μM) in the presence or absence of MG for 24 h; then, the production of AVOs showing bright red fluorescence was evaluated via AO staining using Leica Q Fluoro software. (<b>B</b>) PFT-α/MG co-treatment enhanced the levels of the autophagic protein LC3. For the analysis of the LC3 forms, cells were treated with MG in the presence or absence of PFT-α (20 μM), followed by a Western blot analysis. The relative quantification was assessed after a densitometric analysis of bands and normalization to γ-tubulin. Data reported in the histograms were the average of three independent experiments. (**) <span class="html-italic">p</span> &lt; 0.01 compared with the untreated sample. (#) <span class="html-italic">p</span> &lt; 0.05 compared with the MG-treated sample.</p>
Full article ">Figure 11
<p>PFT-α, which is a p53 inhibitor, negatively affected the MG-induced apoptosis in HCT116 cells. (<b>A</b>) PFT-α inhibited the MG-induced cytotoxic effect. The pre-incubation of cells with PFT-α was performed as reported in the Results section; then, different MG concentrations were added and incubation was protracted for 48 h. Cell viability was analyzed using an MTT assay. Values reported in the line chart represent the mean of three independent experiments ± SD; (*) <span class="html-italic">p</span> &lt; 0.05 and (**) <span class="html-italic">p</span> &lt; 0.01 compared with the untreated conditions. (#) <span class="html-italic">p</span> &lt; 0.05 and (##) <span class="html-italic">p</span> &lt; 0.01 compared with the MG-treated sample. (<b>B</b>) Micrographs showing the PFT-α effect on morphological changes induced by the MG treatment. Pictures were taken using a Leica inverted microscope as reported in the Materials and Methods section. (<b>C</b>) Effect of PFT-α on the p53 and caspase-3 levels. Protein lysates were prepared as reported in the Materials and Methods section and resolved using SDS-PAGE. Blots were detected using specific antibodies directed against the proteins of interest and their level was normalized to γ-tubulin, which was used as the loading control. The blots and histograms of densitometric analyses reported are representative of three independent experiments. (*) <span class="html-italic">p</span> &lt; 0.05 and (**) <span class="html-italic">p</span> &lt; 0.01 compared with the untreated sample; (#) <span class="html-italic">p</span> &lt; 0.05 and (##) <span class="html-italic">p</span> &lt; 0.01 compared with the MG-treated sample.</p>
Full article ">Figure 12
<p>Timeline of the antitumor signaling pathway activated by MG in colon cancer cells. MG exposure triggered autophagy and apoptotic cell demise. The represented processes are orchestrated by an intertwined liaison between oxidative injury and p53. The early generation of ROS, accompanied with intracellular calcium increase and ER stress, stimulated p53 to switch the autophagy toward apoptotic cell death.</p>
Full article ">
15 pages, 2098 KiB  
Article
Red Quinoa Hydrolysates with Antioxidant Properties Improve Cardiovascular Health in Spontaneously Hypertensive Rats
by Miguel López-Moreno, Estefanía Jiménez-Moreno, Antonio Márquez Gallego, Gema Vera Pasamontes, José Antonio Uranga Ocio, Marta Garcés-Rimón and Marta Miguel-Castro
Antioxidants 2023, 12(6), 1291; https://doi.org/10.3390/antiox12061291 - 16 Jun 2023
Cited by 7 | Viewed by 2158
Abstract
In recent years, quinoa has been postulated as an emerging crop for the production of functional foods. Quinoa has been used to obtain plant protein hydrolysates with in vitro biological activity. The aim of the present study was to evaluate the beneficial effect [...] Read more.
In recent years, quinoa has been postulated as an emerging crop for the production of functional foods. Quinoa has been used to obtain plant protein hydrolysates with in vitro biological activity. The aim of the present study was to evaluate the beneficial effect of red quinoa hydrolysate (QrH) on oxidative stress and cardiovascular health in an in vivo experimental model of hypertension (HTN) in the spontaneously hypertensive rat (SHR). The oral administration of QrH at 1000 mg/kg/day (QrHH) showed a significant reduction in SBP from baseline (−9.8 ± 4.5 mm Hg; p < 0.05) in SHR. The mechanical stimulation thresholds did not change during the study QrH groups, whereas in the case of SHR control and SHR vitamin C, a significant reduction was observed (p < 0.05). The SHR QrHH exhibited higher antioxidant capacity in the kidney than the other experimental groups (p < 0.05). The SHR QrHH group showed an increase in reduced glutathione levels in the liver compared to the SHR control group (p < 0.05). In relation to lipid peroxidation, SHR QrHH exhibited a significant decrease in plasma, kidney and heart malondialdehyde (MDA) values compared to the SHR control group (p < 0.05). The results obtained revealed the in vivo antioxidant effect of QrH and its ability to ameliorate HTN and its associated complications. Full article
(This article belongs to the Special Issue Dietary Antioxidants and Cardiovascular Health)
Show Figures

Figure 1

Figure 1
<p>Main modifiable and non-modifiable risk factors that can increase the likelihood of developing CVDs and hypertension.</p>
Full article ">Figure 2
<p>Body weight gain at the end of the treatment. Data represent mean values ± standard error of the mean (SEM) for each group. One-way ANOVA followed by Bonferroni post hoc test. Different letters indicate significant differences (<span class="html-italic">p</span> &lt; 0.05).</p>
Full article ">Figure 3
<p>Evolution of water intake in the different experimental groups during the study. Data represent mean values ± standard error of the mean (SEM) for each group. Two-way ANOVA followed by Bonferroni post hoc test. Different letters indicate significant differences between groups (<span class="html-italic">p</span> &lt; 0.05).</p>
Full article ">Figure 4
<p>Increase in systolic blood pressure (SBP) in the different experimental groups during the 8-week experimental period. Data represent mean values ± standard error of the mean (SEM) for each group. One-way ANOVA followed by Bonferroni post hoc test. Different letters indicate significant differences between groups (<span class="html-italic">p</span> &lt; 0.05).</p>
Full article ">Figure 5
<p>Mechanical stimulation threshold at weeks 2, 6 and 10 of the study, using the Von Frey filament test. Data represent mean values ± standard error of the mean (SEM) for each group. Two-way ANOVA followed by Bonferroni post hoc test. Different letters indicate significant differences between groups (<span class="html-italic">p</span> &lt; 0.05).</p>
Full article ">Figure 6
<p>Antioxidant capacity of kidney assessed by oxygen radical absorbance capacity (ORAC). Data represent mean values ± standard error of the mean (SEM) for each group. One-way ANOVA followed by Bonferroni post hoc test. Different letters indicate significant differences between groups (<span class="html-italic">p</span> &lt; 0.05).</p>
Full article ">Figure 7
<p>Reduced glutathione levels in the liver of the different experimental groups. Data represent mean values ± standard error of the mean (SEM) for each group. One-way ANOVA followed by Bonferroni post hoc test. Different letters indicate significant differences between groups (<span class="html-italic">p</span> &lt; 0.05).</p>
Full article ">Figure 8
<p>Malondialdehyde (MDA) concentration in plasma (<b>a</b>), heart (<b>b</b>) and kidney (<b>c</b>) at the end of the experimental period. Data represent mean values ± standard error of the mean (SEM) for each group. One-way ANOVA followed by Bonferroni post hoc test. Different letters indicate significant differences between groups (<span class="html-italic">p</span> &lt; 0.05).</p>
Full article ">
20 pages, 1060 KiB  
Review
Insights into the Role of Plasmatic and Exosomal microRNAs in Oxidative Stress-Related Metabolic Diseases
by Ayauly Duisenbek, Gabriela C. Lopez-Armas, Miguel Pérez, María D. Avilés Pérez, José Miguel Aguilar Benitez, Víctor Roger Pereira Pérez, Juan Gorts Ortega, Arailym Yessenbekova, Nurzhanyat Ablaikhanova, Germaine Escames, Darío Acuña-Castroviejo and Iryna Rusanova
Antioxidants 2023, 12(6), 1290; https://doi.org/10.3390/antiox12061290 - 16 Jun 2023
Cited by 5 | Viewed by 2057
Abstract
A common denominator of metabolic diseases, including type 2 diabetes Mellitus, dyslipidemia, and atherosclerosis, are elevated oxidative stress and chronic inflammation. These complex, multi-factorial diseases are caused by the detrimental interaction between the individual genetic background and multiple environmental stimuli. The cells, including [...] Read more.
A common denominator of metabolic diseases, including type 2 diabetes Mellitus, dyslipidemia, and atherosclerosis, are elevated oxidative stress and chronic inflammation. These complex, multi-factorial diseases are caused by the detrimental interaction between the individual genetic background and multiple environmental stimuli. The cells, including the endothelial ones, acquire a preactivated phenotype and metabolic memory, exhibiting increased oxidative stress, inflammatory gene expression, endothelial vascular activation, and prothrombotic events, leading to vascular complications. There are different pathways involved in the pathogenesis of metabolic diseases, and increased knowledge suggests a role of the activation of the NF-kB pathway and NLRP3 inflammasome as key mediators of metabolic inflammation. Epigenetic-wide associated studies provide new insight into the role of microRNAs in the phenomenon of metabolic memory and the development consequences of vessel damage. In this review, we will focus on the microRNAs related to the control of anti-oxidative enzymes, as well as microRNAs related to the control of mitochondrial functions and inflammation. The objective is the search for new therapeutic targets to improve the functioning of mitochondria and reduce oxidative stress and inflammation, despite the acquired metabolic memory. Full article
Show Figures

Figure 1

Figure 1
<p>Metabolic changes in the endothelial cell. The endogenous increased cytokines act through the Toll-like receptors (TLRs) to activate NF-kB—NLRP3 inflammasome pathway. Under conditions of hyperglycemia, the accumulation of reducing equivalents (NADH+H, FADH2) feeds the respiratory chain with electrons, which leads to greater electron escape and increased superoxide radical (O<sub>2<sup>−</sup></sub>•) production. The increment of polyol pathway leads to Advanced glycation end products (AGEs) accumulation, which act on the AGEs receptors (RAGE), leading to the activation of pro-oxidative enzymes, such as NAD(P)H oxidase, and decrement of NRF2, one important component of the anti-oxidative system of the cell. Increased β-oxidation upon the excess of fatty acid contributes to the accumulation of D-acyl glycerol, thus conducing to the activation of protein kinase C (PKC). The translocation of NF-κB to the nucleus activates the transcription of its target genes, including pro-IL-1, pro-il-18, and Pro-Caspase 1. The activation of NLRP3 inflammasome facilities the cleavage of the pro-Caspase 1 into its active form Caspase1, which in turn transforms IL-1β and IL-18 into their active forms. Subsequently, IL-18 induces the production of TNF-λ, which in turn promotes the synthesis and release of IL-6 and C reactive protein (CRP) (not shown). At the same time, the production of adhesion molecules is activated: vascular adhesion molecule-1 (VCAM-1), platelet-derived grown factor (PDGF), and vascular endothelial grown factor (VEGF). The endoplasmic reticulum (ER) stress contributes to the trigger for NLRP3 inflammasome activation and potentiating oxidative stress.</p>
Full article ">Figure 2
<p>Some examples of the exosome’s microRNAs’ participation in the pathogenesis of metabolic diseases. Exosomes derived from adipose tissue macrophages, designated ATM-EXO, isolated from obese mice have been shown to confer insulin resistance and glucose intolerance when injected into lean mice (achieved in miR-155). ATM-derived miR-29 can transfer to myocytes, hepatocytes, and adipocytes, causing insulin resistance in vivo. miR-690, abundant in exosomes of anti-inflammatory M2-like macrophages, repolarized ATM-EXO towards an anti-inflammatory M2-like phenotype.</p>
Full article ">
34 pages, 3068 KiB  
Review
Brain Iron Metabolism, Redox Balance and Neurological Diseases
by Guofen Gao, Linhao You, Jianhua Zhang, Yan-Zhong Chang and Peng Yu
Antioxidants 2023, 12(6), 1289; https://doi.org/10.3390/antiox12061289 - 16 Jun 2023
Cited by 17 | Viewed by 2455
Abstract
The incidence of neurological diseases, such as Parkinson’s disease, Alzheimer’s disease and stroke, is increasing. An increasing number of studies have correlated these diseases with brain iron overload and the resulting oxidative damage. Brain iron deficiency has also been closely linked to neurodevelopment. [...] Read more.
The incidence of neurological diseases, such as Parkinson’s disease, Alzheimer’s disease and stroke, is increasing. An increasing number of studies have correlated these diseases with brain iron overload and the resulting oxidative damage. Brain iron deficiency has also been closely linked to neurodevelopment. These neurological disorders seriously affect the physical and mental health of patients and bring heavy economic burdens to families and society. Therefore, it is important to maintain brain iron homeostasis and to understand the mechanism of brain iron disorders affecting reactive oxygen species (ROS) balance, resulting in neural damage, cell death and, ultimately, leading to the development of disease. Evidence has shown that many therapies targeting brain iron and ROS imbalances have good preventive and therapeutic effects on neurological diseases. This review highlights the molecular mechanisms, pathogenesis and treatment strategies of brain iron metabolism disorders in neurological diseases. Full article
(This article belongs to the Special Issue Iron Metabolism, Redox Balance and Neurological Diseases)
Show Figures

Figure 1

Figure 1
<p>Roles of different cells in brain iron metabolism. The main route of brain iron uptake is where the iron in the blood crosses the blood–brain barrier (BBB) via Tf-TfR1 in the apical surface of brain microvascular epithelial cells (BMVECs) and FPN1 in the basal surface of BMVECs. Iron can also enter the brain through the transcytosis of ferritin by its receptors at BBB. After iron influxes into the brain parenchymal tissue, it can enter astrocytes through their end feet surrounding BBB and then be transferred to neurons. Iron across the BBB can also directly enter the interstitial fluid of the brain and be transferred to neurons and other cells without passing through astrocytes (see black lines). Astrocytes hepcidin secreted through its end feet to directly decrease FPN1 level of BMVECs, which decreased the iron influx into brain tissues. GPI-CP expressed by astrocytes assists FPN1 in releasing iron into the brain. Astrocyte-specific <span class="html-italic">Cp</span> knockout blocks iron influx FPN1-CP pathway into the brain (see black lines and crosses). Neurons acquire both trivalent and divalent iron through TfR1, TCT1 and DMT1, while those astrocytes that are not part of the BBB acquire iron via DMT1 and ZIP molecules. Oligodendrocytes mainly uptake iron via DMT1 and Tim2. Oligodendrocytes can secrete Tf, while the activated microglia can secrete Lf. Neurons and glia store iron in ferritin and release iron through FPN1 with the coordination of CP/hephaestin or hepcidin, thereby further promoting cross-talk and interaction with other types of cells.</p>
Full article ">Figure 2
<p>Interplay between ferroptosis and iron homeostasis. Lipid peroxides that induce ferroptosis are produced through auto-oxidation and/or enzymatic activity of LOX on lipid esters generated from lipids via the activity of ACSL4 and lysophosphatidylcholine acyltransferase 3 (LPCAT3). GPx4 blocks ferroptosis by converting lipid peroxides to lipid alcohols, whereas reductions in GSH or GPx4 activity by blocking of xCT antiporter (e.g., by erastin) or inhibiting of GPx4 (e.g., RSL3) can trigger ferroptosis. The increase in labile iron pool in the cytosol via an increased iron uptake through TfR1 and/or autophagic degradation of ferritin can exacerbate ferroptosis via facilitating lipid peroxidation, and, thus, iron chelators, such as DFO and ROS scavengers (e.g., ferrostatin-1), suppress ferroptosis.</p>
Full article ">Figure 3
<p>The interplay of iron and ROS in the pathogenesis of neurological diseases. Dysregulation of iron content in neurological system and the associated generation of ROS participate in the pathological processes of Alzheimer’s disease (AD), Parkinson’s disease (PD), stroke, neuropsychiatric disorders and abnormal neurodevelopment.</p>
Full article ">Figure 4
<p>Iron accumulation is important in the pathogenesis of PD. Dysregulation of CP, IRP2, Nrf2 and FtMt alters brain iron levels, which, in turn, affects the expression of iron metabolism proteins. Iron overload and increased ROS aggravate the development and progression of PD, and their interactions with α-synuclein, dopamine, neuromelanin, Parkin and LRRK2 contribute to dopaminergic neuronal cell death and the onset of PD.</p>
Full article ">Figure 5
<p>Iron accumulation participates in the pathogenesis of AD. Elevated cellular iron is related to Aβ<sub>1-42</sub> production and tau phosphorylation. Excessive iron leads to mitochondrial dysfunction. The pre-oxidant effects of iron induce DNA damage and lipid ROS generation, contributing to cell death.</p>
Full article ">
15 pages, 3475 KiB  
Article
Effect of Copper Sulphate Exposure on the Oxidative Stress, Gill Transcriptome and External Microbiota of Yellow Catfish, Pelteobagrus fulvidraco
by Shun Zhou, Qiuhong Yang, Yi Song, Bo Cheng and Xiaohui Ai
Antioxidants 2023, 12(6), 1288; https://doi.org/10.3390/antiox12061288 - 16 Jun 2023
Cited by 4 | Viewed by 2152
Abstract
This study aimed to investigate the potential adverse effects of the practical application of copper sulfate on yellow catfish (Pelteobagrus fulvidraco) and to provide insights into the gill toxicity induced by copper sulphate. Yellow catfish were exposed to a conventional anthelmintic [...] Read more.
This study aimed to investigate the potential adverse effects of the practical application of copper sulfate on yellow catfish (Pelteobagrus fulvidraco) and to provide insights into the gill toxicity induced by copper sulphate. Yellow catfish were exposed to a conventional anthelmintic concentration of copper sulphate (0.7 mg/L) for seven days. Oxidative stress biomarkers, transcriptome, and external microbiota of gills were examined using enzymatic assays, RNA-sequencing, and 16S rDNA analysis, respectively. Copper sulphate exposure led to oxidative stress and immunosuppression in the gills, with increased levels of oxidative stress biomarkers and altered expression of immune-related differentially expressed genes (DEGs), such as IL-1β, IL4Rα, and CCL24. Key pathways involved in the response included cytokine–cytokine receptor interaction, NOD-like receptor signaling pathway, and Toll-like receptor signaling pathway. The 16S rDNA analysis revealed copper sulphate altered the diversity and composition of gill microbiota, as evidenced by a significant decrease in the abundance of Bacteroidotas and Bdellovibrionota and a significant increase in the abundance of Proteobacteria. Notably, a substantial 8.5-fold increase in the abundance of Plesiomonas was also observed at the genus level. Our findings demonstrated that copper sulphate induced oxidative stress, immunosuppression, and gill microflora dysbiosis in yellow catfish. These findings highlight the need for sustainable management practices and alternative therapeutic strategies in the aquaculture industry to mitigate the adverse effects of copper sulphate on fish and other aquatic organisms. Full article
(This article belongs to the Section Health Outcomes of Antioxidants and Oxidative Stress)
Show Figures

Figure 1

Figure 1
<p>Summary of differentially expressed genes (DEGs) in gills of yellow catfish (<span class="html-italic">Pelteobagrus fulvidraco</span>) after exposure to 0.7 mg/L copper sulphate. (<b>A</b>) Volcano plot of DEGs. Red and blue dots represent significantly up-regulated and down-regulated DEGs, respectively. (<b>B</b>) Transcriptome analysis of the number and expression of DEGs.</p>
Full article ">Figure 2
<p>GO and KEGG enrichment analysis of the differentially expressed genes in the gills of yellow catfish (<span class="html-italic">Pelteobagrus fulvidraco</span>) after exposure to 0.7 mg/L copper sulphate. The vertical axis represents different GO terms (<b>A</b>–<b>C</b>) or pathways (<b>D</b>), and the horizontal axis represents rich factor.</p>
Full article ">Figure 3
<p>QRT−PCR verification of differentially expressed genes in the gills of yellow catfish (<span class="html-italic">Pelteobagrus fulvidraco</span>) after exposure to 0.7 mg/L copper sulphate. The <span class="html-italic">X</span>−axis displays 9 DEGs, and the <span class="html-italic">Y</span>−axis represents relative fold change. The data are expressed as the means ± SD (n = 3).</p>
Full article ">Figure 4
<p>Simpson index and Shannon index of bacterial communities in the gills of yellow catfish (<span class="html-italic">Pelteobagrus fulvidraco</span>) after exposure to 0.7 mg/L copper sulphate. C: 0 mg/L copper sulphate; T: 0.7 mg/L copper sulphate.</p>
Full article ">Figure 5
<p>Relative abundances of dominant microbial phyla (<b>A</b>) and genera (<b>B</b>) in the gills of yellow catfish (<span class="html-italic">Pelteobagrus fulvidraco</span>) after exposure to 0.7 mg/L copper sulphate. C: 0 mg/L copper sulphate; T: 0.7 mg/L copper sulphate.</p>
Full article ">Figure 6
<p>(<b>A</b>) The abundance differences in microbial taxa of yellow catfish (<span class="html-italic">Pelteobagrus fulvidraco</span>) between the control group and copper sulphate exposure group at the genus level, ** <span class="html-italic">p</span> &lt; 0.01; (<b>B</b>) KEGG analysis identified various enriched pathways between the control group and copper sulphate exposure group. C: 0 mg/L copper sulphate; T: 0.7 mg/L copper sulphate.</p>
Full article ">
14 pages, 5304 KiB  
Article
The Beneficial Effect of Lomitapide on the Cardiovascular System in LDLr−/− Mice with Obesity
by Undral Munkhsaikhan, Young In Kwon, Amal M. Sahyoun, María Galán, Alexis A. Gonzalez, Karima Ait-Aissa, Ammaar H. Abidi, Adam Kassan and Modar Kassan
Antioxidants 2023, 12(6), 1287; https://doi.org/10.3390/antiox12061287 - 16 Jun 2023
Viewed by 2348
Abstract
Objectives: Homozygous familial hypercholesteremia (HoFH) is a rare, life-threatening metabolic disease, mainly caused by a mutation in the LDL receptor. If untreated, HoFH causes premature death from acute coronary syndrome. Lomitapide is approved by the FDA as a therapy to lower lipid levels [...] Read more.
Objectives: Homozygous familial hypercholesteremia (HoFH) is a rare, life-threatening metabolic disease, mainly caused by a mutation in the LDL receptor. If untreated, HoFH causes premature death from acute coronary syndrome. Lomitapide is approved by the FDA as a therapy to lower lipid levels in adult patients with HoFH. Nevertheless, the beneficial effect of lomitapide in HoFH models remains to be defined. In this study, we investigated the effect of lomitapide on cardiovascular function using LDL receptor-knockout mice (LDLr/). Methods: Six-week-old LDLr/ mice were fed a standard diet (SD) or a high-fat diet (HFD) for 12 weeks. Lomitapide (1 mg/Kg/Day) was given by oral gavage for the last 2 weeks in the HFD group. Body weight and composition, lipid profile, blood glucose, and atherosclerotic plaques were measured. Vascular reactivity and markers for endothelial function were determined in conductance arteries (thoracic aorta) and resistance arteries (mesenteric resistance arteries (MRA)). Cytokine levels were measured by using the Mesoscale discovery V-Plex assays. Results: Body weight (47.5 ± 1.5 vs. 40.3 ± 1.8 g), % of fat mass (41.6 ± 1.9% vs. 31.8 ± 1.7%), blood glucose (215.5 ± 21.9 vs. 142.3 ± 7.7 mg/dL), and lipid levels (cholesterol: 600.9 ± 23.6 vs. 451.7 ± 33.4 mg/dL; LDL/VLDL: 250.6 ± 28.9 vs. 161.1 ± 12.24 mg/dL; TG: 299.5 ± 24.1 vs. 194.1 ± 28.1 mg/dL) were significantly decreased, and the % of lean mass (56.5 ± 1.8% vs. 65.2 ± 2.1%) was significantly increased in the HFD group after lomitapide treatment. The atherosclerotic plaque area also decreased in the thoracic aorta (7.9 ± 0.5% vs. 5.7 ± 0.1%). After treatment with lomitapide, the endothelium function of the thoracic aorta (47.7 ± 6.3% vs. 80.7 ± 3.1%) and mesenteric resistance artery (66.4 ± 4.3% vs. 79.5 ± 4.6%) was improved in the group of LDLr/ mice on HFD. This was correlated with diminished vascular endoplasmic (ER) reticulum stress, oxidative stress, and inflammation. Conclusions: Treatment with lomitapide improves cardiovascular function and lipid profile and reduces body weight and inflammatory markers in LDLr/ mice on HFD. Full article
(This article belongs to the Special Issue Natural Antioxidants in Obesity and Related Diseases)
Show Figures

Figure 1

Figure 1
<p>Lomitapide reduced body weight and blood glucose and ameliorated body composition profile in LDLr<sup>−</sup>/<sup>−</sup> mice on HFD. Body weight (BW) (<b>A</b>), blood glucose (BG) (<b>B</b>), percentage of fat mass (<b>C</b>), and lean mass (<b>D</b>) in LDLr<sup>−</sup>/<sup>−</sup> control mice and mice fed with a high-fat diet (HFD) in the presence and absence of lomitapide treatment (<span class="html-italic">n</span> = 6–10). * <span class="html-italic">p</span> &lt; 0.05; ** <span class="html-italic">p</span> &lt; 0.01; *** <span class="html-italic">p</span> &lt; 0.001; **** <span class="html-italic">p</span> &lt; 0.0001.</p>
Full article ">Figure 2
<p>Lomitapide enhanced the lipid profile in LDLr<sup>−</sup>/<sup>−</sup> mice on HFD. Total cholesterol (<b>A</b>), LDL/VLDL (<b>B</b>), HDL (<b>C</b>), and TG (<b>D</b>) in plasma from LDLr<sup>−</sup>/<sup>−</sup> control mice and mice fed with a high-fat diet (HFD) treated with vehicle or lomitapide (<span class="html-italic">n</span> = 5–6). LDL/VLDL: low-density lipoprotein/very low-density lipoprotein; HDL: high-density lipoprotein; TG: triglyceride. ns &gt; 0.05; * <span class="html-italic">p</span> &lt; 0.05; ** <span class="html-italic">p</span> &lt; 0.01; *** <span class="html-italic">p</span> &lt; 0.001; **** <span class="html-italic">p</span> &lt; 0.0001.</p>
Full article ">Figure 3
<p>Oil O Red staining and quantification of plaque lesion area. Lomitapide decreased plaque surface area in the thoracic aorta of LDLr<sup>−</sup>/<sup>−</sup> mice with obesity (<span class="html-italic">n</span> = 3). Representative images of en face staining by Oil O red of the entire aorta are shown (<b>A</b>), the whole aorta quantification of the lesion area (<b>B</b>), the aortic arch area (<b>C</b>), and the descending aorta (<b>D</b>). ns &gt; 0.05; * <span class="html-italic">p</span> &lt; 0.05; ** <span class="html-italic">p</span> &lt; 0.01; *** <span class="html-italic">p</span> &lt; 0.001; **** <span class="html-italic">p</span> &lt; 0.0001 assessed by ANOVA followed by the Tukey test for multiple comparisons.</p>
Full article ">Figure 4
<p>Lomitapide prevented endothelial dysfunction in the thoracic aorta from LDLr<sup>−</sup>/<sup>−</sup> mice on HFD. Endothelium-dependent dilation to Ach (<b>A</b>), endothelium-independent dilation to SNP (<b>B</b>) (<span class="html-italic">n</span> = 11–14), immunoblots showing (T-eNOS, p-eNOS, BIP, CHOP, P65, TNFα, and GAPDH) (<b>C</b>), and quantification (<b>D</b>) in thoracic aorta from LDLr<sup>−</sup>/<sup>−</sup> lean mice and LDLr<sup>−</sup>/<sup>−</sup> obese mice treated with vehicle or lomitapide (<span class="html-italic">n</span> = 3–6). Ach: acetylcholine; SNP: sodium nitroprusside; T-eNOS: total endothelial nitric oxide synthase; p-eNOS: phosphorylated endothelial nitric oxide synthase; BIP: GRP78; CHOP: he C/EBP homologous protein; P65: NF-kappa-B; TNFα: tumor necrosis factor alpha. ns &gt; 0.05; * <span class="html-italic">p</span> &lt; 0.05; ** <span class="html-italic">p</span> &lt; 0.01; *** <span class="html-italic">p</span> &lt; 0.001; **** <span class="html-italic">p</span> &lt; 0.0001.</p>
Full article ">Figure 5
<p>Lomitapide prevented endothelial dysfunction in MRA from LDLr<sup>−</sup>/<sup>−</sup> mice on HFD. Endothelium-dependent dilation to Ach (<b>A</b>), endothelium-independent dilation to SNP (<b>B</b>) (<span class="html-italic">n</span> = 11–18), immunoblots showing (T-eNOS, p-eNOS, BIP, CHOP, P65, TNFα, and GAPDH) (<b>C</b>), and quantification (<b>D</b>) in MRA from LDLr<sup>−</sup>/<sup>−</sup> lean mice and LDLr<sup>−</sup>/<sup>−</sup> obese mice treated with vehicle or lomitapide (<span class="html-italic">n</span> = 3–6). Ach: acetylcholine; SNP: sodium nitroprusside; T-eNOS: total endothelial nitric oxide synthase; p-eNOS: phosphorylated endothelial nitric oxide synthase; BIP: GRP78; CHOP: the C/EBP homologous protein; P65: NF-kappa-B; TNFα: tumor necrosis factor alpha. ns &gt; 0.05; * <span class="html-italic">p</span> &lt; 0.05; ** <span class="html-italic">p</span> &lt; 0.01; *** <span class="html-italic">p</span> &lt; 0.001; **** <span class="html-italic">p</span> &lt; 0.0001.</p>
Full article ">Figure 6
<p>Lomitapide treatment downregulated oxidative and endoplasmic reticulum (ER) stress and inflammatory markers in thoracic aorta from LDLr<sup>−</sup>/<sup>−</sup> mice on HFD. mRNA levels for ER stress markers (BIP, ATF6, ATF4) (<b>A</b>–<b>C</b>), oxidative stress marker (NOX2) (<b>D</b>), and inflammation markers (NFkB-p65, -p50, TNFα, VCAM1) (<b>E</b>–<b>H</b>) in thoracic aorta from LDLr<sup>−</sup>/<sup>−</sup> lean mice and LDLr<sup>−</sup>/<sup>−</sup> obese mice treated with vehicle or lomitapide (<span class="html-italic">n</span> = 8–12). BIP: GRP78; ATF6 and ATF4: activating transcription factor; p65 and p50: a subunit of NF-kappa B transcription complex; TNFα: tumor necrosis factor alpha; VCAM1: vascular cell adhesion molecule 1. ns &gt; 0.05; * <span class="html-italic">p</span> &lt; 0.05; ** <span class="html-italic">p</span> &lt; 0.01; *** <span class="html-italic">p</span> &lt; 0.001; **** <span class="html-italic">p</span> &lt; 0.0001.</p>
Full article ">Figure 7
<p>Lomitapide treatment showed an anti-inflammatory effect in LDLr<sup>−</sup>/<sup>−</sup> mice with obesity. Lomitapide treatment significantly decreased inflammatory cytokines IL-6 (<b>A</b>), KC/GRO (<b>B</b>), and TNFα (<b>C</b>) in LDLr<sup>−</sup>/<sup>−</sup> obese mice (<span class="html-italic">n</span> = 12–14). IL-6: interleukin 6; KC/GRO: neutrophil-activating protein 3; TNFα: tumor necrosis factor alpha. ns &gt; 0.05; * <span class="html-italic">p</span> &lt; 0.05; *** <span class="html-italic">p</span> &lt; 0.001; **** <span class="html-italic">p</span> &lt; 0.0001.</p>
Full article ">Figure 8
<p>Schema summarizing the findings in this paper.</p>
Full article ">
22 pages, 2309 KiB  
Review
Therapeutic Applications of Plant-Derived Extracellular Vesicles as Antioxidants for Oxidative Stress-Related Diseases
by Manho Kim, Hyejun Jang, Wijin Kim, Doyeon Kim and Ju Hyun Park
Antioxidants 2023, 12(6), 1286; https://doi.org/10.3390/antiox12061286 - 16 Jun 2023
Cited by 8 | Viewed by 3061
Abstract
Extracellular vesicles (EVs) composed of a lipid bilayer are released from various cell types, including animals, plants, and microorganisms, and serve as important mediators of cell-to-cell communication. EVs can perform a variety of biological functions through the delivery of bioactive molecules, such as [...] Read more.
Extracellular vesicles (EVs) composed of a lipid bilayer are released from various cell types, including animals, plants, and microorganisms, and serve as important mediators of cell-to-cell communication. EVs can perform a variety of biological functions through the delivery of bioactive molecules, such as nucleic acids, lipids, and proteins, and can also be utilized as carriers for drug delivery. However, the low productivity and high cost of mammalian-derived EVs (MDEVs) are major barriers to their practical clinical application where large-scale production is essential. Recently, there has been growing interest in plant-derived EVs (PDEVs) that can produce large amounts of electricity at a low cost. In particular, PDEVs contain plant-derived bioactive molecules such as antioxidants, which are used as therapeutic agents to treat various diseases. In this review, we discuss the composition and characteristics of PDEVs and the appropriate methods for their isolation. We also discuss the potential use of PDEVs containing various plant-derived antioxidants as replacements for conventional antioxidants. Full article
(This article belongs to the Special Issue Plant Materials and Their Antioxidant Potential)
Show Figures

Graphical abstract

Graphical abstract
Full article ">Figure 1
<p>Schematic of disease development through damage to cellular components of ROS increased by various environmental stresses (created with BioRender.com).</p>
Full article ">Figure 2
<p><b>A</b> schematic diagram for the isolation of EVs from various sources using different isolation methods (created with BioRender.com).</p>
Full article ">Figure 3
<p>Effect of ASEVs on chronic skin wounds. (<b>A</b>) A schematic illustration of the ASEVs isolation procedure using the PEG-based precipitation method. (<b>B</b>) Effect of ASEVs on mRNA expression levels of pro-inflammatory cytokines IL-6 and IL-1b in LPS-stimulated RAW 264.7 cells. (<b>C</b>) Proliferation promoting effect of ASEVs in HDFs (*** <span class="html-italic">p</span> &lt; 0.005). Adapted from Ref. [<a href="#B84-antioxidants-12-01286" class="html-bibr">84</a>].</p>
Full article ">Figure 4
<p>Preventive effect of GEVs on colitis-associated cancer (CAC). (<b>A</b>) Structural integrity and morphological characteristics of GEVs confirmed via transmission electron microscopy (TEM) and atomic force microscopy (AFM). (<b>B</b>) Effect of GEVs on inhibition of CAC formation. Colon tumors were obtained at the end of the CAC model mouse induction (* <span class="html-italic">p</span> &lt; 0.05, ** <span class="html-italic">p</span> &lt; 0.01, ns, not significant). Adapted with permission from Ref. [<a href="#B139-antioxidants-12-01286" class="html-bibr">139</a>]. Copyright 2016 Elsevier.</p>
Full article ">Figure 5
<p>Inhibitory effect of PLEVs on UV-induced senescence in HaCaT cells. (<b>A</b>) Effect of PLEVs on mRNA expression levels of MMP-1 and COL1A2, markers of aging. (<b>B</b>) Effect of PLEVs on protein expression levels of MMP-1 and COL1A2, markers of aging. (<b>C</b>) Effect of PLEVs on SA-β-Gal levels increased by UV treatment (* <span class="html-italic">p</span> &lt; 0.05, ** <span class="html-italic">p</span> &lt; 0.01). Adapted from Ref. [<a href="#B140-antioxidants-12-01286" class="html-bibr">140</a>].</p>
Full article ">
16 pages, 1479 KiB  
Article
From Grape By-Products to Enriched Yogurt Containing Pomace Extract Loaded in Nanotechnological Nutriosomes Tailored for Promoting Gastro-Intestinal Wellness
by Ines Castangia, Federica Fulgheri, Francisco Javier Leyva-Jimenez, Maria Elena Alañón, Maria de la Luz Cádiz-Gurrea, Francesca Marongiu, Maria Cristina Meloni, Matteo Aroffu, Matteo Perra, Mohamad Allaw, Rita Abi Rached, Rodrigo Oliver-Simancas, Elvira Escribano Ferrer, Fabiano Asunis, Maria Letizia Manca and Maria Manconi
Antioxidants 2023, 12(6), 1285; https://doi.org/10.3390/antiox12061285 - 15 Jun 2023
Cited by 2 | Viewed by 1598
Abstract
Grape pomace is the main by-product generated during the winemaking process; since it is still rich in bioactive molecules, especially phenolic compounds with high antioxidant power, its transformation in beneficial and health-promoting foods is an innovative challenge to extend the grape life cycle. [...] Read more.
Grape pomace is the main by-product generated during the winemaking process; since it is still rich in bioactive molecules, especially phenolic compounds with high antioxidant power, its transformation in beneficial and health-promoting foods is an innovative challenge to extend the grape life cycle. Hence, in this work, the phytochemicals still contained in the grape pomace were recovered by an enhanced ultrasound assisted extraction. The extract was incorporated in liposomes prepared with soy lecithin and in nutriosomes obtained combining soy lecithin and Nutriose FM06®, which were further enriched with gelatin (gelatin-liposomes and gelatin-nutriosomes) to increase the samples’ stability in modulated pH values, as they were designed for yogurt fortification. The vesicles were sized ~100 nm, homogeneously dispersed (polydispersity index < 0.2) and maintained their characteristics when dispersed in fluids at different pH values (6.75, 1.20 and 7.00), simulating salivary, gastric and intestinal environments. The extract loaded vesicles were biocompatible and effectively protected Caco-2 cells against oxidative stress caused by hydrogen peroxide, to a better extent than the free extract in dispersion. The structural integrity of gelatin-nutriosomes, after dilution with milk whey was confirmed, and the addition of vesicles to the yogurt did not modify its appearance. The results pointed out the promising suitability of vesicles loading the phytocomplex obtained from the grape by-product to enrich the yogurt, offering a new and easy strategy for healthy and nutritional food development. Full article
(This article belongs to the Section Extraction and Industrial Applications of Antioxidants)
Show Figures

Figure 1

Figure 1
<p>Representative cryo-TEM images of grape pomace extract loaded liposomes (<b>A</b>), nutriosomes (<b>B</b>), gelatin-liposomes (<b>C</b>) and gelatin-nutriosomes (<b>D</b>).</p>
Full article ">Figure 2
<p>(<b>A</b>) Viability of Caco-2 cells incubated for 48 h with grape pomace extract in dispersion or loaded in liposomes, nutriosomes, gelatin-liposomes and gelatin-nutriosomes, diluted at different concentrations (40, 4 and 0.4 µg/mL). Mean values ± standard deviations were reported (n = 8). Same letter (a) indicates not statistically different values (<span class="html-italic">p</span> &gt; 0.05 versus untreated cells) and statistically different from other (<span class="html-italic">p</span> &lt; 0.05 versus other formulations). (<b>B</b>) Viability of Caco-2 cells stressed with hydrogen peroxide (1:40,000) and treated with grape pomace extract (4 µg/mL) in dispersions or loaded in liposomes, nutriosomes, gelatin-liposomes and gelatin-nutriosomes. Mean values ± standard deviations were reported (n = 8). Same letter (a and b) indicates not statistically different values (<span class="html-italic">p</span> &gt; 0.05 versus untreated cells) and statistically different from other (<span class="html-italic">p</span> &lt; 0.05 versus other formulations).</p>
Full article ">Figure 3
<p>Viscosity values of yogurt containing different concentrations (0, 5, 10 and 20%) of liposomes, nutriosomes, gelatin-liposomes and gelatin-nutriosomes. Mean values ± standard deviations (error bars) were reported. Same letter (a, b, c, d, e, f, g and h) indicates not statistically different values (<span class="html-italic">p</span> &gt; 0.05), while different letters indicates statistically different values (<span class="html-italic">p</span> &lt; 0.05).</p>
Full article ">
16 pages, 10320 KiB  
Article
Caffeic Acid, a Polyphenolic Micronutrient Rescues Mice Brains against Aβ-Induced Neurodegeneration and Memory Impairment
by Amjad Khan, Jun Sung Park, Min Hwa Kang, Hyeon Jin Lee, Jawad Ali, Muhammad Tahir, Kyonghwan Choe and Myeong Ok Kim
Antioxidants 2023, 12(6), 1284; https://doi.org/10.3390/antiox12061284 - 15 Jun 2023
Cited by 7 | Viewed by 2159
Abstract
Oxidative stress plays an important role in cognitive dysfunctions and is seen in neurodegeneration and Alzheimer’s disease (AD). It has been reported that the polyphenolic compound caffeic acid possesses strong neuroprotective and antioxidant effects. The current study was conducted to investigate the therapeutic [...] Read more.
Oxidative stress plays an important role in cognitive dysfunctions and is seen in neurodegeneration and Alzheimer’s disease (AD). It has been reported that the polyphenolic compound caffeic acid possesses strong neuroprotective and antioxidant effects. The current study was conducted to investigate the therapeutic potential of caffeic acid against amyloid beta (Aβ1–42)-induced oxidative stress and memory impairments. Aβ1–42 (5 μL/5 min/mouse) was administered intracerebroventricularly (ICV) into wild-type adult mice to induce AD-like pathological changes. Caffeic acid was administered orally at 50 mg/kg/day for two weeks to AD mice. Y-maze and Morris water maze (MWM) behavior tests were conducted to assess memory and cognitive abilities. Western blot and immunofluorescence analyses were used for the biochemical analyses. The behavioral results indicated that caffeic acid administration improved spatial learning, memory, and cognitive abilities in AD mice. Reactive oxygen species (ROS) and lipid peroxidation (LPO) assays were performed and showed that the levels of ROS and LPO were markedly reduced in the caffeic acid-treated mice, as compared to Aβ-induced AD mice brains. Moreover, the expression of nuclear factor erythroid 2–related factor 2 (Nrf2) and heme oxygenase-1 (HO-1) were regulated with the administration of caffeic acid, compared to the Aβ-injected mice. Next, we checked the expression of ionized calcium-binding adaptor molecule 1 (Iba-1), glial fibrillary acidic proteins (GFAP), and other inflammatory markers in the experimental mice, which suggested enhanced expression of these markers in AD mice brains, and were reduced with caffeic acid treatment. Furthermore, caffeic acid enhanced synaptic markers in the AD mice model. Additionally, caffeic acid treatment also decreased Aβ and BACE-1 expression in the Aβ-induced AD mice model. Full article
Show Figures

Figure 1

Figure 1
<p>Chemical structure of caffeic acid.</p>
Full article ">Figure 2
<p>Neuroprotective effect of caffeic acid on memory functions Aβ and BACE-1 expression. (<b>a</b>) Percentage of spontaneous alteration behavior of experimental mice in the Y-maze. (<b>b</b>) Average escape latency to reach the hidden platform until day 5. (<b>c</b>) Time spent by the mice in the quadrant where the platform was present in training; (<b>d</b>) represents the probe test. (<b>e</b>) Western blot analysis of Aβ and BACE-1 in the experimental animal’s hippocampus. The bands were quantified using ImageJ software and the difference is shown by their respective histogram. β-actin was used as a loading control. (<b>f</b>) Confocal microscopy of Aβ and representative histogram and DAPI staining (blue) in the hippocampus (Cornu Ammonis (CA1) and Dentate Gyrus (DG regions) of adult mice. The density values are relative to those in the control group and are expressed in arbitrary units (AU), magnification 10×, scale bar = 50 μm. The data are presented as the mean ± SEM of 4 mice/group for Western blot and 4 mice/group for confocal microscopy, and are representative of three independent experiments. ω indicates a significant difference from saline-injected control mice; σ indicates a significant difference from Aβ<sub>1–42</sub>-injected mice. Significance = ω <span class="html-italic">p</span> &lt; 0.05, σ <span class="html-italic">p</span> &lt; 0.05.</p>
Full article ">Figure 3
<p>Effects of caffeic acid on oxidative stress. (<b>a</b>,<b>b</b>) Representative histograms showing ROS and LPO in the experimental mice brain hippocampus. (<b>c</b>) Analysis of the protein expression in Nrf2 and HO-1 by immunoblotting in the mouse hippocampus. Using ImageJ software, the bands were quantified, and the difference is displayed by the histogram for each band. β-actin was used as a loading control. (<b>d</b>) Immunofluorescence of Nrf2 and representative histogram and DAPI staining (blue) in the hippocampus (Cornu Ammonis (CA1) and Dentate Gyrus (DG regions) of adult mice. The density values are relative to those in the control group and are expressed in arbitrary units (AU), magnification 10×, scale bar = 50 μm. The data are presented as the mean ± SEM of 4 mice/group for Western blot and 4 mice/group for confocal microscopy and are representative of three independent experiments. ω indicates a significant difference from saline-injected control mice; σ indicates a significant difference from Aβ<sub>1–42</sub>-injected mice. Significance = ω <span class="html-italic">p</span> &lt; 0.05, σ <span class="html-italic">p</span> &lt; 0.05.</p>
Full article ">Figure 4
<p>Caffeic acid reduced the activation of glial cells. (<b>a</b>) Western blot examination of GFAP and Iba-1 of mice’s hippocampus. The bands were quantified using ImageJ software and the difference showed by their respective histogram. β-actin was used as a loading control. (<b>b</b>) Confocal microscopy analysis of GFAP (green) and representative histogram and DAPI staining (blue) in the hippocampus of adult mice. The density values are relative to those in the control group and are expressed in arbitrary units (AU), magnification 10×, scale bar = 50 μm. The data are presented as the mean ± SEM of 4 mice/group for Western blot and 4 mice/group for confocal microscopy, and are representative of three independent experiments. ω indicates a significant difference from saline-injected control mice; σ indicates a significant difference from Aβ<sub>1–42</sub>-injected mice. Significance = ω <span class="html-italic">p</span> &lt; 0.05, σ <span class="html-italic">p</span> &lt; 0.05.</p>
Full article ">Figure 5
<p>Effects of caffeic acid on inflammatory markers. (<b>a</b>) Western blot analysis of p-NFκB, TNF-α, and IL-1β in the brain of adult mice. The bands were quantified using ImageJ software and the difference showed by their respective histogram. β-actin was used as a loading control. (<b>b</b>) Immunofluorescence analysis of TLR4 (green), along with its respective histogram, and DAPI staining (blue) in the hippocampal of adult mice. The density values are relative to those in the control group and are expressed in arbitrary units (AU), magnification 10×, scale bar = 50 μm. The data are presented as the mean ± SEM of 4 mice/group for Western blot and 4 mice/group for confocal microscopy, and are representative of three independent experiments. ω indicates a significant difference from saline-injected control mice; σ indicates a significant difference from Aβ<sub>1–42</sub>-injected mice. Significance = ω <span class="html-italic">p</span> &lt; 0.05, σ <span class="html-italic">p</span> &lt; 0.05.</p>
Full article ">Figure 6
<p>Effects of caffeic acid on p-PI3k/p-AKT and BDNF in Aβ-injected mice brains. Western blot analysis of p-PI3k, p-AKT, and BDNF, in different experimental mice groups. Using ImageJ software, the bands were quantified and the difference is shown by their respective histogram. β-actin was used as a loading control. The data are presented as the mean ± SEM of 4 mice/group for Western blot and are representative of three independent experiments. ω indicates a significant difference from saline-injected control mice; σ indicates a significant difference from Aβ<sub>1–42</sub>-injected mice. Significance = ω <span class="html-italic">p</span> &lt; 0.05, σ <span class="html-italic">p</span> &lt; 0.05.</p>
Full article ">Figure 7
<p>Neuroprotective properties of caffeic acid on synaptic protein expression. (<b>a</b>) Western blot analysis of synaptic proteins synaptosomal-associated protein 25 (SNAP-25) synaptophysin, and SNAP-23 proteins. ImageJ software was used to quantify the bands, and the difference is shown by their respective histogram. β-actin was used as a loading control. (<b>b</b>) Immunofluorescence analysis of postsynaptic density protein-95 (PSD-95) (green), along with its respective histogram, and DAPI staining (blue) in the brain (hippocampus). The density values are relative to those in the control group and are expressed in arbitrary units (AU), magnification 10×, scale bar = 50 μm. The data are presented as the mean ± SEM of 4 mice/group for Western blot and 4 mice/group for confocal microscopy, and are representative of three independent experiments. ω indicates a significant difference from saline-injected control mice; σ indicates a significant difference from Aβ<sub>1–42</sub>-injected mice. Significance = ω <span class="html-italic">p</span> &lt; 0.05, σ <span class="html-italic">p</span> &lt; 0.05.</p>
Full article ">
Previous Issue
Next Issue
Back to TopTop