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

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

Years

Between: -

Subjects

remove_circle_outline
remove_circle_outline
remove_circle_outline
remove_circle_outline
remove_circle_outline
remove_circle_outline
remove_circle_outline
remove_circle_outline
remove_circle_outline

Journals

remove_circle_outline
remove_circle_outline
remove_circle_outline
remove_circle_outline
remove_circle_outline
remove_circle_outline
remove_circle_outline
remove_circle_outline
remove_circle_outline
remove_circle_outline
remove_circle_outline
remove_circle_outline
remove_circle_outline
remove_circle_outline
remove_circle_outline
remove_circle_outline
remove_circle_outline
remove_circle_outline
remove_circle_outline
remove_circle_outline
remove_circle_outline
remove_circle_outline
remove_circle_outline
remove_circle_outline
remove_circle_outline
remove_circle_outline
remove_circle_outline
remove_circle_outline
remove_circle_outline
remove_circle_outline
remove_circle_outline
remove_circle_outline
remove_circle_outline
remove_circle_outline
remove_circle_outline
remove_circle_outline

Article Types

Countries / Regions

remove_circle_outline
remove_circle_outline
remove_circle_outline
remove_circle_outline
remove_circle_outline
remove_circle_outline
remove_circle_outline
remove_circle_outline
remove_circle_outline
remove_circle_outline
remove_circle_outline
remove_circle_outline

Search Results (28,135)

Search Parameters:
Keywords = oxidative stress

Order results
Result details
Results per page
Select all
Export citation of selected articles as:
13 pages, 1525 KiB  
Article
MiR-146a Is Mutually Regulated by High Glucose-Induced Oxidative Stress in Human Periodontal Ligament Cells
by Chihiro Fumimoto, Nobuhiro Yamauchi, Emika Minagawa and Makoto Umeda
Int. J. Mol. Sci. 2024, 25(19), 10702; https://doi.org/10.3390/ijms251910702 (registering DOI) - 4 Oct 2024
Abstract
The high-glucose conditions caused by diabetes mellitus (DM) exert several effects on cells, including inflammation. miR-146a, a kind of miRNA, is involved in inflammation and may be regulated mutually with reactive oxygen species (ROS), which are produced under high-glucose conditions. In the present [...] Read more.
The high-glucose conditions caused by diabetes mellitus (DM) exert several effects on cells, including inflammation. miR-146a, a kind of miRNA, is involved in inflammation and may be regulated mutually with reactive oxygen species (ROS), which are produced under high-glucose conditions. In the present study, we used human periodontal ligament cells (hPDLCs) to determine the effects of the high-glucose conditions of miR-146a and their involvement in the regulation of oxidative stress and inflammatory cytokines using Western blotting, PCR, ELISA and other methods. When hPDLCs were subjected to high glucose (24 mM), cell proliferation was not affected; inflammatory cytokine expression, ROS induction, interleukin-1 receptor-associated kinase 1 (IRAK1) and TNF receptor-associated factor 6 (TRAF6) expression increased, but miR-146a expression decreased. Inhibition of ROS induction with the antioxidant N-acetyl-L-cysteine restored miR-146a expression and decreased inflammatory cytokine expression compared to those under high-glucose conditions. In addition, overexpression of miR-146a significantly suppressed the expression of the inflammatory cytokines IRAK1 and TRAF6, regardless of the glucose condition. Our findings suggest that oxidative stress and miR-146a expression are mutually regulated in hPDLCs under high-glucose conditions. Full article
(This article belongs to the Special Issue Molecular Insight into Oral Diseases)
Show Figures

Figure 1

Figure 1
<p>High glucose concentrations increase inflammatory cytokine production without affecting the proliferation of human periodontal ligament cells (hPDLCs). (<b>A</b>) hPDLCs stained with calcein-AM were imaged by a fluorescence microscope at 24, 48 and 72 h after incubation (scale bars: 500 μm). (<b>B</b>) The data for live cell staining are shown as the percentage of the area stained with calcein. (<b>C</b>) Cell proliferation was measured 24, 48 and 72 h after incubation. (<b>D</b>–<b>F</b>) <span class="html-italic">IL-6</span> mRNA gene expression was assessed at the point of 24, 48 and 72 h after stimulation. (<b>G</b>) IL-6 production was measured at 24, 48 and 72 h. (<b>H</b>–<b>J</b>) <span class="html-italic">IL-8</span> mRNA gene expression was assessed at the point of 24, 48 and 72 h after stimulation. (<b>K</b>) IL-8 production was measured at the point of 24, 48 and 72 h after stimulation. Significant increase control: *<span class="html-italic">p</span> &lt; 0.05, **<span class="html-italic"> p</span> &lt; 0.01.</p>
Full article ">Figure 2
<p>ROS induction was enhanced under high-glucose conditions. (<b>A</b>) The fluorescence intensity of ROS levels was measured using a plate reader and the data were compared to those of the control (scale bar: 200 μm). (<b>B</b>) Fluorescent staining was performed using a fluorescence microscope. (<b>C</b>) The mitochondria were stained after stimulation using MitoTracker (scale bar: 50 μm). (<b>D</b>–<b>F</b>) NO production was measured using a Griess assay at 24, 48 and 72 h after stimulation. Significant increases compared with the control: *<span class="html-italic"> p</span> &lt; 0.05, ††<span class="html-italic">p</span> &lt; 0.05.</p>
Full article ">Figure 3
<p>High-glucose conditions decrease miR-146a expression and increase IRAK1, TRAF6 and NF-kB expression. (<b>A</b>) MiR-146a expression was measured at 72 h. (<b>B</b>) Immunofluorescence staining of IRAK1 and TRAF6 was visualized by confocal laser microscopy after 72 h of incubation (scale bar: 50 μm). (<b>C</b>) The levels of IRAK1 and TRAF6 were analyzed using Western blotting. Western blotting was performed on protein extracts of these cells with antibodies against the indicated proteins, using β-actin as a loading control. (<b>D</b>) TRAF6 expression was quantified using ImageJ software. (<b>E</b>) IRAK1 expression was quantified using ImageJ software (version 1.53e). (<b>F</b>) NF-kB and pNF-kB were analyzed using Western blotting. Western blotting was performed on protein extracts of these cells with antibodies against the indicated proteins, using β-actin as a loading control. (<b>G</b>) NF-kB and pNF-kB expression was quantified using ImageJ software. Significant increases compared with the control: *<span class="html-italic"> p</span> &lt; 0.05, **<span class="html-italic"> p</span> &lt; 0.01. Significant decreases compared with those at 5.5 mM: ††<span class="html-italic"> p</span> &lt; 0.01.</p>
Full article ">Figure 4
<p>MiR-146a expression and inflammatory cytokine production of hPDLCs are controlled by the production of ROS, which is influenced by NAC. NAC (1, 5 and 10 mM) was added to the medium, and ROS levels were examined under a high glucose concentration. (<b>A</b>) Fluorescence staining was performed using a fluorescence microscope. Fluorescence intensity was measured using a microplate reader and the data were compared with those of the control (scale bar: 200 μm). Significant increases were observed compared with those at 5.5 mM glucose: *<span class="html-italic"> p</span> &lt; 0.05. Significant decreases were observed compared with those at 24 mM glucose: ††<span class="html-italic">p</span> &lt; 0.01. (<b>B</b>) Expression of miR-146a was measured at 72 h. (<b>C</b>) <span class="html-italic">IL-6</span> mRNA gene expression was measured at 72 h. (<b>D</b>) IL-6 production was measured at 72 h. (<b>E</b>) <span class="html-italic">IL-8</span> mRNA gene expression was measured at 72 h. (<b>F</b>) IL-8 production was measured at 72 h. Significant increases were compared with those at 24 mM glucose: *<span class="html-italic"> p</span> &lt; 0.05, **<span class="html-italic"> p</span> &lt; 0.01. Significant decreases were compared with those at 24 mM glucose: †<span class="html-italic"> p</span> &lt; 0.05, ††<span class="html-italic"> p</span> &lt; 0.01.</p>
Full article ">Figure 5
<p>miR-146a down-regulates the inflammatory response and ROS induction in hPDLCs. (<b>A</b>,<b>C</b>) After 24 h of transfection, cultures were stimulated with 5.5 mM or 24 mM glucose. <span class="html-italic">IL-6</span> and <span class="html-italic">IL-8</span> mRNA expression was measured at 72 h post-stimulation. (<b>B</b>,<b>D</b>) After 24 h of transfection, cultures were stimulated with 5.5 mM or 24 mM glucose. IL-6 and IL-8 production was measured at 72 h post-stimulation. (<b>E</b>,<b>F</b>) After 24 h of transfection, cultures were stimulated with 5.5 mM or 24 mM glucose. Fluorescence staining was performed using a fluorescence microscope. Fluorescence intensity was measured using a microplate reader, and the data were compared with those of the control (scale bar: 200 μm). Significant decreases compared with miR-146a NC: †<span class="html-italic"> p</span> &lt; 0.05, ††<span class="html-italic"> p</span> &lt; 0.01.</p>
Full article ">Figure 6
<p>MiR-146a down-regulates IRAK1, TRAF6 and NF-κB in hPDLCs. (<b>A</b>) After 24 h of transfection, cultures were stimulated with 5.5 mM or 24 mM glucose. Immunofluorescence staining of IRAK1 and TRAF6 was visualized by confocal laser microscopy after 72 h of incubation (scale bar: 50 μm). (<b>B</b>) After 24 h of transfection, cultures were stimulated with 5.5 mM or 24 mM glucose. The levels of IRAK1 and TRAF6 were analyzed using Western blotting. Western blotting was performed on the protein extracts of these cells with antibodies against the indicated proteins, using β-actin as a loading control. (<b>C</b>) IRAK1 expression was quantified using ImageJ software. (<b>D</b>) TRAF6 expression was quantified using ImageJ software. (<b>E</b>) After 24 h of transfection, cultures were stimulated with 5.5 mM or 24 mM glucose. The levels of NF-κB and pNF-κB were analyzed using Western blotting. Western blotting was performed on the protein extracts of these cells with antibodies against the indicated proteins, using β-actin as a loading control. (<b>F</b>) NF-κB and p-NF-κB expression was quantified using ImageJ software. Significant decreases compared with miR-146a NC: ††<span class="html-italic"> p</span> &lt; 0.01.</p>
Full article ">
31 pages, 6195 KiB  
Article
Evaluating the Root Extract of Reynoutria ciliinervis (Nakai) Moldenke: An Analysis of Active Constituents, Antioxidant Potential, and Investigation of Hepatoprotective Effects in Rats
by Zheng Xing, Yang Han, Hao Pang, Li Li, Guangqing Xia, Junyi Zhu, Jing Han and Hao Zang
Molecules 2024, 29(19), 4701; https://doi.org/10.3390/molecules29194701 - 4 Oct 2024
Abstract
Reynoutria ciliinervis (Nakai) Moldenke (R. ciliinervis) root, a traditional Chinese medicine, was found to exhibit remarkable pharmacological properties through a series of comprehensive investigations. Our study commenced with a qualitative phytochemical analysis that identified 12 bioactive compounds within the plant. Subsequently, [...] Read more.
Reynoutria ciliinervis (Nakai) Moldenke (R. ciliinervis) root, a traditional Chinese medicine, was found to exhibit remarkable pharmacological properties through a series of comprehensive investigations. Our study commenced with a qualitative phytochemical analysis that identified 12 bioactive compounds within the plant. Subsequently, utilizing ultraviolet-visible spectrophotometry, the methanol extract emerged as the optimal solvent extract, which was abundant in diverse classes of compounds such as carbohydrates, phenolics, steroids, alkaloids, phenolic acids, and tannins. In vitro antioxidant assays underscored the exceptional free radical scavenging, metal ion chelation, hydrogen peroxide scavenging, singlet oxygen quenching, and β-carotene bleaching capabilities of the methanol extract, significantly outperforming other solvent extracts. Further ultra high-performance liquid chromatography–electrospray ionization–quadrupole time of flight–mass spectrometry analysis revealed the presence of 45 compounds, predominantly anthraquinones and phenolics, in the methanol extract. The extract demonstrated robust stability under various conditions, including high temperatures, varying pH levels, and simulated gastrointestinal digestion as well as efficacy in inhibiting the oxidation in edible oils. Acute toxicity tests in mice confirmed the safety of the methanol extract and provided a valuable dosage reference for future studies. Importantly, high-dose methanol extract exhibited a significant pre-protective effect against D-galactosamine-induced liver injury in rats, as evidenced by reduced alanine aminotransferase, aspartate aminotransferase, γ-glutamyl transpeptidase, malondialdehyde levels, and elevated catalase and albumin levels. These findings suggest a potential role for the methanol extract of R. ciliinervis root in treating oxidative stress-related disorders, highlighting the plant’s immense medicinal potential. Our research offers a thorough evaluation of the bioactive components, antioxidant properties, stability, and liver-protecting effects of the methanol extract, setting the stage for deeper investigation and potential clinical applications. Full article
(This article belongs to the Special Issue Medicinal Value of Natural Bioactive Compounds and Plant Extracts II)
9 pages, 3846 KiB  
Article
Short-Term Metformin Therapy in Clomiphene Citrate Resistant PCOS Patients Improves Fertility Outcome by Regulating Follicular Fluid Redox Balance: A Case-Controlled Study
by Mustafa Tas
Diagnostics 2024, 14(19), 2215; https://doi.org/10.3390/diagnostics14192215 - 4 Oct 2024
Abstract
Objectives: To determine the effect of short-term metformin administration on follicular fluid (FF) total oxidant status (TOS), total antioxidant status (TAS), oxidative stress index (OSI) and nuclear factor kappa B (NF-kB) in women with clomiphene citrate-resistant polycystic ovary syndrome (PCOS). Methods: Fifty-eight patients [...] Read more.
Objectives: To determine the effect of short-term metformin administration on follicular fluid (FF) total oxidant status (TOS), total antioxidant status (TAS), oxidative stress index (OSI) and nuclear factor kappa B (NF-kB) in women with clomiphene citrate-resistant polycystic ovary syndrome (PCOS). Methods: Fifty-eight patients aged 23–34 who were planned to have intracytoplasmic sperm injection due to clomiphene citrate-resistant PCOS were included in the study. Participants were divided into two groups according to whether they used metformin or not. While 30 of 58 PCOS patients were using short-term metformin in combination with controlled ovarian stimulation, 28 PCOS patients were not using metformin. Metformin was started in the mid-luteal period and continued until the day before oocyte retrieval at 850 mg twice daily. To determine FF-NF-kB, TAS, TOS and OSI values, a dominant follicle ≥17–18 mm in diameter was selected for aspiration. Results: The number of mature follicles and fertilization rates of the metformin group were significantly higher than those not taking metformin. FF-TOS and OSI of the metformin group were significantly lower than those of the group not receiving metformin. Patients receiving metformin had higher FF-TAS than the group not receiving metformin. FF-NF-kB levels of the metformin group were significantly lower than the group not receiving metformin. Insulin resistance, FF-NF-kB and FF-TOS were negatively correlated with the number of mature oocytes. FF-TAS was positively correlated with the number of oocytes. Conclusions: Short-term metformin treatment in clomiphene-resistant PCOS improves the number of mature follicles and fertilization rates by regulating the intra-follicle redox balance. Full article
(This article belongs to the Section Pathology and Molecular Diagnostics)
Show Figures

Figure 1

Figure 1
<p>Graphical representation of follicular fluid NF-kB (<b>A</b>), TOS (<b>B</b>), TAS (<b>C</b>) and OSI (<b>D</b>) levels in groups that received and did not receive short-term metformin treatment. Note that inflammatory markers (NF-KB) and oxidant markers (TOS, OSI) decreased in the metformin group, while antioxidant TAS increased.</p>
Full article ">Figure 2
<p>Graphical representation of correlation matrix of variables. Blue colors indicate positive correlations, and red colors indicate negative correlations.</p>
Full article ">
13 pages, 3957 KiB  
Article
Molecular Mechanism of 5,6-Dihydroxyflavone in Suppressing LPS-Induced Inflammation and Oxidative Stress
by Yujia Cao, Yee-Joo Tan and Dejian Huang
Int. J. Mol. Sci. 2024, 25(19), 10694; https://doi.org/10.3390/ijms251910694 - 4 Oct 2024
Abstract
5,6-dihydroxyflavone (5,6-DHF), a flavonoid that possesses potential anti-inflammatory and antioxidant activities owing to its special catechol motif on the A ring. However, its function and mechanism of action against inflammation and cellular oxidative stress have not been elucidated. In the current study, 5,6-DHF [...] Read more.
5,6-dihydroxyflavone (5,6-DHF), a flavonoid that possesses potential anti-inflammatory and antioxidant activities owing to its special catechol motif on the A ring. However, its function and mechanism of action against inflammation and cellular oxidative stress have not been elucidated. In the current study, 5,6-DHF was observed inhibiting lipopolysaccharide (LPS)-induced nitric oxide (NO) and cytoplasmic reactive oxygen species (ROS) production with the IC50 of 11.55 ± 0.64 μM and 0.8310 ± 0.633 μM in murine macrophages, respectively. Meanwhile, 5,6-DHF suppressed the overexpression of pro-inflammatory mediators such as proteins and cytokines and eradicated the accumulation of mitochondrial ROS (mtROS). The blockage of the activation of cell surface toll-like receptor 4 (TLR4), impediment of the phosphorylation of c-Jun N-terminal kinase (JNK) and p38 from the mitogen-activated protein kinases (MAPK) pathway, Janus kinase 2 (JAK2) and signal transducer and activator of transcription 3 (STAT3) from the JAK-STAT pathway, and p65 from nuclear factor-κB (NF-κB) pathways were involved in the process of 5,6-DHF suppressing inflammation. Furthermore, 5,6-DHF acted as a cellular ROS scavenger and heme-oxygenase 1 (HO-1) inducer in relieving cellular oxidative stress. Importantly, 5,6-DHF exerted more potent anti-inflammatory activity than its close structural relatives, such as baicalein and chrysin. Overall, our findings pave the road for further research on 5,6-DHF in animal models. Full article
(This article belongs to the Special Issue Cellular Redox Mechanisms in Inflammation and Programmed Cell Death)
Show Figures

Figure 1

Figure 1
<p>Chemical structure of 5,6-DHF (<b>A</b>). Dose response curves of 5,6-DHF on LPS-induced NO production (<b>B</b>) and cell cytotoxicity of 5,6-DHF on RAW 264.7 cells (<b>C</b>). Comparison of the anti-NO activity of 5,6-DHF and its structural analogues at 25 μM (<b>D</b>). The values shown are the mean ± SD of three independent experiments in duplicate. Different letters indicate statistical significance (<span class="html-italic">p</span> &lt; 0.05).</p>
Full article ">Figure 2
<p>The suppressive effects of 5,6-DHF on pro-inflammatory protein expression in LPS-stimulated RAW 264.7 models (<b>A</b>). The expression levels of COX-2 (<b>B</b>), iNOS (<b>C</b>), and TLR4 (<b>D</b>) were determined by Western blot. The inhibitory effects of 5,6-DHF on mRNA expression levels of IL-1β (<b>E</b>), IL-6 (<b>F</b>), and TNF-α (<b>G</b>) tested by qRT-PCR. Data points and bar represent arithmetic means ± SD. Ns, not significant. ** <span class="html-italic">p</span> &lt; 0.01 and **** <span class="html-italic">p</span> &lt; 0.0001 as compared to DMSO group.</p>
Full article ">Figure 3
<p>Effects of 5,6-DHF on MAPK pathway in LPS-induced RAW 264.7 cells (<b>A</b>). Suppressive effects of 5,6-DHF on the LPS-induced p-p38 (<b>B</b>), p-JNK, (<b>C</b>) and p-ERK1/2 (<b>D</b>) expressions and the phosphorylation level of p38 (<b>E</b>), JNK (<b>F</b>), and ERK1/2 (<b>G</b>). All expressions were normalized to that of DMSO treatment. Data points and bar represent arithmetic means ± SD. Ns, not significant. ** <span class="html-italic">p</span> &lt; 0.01 and **** <span class="html-italic">p</span> &lt; 0.0001 as compared to DMSO group.</p>
Full article ">Figure 4
<p>Effects of 5,6-DHF on JAK-STAT pathway in LPS-induced RAW 264.7 cells (<b>A</b>). Suppressive effects of 5,6-DHF on the LPS-induced p-JAK2 (<b>B</b>) and p-STAT3 (<b>C</b>) expression and the phosphorylation level of JAK2 (<b>D</b>) and STAT3 (<b>E</b>). All expressions were normalized to that of DMSO treatment. Data points and bar represent arithmetic means ± SD. Ns, not significant. ** <span class="html-italic">p</span> &lt; 0.01, *** <span class="html-italic">p</span> &lt; 0.001, and **** <span class="html-italic">p</span> &lt; 0.0001 as compared to DMSO group.</p>
Full article ">Figure 5
<p>Effects of 5,6-DHF on NF-κB pathway in LPS-induced RAW 264.7 cells (<b>A</b>); suppressive effects of 5,6-DHF on the LPS-induced p-p65 expression (<b>B</b>); and the phosphorylation level of p65 (<b>C</b>). All expressions were normalized to that of DMSO treatment. Data points and bar represent arithmetic means ± SD. Ns, not significant. * <span class="html-italic">p</span> &lt; 0.05 and ** <span class="html-italic">p</span> &lt; 0.01 as compared to DMSO group.</p>
Full article ">Figure 6
<p>Cellular ROS-scavenging activity of 5,6-DHF on LPS-induced RAW264.7 cells, images were captured by fluorescent microscope (<b>A</b>); blue: cell nuclei stained by H33342; and green: cellular ROS stained by H<sub>2</sub>DCFDA. Dose-response curve of ROS-scavenging activity of 5,6-DHF; scale bar 20 μm (<b>B</b>). Mitochondrial ROS-scavenging activity of 5,6-DHF on LPS-induced RAW264.7 cells (<b>C</b>). The fluorescence was visualized by fluorescent microscopy (blue: cell nuclei stained by H33342; and red: mitochondrial ROS stained by MitoSOX red). The influence of 5,6-DHF on the expression levels of HO-1; scale bar 20 μm (<b>D</b>) Data points and bar represent arithmetic means ± SD. Ns, not significant. *** <span class="html-italic">p</span> &lt; 0.001 and **** <span class="html-italic">p</span> &lt; 0.0001 as compared to DMSO group.</p>
Full article ">Figure 7
<p>Schematic diagram of 5,6-DHF mechanisms in inhibiting LPS-induced inflammatory responses and oxidative stress in RAW 264.7 cells.</p>
Full article ">
17 pages, 4163 KiB  
Article
Canola Oil Ameliorates Obesity by Suppressing Lipogenesis and Reprogramming the Gut Microbiota in Mice via the AMPK Pathway
by Jing Gao, Li Ma, Jie Yin, Tiejun Li, Yulong Yin and Yongzhong Chen
Nutrients 2024, 16(19), 3379; https://doi.org/10.3390/nu16193379 - 4 Oct 2024
Abstract
Background: obesity is a worldwide problem that seriously endangers human health. Canola oil (Col) has been reported to regulate hepatic steatosis by influencing oxidative stress and lipid metabolism in Kunming mice. However, whether Col exhibits an anti-obesity effect by altering the gut microbiota [...] Read more.
Background: obesity is a worldwide problem that seriously endangers human health. Canola oil (Col) has been reported to regulate hepatic steatosis by influencing oxidative stress and lipid metabolism in Kunming mice. However, whether Col exhibits an anti-obesity effect by altering the gut microbiota remains unknown. Methods: in this study, we observed that a high-fat diet increased lipogenesis and gut microbiota disorder in C57BL/6J male mice, while the administration of Col suppressed lipogenesis and improved gut microbiota disorder. Results: the results show that Col markedly reduced the final body weight and subcutaneous adipose tissue of C57BL/6J male mice fed a high-fat diet (HFD) after 6 weeks of administration. However, although Col did not effectively increase the serum concentration of HDL, we found that treatment with Col notably inhibited the low-density lipoprotein (LDL), total cholesterol (TC), and triglycerides (TGs) in HFD mice. Furthermore, Col ameliorated obesity in the liver compared to mice that were only fed a high-fat diet. We also found that Col significantly inhibited the relative expression of sterol regulatory element binding protein (SREBP1/2), peroxisome proliferator-activated receptor γ (PPARγ), and insulin-induced genes (Insig1/2) that proved to be closely associated with lipogenesis in HFD mice. In addition, the concentration of acetic acid was significantly increased in Col-treatment HFD mice. Further, we noted that Col contributed to the reprogramming of the intestinal microbiota. The relative abundances of Akkermansia, Dubosiella, and Alistipes were enhanced under treatment with Col in HFD mice. The results also imply that Col markedly elevated the phosphorylation level of the AMP-activated protein kinase (AMPK) pathway in HFD mice. Conclusions: the results of our study show that Col ameliorates obesity and suppresses lipogenesis in HFD mice. The underlying mechanisms are possibly associated with the reprogramming of the gut microbiota, in particular, the acetic acid-mediated increased expression of Alistipes via the AMPK signaling pathway. Full article
(This article belongs to the Section Phytochemicals and Human Health)
Show Figures

Figure 1

Figure 1
<p>Canola oil inhibits the body weight gain of obese mice. (<b>A</b>) Body weight change during the no-treatment experiment stage (high-fat diet-induced obesity stage); (<b>B</b>) body weight gain during the no-treatment experiment stage; (<b>C</b>) body weight change during the Col-treatment experiment stage; (<b>D</b>) body weight gain during the Col-treatment experiment stage. Data are expressed as the mean ± SEM (<span class="html-italic">n</span> = 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, and ns <span class="html-italic">p</span> &gt; 0.05.</p>
Full article ">Figure 2
<p>Canola oil inhibits the subcutaneous adipose tissue accumulation in obese mice. (<b>A</b>) Relative weight of liver to body weight; (<b>B</b>) relative weight of subcutaneous adipose tissue to body weight. Data are expressed as the mean ± SEM (<span class="html-italic">n</span> = 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, and ns <span class="html-italic">p</span> &gt; 0.05.</p>
Full article ">Figure 3
<p>Canola oil inhibits the biochemical indexes of lipids in obese mice. Serum HDL (<b>A</b>), LDL (<b>B</b>), TC (<b>C</b>), TG (<b>D</b>), and Glu (<b>E</b>) levels in serum; HDL (<b>F</b>), LDL (<b>G</b>), TC (<b>H</b>), TG (<b>I</b>), and Glu (<b>J</b>) in liver. Data are expressed as the mean ± SEM (<span class="html-italic">n</span> = 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, and ns <span class="html-italic">p</span> &gt; 0.05.</p>
Full article ">Figure 4
<p>Canola oil regulates the hepatic lipase bioactivity of obese mice. (<b>A</b>) Activity of acetyl-CoA carboxylase; (<b>B</b>) hepatic lipase; (<b>C</b>) lipase synthase; (<b>D</b>) lipoprotein lipase; (<b>E</b>) lipase. Data are expressed as the mean ± SEM (<span class="html-italic">n</span> = 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, and ns <span class="html-italic">p</span> &gt; 0.05.</p>
Full article ">Figure 5
<p>Canola oil inhibits the hepatic lipid anabolism of obese mice. (<b>A</b>,<b>B</b>,<b>H</b>,<b>I</b>) mRNA and protein expression of SREBP1/2; (<b>C</b>,<b>D</b>,<b>J</b>,<b>K</b>) mRNA and protein expression of Insig1/2; (<b>E</b>,<b>L</b>) mRNA and protein expression of ACC; (<b>F</b>,<b>G</b>,<b>M</b>,<b>N</b>) mRNA and protein expression of PPAR/PPARγ; (<b>O</b>–<b>R</b>) mRNA and protein expression of LXRα/LXRβ. Data are expressed as the mean ± SEM (<span class="html-italic">n</span> = 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, and ns <span class="html-italic">p</span> &gt; 0.05.</p>
Full article ">Figure 6
<p>Canola oil inhibits the AMPK pathway activity of obese mice. (<b>A</b>) mRNA expression of AMPK; (<b>B</b>) mRNA expression of pAMPK. (<b>C</b>) Protein expression of pAMPK; (<b>D</b>) Protein expression of pAMPK. Data are expressed as the mean ± SEM (<span class="html-italic">n</span> = 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, and ns <span class="html-italic">p</span> &gt; 0.05.</p>
Full article ">Figure 7
<p>Canola oil reverses dysbiosis with respect to the composition and relative abundance of the gut microbiota. (<b>A</b>) OTU annotation of fecal microbiota and the alpha-diversity of fecal microbiota; (<b>B</b>) composition of microbiota at the phylum level; (<b>C</b>) composition of microbiota at the order level; (<b>D</b>) composition of microbiota at the genus level. Data are expressed as the mean ± SEM (<span class="html-italic">n</span> = 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, and ns <span class="html-italic">p</span> &gt; 0.05.</p>
Full article ">Figure 8
<p>Canola oil influences the SCFAs concentration of obese mice. Data are expressed as the mean ± SEM (<span class="html-italic">n</span> = 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, and ns <span class="html-italic">p</span> &gt; 0.05.</p>
Full article ">Figure 9
<p>Col may exhibit the anti-obesity roles via microbiota-mediated acetic acid of obese mice (<b>A</b>) The correlation analyses between <span class="html-italic">Akkermansia</span> and acetic acid content; (<b>B</b>) The correlation analyses between <span class="html-italic">Dubosiella</span> and acetic acid content; (<b>C</b>) The correlation analyses between <span class="html-italic">Alistipes</span> and acetic acid content. Dot: samples in group; red dash line: the correlation analyses line. Data are expressed as the mean ± SEM (<span class="html-italic">n</span> = 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, and ns <span class="html-italic">p</span> &gt; 0.05.</p>
Full article ">Figure 10
<p>Col ameliorated obesity and suppressed lipogenesis in HFD mice. The underlying mechanisms are possibly associated with the reprogramming of the gut microbiota, in particular, the acetic acid-mediated increased expression of <span class="html-italic">Alistipes</span> via the AMPK signaling pathway. ↑—increased expression.</p>
Full article ">
17 pages, 3209 KiB  
Article
Effects of High-Linear-Energy-Transfer Heavy Ion Radiation on Intestinal Stem Cells: Implications for Gut Health and Tumorigenesis
by Santosh Kumar, Shubhankar Suman, Jerry Angdisen, Bo-Hyun Moon, Bhaskar V. S. Kallakury, Kamal Datta and Albert J. Fornace
Cancers 2024, 16(19), 3392; https://doi.org/10.3390/cancers16193392 - 4 Oct 2024
Abstract
Heavy ion radiation, prevalent in outer space and relevant for radiotherapy, is densely ionizing and poses a risk to intestinal stem cells (ISCs), which are vital for maintaining intestinal homeostasis. Earlier studies have shown that heavy-ion radiation can cause chronic oxidative stress, persistent [...] Read more.
Heavy ion radiation, prevalent in outer space and relevant for radiotherapy, is densely ionizing and poses a risk to intestinal stem cells (ISCs), which are vital for maintaining intestinal homeostasis. Earlier studies have shown that heavy-ion radiation can cause chronic oxidative stress, persistent DNA damage, cellular senescence, and the development of a senescence-associated secretory phenotype (SASP) in mouse intestinal mucosa. However, the specific impact on different cell types, particularly Lgr5+ intestinal stem cells (ISCs), which are crucial for maintaining cellular homeostasis, GI function, and tumor initiation under genomic stress, remains understudied. Using an ISCs-relevant mouse model (Lgr5+ mice) and its GI tumor surrogate (Lgr5+Apc1638N/+ mice), we investigated ISCs-specific molecular alterations after high-LET radiation exposure. Tissue sections were assessed for senescence and SASP signaling at 2, 5 and 12 months post-exposure. Lgr5+ cells exhibited significantly greater oxidative stress following 28Si irradiation compared to γ-ray or controls. Both Lgr5+ cells and Paneth cells showed signs of senescence and developed a senescence-associated secretory phenotype (SASP) after 28Si exposure. Moreover, gene expression of pro-inflammatory and pro-growth SASP factors remained persistently elevated for up to a year post-28Si irradiation. Additionally, p38 MAPK and NF-κB signaling pathways, which are critical for stress responses and inflammation, were also upregulated after 28Si radiation. Transcripts involved in nutrient absorption and barrier function were also altered following irradiation. In Lgr5+Apc1638N/+ mice, tumor incidence was significantly higher in those exposed to 28Si radiation compared to the spontaneous tumorigenesis observed in control mice. Our results indicate that high-LET 28Si exposure induces persistent DNA damage, oxidative stress, senescence, and SASP in Lgr5+ ISCs, potentially predisposing astronauts to altered nutrient absorption, barrier function, and GI carcinogenesis during and after a long-duration outer space mission. Full article
(This article belongs to the Special Issue Radiation Exposure, Inflammation and Cancers)
Show Figures

Figure 1

Figure 1
<p>Heavy ion <sup>28</sup>Si irradiation leads to persistent ROS and DNA damage in Lgr5<sup>+</sup> ISCs two months after exposure. (<b>A</b>) Representative flow cytometry histogram showing increased CellROX fluorescence intensity in ISCs after <sup>28</sup>Si radiation. (<b>B</b>) Quantifying fluorescence intensity data from five mice are presented as percent change in mean fluorescence in irradiated samples relative to controls. (<b>C</b>) Representative flow cytometry histogram showing increased MitoSOX fluorescence intensity in ISCs after <sup>28</sup>Si radiation. (<b>D</b>) Quantified fluorescence intensity data from five mice is presented as a percent change in mean fluorescence in irradiated samples relative to control. (<b>E</b>) Representative IF images of intestinal sections co-stained for γH2AX and Lgr5-GFP showing DNA damage as foci. Nuclei were counterstained with DAPI. Scale bar 10 μm. (<b>F</b>) Images were quantified from 10 FOVs and analyzed statistically. Bar graph representing increased average γH2AX foci/HPF in the irradiated group. (<b>G</b>) Graphical representation of Lgr5-GFP mean fluorescent intensity showing a modest decrease expression after radiation relative to control. *, significant relative to control; **, significant relative to γ-rays.</p>
Full article ">Figure 2
<p><sup>28</sup>Si ion exposure promotes crypt cell senescence 2 months after irradiation. Intestinal tissue sections were stained with phospho-Histone H3 (PHH3) to asses proliferation after irradiation. (<b>A</b>) Bar graph showing average PHH3 positive cells in each treatment group. (<b>B</b>) Representative images from each group showed increased SA-β-Gal positive cells (Blue color) at the crypt base after irradiation. Nuclei were counterstained with nuclear red stain. (Scale bar, 20 μm.) (<b>C</b>) SA-β-Gal positive cells were counted from at least 10 FOVs from each treatment group, analyzed statistically and represented in the form of a bar graph. (<b>D</b>) Senescent cells were detected using SA-β-Gal staining, and sections were co-stained with an anti-lysozyme (Paneth cell markers) antibody. Representative images showing Paneth cells (brown) senescence (blue) 2 months after irradiation. (Scale bar 20 μm.) (<b>E</b>) Average SA-β-Gal positive Paneth cells presented as a bar graph. (<b>F</b>) Representative images showed increased ISCs senescence after <sup>28</sup>Si exposure. (Scale bar, 10 μm.) (<b>G</b>) Average SA-β-Gal positive ISCs presented in the form of a bar graph. Average count of senescent cells determined from at least 10 different FOVs, analyzed statically represented in the form of a bar graph. Nuclei were counterstained with hematoxylin. *, significant relative to control; **, significant relative to γ-ray.</p>
Full article ">Figure 3
<p>Heavy ion <sup>28</sup>Si promotes accelerated SASP in mouse intestinal crypt 2 months post-exposure. (<b>A</b>) Fluorescent images showing IL8 (grey), p21 (red), Lgr5-GFP (green) in intestine sections. Nuclei were counterstained with DAPI. (<b>B</b>) Bar graph showing average pixel intensity/field of Lgr5-GFP, IL8 and p21. (<b>C</b>) Representative images show IL1β (brown) and senescent cells (blue) in intestinal tissue sections. (<b>D</b>) Graphical representation average DAB or β-Gal stain pixel intensity was quantified from at least ten different FOVs. (<b>E</b>) Representative images show senescent cells (blue) with IL1R (brown) expression in intestine sections. (<b>F</b>) Bar graph showing average pixel intensity of IL1β (left panel) or β-gal (right panel) in same sections. (<b>G</b>) Representative images show phospho-p38 positive (brown) cells and senescent (blue) cells. (<b>H</b>) Bar graph representing mean pixel intensity of p-p38 (left panel) and β-gal (right panel). (<b>I</b>) Representative images showing phospho-NFκB (brown) and senescent cells (blue) across the treatment group. (<b>J</b>) Bar graph showing average pixel intensity of p-NFκB (left panel) and SA-β-gal (right panel). Nuclei were counterstained with Hematoxylin. A FOV of at least 10 was used for statistical quantification and analysis. Scale bar, 10 μm. *, significant relative to control; **, significant relative to γ-ray.</p>
Full article ">Figure 4
<p><sup>28</sup>Si exposure leads to persistent increased DNA damage and senescence at 5 or 12 months after irradiation. (<b>A</b>) Fluorescent images of γH2AX, p16, or Lgr5-GFP showing increased DNA damage and senescence in ISCs at 5 months (left panel) and 12 months (right panel) after radiation exposure. Nuclei were counterstained with DAPI. At least ten different FOVs were captured, γH2AX foci were counted and Lgr5-GFP fluorescent intensity was quantified. Scale bar, 10 μm. (<b>B</b>) γH2AX foci were counted and represented graphically, showing time-dependent increased DNA damage response. Graphical representation of p16 or Lgr5-GFP fluorescent intensity at 5 months (<b>C</b>) and 12 months (<b>D</b>) post radiation exposure. (<b>E</b>) Graphical representation of average γH2AX foci per HPF at 2 m, 5 m and 12 m post-irradiation in intestinal sections. *, significant relative to control; **, significant relative to γ-ray.</p>
Full article ">Figure 5
<p>Space radiation exposure induces senescence and promotes SASP acquisition at the 5- and 12-month time points. Intestinal tissue sections were co-stained for IL8, p21, and Lgr5-GFP, and signals were detected under a fluorescent microscope. Representative fluorescent images of IL8, p21 and Lgr5-GFP staining showing senescence and SASP at 5 months (<b>A</b>) and 12 months (<b>B</b>) post-radiation. Nuclei were visualized using DAPI. Scale bar, 10 μm. At least ten different FOV images captured were fluorescent intensity quantified and represented graphically for 5 months (<b>C</b>) and 12 months (<b>D</b>) post-irradiation. *, significant relative to control; **, significant relative to γ-ray.</p>
Full article ">Figure 6
<p>Space radiation exposure perturbs barrier functions and genes regulating nutrient absorption up to 1 year after irradiation. Bar graph showing relative mRNA levels of nutrient transporters at 2 months (<b>A</b>–<b>C</b>) and 12 months (<b>D</b>–<b>F</b>) after irradiation showing altered expression of glucose transporter, gut hormones and Na/H exchanger, cholesterol and fatty acid transporters genes. Serum citrulline levels were detected using an ELISA assay kit. Bar graph showing serum citrulline level (μg/mL) at 2 months (<b>G</b>) and 12 months (<b>I</b>) after irradiation. Serum IFABP was detected using an I-FABP1 ELISA assay kit. Bar graph showing levels of I-FABP1 (ng/mL) in serum at 2 months (<b>H</b>) and 12 months (<b>J</b>) after irradiation. *, significant relative to control; **, significant relative to γ-ray.</p>
Full article ">Figure 7
<p><sup>28</sup>Si exposure leads to persistent increased DNA damage, senescence and SASP after 5 months of radiation in intestinal tumor from <span class="html-italic">Lgr5</span><sup>+</sup><span class="html-italic">Apc</span><sup>1638N/+</sup> mice. Intestinal tumor sections were co-stained with p16, γH2AX, and Lgr5-GFP to detect DNA damage response and senescence after 5 months of radiation exposure. Representative fluorescent images show increased p16 and γH2AX expression in tumor tissue sections. (<b>A</b>) Representative fluorescent images showing increased senescence (p16) and DNA damage response (γH2AX) expression in tumor intestine tissue sections after 5 months of radiation exposure. (<b>B</b>) γH2AX foci were counted and represented graphically, showing higher DNA damage response in tumors relative to unirradiated control. Graphical representation of Lgr5-GFP or p16 fluorescent intensity in the tumor samples. (<b>C</b>) Representative fluorescent images of IL8, p21, and Lgr5-GFP in the tumor intestine show increased senescence and SASP activity after radiation. (<b>D</b>). The fluorescent intensity of IL8, p21, and Lgr5-GFP was quantified and represented as a bar graph. Nuclei were visualized using DAPI. (<b>E</b>) Schematic representation and summary of observations after space radiation exposure in mouse ISCs. Red arrow indicates increased levels after irradiation. Scale bar, 10 μm. *, significant relative to control; **, significant relative to γ-ray. Statistical significance is set at <span class="html-italic">p</span> &lt; 0.05, and error bars represent mean ± SEM.</p>
Full article ">
24 pages, 3554 KiB  
Review
Potential Neuroprotective Effects of Alpinia officinarum Hance (Galangal): A Review
by Izzat Zulhilmi Abd Rahman, Siti Hajar Adam, Adila A. Hamid, Mohd Helmy Mokhtar, Ruslinda Mustafar, Mohd Izhar Ariff Mohd Kashim, Ami Febriza and Nur Izzati Mansor
Nutrients 2024, 16(19), 3378; https://doi.org/10.3390/nu16193378 - 4 Oct 2024
Abstract
Background/Objectives: This review aims to provide a detailed understanding of the current evidence on Alpinia officinarum Hance (A. officinarum) and its potential therapeutic role in central nervous system (CNS) disorders. CNS disorders encompass a wide range of disorders affecting the brain [...] Read more.
Background/Objectives: This review aims to provide a detailed understanding of the current evidence on Alpinia officinarum Hance (A. officinarum) and its potential therapeutic role in central nervous system (CNS) disorders. CNS disorders encompass a wide range of disorders affecting the brain and spinal cord, leading to various neurological, cognitive and psychiatric impairments. In recent years, natural products have emerged as potential neuroprotective agents for the treatment of CNS disorders due to their outstanding bioactivity and favourable safety profile. One such plant is A. officinarum, also known as lesser galangal, a perennial herb from the Zingiberaceae family. Its phytochemical compounds such as flavonoids and phenols have been documented to have a powerful antioxidants effect, capable of scavenging free radicals and preventing oxidative damage. Methods: In this review, we critically evaluate the in vitro and in vivo studies and examine the mechanisms by which A. officinarum exerts its neuroprotective effect. Results: Several studies have confirmed that A. officinarum exerts its neuroprotective effects by reducing oxidative stress and cell apoptosis, promoting neurite outgrowth, and modulating neurotransmitter levels and signalling pathways. Conclusions: Although previous studies have shown promising results in various models of neurological disorders, the underlying mechanisms of A. officinarum in Alzheimer’s (AD) and Parkinson’s disease (PD) are still poorly understood. Further studies on brain tissue and cognitive and motor functions in animal models of AD and PD are needed to validate the results observed in in vitro studies. In addition, further clinical studies are needed to confirm the safety and efficacy of A. officinarum in CNS disorders. Full article
(This article belongs to the Special Issue Nutritional Regulation of Plant Extracts on Human Health)
Show Figures

Figure 1

Figure 1
<p>The various parts of <span class="html-italic">A. officinarum</span>. (<b>A</b>) the whole plant; (<b>B</b>) roots and rhizome; (<b>C</b>) rhizome; (<b>D</b>) leaves; and (<b>E</b>) illustration of the flower of <span class="html-italic">A. officinarum</span>.</p>
Full article ">Figure 2
<p>Flowchart of the article selection process.</p>
Full article ">Figure 3
<p>Chemical structures of flavonoids and diarylheptanoids from <span class="html-italic">A. officinarum</span> which are commonly reported for their biological activities.</p>
Full article ">Figure 4
<p>Proposed therapeutic effects of <span class="html-italic">Alpinia officinarum</span> in central nervous system (CNS).</p>
Full article ">
15 pages, 2044 KiB  
Review
HPV Proteins as Therapeutic Targets for Phytopharmaceuticals Related to Redox State in HPV-Related Cancers
by Alfredo Cruz-Gregorio, Ana Karina Aranda-Rivera and José Pedraza-Chaverri
Future Pharmacol. 2024, 4(4), 716-730; https://doi.org/10.3390/futurepharmacol4040038 - 4 Oct 2024
Abstract
The high-risk Human Papillomavirus (HR-HPV) is the causal agent of different human cancers such as cervical, vulvar, and oropharynx cancer. This is because persistent HR-HPV infection alters several cellular processes involved in cell proliferation, apoptosis, immune evasion, genomic instability, and cellular transformation. The [...] Read more.
The high-risk Human Papillomavirus (HR-HPV) is the causal agent of different human cancers such as cervical, vulvar, and oropharynx cancer. This is because persistent HR-HPV infection alters several cellular processes involved in cell proliferation, apoptosis, immune evasion, genomic instability, and cellular transformation. The above is mainly due to the expression of early expression proteins of HR-HPV, which interact and alter these processes. HR-HPV proteins have even been shown to regulate redox state and mitochondrial metabolism, which has been suggested as a risk factor for cancer development. Redox state refers to a balance between reactive oxygen species (ROS) and antioxidants. Although ROS regulates cell signaling, high levels of ROS generate oxidative stress (OS). OS promotes damage to DNA, proteins, carbohydrates, and lipids, which causes mutation accumulation and genome instability associated with cancer development. Thus, OS has been associated with the establishment and development of different types of cancer and has recently been proposed as a cofactor in HR-HPV-associated cancers. However, OS also induces cell death, which can be used as a target for different molecules, such as phytochemicals. Furthermore, phytochemicals target HPV oncoproteins E6 and E7, causing their degradation. Because phytochemicals could induce OS and target HPV oncoproteins, we hypothesize that these compounds induce cell death in HPV-associated cancers. Since the redox state is crucial in developing, establishing, and clearing HR-HPV-associated cancer, this review focuses on evidence for using phytochemicals as therapeutic agents that target HPV proteins and the redox state to induce the elimination of HPV-related cancers. Full article
Show Figures

Figure 1

Figure 1
<p>Phytopharmaceutical molecules with anticancer effects, inducing cell death via apoptosis. Cyanidrical derivative of 11-keto-β-boswellic acid, known as 2-cyano-3,11-dioxide-1,12-dien-24-oate butyl (BCDD); human papillomavirus (HPV); p53-upregulated modulator of apoptosis (PUMA); nuclear factor kappa-light-chain-enhancer of activated B cells (NF-kB); kinase B (AKT); estrogen receptor (ER); Epigallocatechin-3-gallate (EGCG); lipid from Pinellia pedatisecta Schott (PE).</p>
Full article ">Figure 2
<p>Phytopharmaceutical molecules with anticancer effects, targeting HPV E6/E7 oncoproteins. Staurosporine (ST); mouse double minute 2 homolog (MDM2); nuclear factor kappa-light-chain-enhancer of activated B cells (NF-kB); activator protein 1 (AP-1); inducible nitric oxide synthase (iNOS); cyclooxygenase-2 (COX-2); interleukin (IL); peroxisome proliferator-activated receptor γ (PPARγ); B-cell lymphoma 2 (BCL-2).</p>
Full article ">Figure 3
<p>Quercetin and DHA decrease HPV E6/E7 oncoproteins by activating proteasome and p53. Staurosporine (ST); Docosahexaenoic acid (DHA); ubiquitination-proteasome system (UPS); reactive oxygen species (ROS); growth phase 2 (G2).</p>
Full article ">
19 pages, 21418 KiB  
Article
Genetic Transformation of Triticum dicoccum and Triticum aestivum with Genes of Jasmonate Biosynthesis Pathway Affects Growth and Productivity Characteristics
by Dmitry N. Miroshnichenko, Alexey V. Pigolev, Alexander S. Pushin, Valeria V. Alekseeva, Vlada I. Degtyaryova, Evgeny A. Degtyaryov, Irina V. Pronina, Andrej Frolov, Sergey V. Dolgov and Tatyana V. Savchenko
Plants 2024, 13(19), 2781; https://doi.org/10.3390/plants13192781 - 4 Oct 2024
Viewed by 72
Abstract
The transformation protocol based on the dual selection approach (fluorescent protein and herbicide resistance) has been applied here to produce transgenic plants of two cereal species, emmer wheat and bread wheat, with the goal of activating the synthesis of the stress hormone jasmonates [...] Read more.
The transformation protocol based on the dual selection approach (fluorescent protein and herbicide resistance) has been applied here to produce transgenic plants of two cereal species, emmer wheat and bread wheat, with the goal of activating the synthesis of the stress hormone jasmonates by overexpressing ALLENE OXIDE SYNTHASE from Arabidopsis thaliana (AtAOS) and bread wheat (TaAOS) and OXOPHYTODIENOATE REDUCTASE 3 from A. thaliana (AtOPR3) under the strong constitutive promoter (ZmUbi1), either individually or both genes simultaneously. The delivery of the expression cassette encoding AOS was found to affect morphogenesis in both wheat species negatively. The effect of transgene expression on the accumulation of individual jasmonates in hexaploid and tetraploid wheat was observed. Among the introduced genes, overexpression of TaAOS was the most successful in increasing stress-inducible phytohormone levels in transgenic plants, resulting in higher accumulations of JA and JA-Ile in emmer wheat and 12-OPDA in bread wheat. In general, overexpression of AOS, alone or together with AtOPR3, negatively affected leaf lamina length and grain numbers per spike in both wheat species. Double (AtAOS + AtOPR3) transgenic wheat plants were characterized by significantly reduced plant height and seed numbers, especially in emmer wheat, where several primary plants failed to produce seeds. Full article
Show Figures

Figure 1

Figure 1
<p>Emmer wheat plants transformed with <span class="html-italic">AtAOS</span> and <span class="html-italic">AtOPR3</span> genes around the flowering stage as grown in the greenhouse; note developmental differences between the primary T0 plants, RAB4, which is silenced for expression of introduced genes and the RAB2a, RAB5a, and RAB5b plants with a high level of constitutive expression of both the <span class="html-italic">AtAOS</span> and <span class="html-italic">AtOPR3</span> genes.</p>
Full article ">Figure 2
<p>Relative expression levels of <span class="html-italic">AtAOS</span> and <span class="html-italic">AtOPR3</span> in leaves of transgenic emmer (<b>a</b>) and bread (<b>b</b>) wheat lines; T4 homozygous plants, with the exceptions of RAB2 and RAR5, where leaf extracts of T0 plants are analyzed; data are means of at least three biological replicates ± SE; (<b>a</b>,<b>b</b>) expression levels in plants of ‘double’ transgenic lines carrying <span class="html-italic">AtAOS</span> and <span class="html-italic">AtOPR3</span> genes; (<b>c</b>) expression levels of <span class="html-italic">AtAOS</span> gene in transgenic lines of bread wheat Sar-60, for normalization, the relative expression level detected in SAB1 plants (panel (<b>b</b>)) is used.</p>
Full article ">Figure 3
<p>Production of transgenic wheat plants constitutively overexpressing <span class="html-italic">TaAOS</span> gene. (<b>a</b>) Transient <span class="html-italic">RFP</span> gene expression; morphogenic explant 24 h after the delivery of pANIC-<span class="html-italic">TaAOS</span> plasmid to Runo cells; (<b>b</b>) aging and necrosis of Runo wheat tissue with <span class="html-italic">RFP</span> expression; 45 days of in vitro culture; (<b>c</b>) early stage of transgenic somatic embryo formation of emmer wheat Runo, 60 days after bombardment with decreased concentration of herbicide; (<b>d</b>) formation of the RFP-positive single embryo-like structure of Sar-60 surrounded by leafy structures with RFP fluorescence on the medium with decreased herbicide concentration, 80 days after bombardment; (<b>e</b>) segregation of introduced expression cassette in T1 embryos germinated in vitro; 5 days of culture; transgenic line SD3 (<b>f</b>) RFP fluorescence in T2 kernels of homozygous sub-line RD1 in comparison with non-transgenic kernels of emmer wheat Runo. Bright field images are shown on the left side and fluorescent images are shown on the right side.</p>
Full article ">Figure 4
<p>Expression levels of <span class="html-italic">TaAOS</span> gene in leaves of transgenic wheat lines; (<b>a</b>) emmer wheat (cv. Runo) transgenic lines; (<b>b</b>) bread wheat (Sar-60) transgenic lines; data are means of at least five biological replicates ± SE; stars above the graphs indicate statistically significant differences with non-transgenic wheat (* <span class="html-italic">p</span> ≤ 0.05, ** <span class="html-italic">p</span> ≤ 0.01, **** <span class="html-italic">p</span> ≤ 0.001).</p>
Full article ">Figure 5
<p>Analysis of leaf length of non-transgenic emmer wheat (cv. Runo) and transgenic plants with overexpression of <span class="html-italic">AtAOS</span> and <span class="html-italic">AtOPR3</span> (RAB1) or <span class="html-italic">TaAOS</span> (RD1 and RD4). Values represent the lengths of 1st, 2nd, 3rd, and 4th leaves measured in 22–25 plants (transgenic lines) or 38 plants (non-transgenic (Runo)) (average ± sd). Stars indicate statistically significant differences calculated according Dunnett’s multiple comparison test: (“****”, <span class="html-italic">p</span> &lt; 0.001), (NS, non-significant).</p>
Full article ">Figure 6
<p>Analysis of leaf length of non-transgenic bread emmer wheat Sar-60 and transgenic lines with overexpression of <span class="html-italic">AtAOS</span> (SA7), <span class="html-italic">TaAOS</span> (SD2, SD3), or <span class="html-italic">AtAOS</span> and <span class="html-italic">AtOPR3</span> simultaneously (SAB1, SAB3). Values represent the lengths of 1st, 2nd, 3rd, and 4th leaves measured in 22–25 plants (average ± sd). Stars indicate statistically significant differences calculated according to Dunnett’s multiple comparisons 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; 0.005), (“****”, <span class="html-italic">p</span> &lt; 0.001), (NS, non-significant).</p>
Full article ">Figure 7
<p>The morphology of transgenic bread wheat lines of plants transformed with <span class="html-italic">AtAOS</span> (SA7), <span class="html-italic">TaAOS</span> (SD2, SD3) and with both <span class="html-italic">AtAOS</span> and <span class="html-italic">AtOPR3</span> (SAB1, SAB3) genes. (<b>a</b>,<b>b</b>) plants are in boot developmental stage; (<b>c</b>,<b>d</b>) plants are in early ripening developmental stage.</p>
Full article ">Figure 8
<p>Average plant height and the productivity of transgenic wheat plants transformed with <span class="html-italic">AtAOS</span> (SA7), <span class="html-italic">TaAOS</span> (RD1, RD4, SD2, SD3), and <span class="html-italic">AtAOS</span> and <span class="html-italic">AtOPR3</span> simultaneously (RAB1, SAB1, SAB3). (<b>a</b>,<b>b</b>), average plant height; (<b>c</b>,<b>d</b>), mean number of seeds per spike; stars indicate statistically significant differences with corresponding non-transgenic wheat cultivar calculated according to Dunnett’s multiple comparisons 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), (ns, not significant).</p>
Full article ">Figure 9
<p>Schematic representation of the pANIC-<span class="html-italic">TaAOS</span> expression cassette used for emmer wheat and bread wheat transformation. <span class="html-italic">OsAct1</span>, rice <span class="html-italic">Actin 1</span> promoter; <span class="html-italic">BAR</span>, BASTA resistance gene (phosphinothricin acetyl transferase); 35ST, CaMV 35S terminator; PvUbi1, <span class="html-italic">Ubiquitin 1</span> promoter from <span class="html-italic">Panicum virgatum</span>; pporRFP, Red Fluorescent Protein gene from <span class="html-italic">Porites porites</span>; <span class="html-italic">NosT</span>, <span class="html-italic">Nopaline Synthase</span> terminator; <span class="html-italic">ZmUbi1</span>, maize <span class="html-italic">Ubiquitin 1</span> promoter; OCS T, octopine synthase terminator sequence; attB1 and attB2—site-specific recombination sequences; <span class="html-italic">Amp<sup>R</sup></span>, ampicillin resistance gene; <span class="html-italic">Kan<sup>R</sup></span>, kanamycin resistance gene. Arrows indicate promoters; regions controlling the expression of <span class="html-italic">TaAOS</span> gene are highlighted in green color.</p>
Full article ">
9 pages, 2935 KiB  
Case Report
G6PD Potenza: A Novel Pathogenic Variant Broadening the Mutational Landscape in the Italian Population
by Claudio Ricciardi Tenore, Eugenia Tulli, Claudia Calò, Roberto Bertozzi, Jessica Evangelista, Giulia Maneri, Martina Rinelli, Francesca Brisighelli, Alessia Perrucci, Elisa De Paolis, Andrea Urbani, Maria De Bonis and Angelo Minucci
Genes 2024, 15(10), 1298; https://doi.org/10.3390/genes15101298 - 4 Oct 2024
Viewed by 134
Abstract
Background: Glucose 6 phosphate dehydrogenase (G6PD) is a rate-limiting enzyme of the pentose phosphate pathway. The loss of G6PD activity in red blood cells increases the risk of acute haemolytic anaemia under oxidative stress induced by infections, some medications, or fava beans. [...] Read more.
Background: Glucose 6 phosphate dehydrogenase (G6PD) is a rate-limiting enzyme of the pentose phosphate pathway. The loss of G6PD activity in red blood cells increases the risk of acute haemolytic anaemia under oxidative stress induced by infections, some medications, or fava beans. More than 200 single missense mutations are known in the G6PD gene. A 41-year-old woman with a family history of favism coming from the Basilicata region (Italy) was evaluated at our hospital for G6PD abnormalities. Methods: DNA was extracted from a peripheral blood sample and genotyped for the most common G6PD pathogenic variants (PVs). Positive results obtained by Restriction Fragment Length Polymorphism (RFLP), as per practice in our laboratory, were then reconfirmed in Sanger sequencing. Results: RFLP analysis highlighted a variant compatible with the G6PD Cassano variant. Confirmatory testing by Sanger unexpectedly identified a novel variant: c.1357G>A, p.(Val453Met) (NM_001360016.2); the same variant was found in the patient’s mother. In silico models predicted a deleterious effect of this variant at the protein level. The novel G6PD variant was named “G6PD Potenza” on the basis of the patient’s regional origin. Conclusions: This case describes a novel G6PD variant. It also highlights how the Sanger sequencing technique still represents an indispensable confirmatory standard method for variants that could be misinterpreted by only using a “first-level” approach, such as the RFLP. We stress that the evaluation of clinical manifestations in G6PD-deficient patients is of primary importance for the classification of each new G6PD mutation, in agreement with the new WHO guidelines. Full article
(This article belongs to the Section Molecular Genetics and Genomics)
Show Figures

Figure 1

Figure 1
<p>The sequencing results of the <span class="html-italic">G6PD</span> gene analysis. The blue arrows indicate the position of the <span class="html-italic">G6PD Cassano</span>, instead the red arrows indicate the position of the novel nucleotide change identified in this study. The patient resulted as a heterozygote for the c.1357 G&gt;A variant in the <span class="html-italic">G6PD</span> gene; her mother was a heterozygote for the same variant. The figure demonstrates the proximity of the new variant to <span class="html-italic">G6PD Cassano</span> in a heterozygous patient shown as an example.</p>
Full article ">Figure 2
<p>Capillary electrophoresis analysis of the enzymatic digestions of <span class="html-italic">Potenza</span> and <span class="html-italic">Cassano</span> PCR products. The figure shows the results obtained from the tape station capillary electrophoresis of the enzymatic digestions, performed with samples of the wild-type control (panel <b>A</b>), the patient (<b>B</b>), heterozygous <span class="html-italic">G6PD Cassano</span> (panel <b>C</b>), and the patient’s mother (panel <b>D</b>). Clear different profiles emerged from the wild-type and <span class="html-italic">Potenza</span> or <span class="html-italic">Cassano</span> analyses. However, the presence of two independent fragment patterns belonging to the presence of <span class="html-italic">Potenza</span> and <span class="html-italic">Cassano</span> variants is not evident. The altered pattern was also identified in the patient’s mother’s sample (panel <b>D</b>).</p>
Full article ">Figure 3
<p>Schematic representation of restriction sites of NlaIII enzyme, obtained with NEBcutter V2.0.</p>
Full article ">Figure 4
<p>A schematic representation of the alteration in the <span class="html-italic">G6PD</span> protein functional domains. The variants <span class="html-italic">Cassano</span> (c.1347G&gt;C p.Q449H) and <span class="html-italic">Potenza</span> (c.1357G&gt;A p.Val453Met) are highlighted in black in the protein domains. The crystallographic structure of the human <span class="html-italic">G6PD</span> enzyme, assembled from <a href="https://swissmodel.expasy.org" target="_blank">https://swissmodel.expasy.org</a>, (accessed on 10 August 2024).</p>
Full article ">
12 pages, 4531 KiB  
Article
Adverse Outcomes Following Exposure to Perfluorooctanesulfonamide (PFOSA) in Larval Zebrafish (Danio rerio): A Neurotoxic and Behavioral Perspective
by Nikita David, Emma Ivantsova, Isaac Konig, Cole D. English, Lev Avidan, Mark Kreychman, Mario L. Rivera, Camilo Escobar, Eliana Maira Agostini Valle, Amany Sultan and Christopher J. Martyniuk
Toxics 2024, 12(10), 723; https://doi.org/10.3390/toxics12100723 - 4 Oct 2024
Viewed by 115
Abstract
Toxicity mechanisms of per- and polyfluoroalkyl substances (PFASs), a chemical class present in diverse ecosystems, as well as many of their precursors, have been increasingly characterized in aquatic species. Perfluorooctanesulfonamide (PFOSA, C8H2F17NO2S) is a common [...] Read more.
Toxicity mechanisms of per- and polyfluoroalkyl substances (PFASs), a chemical class present in diverse ecosystems, as well as many of their precursors, have been increasingly characterized in aquatic species. Perfluorooctanesulfonamide (PFOSA, C8H2F17NO2S) is a common precursor of perfluorooctane sulfonic acid (PFOS), a long-chain PFAS. Here, we assessed sub-lethal endpoints related to development, oxidative stress, transcript levels, and distance moved in zebrafish embryos and larvae following continuous exposure to PFOSA beginning at 6 h post-fertilization (hpf). PFOSA decreased survival in fish treated with 1 µg/L PFOSA; however, the effect was modest relative to the controls (difference of 10%). Exposure up to 10 µg/L PFOSA did not affect hatch rate, nor did it induce ROS in 7-day-old larvae fish. The activity of larval fish treated with 100 µg/L PFOSA was reduced relative to the solvent control. Transcripts related to oxidative stress response and apoptosis were measured and BCL2-associated X, apoptosis regulator (bax), cytochrome c, somatic (cycs), catalase (cat), superoxide dismutase 2 (sod2) were induced with high concentrations of PFOSA. Genes related to neurotoxicity were also measured and transcript levels of acetylcholinesterase (ache), elav-like RNA binding protein 3 (elavl3), growth-associated protein 43 (gap43), synapsin II (syn2a), and tubulin 3 (tubb3) were all increased in larval fish with higher PFOSA exposure. These data improve our understanding of the potential sub-lethal toxicity of PFOSA in fish species. Full article
Show Figures

Figure 1

Figure 1
<p>Percent of surviving zebrafish following exposure to one of either ERM, 0.1% DMSO or 0.1, 1, 10, or 100 µg/L PFOSA over 7 days. Error bars (S.E.M.) are small and within symbols in some cases.</p>
Full article ">Figure 2
<p>Total percent of hatched zebrafish embryos/larvae following exposure to one of either ERM, 0.1% DMSO or 0.1, 1, 10, or 100 µg/L PFOSA over 7 days.</p>
Full article ">Figure 3
<p>Normalized reactive oxygen species (to µg/mL media/protein). Each circle represents a biological replicate; the mean (±S.D.) is represented by the horizontal line (One-Way ANOVA and Dunnett’s multiple comparisons test; n = 5/treatment). ns = not significant.</p>
Full article ">Figure 4
<p>The distance moved in each of the light and dark zones (10 min bins) of 7-day zebrafish larvae exposed to 0.1% DMSO, ERM or 0.1, 1, 10, or 100 µg/L PFOSA. Graphs are the combined output from three independent VMR runs. Columns depict mean (±S.D.) (Kruskal–Wallis test and Dunn’s multiple comparisons test; n = 8–12 fish/treatment/run). Asterisk indicates difference at ** <span class="html-italic">p</span> &lt; 0.01.</p>
Full article ">Figure 5
<p>Linear regression for relative expression of (<b>A</b>) <span class="html-italic">bax</span>, (<b>B</b>) <span class="html-italic">cat,</span> and (<b>C</b>) <span class="html-italic">sod2</span> in 7-day old zebrafish. The DMSO and ERM group were combined (“Cntl”). Each circle indicates a biological replicate or beaker of pooled fish (n = 4 to 6). The solid line indicates the relationship between expression and the region between the two outermost dotted lines is the 95% confidence interval of the X intercept.</p>
Full article ">Figure 6
<p>Relative expression of (<b>A</b>) <span class="html-italic">bax</span> and (<b>B</b>) <span class="html-italic">elavl3</span> in 7-day old larval zebrafish exposed to 0.1% DMSO, ERM or 0.1, 1, or 10 µg/L PFOSA. Each circle is a beaker of pooled fish (biological replicate), and the mean (±S.D.) is indicated by the horizontal line (One-Way ANOVA, Dunnett’s multiple comparisons test; n = 4 to 6). Asterisk indicates * <span class="html-italic">p</span> &lt; 0.05 or ** <span class="html-italic">p</span> &lt; 0.01.</p>
Full article ">Figure 7
<p>Linear regression for relative expression of (<b>A</b>) <span class="html-italic">ache</span>, (<b>B</b>) <span class="html-italic">elavl3</span>, (<b>C</b>) <span class="html-italic">gap43</span>, (<b>D</b>) <span class="html-italic">syn2a</span>, and (<b>E</b>) <span class="html-italic">tubulin (tubb3)</span> in 7-day old zebrafish. The DMSO and ERM group were combined (“Cntl”). Each circle indicates a biological replicate or beaker of pooled fish (n = 4 to 6). The solid line indicates the relationship between expression and the region between the two outermost dotted lines is the 95% confidence interval of the X intercept.</p>
Full article ">
27 pages, 1447 KiB  
Review
Bioactive Potential of Algae and Algae-Derived Compounds: Focus on Anti-Inflammatory, Antimicrobial, and Antioxidant Effects
by Maima Matin, Magdalena Koszarska, Atanas G. Atanasov, Karolina Król-Szmajda, Artur Jóźwik, Adrian Stelmasiak and Monika Hejna
Molecules 2024, 29(19), 4695; https://doi.org/10.3390/molecules29194695 - 3 Oct 2024
Viewed by 282
Abstract
Algae, both micro- and macroalgae, are recognized for their rich repository of bioactive compounds with potential therapeutic applications. These marine organisms produce a variety of secondary metabolites that exhibit significant anti-inflammatory, antioxidant, and antimicrobial properties, offering promising avenues for the development of new [...] Read more.
Algae, both micro- and macroalgae, are recognized for their rich repository of bioactive compounds with potential therapeutic applications. These marine organisms produce a variety of secondary metabolites that exhibit significant anti-inflammatory, antioxidant, and antimicrobial properties, offering promising avenues for the development of new drugs and nutraceuticals. Algae-derived compounds, including polyphenols, carotenoids, lipids, and polysaccharides, have demonstrated efficacy in modulating key inflammatory pathways, reducing oxidative stress, and inhibiting microbial growth. At the molecular level, these compounds influence macrophage activity, suppress the production of pro-inflammatory cytokines, and regulate apoptotic processes. Studies have shown that algae extracts can inhibit inflammatory signaling pathways such as NF-κB and MAPK, reduce oxidative damage by activating Nrf2, and offer an alternative to traditional antibiotics by combatting bacterial infections. Furthermore, algae’s therapeutic potential extends to addressing diseases such as cardiovascular disorders, neurodegenerative conditions, and cancer, with ongoing research exploring their efficacy in preclinical animal models. The pig model, due to its physiological similarities to humans, is highlighted as particularly suitable for validating the bioactivities of algal compounds in vivo. This review underscores the need for further investigation into the specific mechanisms of action and clinical applications of algae-derived biomolecules. Full article
(This article belongs to the Special Issue Marine Bioactives for Human Health)
23 pages, 5488 KiB  
Article
Alleviative Effect of Exogenous Application of Fulvic Acid on Nitrate Stress in Spinach (Spinacia oleracea L.)
by Kangning Han, Jing Zhang, Cheng Wang, Youlin Chang, Zeyu Zhang and Jianming Xie
Agronomy 2024, 14(10), 2280; https://doi.org/10.3390/agronomy14102280 - 3 Oct 2024
Viewed by 127
Abstract
Salt stress could be a significant factor limiting the growth and development of vegetables. In this study, Fulvic Acid (FA) (0.05%, 0.1%, 0.15%, 0.2%, and 0.25%) was applied under nitrate stress (150 mM), with normal Hoagland nutrient solution as a control to investigate [...] Read more.
Salt stress could be a significant factor limiting the growth and development of vegetables. In this study, Fulvic Acid (FA) (0.05%, 0.1%, 0.15%, 0.2%, and 0.25%) was applied under nitrate stress (150 mM), with normal Hoagland nutrient solution as a control to investigate the influence of foliar spray FA on spinach growth, photosynthesis, and oxidative stress under nitrate stress. The results showed that nitrate stress significantly inhibited spinach growth, while ROS (reactive oxygen species) accumulation caused photosystem damage, which reduced photosynthetic capacity. Different concentrations of FA alleviated the damage caused by nitrate stress in spinach to varying degrees in a concentration-dependent manner. The F3 treatment (0.15% FA + 150 mM NO3) exhibited the most significant mitigating effect. FA application promoted the accumulation of biomass in spinach under nitrate stress and increased chlorophyll content, the net photosynthetic rate, the maximum photochemical quantum yield of PSII (Photosystem II) (Fv/Fm), the quantum efficiency of PSII photochemistry [Y(II)], the electron transport rate, and the overall functional activity index of the electron transport chain between the PSII and PSI systems (PItotal); moreover, FA decreased PSII excitation pressure (1 − qP), quantum yields of regulated energy dissipation of PSII [Y(NPQ)], and the relative variable initial slope of fluorescence. FA application increased superoxide dismutase, peroxidase, and catalase activities and decreased malondialdehyde, H2O2, and O2 levels in spinach under nitrate stress. FA can enhance plant resistance to nitrate by accelerating the utilization of light energy in spinach to mitigate excess light energy and ROS-induced photosystem damage and increase photosynthetic efficiency. Full article
(This article belongs to the Special Issue Crop and Vegetable Physiology under Environmental Stresses)
18 pages, 1371 KiB  
Review
Therapeutic Potential of Pomegranate Extract for Women’s Reproductive Health and Breast Cancer
by Jung Yoon Jang, Donghwan Kim, Eunok Im and Nam Deuk Kim
Life 2024, 14(10), 1264; https://doi.org/10.3390/life14101264 - 3 Oct 2024
Viewed by 179
Abstract
Pomegranate extract has potential benefits for women’s reproductive health, including fertility enhancement, menstrual cycle regulation, pregnancy support, and polycystic ovary syndrome (PCOS) treatment. It possesses antioxidant properties, reducing oxidative stress and improving fertility. Pomegranate extract may help regulate hormonal imbalances and promote regular [...] Read more.
Pomegranate extract has potential benefits for women’s reproductive health, including fertility enhancement, menstrual cycle regulation, pregnancy support, and polycystic ovary syndrome (PCOS) treatment. It possesses antioxidant properties, reducing oxidative stress and improving fertility. Pomegranate extract may help regulate hormonal imbalances and promote regular menstrual cycles. The extract’s rich nutrient profile supports placental development and fetal growth and may reduce the risk of preterm birth. Additionally, pomegranate extract shows promise in improving insulin sensitivity and reducing inflammation and oxidative damage in PCOS. Some studies suggest its potential anticancer properties, particularly against breast cancer. However, further research, including human clinical trials, is necessary to establish its effectiveness and safety. The current evidence is limited and primarily based on in vitro studies, animal studies, and clinical trials. This review provides a comprehensive summary of the benefits of pomegranate extract for women’s reproductive health and breast cancer, serving as a reference for future research. Full article
(This article belongs to the Special Issue Advances in the Biomedical Applications of Plants and Plant Extracts)
Show Figures

Figure 1

Figure 1
<p>Physiological benefits of pomegranate extract.</p>
Full article ">Figure 2
<p>Major nutraceuticals found in pomegranate extract. (<b>A</b>–<b>C</b>) <b>Flavonoids:</b> (<b>A</b>) Catechin; (<b>B</b>) Kaempferol; (<b>C</b>) Quercetin; (<b>D</b>–<b>F</b>) <b>Anthocyanins:</b> (<b>D</b>) Cyanidin-3-glucoside; (<b>E</b>) Delfinidin-3-glucoside; (<b>F</b>) Pelargonidin-3-glucoside; (<b>G</b>–<b>I</b>) <b>Tannins:</b> (<b>G</b>) Punicalagin; (<b>H</b>) Punicalin; (<b>I</b>) Ellagic acid; (<b>J</b>–<b>L</b>) <b>Fatty acids and organic acids:</b> (<b>J</b>) Oleic acid; (<b>K</b>) Linoleic acid; (<b>L</b>) Citric acid; (<b>M</b>,<b>N</b>) <b>Sterols:</b> (<b>M</b>) β-Sitosterol; (<b>N</b>) Stigmasterol.</p>
Full article ">Figure 2 Cont.
<p>Major nutraceuticals found in pomegranate extract. (<b>A</b>–<b>C</b>) <b>Flavonoids:</b> (<b>A</b>) Catechin; (<b>B</b>) Kaempferol; (<b>C</b>) Quercetin; (<b>D</b>–<b>F</b>) <b>Anthocyanins:</b> (<b>D</b>) Cyanidin-3-glucoside; (<b>E</b>) Delfinidin-3-glucoside; (<b>F</b>) Pelargonidin-3-glucoside; (<b>G</b>–<b>I</b>) <b>Tannins:</b> (<b>G</b>) Punicalagin; (<b>H</b>) Punicalin; (<b>I</b>) Ellagic acid; (<b>J</b>–<b>L</b>) <b>Fatty acids and organic acids:</b> (<b>J</b>) Oleic acid; (<b>K</b>) Linoleic acid; (<b>L</b>) Citric acid; (<b>M</b>,<b>N</b>) <b>Sterols:</b> (<b>M</b>) β-Sitosterol; (<b>N</b>) Stigmasterol.</p>
Full article ">
18 pages, 2442 KiB  
Article
Cytotoxic Potencies of Zinc Oxide Nanoforms in A549 and J774 Cells
by Nazila Nazemof, Dalibor Breznan, Yasmine Dirieh, Erica Blais, Linda J. Johnston, Azam F. Tayabali, James Gomes and Premkumari Kumarathasan
Nanomaterials 2024, 14(19), 1601; https://doi.org/10.3390/nano14191601 - 3 Oct 2024
Viewed by 361
Abstract
Zinc oxide nanoparticles (NPs) are used in a wide range of consumer products and in biomedical applications, resulting in an increased production of these materials with potential for exposure, thus causing human health concerns. Although there are many reports on the size-related toxicity [...] Read more.
Zinc oxide nanoparticles (NPs) are used in a wide range of consumer products and in biomedical applications, resulting in an increased production of these materials with potential for exposure, thus causing human health concerns. Although there are many reports on the size-related toxicity of ZnO NPs, the toxicity of different nanoforms of this chemical, toxicity mechanisms, and potency determinants need clarification to support health risk characterization. A set of well-characterized ZnO nanoforms (e.g., uncoated ca. 30, 45, and 53 nm; coated with silicon oil, stearic acid, and (3-aminopropyl) triethoxysilane) were screened for in vitro cytotoxicity in two cell types, human lung epithelial cells (A549), and mouse monocyte/macrophage (J774) cells. ZnO (bulk) and ZnCl2 served as reference particles. Cytotoxicity was examined 24 h post-exposure by measuring CTB (viability), ATP (energy metabolism), and %LDH released (membrane integrity). Cellular oxidative stress (GSH-GSSG) and secreted proteins (targeted multiplex assay) were analyzed. Zinc oxide nanoform type-, dose-, and cell type-specific cytotoxic responses were seen, along with cellular oxidative stress. Cell-secreted protein profiles suggested ZnO NP exposure-related perturbations in signaling pathways relevant to inflammation/cell injury and corresponding biological processes, namely reactive oxygen species generation and apoptosis/necrosis, for some nanoforms, consistent with cellular oxidative stress and ATP status. The size, surface area, agglomeration state and metal contents of these ZnO nanoforms appeared to be physicochemical determinants of particle potencies. These findings warrant further research on high-content “OMICs” to validate and resolve toxicity pathways related to exposure to nanoforms to advance health risk-assessment efforts and to inform on safer materials. Full article
Show Figures

Figure 1

Figure 1
<p>Cell morphology observed after exposure of A549 and J774 to the different doses of ZnO nanoparticles (e.g., (<b>A</b>) UC-2 and (<b>B</b>) AM): light microscopy images (40× magnification).</p>
Full article ">Figure 2
<p>Cytotoxicity in A549 cells (mean ± SEM) after exposure (24 h) to ZnO nanoforms and the reference particles. Exposure experiments were conducted three times (n = 3), with duplicate samples per treatment group in each exposure experiment. (<b>A</b>) LDH Release, (<b>B</b>) CTB (Resazurin) Reduction, (<b>C</b>) Cellular ATP Levels.</p>
Full article ">Figure 3
<p>Cytotoxicity in J774 cells (mean ± SEM) after exposure (24 h) to ZnO nanoforms and reference particles. Exposure experiments were conducted three times (n = 3), with duplicate samples per treatment group in each exposure experiment. (<b>A</b>) LDH Release, (<b>B</b>) CTB (Resazurin) Reduction, (<b>C</b>) Cellular ATP Levels.</p>
Full article ">Figure 4
<p>Cellular oxidative stress status in (<b>A</b>) A549 and (<b>B</b>) J774 cells after exposure to ZnO NPs, as well as to the reference particles (30 µg/cm<sup>2</sup>).</p>
Full article ">Figure 5
<p>Heatmap and hierarchical clustering of secreted protein fold changes normalized to control (24 h post exposure of cells to ZnO nanoforms and reference particles: (<b>A</b>) A549 and (<b>B</b>) J774). Red—increased; green—decreased.</p>
Full article ">Figure 6
<p>Pathway analysis results for in vitro cellular exposure (24 h) to ZnO nanoforms and the reference particles ((<b>A</b>) A549 and (<b>B</b>) J774). Orange—increased; blue—decreased.</p>
Full article ">
Back to TopTop