[go: up one dir, main page]

 
 
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

Article Types

Countries / Regions

remove_circle_outline
remove_circle_outline
remove_circle_outline
remove_circle_outline
remove_circle_outline
remove_circle_outline
remove_circle_outline
remove_circle_outline

Search Results (3,055)

Search Parameters:
Keywords = HepG2 cells

Order results
Result details
Results per page
Select all
Export citation of selected articles as:
21 pages, 18833 KiB  
Article
Mitigating Hyperglycaemic Oxidative Stress in HepG2 Cells: The Role of Carica papaya Leaf and Root Extracts in Promoting Glucose Uptake and Antioxidant Defence
by Mthokozisi Bongani Nxumalo, Nosipho Ntanzi, Hezekiel Mathambo Kumalo and Rene Bernadette Khan
Nutrients 2024, 16(20), 3496; https://doi.org/10.3390/nu16203496 (registering DOI) - 15 Oct 2024
Viewed by 334
Abstract
Background/Objectives: Diabetes often goes undiagnosed, with 60% of people in Africa unaware of their condition. Type 2 diabetes mellitus (T2DM) is associated with insulin resistance and is treated with metformin, despite the undesirable side effects. Medicinal plants with therapeutic potential, such as Carica [...] Read more.
Background/Objectives: Diabetes often goes undiagnosed, with 60% of people in Africa unaware of their condition. Type 2 diabetes mellitus (T2DM) is associated with insulin resistance and is treated with metformin, despite the undesirable side effects. Medicinal plants with therapeutic potential, such as Carica papaya, have shown promising anti-diabetic properties. This study explored the role of C. papaya leaf and root extracts compared to metformin in reducing hyperglycaemia-induced oxidative stress and their impact on liver function using HepG2 as a reference. Methods: The cytotoxicity was assessed through the MTT assay. At the same time, glucose uptake and metabolism (ATP and ∆Ψm) in HepG2 cells treated with C. papaya aqueous leaf and root extract were evaluated using a luminometry assay. Additionally, antioxidant properties (SOD2, GPx1, GSH, and Nrf2) were measured using qPCR and Western blot following the detection of MDA, NO, and iNOS, indicators of free radicals. Results: The MTT assay showed that C. papaya extracts did not exhibit toxicity in HepG2 cells and enhanced glucose uptake compared to the hyperglycaemic control (HGC) and metformin. The glucose levels in C. papaya-treated cells increased ATP production (p < 0.05), while the ∆Ψm was significantly increased in HGR1000-treated cells (p < 0.05). Furthermore, C. papaya leaf extract upregulated GPx1 (p < 0.05), GSH, and Nrf2 gene (p < 0.05), while SOD2 and Nrf2 proteins were reduced (p > 0.05), ultimately lowering ROS (p > 0.05). Contrarily, the root extract stimulated SOD2 (p > 0.05), GPx1 (p < 0.05), and GSH levels (p < 0.05), reducing Nrf2 gene and protein expression (p < 0.05) and resulting in high MDA levels (p < 0.05). Additionally, the extracts elevated NO levels and iNOS expression (p < 0.05), suggesting potential RNS activation. Conclusion: Taken together, the leaf extract stimulated glucose metabolism and triggered ROS production, producing a strong antioxidant response that was more effective than the root extract and metformin. However, the root extract, particularly at high concentrations, was less effective at neutralising free radicals as it did not stimulate Nrf2 production, but it did maintain elevated levels of SOD2, GSH, and GPx1 antioxidants. Full article
(This article belongs to the Section Phytochemicals and Human Health)
Show Figures

Figure 1

Figure 1
<p>The effects of <span class="html-italic">C. papaya</span> leaf and root extracts on cell viability and cytotoxicity in HepG2 cells induced with NG and hyperglycaemia. <span class="html-italic">C. papaya</span> leaf extract had fluctuating effects on cell viability, ranging from 85% to 103%, in both NG and hyperglycaemia (HG) treated cells (<b>A</b>,<b>B</b>), respectively. However, cell viability remained similar to the control (100%) in NG cells treated with <span class="html-italic">C. papaya</span> root (<b>C</b>). In contrast, the viability was increased to 110% in HG-treated cells (<b>D</b>). All experiments were conducted in triplicates.</p>
Full article ">Figure 2
<p>In hyperglycemic conditions, the concentration of <span class="html-italic">C. papaya</span> leaf and root extracts at 500 and 1000 µg/mL maintained cell viability similar to normal and HG controls. Metformin slightly reduced it to 81%.</p>
Full article ">Figure 3
<p>The levels of glucose in HGMet- and HGL500-treated cells were non-significantly decreased in relation to the HGC-treated cells, while were increased significantly in the HGL1000-, HGR500-, and HGR1000-treated cells (##, *** <span class="html-italic">p</span> &lt; 0.05, unpaired student <span class="html-italic">t</span>-test with Welch’s correction).</p>
Full article ">Figure 4
<p>The effect of <span class="html-italic">C. papaya</span> on hyperglycaemia-induced HepG2 was shown through two measures: mitochondrial membrane potential and ATP activity levels. <span class="html-italic">C. papaya</span> treatment showed a slight increase in ∆Ψm except for HGR1000, where a significant increase was observed (<b>A</b>). Also, ATP levels in HG-treated HepG2 cells were significantly higher compared to HG-control (<b>B</b>) (###, *, **, ***, <span class="html-italic">p</span> &lt; 0.05, unpaired student <span class="html-italic">t</span>-test with Welch’s correction).</p>
Full article ">Figure 5
<p>The effect of <span class="html-italic">C. papaya</span> leaf and root extracts on lipid peroxidation induced by HG (<b>A</b>) and nitrate/nitrite concentrations (<b>B</b>). Results showed that HGMet treatment led to a decrease in MDA levels, while <span class="html-italic">C. papaya</span> treatment resulted in an increased MDA concentration (<b>A</b>). Moreover, HGMet treatment significantly lowered nitrate/nitrile levels in cells, whereas <span class="html-italic">C. papaya</span> treatment led to a significant increase in nitrate/nitrile levels in HepG2 cells (<b>B</b>) (*, **, *** <span class="html-italic">p</span> &lt; 0.05, unpaired student <span class="html-italic">t</span>-test with Welch’s correction).</p>
Full article ">Figure 6
<p>The presence of hyperglycaemia significantly reduced GSH levels from normal glucose to HGC. However, treatments effectively restored intracellular GSH levels above the HGC (<b>A</b>). The GSH/GSSG ratio decreased in HGMet-treated cells. At the same time, it increased in HG-induced cells treated with <span class="html-italic">C. papaya</span> leaf and root extracts, indicating the antioxidant effect of <span class="html-italic">C. papaya</span> extracts on GSH/GSSG in HG HepG2 cells (<b>B</b>) (##, ** <span class="html-italic">p</span> &lt; 0.05, unpaired student <span class="html-italic">t</span>-test with Welch’s correction).</p>
Full article ">Figure 7
<p>The effects of <span class="html-italic">C. papaya</span> on the iNOS, SOD2, and GPx1 protein expression of hyperglycaemia-induced HepG2 cells (<b>A</b>) shows a non-significant increase in iNOS protein expression from the NG-control to the HG-control; however, a significant increase in the protein expression was seen in HGMet and <span class="html-italic">C. papaya</span>-treated cells compared to the HG-control. (<b>B</b>) SOD2 expression was upregulated in HG cells treated with metformin but downregulated in <span class="html-italic">C. papaya</span> HGL500 and HGL1000. SOD2 expression remained slightly higher in HGR-treated cells compared to HGC. (<b>C</b>) GPx1 protein expression decreased in HG-control when compared to the NG-control, while it was increased in HGMet and <span class="html-italic">C. papaya</span>-treated HepG2 (##, *, **, *** <span class="html-italic">p</span> &lt; 0.05, unpaired student <span class="html-italic">t</span>-test with Welch’s correction).</p>
Full article ">Figure 8
<p>The gene expression of <span class="html-italic">Nrf2</span> was found to be lower in HGC compared to NGC, while its protein expression was significantly higher in HGC. However, when metformin, HGL500, and HGL1000 were administered, there was an increase in the gene expression of <span class="html-italic">Nrf2</span>, while the protein expression decreased as a result of the same treatments (<b>A</b>). On the other hand, treating hyperglycemic HepG2 cells with <span class="html-italic">C. papaya</span> root extract led to a downregulation of the Nrf2 gene expression, while HGR500 increased the protein expression. However, the protein expression was significantly reduced in cells treated with HGR1000 (<b>B</b>) (##, *, **, *** <span class="html-italic">p</span> &lt; 0.05, unpaired student <span class="html-italic">t</span>-test with Welch’s correction).</p>
Full article ">
17 pages, 3905 KiB  
Article
Oilseed Cakes: A Promising Source of Antioxidant, and Anti-Inflammatory Agents—Insights from Lactuca sativa
by Mayye Majed, Amal A. Galala, Mohamed M. Amer, Dirk Selmar and Sara Abouzeid
Int. J. Mol. Sci. 2024, 25(20), 11077; https://doi.org/10.3390/ijms252011077 (registering DOI) - 15 Oct 2024
Viewed by 259
Abstract
This study evaluated the antioxidant and antibacterial properties of methanolic extracts derived from oilseed cakes of Lactuca sativa (lettuce), Nigella sativa (black seed), Eruca sativa (rocket), and Linum usitatissimum (linseed). Lettuce methanolic extract showed the highest potential, so it was selected for further [...] Read more.
This study evaluated the antioxidant and antibacterial properties of methanolic extracts derived from oilseed cakes of Lactuca sativa (lettuce), Nigella sativa (black seed), Eruca sativa (rocket), and Linum usitatissimum (linseed). Lettuce methanolic extract showed the highest potential, so it was selected for further investigation. High-performance liquid chromatography (HPLC-DAD) analysis and bioassay-guided fractionation of lettuce seed cake extract led to the isolation of five compounds: 1,3-propanediol-2-amino-1-(3′,4′-methylenedioxyphenyl) (1), luteolin (2), luteolin-7-O-β-D-glucoside (3), apigenin-7-O-β-D-glucoside (4), and β-sitosterol 3-O-β-D-glucoside (5). Compound (1) was identified from Lactuca species for the first time, with high yield. The cytotoxic effects of the isolated compounds were tested on liver (HepG2) and breast (MCF-7) cancer cell lines, compared to normal cells (WI-38). Compounds (2), (3), and (4) exhibited strong activity in all assays, while compound (1) showed weak antioxidant, antimicrobial, and cytotoxic effects. The anti-inflammatory activity of lettuce seed cake extract and compound (1) was evaluated in vivo using a carrageenan-induced paw oedema model. Compound (1) and its combination with ibuprofen significantly reduced paw oedema, lowered inflammatory mediators (IL-1β, TNF-α, PGE2), and restored antioxidant enzyme activity. Additionally, compound (1) showed promising COX-1 and COX-2 inhibition in an in vitro enzymatic anti-inflammatory assay, with IC50 values of 17.31 ± 0.65 and 4.814 ± 0.24, respectively. Molecular docking revealed unique interactions of compound (1) with COX-1 and COX-2, suggesting the potential for targeted inhibition. These findings underscore the value of oilseed cakes as a source of bioactive compounds that merit further investigation. Full article
Show Figures

Figure 1

Figure 1
<p>IC<sub>50</sub> of different fractions of lettuce seedcake extract in ABTS assay using ascorbic acid as standard.</p>
Full article ">Figure 2
<p>HPLC of the major compounds determined in lettuce seed cake extract. Compounds were monitored using a photo diode array (PDA) detector at 280, 350 nm.</p>
Full article ">Figure 3
<p>Structures of the isolated compounds from <span class="html-italic">Lactuca sativa</span> seedcake active fractions: 1,3-propanediol-2-amino-1-(3′,4′-methylenedioxyphenyl) (<b>1</b>), luteolin (<b>2</b>), luteolin 7-O-<span class="html-italic">β</span>-D glucoside (<b>3</b>), apigenin 7-O-<span class="html-italic">β</span>-D glucoside (<b>4</b>), and <span class="html-italic">β</span>-sitosterol 3-O-<span class="html-italic">β</span>-D-glucoside (<b>5</b>).</p>
Full article ">Figure 4
<p>IC<sub>50</sub> of isolated compounds from lettuce seedcake in ABTS assay using ascorbic acid as standard.</p>
Full article ">Figure 5
<p>Microscopic pictures of H&amp;E-stained skin sections. Magnifications X: 40 bar 200 are represented in the first column, thin black arrows refer to dermal inflammation and X: 400 bar 50 are represented in the second column where thin black arrows indicate leukocytic cells infiltration in dermis.</p>
Full article ">Figure 6
<p>Effect of treatments on TNF-α (<b>A</b>), IL-1β (<b>B</b>), and PGE2 (<b>C</b>) in carrageenan-induced oedema in rat hind paws. Values represent mean ± SD for each group (<span class="html-italic">n</span> = 6). * <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.</p>
Full article ">Figure 7
<p>IC<sub>50</sub> of compound <b>1</b> and Ibuprofen standard for COX-1 and COX-2 inhibition assay.</p>
Full article ">Figure 8
<p>The 3D and 2D representations of compound <b>1</b> and ibuprofen interacting with the COX-1 and COX-2 enzymes are shown. The 2D diagrams illustrate color-coded interactions between the ligand and specific amino acid residues, while the 3D models display the ligand within each enzyme’s binding pocket, highlighting key structural features. In 3D model of protein (red: alpha helices; cyan: beta sheets; green: loops) complexed with drug (stick model).</p>
Full article ">
12 pages, 2769 KiB  
Article
(E)-5-hydroxy-7-methoxy-3-(2-hydroxybenzyl)-4-chromanone, a Major Homoisoflavonoid, Attenuates Free Fatty Acid-Induced Hepatic Steatosis by Activating AMPK and PPARα Pathways in HepG2 Cells
by Jae-Eun Park and Ji-Sook Han
Nutrients 2024, 16(20), 3475; https://doi.org/10.3390/nu16203475 - 14 Oct 2024
Viewed by 309
Abstract
Background: (E)-5-hydroxy-7-methoxy-3-(2-hydroxybenzyl)-4-chromanone (HMC), a homoisoflavonoid isolated from Portulaca oleracea, has significant anti-adipogenesis potential; it regulates adipogenic transcription factors. However, whether HMC improves hepatic steatosis in hepatocytes remains vague. This study investigated whether HMC ameliorates hepatic steatosis in free fatty acid-treated [...] Read more.
Background: (E)-5-hydroxy-7-methoxy-3-(2-hydroxybenzyl)-4-chromanone (HMC), a homoisoflavonoid isolated from Portulaca oleracea, has significant anti-adipogenesis potential; it regulates adipogenic transcription factors. However, whether HMC improves hepatic steatosis in hepatocytes remains vague. This study investigated whether HMC ameliorates hepatic steatosis in free fatty acid-treated human hepatocellular carcinoma (HepG2) cells, and if so, its mechanism of action was analyzed. Methods: Hepatic steatosis was induced by a free fatty acid mixture in HepG2 cells. Thereafter, different HMC concentrations (10, 30, and 50 µM) or fenofibrate (10 µM, a PPARα agonist, positive control) was treated in HepG2 cells.Results: HMC markedly decreased lipid accumulation and triglyceride content in free fatty acid-treated HepG2 cell; it (10 and 50 μM) markedly upregulated protein expressions of pAMP-activated protein kinase (AMPK) and acetyl-CoA carboxylase. HMC (10 and 50 μM) markedly inhibited the expression of sterol regulatory element-binding protein-1c, fatty acid synthase, and stearoyl-coA desaturase 1, which are the enzymes involved in lipid synthesis. Furthermore, HMC (10 and 50 μM) markedly upregulated the protein expression of peroxisome proliferator-activated receptor alpha (PPARα) and enhanced the protein expressions of carnitine palmitoyl transferase 1 and acyl-CoA oxidase 1. Conclusion: HMC inhibits lipid accumulation and promotes fatty acid oxidation by AMPK and PPARα pathways in free fatty acid-treated HepG2 cells, thereby attenuating hepatic steatosis. Full article
(This article belongs to the Special Issue Effects of Phytochemicals on Metabolic Disorders and Human Health)
Show Figures

Figure 1

Figure 1
<p>Chemical structure of HMC isolated from <span class="html-italic">P. oleracea</span>.</p>
Full article ">Figure 2
<p>Cell Viability with HMC. Cell viability was stimulatied with 1.0 mM FFA stock solution (oleate/palmitate = 2:1) and then treatment with HMC (10, 30, and 50 μM) for 24 h. Each value is expressed as the mean ± standard deviation (<span class="html-italic">n</span> = 3), and values with different superscript letters were markedly different with <span class="html-italic">p</span> &lt; 0.05 as analyzed using ANOVA, followed by Tukey’s HSD test. *: Statistically significant differences.</p>
Full article ">Figure 3
<p>HMC inhibits lipid accumulation in FFA-induced hepatic steatosis. (<b>A</b>) Cells were fixed and stained with Oil Red O staining to visualize the lipid droplets by microscopy (magnification = 100X). Scale bar = 100 μM. (<b>B</b>) The effects of HMC on the inhibition of lipid accumulation. Each value is expressed as the mean ± standard deviation (<span class="html-italic">n</span> = 3), and values with different letters were markedly different with <span class="html-italic">p</span> &lt; 0.05 as analyzed using ANOVA, followed by Tukey’s HSD test. *: Statistically significant differences.</p>
Full article ">Figure 4
<p>HMC inhibits the phosphorylation of SREBP-1c, SCD-1, and FAS in FFA-induced hepatic steatosis. (<b>A</b>) The protein expression of SREBP-1c. (<b>B</b>) The protein expression of SCD-1 and FAS. Each value is expressed as the mean ± standard deviation (<span class="html-italic">n</span> = 3), and values with different superscript letters were markedly different with <span class="html-italic">p</span> &lt; 0.05 as analyzed using ANOVA, followed by Tukey’s HSD test. *: Statistically significant differences.</p>
Full article ">Figure 5
<p>HMC promotes AMPK and ACC in FFA-induced hepatic steatosis. (<b>A</b>) The protein expression of AMPK. (<b>B</b>) The protein expression of ACC. Each value is expressed as the mean ± standard deviation (<span class="html-italic">n</span> = 3), and values with different superscript letters were markedly different with <span class="html-italic">p</span> &lt; 0.05 as analyzed using one-way analysis of variance (ANOVA), followed by Tukey’s HSD test. *: Statistically significant differences.</p>
Full article ">Figure 6
<p>HMC inhibits the phosphorylation of SREBP-1c, SCD-1, and FAS via activating AMPK in FFA-induced hepatic steatosis. (<b>A</b>) The protein expression of SREBP-1c. (<b>B</b>) The protein expression of SCD-1 and FAS. Each value is expressed as the mean ± standard deviation (<span class="html-italic">n</span> = 3), and values with different superscript letters were markedly different with <span class="html-italic">p</span> &lt; 0.05 as analyzed using one-way analysis of variance (ANOVA), followed by Tukey’s HSD test. *: Statistically significant differences.</p>
Full article ">Figure 7
<p>HMC increases the protein expression of PPARα, CPT1, and ACOX1. (<b>A</b>) The protein expression of PPARα, CPT1, and ACOX1. (<b>B</b>) Expression of PPARα, CPT1, and ACOX1. Each value is expressed as the mean ± standard deviation (<span class="html-italic">n</span> = 3), and values with different superscript letters were markedly different with <span class="html-italic">p</span> &lt; 0.05 as analyzed using one-way analysis of variance (ANOVA), followed by Tukey’s HSD test. *: Statistically significant differences.</p>
Full article ">Figure 8
<p>Proposed mechanism of attenuation of hepatic steatosis.</p>
Full article ">
19 pages, 2556 KiB  
Article
Anti-Hyperglycemic Effects of Thai Herbal Medicines
by Athit Bunyakitcharoen, Weerakit Taychaworaditsakul, Seewaboon Sireeratawong and Sunee Chansakaow
Plants 2024, 13(20), 2862; https://doi.org/10.3390/plants13202862 - 13 Oct 2024
Viewed by 434
Abstract
This study aims to investigate selected medicinal plants’ anti-oxidative and antihyperglycemic activities to develop an effective remedy for lowering blood glucose levels and/or reducing diabetes complications. Thai medicinal plants, reported to have blood sugar-lowering effects, were selected for the study: Coccinia grandis, [...] Read more.
This study aims to investigate selected medicinal plants’ anti-oxidative and antihyperglycemic activities to develop an effective remedy for lowering blood glucose levels and/or reducing diabetes complications. Thai medicinal plants, reported to have blood sugar-lowering effects, were selected for the study: Coccinia grandis, Gymnema inodorum, Gynostemma pentaphyllum, Hibiscus sabdariffa, Momordica charantia, Morus alba, and Zingiber officinale. Each species was extracted by Soxhlet’s extraction using ethanol as solvent. The ethanolic crude extract of each species was then evaluated for its phytochemicals, anti-oxidant, and antihyperglycemic activities. The results showed that the extract of Z. officinale gave the highest values of total phenolic and total flavonoid content (167.95 mg gallic acid equivalents (GAE)/g and 81.70 mg CE/g, respectively). Anti-oxidant activity was determined using DPPH and ABTS radical scavenging activity. Among the ethanolic extracts, Z. officinale exhibited the highest anti-oxidant activity with IC50 values of 19.16 and 8.53 µg/mL, respectively. The antihyperglycemic activity was assessed using α-glucosidase inhibitory and glucose consumption activities. M. alba and G. pentaphyllum demonstrated the highest α-glucosidase inhibitory activity among the ethanolic extracts, with IC50 values of 134.40 and 329.97 µg/mL, respectively. Z. officinale and H. sabdariffa showed the highest percentage of glucose consumption activity in induced insulin-resistant HepG2 cells at a concentration of 50 µg/mL with 145.16 and 107.03%, respectively. The results from α-glucosidase inhibitory and glucose consumption activities were developed as an effective antihyperglycemic remedy. Among the remedies tested, the R1 remedy exhibited the highest potential for reducing blood glucose levels, with an IC50 value of 122.10 µg/mL. Therefore, the R1 remedy should be further studied for its effects on animals. Full article
(This article belongs to the Section Phytochemistry)
Show Figures

Figure 1

Figure 1
<p>Effect of medicinal plants on cell viability in insulin-resistant HepG2 cells by SRB assay. The results were represented as mean ± SD (n = 3) and significantly different from the control group (untreated 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, **** <span class="html-italic">p</span> &lt; 0.0001).</p>
Full article ">Figure 2
<p>Glucose consumption percentage of ethanolic extracts; (<b>a</b>) <span class="html-italic">C. grandis</span> extract; (<b>b</b>) <span class="html-italic">G. inodorum</span> extract; (<b>c</b>) <span class="html-italic">G. pentaphyllum</span> extract; (<b>d</b>) <span class="html-italic">H. sabdariffa</span> extract; (<b>e</b>) <span class="html-italic">M. charantia</span> extract; (<b>f</b>) <span class="html-italic">M. alba</span> extract and (<b>g</b>) <span class="html-italic">Z. officinale</span> extract. The results were represented as mean ± SD (n = 3), and significantly differed from the control group (insulin-resistant group) (* <span class="html-italic">p</span> &lt; 0.05).</p>
Full article ">Figure 3
<p>Glucose consumption percentage of remedies: (<b>a</b>) R2 remedy and (<b>b</b>) R3 remedy. The results were represented as mean ± SD (n = 3) and significantly differed from the control group (insulin-resistant group) (* <span class="html-italic">p</span> &lt; 0.05).</p>
Full article ">Figure 4
<p>TLC chromatograms of ethanolic extract of remedy: (<b>a</b>) visible light; (<b>b</b>) UV at 254 nm; (<b>c</b>) UV at 366 nm; (<b>d</b>) 20% sulfuric acid spraying reagent; (1) rutin; (2) quercetin; (3) chlorogenic acid; (4) ginsenoside Rb1; (5) <span class="html-italic">G. pentaphyllum</span> extract; (6) <span class="html-italic">M. alba</span> extract and (7) the R1 remedy.</p>
Full article ">Figure 5
<p>Compound structures of standard references: (<b>a</b>) rutin, (<b>b</b>) quercetin, (<b>c</b>) chlorogenic acid and (<b>d</b>) ginsenoside Rb1.</p>
Full article ">Figure 6
<p>Densitograms of the remedy: (<b>a</b>) UV 254 nm, (<b>b</b>) UV at 366 nm, (1) rutin, (2) quercetin, (3) chlorogenic acid, (4) ginsenoside Rb1, (5) <span class="html-italic">G. pentaphyllum</span> extract, (6) <span class="html-italic">M. alba</span> extract and (7) the R1 remedy.</p>
Full article ">
15 pages, 2336 KiB  
Article
Characterization of Human Breast Milk-Derived Limosilactobacillus reuteri MBHC 10138 with Respect to Purine Degradation, Anti-Biofilm, and Anti-Lipid Accumulation Activities
by Jinhua Cheng, Joo-Hyung Cho and Joo-Won Suh
Antibiotics 2024, 13(10), 964; https://doi.org/10.3390/antibiotics13100964 (registering DOI) - 12 Oct 2024
Viewed by 269
Abstract
Background: Human breast milk is a valuable source of potential probiotic candidates. The bacteria isolated from human breast milk play an important role in the development of the infant gut microbiota, exhibiting diverse biological functions. Methods: In this study, Limosilactobacillus reuteri MBHC 10138 [...] Read more.
Background: Human breast milk is a valuable source of potential probiotic candidates. The bacteria isolated from human breast milk play an important role in the development of the infant gut microbiota, exhibiting diverse biological functions. Methods: In this study, Limosilactobacillus reuteri MBHC 10138 isolated from breast milk was characterized in terms of its probiotic safety characteristics and potential efficacy in hyperuricemia, obesity, lipid liver, and dental caries, conditions which Korean consumers seek to manage using probiotics. Results: Strain MBHC 10138 demonstrated a lack of D-lactate and biogenic amine production as well as a lack of bile salt deconjugation and hemolytic activity. It also exhibited susceptibility to common antibiotics, tolerance to simulated oral–gastric–intestinal conditions, and superior biological activity compared to three L. reuteri reference strains, including KACC 11452 and MJ-1, isolated from feces, and a commercial strain isolated from human breast milk. Notably, L. reuteri MBHC 10138 showed high capabilities in assimilating guanosine (69.48%), inosine (81.92%), and adenosine (95.8%), strongly inhibited 92.74% of biofilm formation by Streptococcus mutans, and reduced lipid accumulation by 32% in HepG2 cells. Conclusions: These findings suggest that strain MBHC 10138, isolated from human breast milk, has potential to be developed as a probiotic for managing hyperuricemia, obesity, and dental caries after appropriate in vivo studies. Full article
(This article belongs to the Section Antibiofilm Strategies)
Show Figures

Figure 1

Figure 1
<p>A phylogenic tree based on the 16S-rRNA gene sequence, showing <span class="html-italic">Limosilactobacillus reuteri</span> MBHC 10138’s relationship with the closely related <span class="html-italic">Limosilactobacillus</span> species. The sequences of related species that were over 97% similar to MBHC 10138 were downloaded from the Ezbiocloud database. The tree was generated by the Mega program (Version 11) using the neighbor-joining method. The numbers at the nodes indicate the percentage levels of bootstrap support based on a neighbor-joining analysis of 1000 replicas. The scale bar denotes 0.002 substitutions per nucleotide position.</p>
Full article ">Figure 2
<p><span class="html-italic">Limosilactobacillus reuteri</span> MBHC 10138, commercial, KACC 11452, and MJ-1 strains’ guanosine (<b>A</b>,<b>D</b>), inosine (<b>B</b>,<b>E</b>), and adenosine (<b>C</b>,<b>F</b>) degradation ability. The cell density is 10<sup>9</sup> CFU/mL (<b>A</b>–<b>C</b>), 10<sup>8</sup> CFU/mL (<b>D</b>,<b>E</b>), and 10<sup>7</sup> CFU/mL (<b>F</b>), for 30 min (<b>A</b>–<b>C</b>) or 15 min (<b>D</b>–<b>F</b>) of incubation. The results are presented as the mean ± standard deviation of triplicate 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.</p>
Full article ">Figure 3
<p><span class="html-italic">Streptococcus mutans</span> KCTC 3065 biofilm formation was inhibited by treatment with <span class="html-italic">Limosilactobacillus reuteri</span> MBHC 10138, commercial, KACC 11452, and MJ-1 strains, at 10<sup>8</sup> cfu/mL concentrations. Biofilm inhibition by <span class="html-italic">L. ruteri</span> was quantified via comparison with the untreated control (<b>A</b>), and the safarin-dyed biofilm was shown on the plates (<b>B</b>). The results are presented as the mean ± standard deviation of triplicate independent experiments. **** <span class="html-italic">p</span> &lt; 0.0001.</p>
Full article ">Figure 4
<p>Lipid accumulation inhibition by <span class="html-italic">Limosilactobacillus reuteri</span> MBHC 10138, commercial, KACC 11452, and MJ-1 treatment in HepG2 cells at a cell density of 10<sup>9</sup> cfu/mL; those stained with Oil Red O were extracted with isopropanol and quantified at OD 510 (<b>A</b>). A typical image of stained HepG2 cells under an optical microscope at 100× magnification (<b>B</b>). The results are presented as the mean ± standard deviation of triplicate independent experiments. ** <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.</p>
Full article ">
20 pages, 1958 KiB  
Article
Assessing the Toxicity of Metal- and Carbon-Based Nanomaterials In Vitro: Impact on Respiratory, Intestinal, Skin, and Immune Cell Lines
by Juliana Carrillo-Romero, Gartze Mentxaka, Adrián García-Salvador, Alberto Katsumiti, Susana Carregal-Romero and Felipe Goñi-de-Cerio
Int. J. Mol. Sci. 2024, 25(20), 10910; https://doi.org/10.3390/ijms252010910 - 10 Oct 2024
Viewed by 467
Abstract
The field of nanotechnology has experienced exponential growth, with the unique properties of nanomaterials (NMs) being employed to enhance a wide range of products across diverse industrial sectors. This study examines the toxicity of metal- and carbon-based NMs, with a particular focus on [...] Read more.
The field of nanotechnology has experienced exponential growth, with the unique properties of nanomaterials (NMs) being employed to enhance a wide range of products across diverse industrial sectors. This study examines the toxicity of metal- and carbon-based NMs, with a particular focus on titanium dioxide (TiO2), zinc oxide (ZnO), silica (SiO2), cerium oxide (CeO2), silver (Ag), and multi-walled carbon nanotubes (MWCNTs). The potential health risks associated with increased human exposure to these NMs and their effect on the respiratory, gastrointestinal, dermal, and immune systems were evaluated using in vitro assays. Physicochemical characterisation of the NMs was carried out, and in vitro assays were performed to assess the cytotoxicity, genotoxicity, reactive oxygen species (ROS) production, apoptosis/necrosis, and inflammation in cell lines representative of the systems evaluated (3T3, Caco-2, HepG2, A549, and THP-1 cell lines). The results obtained show that 3T3 and A549 cells exhibit high cytotoxicity and ROS production after exposure to ZnO NMs. Caco-2 and HepG2 cell lines show cytotoxicity when exposed to ZnO and Ag NMs and oxidative stress induced by SiO2 and MWCNTs. THP-1 cell line shows increased cytotoxicity and a pro-inflammatory response upon exposure to SiO2. This study emphasises the importance of conducting comprehensive toxicological assessments of NMs given their physicochemical interactions with biological systems. Therefore, it is of key importance to develop robust and specific methodologies for the assessment of their potential health risks. Full article
(This article belongs to the Special Issue Toxicity of Nanoparticles)
Show Figures

Figure 1

Figure 1
<p>TEM images of (<b>A</b>) NM 101, (<b>B</b>) NM 110, (<b>C</b>) NM 200, (<b>D</b>) NM 212, (<b>E</b>) NM 300 K, and (<b>F</b>) NM 400 at 0 h in milliQ water (stock) and 0 h and 24 h in DMEM, MEM, and RPMI.</p>
Full article ">Figure 2
<p>ROS production (%) of 3T3 (<b>A</b>), HepG2 (<b>B</b>), Caco-2 (<b>C</b>), and A549 (<b>D</b>) cell lines after exposure to all NMs (100 µg/mL and 50 µg/mL in NM 110 0.5 and NM 300 K 0.5) and to the negative (non-treated cells) and positive (H<sub>2</sub>O<sub>2</sub>) controls for 24 h. Results are expressed as means ± SD of six replicates per tested condition and three independent assays (<span class="html-italic">n</span> = 18). * Significantly different from negative control (C−) (<span class="html-italic">p</span> &lt; 0.05). ** (<span class="html-italic">p</span> &lt; 0.01).</p>
Full article ">Figure 3
<p>Necrosis/apoptosis (%) of THP-1 cell line after exposure to all NMs (100 µg/mL and 50 µg/mL in NM 110 0.5 and NM 300 K 0.5) and to the negative (non-treated cells) and positive (Camptothecin and SDS for apoptosis and necrosis, respectively) controls for 24 h. Results are expressed as means ± SD of six replicates per tested condition and three independent assays (<span class="html-italic">n</span> = 18). * Significantly different from negative control (C−) (<span class="html-italic">p</span> &lt; 0.05).</p>
Full article ">Figure 4
<p>(<b>A</b>) IL-1β, (<b>B</b>) TNF-α, (<b>C</b>) IL-8, and (<b>D</b>) IL-10 release in THP-1 cell line after exposure to all NMs (100 µg/mL and 50 µg/mL in NM 110 0.5 and NM 300 K 0.5) and to the negative (non-treated cells) and positive (LPS) controls for 24 h. Results are expressed as means ± SD of six replicates per tested condition and three independent assays (<span class="html-italic">n</span> = 18). * Significantly different from negative control (C−) (<span class="html-italic">p</span> &lt; 0.05).</p>
Full article ">
10 pages, 3395 KiB  
Communication
Acceleration of Ethanol Metabolism by a Patented Bos taurus Isolated Alcohol Degradation Protein (ADP) on Acute Alcohol Consumption
by Bun Tsoi, Huan Zhang, Chun-Pang So, Angel Ka-Kei Lam, Christina Chui-Wa Poon, Sek-Lun Law, Bing-Lou Wong and Sai-Wang Seto
Foods 2024, 13(19), 3207; https://doi.org/10.3390/foods13193207 - 9 Oct 2024
Viewed by 487
Abstract
Alcoholic beverages are among the most widely enjoyed leisure drinks around the world. However, irresponsible drinking habits can have detrimental effects on human health. Therefore, exploring strategies to alleviate discomfort following alcohol consumption would be beneficial for individuals who inevitably need to consume [...] Read more.
Alcoholic beverages are among the most widely enjoyed leisure drinks around the world. However, irresponsible drinking habits can have detrimental effects on human health. Therefore, exploring strategies to alleviate discomfort following alcohol consumption would be beneficial for individuals who inevitably need to consume alcohol. In this study, three different models were used to determine the efficacy of a patented alcohol degradation protein (ADP) extracted from Bos taurus on ethanol metabolism. In an ethanol-challenged HepG2 cell model, ADP significantly protected the cell from ethanol-induced toxicity. Subsequently, results demonstrated that ADP significantly alleviated the effect of ethanol, as reflected by the increased distance and activity time of zebrafish during the testing period. Additionally, in a rat model, ADP promoted ethanol degradation at 1 and 2 h after ethanol consumption. Mechanistic studies found that ADP treatment increased ADH and ALDH activity in the gastrointestinal tract. ADP also exhibited potent antioxidation effects by lowering HO-1 expression in the liver. In conclusion, we believe that ADP is a promising product for relieving hangover symptoms after ethanol consumption, with demonstrated safety and effectiveness at the recommended dosage. Full article
(This article belongs to the Section Nutraceuticals, Functional Foods, and Novel Foods)
Show Figures

Figure 1

Figure 1
<p>Protective effect of ADP on HepG2 cells challenged with EtOH.</p>
Full article ">Figure 2
<p>ADP showed an anti-hangover effect in zebrafish larvae. (<b>A</b>) Schematic illustration of the design of the zebrafish dark/light assay. The zebrafish larvae received a 2h treatment of 2% EtOH. Then, the zebrafish larvae were exposed to ADP without 2% EtOH for 1 h. After 1 h, the zebrafish larvae were exposed to 3 alternating dark/light cycles (5 min dark–5 min light). Analysis was performed on the average parameters of the three dark phases. (<b>B</b>) The total moving distance of zebrafish larvae during the analysis period. (<b>C</b>) Representative swimming tracks of zebrafish during one of the dark phases. (<b>D</b>–<b>F</b>) Quantitative analysis of average tracking distance, swimming velocity and cumulative moving duration in the three dark phases. All data are presented as mean ± SD (n = 24). The significance of the difference was compared vs. control at ## <span class="html-italic">p</span> &lt; 0.01 and vs. EtOH at ** <span class="html-italic">p</span> &lt; 0.01 and *** <span class="html-italic">p</span> &lt; 0.001 in (<b>D</b>–<b>F</b>).</p>
Full article ">Figure 3
<p>ADP accelerated ethanol elimination after acute ethanol administration. (<b>A</b>) Schematic illustration of the design of ethanol metabolism assay. SD rats received an acute administration of 30% EtOH. ADP and King Drink were pre-administered to respective treatment groups 30 min before EtOH administration. Blood was collected through the tail vein at different time points. (<b>B</b>) Blood ethanol concentration curve following acute intake of EtOH. (<b>C</b>,<b>D</b>) Blood ADH and ALDH activity over the 7 h time course. (<b>E</b>,<b>F</b>) Liver ADH and ALDH activity at 7 h after ethanol administration. (<b>G</b>,<b>H</b>) ADH and ALDH activity in cecum content of animals 7 h after ethanol administration. All data are presented as mean ± SD (n &gt; 8). The significance of the difference in treatment groups was compared vs. control at <sup>#</sup> <span class="html-italic">p</span> &lt; 0.05; <sup>##</sup> <span class="html-italic">p</span> &lt; 0.01; and vs. EtOH at * <span class="html-italic">p</span> &lt; 0.05 in (<b>B</b>,<b>G</b>,<b>H</b>).</p>
Full article ">Figure 4
<p>ADP prevented over-activation of HO-1 after acute ethanol administration. (<b>A</b>) Effects of ADP on histopathological changes in liver tissues after acute ethanol administration. Bar = 50 μm. (<b>B</b>) Representative Western blot bands for protein expressions in liver of animals pretreated with ADP and acute ethanol ingestion. C: control; E: EtOH; K: King Drink; and A: ADP group. (<b>C</b>,<b>D</b>) Statistical analysis of Nrf2 and HO-1 protein expression in liver. All data are presented as mean ± SD (n = 3). The significance of difference in treatment group was compared vs. control at # <span class="html-italic">p</span> &lt; 0.05 and vs. EtOH at * <span class="html-italic">p</span> &lt; 0.05 in (<b>D</b>).</p>
Full article ">
17 pages, 3292 KiB  
Article
NRF2 and Thioredoxin Reductase 1 as Modulators of Interactions between Zinc and Selenium
by Alina Löser, Maria Schwarz and Anna Patricia Kipp
Antioxidants 2024, 13(10), 1211; https://doi.org/10.3390/antiox13101211 - 8 Oct 2024
Viewed by 447
Abstract
Background: Selenium and zinc are essential trace elements known to regulate cellular processes including redox homeostasis. During inflammation, circulating selenium and zinc concentrations are reduced in parallel, but underlying mechanisms are unknown. Accordingly, we modulated the zinc and selenium supply of HepG2 cells [...] Read more.
Background: Selenium and zinc are essential trace elements known to regulate cellular processes including redox homeostasis. During inflammation, circulating selenium and zinc concentrations are reduced in parallel, but underlying mechanisms are unknown. Accordingly, we modulated the zinc and selenium supply of HepG2 cells to study their relationship. Methods: HepG2 cells were supplied with selenite in combination with a short- or long-term zinc treatment to investigate intracellular concentrations of selenium and zinc together with biomarkers describing their status. In addition, the activation of the redox-sensitive transcription factor NRF2 was analyzed. Results: Zinc not only increased the nuclear translocation of NRF2 after 2 to 6 h but also enhanced the intracellular selenium content after 72 h, when the cells were exposed to both trace elements. In parallel, the activity and expression of the selenoprotein thioredoxin reductase 1 (TXNRD1) increased, while the gene expression of other selenoproteins remained unaffected or was even downregulated. The zinc effects on the selenium concentration and TXNRD activity were reduced in cells with stable NRF2 knockdown in comparison to control cells. Conclusions: This indicates a functional role of NRF2 in mediating the zinc/selenium crosstalk and provides an explanation for the observed unidirectional behavior of selenium and zinc. Full article
(This article belongs to the Special Issue Oxidative Stress and NRF2 in Health and Disease—2nd Edition)
Show Figures

Figure 1

Figure 1
<p>Establishment of culture conditions with 2.5% FBS. HepG2 cells were treated with or without 50 nM sodium selenite in combination with or without 100 μM zinc sulfate for 72 h in media containing 10% or 2.5% FCS. The relative cell count (<b>a</b>) was determined by trypan blue exclusion test. Intracellular trace element concentrations of Se (<b>b</b>), and Zn (<b>c</b>) were determined by total reflection X-ray fluorescence spectrometry (TXRF). The measurement was performed for 1000 s with 1 mg/L yttrium as an internal standard. Protein expression of MT (<b>d</b>) was determined by Western blot, normalized to Ponceau staining. Either the −Se/−Zn 10% FBS group (<b>a</b>) or the +Se/+Zn group 10% FBS (<b>b</b>–<b>d</b>) were set as 1. Results are presented as mean + SD (n = 3–4). *** <span class="html-italic">p</span> &lt; 0.001 vs. −Se, # <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. −Zn, <span>$</span><span>$</span> <span class="html-italic">p</span> &lt; 0.01, <span>$</span><span>$</span><span>$</span> <span class="html-italic">p</span> &lt; 0.001 vs. 10% FBS, calculated by two-factorial ANOVA with Bonferroni’s post-test.</p>
Full article ">Figure 2
<p>Zinc increased the activity of TXNRD but not of GPX. HepG2 cells were treated with or without 50 nM sodium selenite in combination with or without 100 µM zinc sulfate for up to 72 h (<b>a</b>–<b>e</b>,<b>g</b>,<b>h</b>) or with indicated Zn concentrations (<b>f</b>) in media containing 2.5% FCS. Enzyme activities of GPX (a) and TXNRD (<b>e</b>,<b>f</b>) and the protein expression levels of GPX1 (<b>b</b>), GPX2 (<b>c</b>), GPX4 (<b>d</b>), TXNRD1 (<b>g</b>), and TXNRD2 (<b>h</b>) were determined photometrically (<b>a</b>,<b>e</b>,<b>f</b>) or by Western blot (<b>b</b>–<b>d</b>,<b>g</b>,<b>h</b>). Protein expression was normalized to Ponceau staining and presented relative to +Se/+Zn treatment. Results are presented as mean + SD (n = 3–4). * <span class="html-italic">p</span> &lt; 0.05, *** <span class="html-italic">p</span> &lt; 0.001 vs. −Se, # <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. −Zn, calculated by two-factorial ANOVA (<b>a</b>–<b>e</b>,<b>g</b>,<b>h</b>) or with one-factorial ANOVA (<b>f</b>) with Bonferroni’s post-test.</p>
Full article ">Figure 3
<p>Zinc enhanced the nuclear translocation of NRF2 and MTF1 and increased the expression of their target genes. HepG2 cells were treated with or without 50 nM sodium selenite in combination with or without 100 μM zinc sulfate up to 72 h (<b>c</b>,<b>d</b>) or treated with selenite for 72 h in combination with or without zinc for the time as indicated (<b>a</b>,<b>b</b>,<b>e</b>–<b>j</b>) in media containing 2.5% FCS. Nuclear protein levels of MTF1 (<b>a</b>), NRF2 (<b>b</b>) and SELENOH (<b>g</b>), enzyme activity of NQO1 (<b>c</b>), and protein expression of NQO1 (<b>d</b>) were determined photometrically (<b>c</b>) or by Western blot (<b>a</b>,<b>b</b>,<b>d</b>,<b>g</b>). Protein expression was normalized to Ponceau staining and presented relative to samples with selenium treatment in combination with 2 h zinc treatment (<b>a</b>,<b>b</b>) or to samples with selenium treatment in combination with 6 h zinc treatment (<b>g</b>) or samples with selenium treatment (<b>d</b>). The mRNA expression levels of NRF2 and MTF1 target genes (<b>e</b>–<b>f</b>,<b>h</b>–<b>j</b>) were analyzed by qPCR. Gene expression was normalized to the normalization factor of the reference genes HPRT, RPL13a, and GAPDH and presented relative to samples with selenium treatment and 6 h zinc treatment. Results are presented as mean + SD (n = 3–4). * <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. −Se, # <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. −Zn, <span>$</span><span>$</span><span>$</span> <span class="html-italic">p</span> &lt; 0.001 vs. 6 h Zn, calculated by two-factorial ANOVA with Bonferroni’s post-test.</p>
Full article ">Figure 4
<p>NRF2 partially mediated the zinc effects on selenium homeostasis. HepG2 cells with a stable small hairpin RNA-mediated NRF2 knockdown (NRF2-KD) and scramble (Scr) control cells were treated with 50 nM sodium selenite for 72 h and 100 μM zinc sulfate for the indicated time (<b>a</b>–<b>c</b>) or for 72 h (<b>d</b>,<b>e</b>) in media containing 2.5% FCS. The nuclear protein levels of NRF2 (<b>a</b>,<b>c</b>) and MTF1 (<b>b</b>,<b>c</b>) were analyzed by Western blot. Protein expression was normalized to the Ponceau staining. Enzyme activities of TNXRD (<b>d</b>) and intracellular Se concentrations (<b>e</b>) were determined photometrically or by total reflection X-ray fluorescence spectrometry, respectively. The measurement was performed for 1000 s with 1 mg/L yttrium as an internal standard. The results are presented as mean + SD (n = 3). ### <span class="html-italic">p</span> &lt; 0.001 vs. –Zn, *** <span class="html-italic">p</span> &lt; 0.001 vs. 6 h Zn, <span>$</span><span>$</span><span>$</span> <span class="html-italic">p</span> &lt; 0.001 vs. Scr (<b>a</b>,<b>b</b>) or *** <span class="html-italic">p</span> &lt; 0.001 vs. –Se, ### <span class="html-italic">p</span> &lt; 0.001 vs. –Zn, <span>$</span><span>$</span> <span class="html-italic">p</span> &lt; 0.01, <span>$</span><span>$</span><span>$</span> <span class="html-italic">p</span> &lt; 0.001 vs. Scr (<b>d</b>,<b>e</b>) calculated by two-factorial ANOVA with Bonferroni’s post-test.</p>
Full article ">Figure 5
<p>Trace element transporters were increased by zinc. HepG2 cells were treated with or without 50 nM sodium selenite for 72 h in combination with or without 100 μM zinc sulfate for indicated time (up to 72 h) in media containing 2.5% FCS. The mRNA expression levels of ZIP8 (<b>a</b>), XCT (<b>b</b>), and APOER2 (<b>d</b>) were analyzed by qPCR. Gene expression was normalized to the normalization factor of the reference genes HPRT, RPL13a, and GAPDH and was presented relative to samples with selenium and 6 h zinc treatment. XCT (<b>c</b>) and APOER2 (<b>e</b>) were analyzed by Western blot. Protein expression was normalized to Ponceau staining and presented relative to samples with selenium and 48 h (<b>c</b>) or 72 h zinc treatment (<b>e</b>). Results are presented as mean + SD (n = 3–4). * <span class="html-italic">p</span> &lt; 0.05 vs. −Se; # <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. −Zn, <span>$</span><span>$</span><span>$</span> <span class="html-italic">p</span> &lt; 0.001 vs. 6 h Zn, calculated by two-factorial ANOVA (<b>a</b>,<b>b</b>,<b>d</b>) or with one-factorial ANOVA (<b>c</b>,<b>e</b>) with Bonferroni’s post-test.</p>
Full article ">
20 pages, 1974 KiB  
Article
Modulatory Effects of Chalcone Thio-Derivatives on NF-κB and STAT3 Signaling Pathways in Hepatocellular Carcinoma Cells: A Study on Selected Active Compounds
by Katarzyna Papierska, Eliza Judasz, Wiktoria Tonińska, Maciej Kubicki and Violetta Krajka-Kuźniak
Int. J. Mol. Sci. 2024, 25(19), 10739; https://doi.org/10.3390/ijms251910739 - 5 Oct 2024
Viewed by 554
Abstract
Our previous studies demonstrated the modulatory effects of new synthetic thio-chalcone derivatives in dishes on the Nrf2, NF-κB, and STAT3 signaling pathways in colon cancer cells. This study aimed to evaluate the effect of four selected active chalcone thio-derivatives on the NF-κB and [...] Read more.
Our previous studies demonstrated the modulatory effects of new synthetic thio-chalcone derivatives in dishes on the Nrf2, NF-κB, and STAT3 signaling pathways in colon cancer cells. This study aimed to evaluate the effect of four selected active chalcone thio-derivatives on the NF-κB and STAT3 signaling pathways involved in inflammatory processes and cell proliferation in human liver cancer cells. Cell survival was assessed for cancer (HepG2) and normal (THLE-2) cell lines. Activation of NF-κB and STAT3 signaling pathways and the expression of proteins controlled by these pathways were estimated by Western blot, and qRT-PCR assessed the expression of NF-κB and STAT3 target genes. We also evaluated the impact on the selected kinases responsible for the phosphorylation of the studied transcription factors by MagneticBead-Based Multiplex Immunoassay. Among the thio-derivatives tested, especially derivatives 1 and 5, there was an impact on cell viability, cell cycle, apoptosis, and activation of NF-κB and STAT3 pathways in hepatocellular carcinoma (HCC), which confirms the possibilities of using them in combinatorial molecular targeted therapy of HCC. The tested synthetic thio-chalcones exhibit anticancer activity by initiating proapoptotic processes in HCC while showing low toxicity to non-cancerous cells. These findings confirm the possibility of using chalcone thio-derivatives in molecularly targeted combination therapy for HCC. Full article
(This article belongs to the Special Issue Advances in Cell Signaling Pathways and Signal Transduction)
Show Figures

Figure 1

Figure 1
<p>The cytotoxicity evaluation of synthetic thio-chalcone derivatives (<b>1</b>, <b>2</b>, <b>4</b>, and <b>5</b>) on: (<b>A</b>) THLE-2 cell line and (<b>B</b>) HepG2 cell line. Control cells were treated with the vehicle. The values are shown as the mean ± SEM calculated from three independent experiments. The colors and the numbers, respectively, mean the tested synthetic chalcone thio-derivatives: <b>1</b>—3-(4-methoxy-3-methylthiophenyl)-1-(3,4,5-trimethoxyphenyl)-prop-2-en-1-one. <b>2</b>—3-(3-methoxy-4-methylthiophenyl)-1-(3-bromo-4,5-dimethoxyphenyl)-prop-2-en-1-one. <b>4</b>—3-(4-methylthiophene)-1-(3-bromo-4,5-dimethoxyphenyl)-prop-2-en-1-one. <b>5</b>—3-(3-methoxy-4-methylthiophenyl)-1-(3-bromo-5-methoxy-4-methylthiophene)-prop-2-en-1-one.</p>
Full article ">Figure 2
<p>Distribution of cell cycle phases in HepG2 cell line after 24 h incubation with synthetic thio-chalcone derivatives (<b>1</b>/5, <b>2</b>/5, <b>4</b>/5, <b>5</b>/5—thio-chalcone/at a concentration of 5 μM; <b>1</b>/15, <b>2</b>/15, <b>4</b>/15, <b>5</b>/15—thio-chalcone/at a concentration of 15 μM). (<b>A</b>) Graphs of the mean ± SEM of the percentage of cells in G1/G0, S, and G2/M phases were calculated from two independent experiments. (*) indicates statistically significant differences compared to the control group for a given phase (<span class="html-italic">p</span> &lt; 0.05). (<b>B</b>) Histograms of the negative (DMSO) and positive (Topotecan 1500 nM) control analysis. (<b>C</b>) Histograms of representative samples for individual compounds.</p>
Full article ">Figure 3
<p>Apoptosis profile of HepG2 cell line treated with the synthetic thio-chalcone derivatives (<b>1</b>/5, <b>2</b>/5, <b>4</b>/5, <b>5</b>/5—thio-chalcone/at a concentration of 5 μM; <b>1</b>/15, <b>2</b>/15, <b>4</b>/15, <b>5</b>/15—thio-chalcone/at a concentration of 15 μM). (<b>A</b>) Graphs of the mean ± SEM of the percentage of apoptotic cells (in the early, late, and complete phases) after 24 h of incubation with the test compounds. Statistically significant differences compared to the control group of early apoptosis * (<span class="html-italic">p</span> &lt; 0.05) and the control group of late apoptosis # (<span class="html-italic">p</span> &lt; 0.05). (<b>B</b>) Histograms of the negative control (DMSO) and positive (Topotecan 1500 nM) analysis. (<b>C</b>) Histograms of representative samples for individual compounds. The four square markers of each graph reflect different cellular states: the top left square contains dead cells (necrosis), the top right contains late apoptosis/dead cells (cells that are positive for both Annexin V and the cell death marker 7-AAD), the left lower corner contains living cells, and the lower right corner contains early apoptosis cells (cells that are positive only for annexin V).</p>
Full article ">Figure 4
<p>Effect of the tested synthetic thio-chalcone derivatives (<b>1</b>, <b>2</b>, <b>4</b>, and <b>5</b>) in the HepG2 cell line on the protein level of: (<b>A</b>) p53 and (<b>B</b>) TNF-α.Protein levels were expressed as relative changes in the protein level with respect to the control. Representative Western blots of cytosolic fraction treated anti-p53 and anti-TNF-α antibodies are presented under the graphs. The order of the bands in the immunoblot image corresponds to the order of the bars in the diagram. Anti-β-actin antibodies were used to normalize the results. The results presented are the mean ± SEM from two separate experiments. Statistically significant differences compared to the control group * (<span class="html-italic">p</span> &lt; 0.05). See also <a href="#app1-ijms-25-10739" class="html-app">Supplementary Figure S1</a>.</p>
Full article ">Figure 5
<p>Effect of synthetic thio-chalcone derivatives (<b>1</b>, <b>2</b>, <b>4</b>, and <b>5</b>) on the regulation of protein controlling several signaling pathways measured by bead-based multiplex immunoassay in HepG2 cell line. The relative changes in the protein level of: (<b>A</b>) CREB and phospho-CREB, (<b>B</b>) JNK and phospho-JNK, (<b>C</b>) ERK and phospho-ERK, (<b>D</b>) Akt and phospho-Akt, (<b>E</b>) p38 and phospho-p38 and (<b>F</b>) p70S6K and phospho-p70S6K were measured. Results are prepared based on the cytosolic fraction of proteins and are shown in comparison to vehicle control. The values are shown as the mean ± SEM calculated from two independent experiments (a fold of control). Statistically significant differences compared to the control group * (<span class="html-italic">p</span> &lt; 0.05).</p>
Full article ">Figure 6
<p>Effect of the tested synthetic thio-chalcone derivatives (<b>1</b>, <b>2</b>, <b>4</b>, and <b>5</b>) on the protein level of: (<b>A</b>) STAT3 in cytosolic fraction from the HepG2 cell line, (<b>B</b>) STAT3 in nuclear fraction from the HepG2 cell line and (<b>C</b>) phospho-STAT3 in nuclear fraction from the HepG2 cell line. Protein levels were expressed as relative changes in the protein level with respect to the control. Representative Western blots of cytosolic fraction-treated anti-STAT3 and nuclear fraction-treated anti-STAT3 and anti-phospho-STAT3 antibodies are presented under the graphs. The order of the bands in the immunoblot image corresponds to the order of the bars in the diagram. Anti-β-actin and anti-lamin antibodies were used to normalize the results. See also <a href="#app1-ijms-25-10739" class="html-app">Supplementary Figure S2</a>. The results presented are the mean ± SEM from two separate experiments. Statistically significant differences compared to the control group * (<span class="html-italic">p</span> &lt; 0.05).</p>
Full article ">Figure 7
<p>Effect of the tested synthetic thio-chalcone derivatives (<b>1</b>, <b>2</b>, <b>4</b>, and <b>5</b>) on the expression of STAT3 and selected target genes in the HepG2 cell line: (<b>A</b>) STAT3 transcript level and (<b>B</b>) Bcl-xL and Bax transcript levels. The transcript level was calculated as the mRNA level compared to control cells, for which expression was considered. The results presented are the mean ± SEM from two separate experiments. Statistically significant differences compared to the control group * (<span class="html-italic">p</span> &lt; 0.05). (<b>C</b>) Bcl-xL and Bax protein levels. Protein levels were expressed as relative changes in the protein level with respect to the control. Representative Western blots of cytosolic fraction-treated anti-Bcl-xL and anti- Bax antibodies are presented under the graphs. The order of the bands in the immunoblot image corresponds to the order of the bars in the diagram. Anti-β-actin antibodies were used to normalize the results. See also <a href="#app1-ijms-25-10739" class="html-app">Supplementary Figure S3</a>. The results presented are the mean ± SEM from two separate experiments. Statistically significant differences compared to the control group * (<span class="html-italic">p</span> &lt; 0.05).</p>
Full article ">Figure 8
<p>Effect of the tested synthetic thio-chalcone derivatives (<b>1</b>, <b>2</b>, <b>4</b>, and <b>5</b>) on the protein level of: (<b>A</b>) NF-κB p50 and p65 subunits in cytosolic fraction from HepG2 cell line, (<b>B</b>) NF-κB p50 and p65 subunits in nuclear fraction from HepG2 cell line and (<b>C</b>) IKKα/β in cytosolic fraction from HepG2 cell line. Protein levels were expressed as relative changes in the protein level with respect to the control. Representative Western blots of cytosolic and nuclear fraction-treated anti-NF-κB p50 and anti-NF-κB p65 and cytosolic fraction-treated anti-IKKα/β antibodies are presented under the graphs. The order of the bands in the immunoblot image corresponds to the bars in the diagram. Anti-β-actin and anti-lamin antibodies were used to normalize the results. See also <a href="#app1-ijms-25-10739" class="html-app">Supplementary Figure S4</a>. The results presented are the mean ± SEM from two separate experiments. Statistically significant differences compared to the control group * (<span class="html-italic">p</span> &lt; 0.05) and compared to the BAY 11-7082 # (<span class="html-italic">p</span> &lt; 0.05).</p>
Full article ">Figure 9
<p>Effect of the tested synthetic thio-chalcone derivatives (<b>1</b>, <b>2</b>, <b>4</b>, and <b>5</b>) on the expression of NF-κB p50 and p65 subunits and selected target genes in the HepG2 cell line. (<b>A</b>) NF-κB p50 and p65 subunit transcript levels, and (<b>B</b>) COX-2 and iNOS transcript levels were measured. The transcript level was calculated as the mRNA level compared to control cells, for which expression was considered 1. The results presented are the mean ± SEM from two separate experiments. Statistically significant differences compared to the control group * (<span class="html-italic">p</span> &lt; 0.05). (<b>C</b>) COX-2 and iNOS protein levels. Protein levels were expressed as relative changes in the protein level with respect to the control. Representative Western blots of cytosolic fraction-treated anti-COX-2 and anti-iNOS antibodies are presented in the graphs. The order of the bands in the immunoblot image corresponds to the order of the bars in the diagram. Anti-β-actin antibodies were used to normalize the results. See also <a href="#app1-ijms-25-10739" class="html-app">Supplementary Figure S5</a>. The results presented are the mean ± SEM from two separate experiments. Statistically significant differences compared to the control group * (<span class="html-italic">p</span> &lt; 0.05).</p>
Full article ">
17 pages, 3598 KiB  
Article
DNA-Protective, Antioxidant and Anti-Carcinogenic Potential of Meadowsweet (Filipendula ulmaria) Dry Tincture
by Tsvetelina Andonova, Yordan Muhovski, Elena Apostolova, Samir Naimov, Silviya Mladenova, Iliya Slavov, Ivayla Dincheva, Vasil Georgiev, Atanas Pavlov and Ivanka Dimitrova-Dyulgerova
Antioxidants 2024, 13(10), 1200; https://doi.org/10.3390/antiox13101200 (registering DOI) - 3 Oct 2024
Viewed by 543
Abstract
Nowadays, interest in natural antioxidants (especially phenolics) for the prevention of oxidative stress-related diseases is increasing due to their fewer side effects and more potent activity than some of their synthetic analogues. New chemical and pharmacological studies of well-known herbal substances are among [...] Read more.
Nowadays, interest in natural antioxidants (especially phenolics) for the prevention of oxidative stress-related diseases is increasing due to their fewer side effects and more potent activity than some of their synthetic analogues. New chemical and pharmacological studies of well-known herbal substances are among the current trends in medicinal plant research. Meadowsweet (Filipendula ulmaria) is a popular herb used in traditional medicine to treat various diseases (including rheumatic-, inflammatory- and tumor-related disease, etc.). The dry tincture of Filipendulae ulmariae herba, collected from the Bulgarian flora, was analyzed using the HPLC method and bioassayed for antioxidant, antiproliferative and DNA-protective activities against oxidative damage. The HPLC phenolic profile showed 12 phenolics, of which salicylic acid (18.84 mg/g dry extract), rutin (9.97 mg/g de), p-coumaric acid (6.80 mg/g de), quercetin (4.47 mg/g de), rosmarinic acid (4.01 mg/g de) and vanillic acid (3.82 mg/g de) were the major components. The high antioxidant potential of the species was confirmed by using four methods, best expressed by the results of the CUPRAC assay (10,605.91 μM TE/g de). The present study reports for the first time the highly protective activities of meadowsweet dry tincture against oxidative DNA damage and its antiproliferative effect against hepatocellular carcinoma (HepG2 cell line). Meadowsweet dry tincture possesses great potential to prevent diseases caused by oxidative stress. Full article
Show Figures

Figure 1

Figure 1
<p><span class="html-italic">Filipendula ulmaria</span>—in the species’ natural habitat (<b>a</b>) and herbarium specimen (<b>b</b>).</p>
Full article ">Figure 2
<p>The main steps of the preparation of <span class="html-italic">Filipendulae ulmaria herba</span> dry tincture: air-dried herba, crushed and powdered (<b>a</b>), maceration with 96% ethanol (<b>b</b>); extract filtration (<b>c</b>); liquid tincture (<b>d</b>); dry tincture obtained (<b>e</b>).</p>
Full article ">Figure 3
<p>In vitro DNA protective activity of <span class="html-italic">Filipendula ulmaria</span> DT: (<b>a</b>) relative concentration of nicked plasmid DNA and (<b>b</b>) agarose gel electrophoresis. Lines 1–3—different Trolox concentrations (25, 50, and 100 µg/mL); line 4—Zip Ruler 1 Express DNA Ladder (Thermo Scientific, Waltham MA USA, cat. No SM1373); line 5—plasmid DNA input; line 6—negative control; lines 7—11 dilutions of tested extract (10 fg; 100 fg; 1 pg; 10 pg; and 100 pg/mL); The number designations refer to parts (<b>a</b>,<b>b</b>). The results are from triplicate measurements.</p>
Full article ">Figure 4
<p>In vitro antiproliferative capacity (%) of <span class="html-italic">F. ulmaria</span> DT on (<b>a</b>) normal HaCaT (human keratinocytes), and (<b>c</b>,<b>e</b>,<b>g</b>) tumor cell lines LnCap clone FGC (prostate cancer), SH-4 (human melanoma), and HepG2 (hepatocellular carcinoma), respectively. Cisplatin was used as positive control (<b>b</b>,<b>d</b>,<b>f</b>,<b>h</b>). Data are presented as means ± standard deviation (SD). The results are from triplicate measurements.</p>
Full article ">
19 pages, 2205 KiB  
Article
The PreS-Based Recombinant Vaccine VVX001 Induces Hepatitis B Virus Neutralizing Antibodies in a Low-Responder to HBsAg-Based HBV Vaccines
by Inna Tulaeva, Felix Lehmann, Nora Goldmann, Alexandra Dubovets, Daria Trifonova, Mikhail Tulaev, Carolin Cornelius, Milena Weber, Margarete Focke-Tejkl, Alexander Karaulov, Rainer Henning, David Niklas Springer, Ursula Wiedermann, Dieter Glebe and Rudolf Valenta
Vaccines 2024, 12(10), 1123; https://doi.org/10.3390/vaccines12101123 - 30 Sep 2024
Viewed by 859
Abstract
Background: Approximately 10–20% of subjects vaccinated with HBsAg-based hepatitis B virus (HBV) vaccines are non-responders. BM32 is a recombinant grass pollen allergy vaccine containing the HBV-derived preS surface antigen as an immunological carrier protein. PreS includes the binding site of HBV to its [...] Read more.
Background: Approximately 10–20% of subjects vaccinated with HBsAg-based hepatitis B virus (HBV) vaccines are non-responders. BM32 is a recombinant grass pollen allergy vaccine containing the HBV-derived preS surface antigen as an immunological carrier protein. PreS includes the binding site of HBV to its receptor on hepatocytes. We investigated whether immunological non-responsiveness to HBV after repeated HBsAg-based vaccinations could be overcome by immunization with VVX001 (i.e., alum-adsorbed BM325, a component of BM32). Methods: A subject failing to develop protective HBV-specific immunity after HBsAg-based vaccination received five monthly injections of 20 µg VVX001. PreS-specific antibody responses were measured by enzyme-linked immunosorbent assay (ELISA) and micro-array technology. Serum reactivity to subviral particles of different HBV genotypes was determined by sandwich ELISA. PreS-specific T cell responses were monitored by carboxyfluorescein diacetate succinimidyl ester (CFSE) staining and subsequent flow cytometry. HBV neutralization was assessed using cultured HBV-infected HepG2 cells. Results: Vaccination with VVX001 induced a strong and sustained preS-specific antibody response composed mainly of the IgG1 subclass. PreS-specific IgG antibodies were primarily directed to the N-terminal part of preS containing the sodium taurocholate co-transporting polypeptide (NTCP) attachment site. IgG reactivity to subviral particles as well as to the N-terminal preS-derived peptides was comparable for HBV genotypes A–H. A pronounced reactivity of CD3+CD4+ lymphocytes specific for preS after the complete injection course remaining up to one year after the last injection was found. Maximal HBV neutralization (98.4%) in vitro was achieved 1 month after the last injection, which correlated with the maximal IgG reactivity to the N-terminal part of preS. Conclusions: Our data suggest that VVX001 may be used as a preventive vaccination against HBV even in non-responders to HBsAg-based HBV vaccines. Full article
(This article belongs to the Special Issue 2nd Edition of Antibody Response to Infection and Vaccination)
Show Figures

Figure 1

Figure 1
<p>Vaccinations with conventional HBsAg-based vaccines in the study subject. Time points and administered vaccines are indicated.</p>
Full article ">Figure 2
<p>Development of antibodies to HBV surface proteins and their virus neutralization capacity. (<b>a</b>) Anti-HBs antibody levels (IU/L) measured after the last booster immunization with Engerix-B and after the preS-based vaccination. (<b>b</b>) PreS-specific IgG levels (OD) measured by ELISA after VVX001 immunization. (<b>c</b>) Percentages reduction in HBeAg secretion of infected NTCP-expressing HepG2 cells after pre-incubation of HBV inocula with sera obtained in the course of immunization compared to the baseline. Neutralization: ≥90% (strong neutralization), ≥50–90% (partial neutralization), ≥10–50% (low neutralization).</p>
Full article ">Figure 3
<p>VVX001-induced IgG antibodies react mainly with the N-terminus of preS1. (<b>a</b>) Localization scheme of preS-derived peptide within the preS sequence (amino acid numbering indicated for genotype A) and the levels of preS/peptide-specific IgG measured by (<b>b</b>) ELISA (OD) and (<b>c</b>) micro-array technology (fluorescence intensity, FI).</p>
Full article ">Figure 4
<p>VVX001-induced antibodies cross-react to the NTCP binding site of preS of HBV genotypes A–H. Shown are the IgG levels to the synthetic peptides representing the NTCP attachment site of HBV genotypes A–H measured by (<b>a</b>) ELISA (OD) and (<b>b</b>) micro-array technology (fluorescence intensity, FI).</p>
Full article ">Figure 5
<p>VVX001-induced antibodies react to SVPs of different HBV genotypes. Shown are IgG levels (OD) to SVPs of HBV genotypes A2, B2, C2, D1, E, F4, H (LHBs, MHBs, SHBs), and D3 (SHBs only) in the serum samples of the study subject before and after immunization as well as in two subjects successfully vaccinated with conventional HBsAg-based vaccines.</p>
Full article ">Figure 6
<p>IC<sub>50</sub> determination. IC<sub>50</sub> values for serum dilutions determined based on the HBeAg results are presented with a 95% confidence interval for (<b>a</b>) the anti-preS-positive human serum obtained in April 2019 and (<b>b</b>) for the anti-S-positive human serum (conventional vaccine, 2600 IU/L anti-HBs).</p>
Full article ">Figure 6 Cont.
<p>IC<sub>50</sub> determination. IC<sub>50</sub> values for serum dilutions determined based on the HBeAg results are presented with a 95% confidence interval for (<b>a</b>) the anti-preS-positive human serum obtained in April 2019 and (<b>b</b>) for the anti-S-positive human serum (conventional vaccine, 2600 IU/L anti-HBs).</p>
Full article ">Figure 7
<p>PreS-specific T cell responses after vaccination with VVX001. Shown are the percentages of proliferated CD3<sup>+</sup>CD4<sup>+</sup> and CD3<sup>+</sup>CD8<sup>+</sup> cells upon stimulation with (<b>a</b>) preS over time, (<b>b</b>) preS, P1-P8, equimolar mix of P1–P8, and (<b>c</b>) peptides representing the NTCP attachment site of genotypes A–H at the time point 4 months after the last injection (26 June 2019).</p>
Full article ">Figure 7 Cont.
<p>PreS-specific T cell responses after vaccination with VVX001. Shown are the percentages of proliferated CD3<sup>+</sup>CD4<sup>+</sup> and CD3<sup>+</sup>CD8<sup>+</sup> cells upon stimulation with (<b>a</b>) preS over time, (<b>b</b>) preS, P1-P8, equimolar mix of P1–P8, and (<b>c</b>) peptides representing the NTCP attachment site of genotypes A–H at the time point 4 months after the last injection (26 June 2019).</p>
Full article ">
14 pages, 5239 KiB  
Article
Unveiling the Mechanism of Compound Ku-Shen Injection in Liver Cancer Treatment through an Ingredient–Target Network Analysis
by Wenkui Zou, Jiazhen Liu, Zexing Wei, Chunhua Peng, Ying Zhao, Yue Ding, Jifan Shi and Juan Zhao
Genes 2024, 15(10), 1278; https://doi.org/10.3390/genes15101278 - 29 Sep 2024
Viewed by 444
Abstract
Background: Compound Ku-Shen Injection (CKI) is a traditional Chinese medicine preparation derived from Ku-Shen and Bai-Tu-Ling, commonly used in the adjunctive treatment of advanced cancers, including liver cancer. However, the underlying mechanisms of CKI’s effectiveness in cancer treatment are not well defined. Methods: [...] Read more.
Background: Compound Ku-Shen Injection (CKI) is a traditional Chinese medicine preparation derived from Ku-Shen and Bai-Tu-Ling, commonly used in the adjunctive treatment of advanced cancers, including liver cancer. However, the underlying mechanisms of CKI’s effectiveness in cancer treatment are not well defined. Methods: This study employs network pharmacology to investigate the traditional Chinese medicine (TCM) compatibility theory underlying CKI’s action in treating liver cancer, with findings substantiated by molecular docking and in vitro experiments. Sixteen active components were identified from CKI, along with 193 potential targets for treating liver cancer. Key therapeutic target proteins, including EGFR and ESR1, were also identified. KEGG enrichment results showed that the neuroactive ligand–receptor interaction, cAMP signaling pathway, and serotonergic synapses make up the key pathway of CKI in the treatment of liver cancer. Molecular docking results confirmed that the key active ingredients effectively bind to the core targets. CCK-8 cytotoxic experiment results show that the CKI key components of oxymatrine and matrine can inhibit the growth of HepG2 liver cancer cell proliferation. A Western blot analysis revealed that oxymatrine suppresses the expression of EGFR, contributing to its therapeutic efficacy against liver cancer. Conclusion: our study elucidated the therapeutic mechanism of CKI in treating liver cancer and unveiled the underlying principles of its TCM compatibility through its mode of action. Full article
(This article belongs to the Section Human Genomics and Genetic Diseases)
Show Figures

Figure 1

Figure 1
<p>(<b>A</b>) The Venn diagram of overlapping target genes of CKI and liver cancer. (<b>B</b>) The ingredient–target network of CKI, where the blue rectangle is CKI, the circular nodes represent the active components of CKI, and the prism nodes denote the potential targets of CKI.</p>
Full article ">Figure 2
<p>Protein–protein interaction network of CKI.</p>
Full article ">Figure 3
<p>(<b>A</b>) Bar graph of GO enrichment results, organized vertically from top to bottom: MF, CC, BP. (<b>B</b>) KEGG pathway enrichment results.</p>
Full article ">Figure 4
<p>Molecular docking profiles of active ingredients and core targets.</p>
Full article ">Figure 5
<p>(<b>A</b>,<b>B</b>) Effects of matrine and oxymatrine on the proliferation of HepG2 cells. (<b>C</b>,<b>D</b>) The effect of oxymatrine on EGFR expression in HepG2 cells. All the results are shown as mean ± SD (n = 3), *** <span class="html-italic">p</span> &lt; 0.001.</p>
Full article ">
19 pages, 6676 KiB  
Article
Isorhamnetin in Quinoa Whole-Grain Flavonoids Intervenes in Non-Alcoholic Fatty Liver Disease by Modulating Bile Acid Metabolism through Regulation of FXR Expression
by Xiaoqin La, Zhaoyan Zhang, Cunli Dong, Hanqing Li, Xiaoting He, Yurui Kang, Changxin Wu and Zhuoyu Li
Foods 2024, 13(19), 3076; https://doi.org/10.3390/foods13193076 - 26 Sep 2024
Viewed by 491
Abstract
Non-alcoholic fatty liver disease (NAFLD) is a severe hepatic health threat with no effective treatment. Based on the results that Chenopodium quinoa Willd. flavonoids eluted with 30% ethanol (CQWF30) can effectively alleviate NAFLD, this study employed ultrahigh-performance liquid chromatography–electrospray ionization–tandem mass spectrometry (UPLC-ESI-MS/MS) [...] Read more.
Non-alcoholic fatty liver disease (NAFLD) is a severe hepatic health threat with no effective treatment. Based on the results that Chenopodium quinoa Willd. flavonoids eluted with 30% ethanol (CQWF30) can effectively alleviate NAFLD, this study employed ultrahigh-performance liquid chromatography–electrospray ionization–tandem mass spectrometry (UPLC-ESI-MS/MS) to analyze the components of CQWF30., and screened for flavonoids with potential NAFLD-mitigating effects through network pharmacology. In vitro models using HepG2 and BEL-7402 cell lines induced with free fatty acid (FFA) showed that isorhamnetin administration reduced intracellular lipid deposition and reversed elevated triglyceride (TG) and total cholesterol (T-CHO) levels. In vivo experiments in high-fat diet (HFD) mice demonstrated that isorhamnetin significantly lowered serum and liver fat content, mitigated liver damage, and modulated bile acid metabolism by upregulating FXR and BSEP and downregulating SLCO1B3. Consequently, isorhamnetin shows promise as a treatment for NAFLD due to its lipid-lowering and hepatoprotective activities. Full article
(This article belongs to the Section Grain)
Show Figures

Graphical abstract

Graphical abstract
Full article ">Figure 1
<p>Identification of CQWF30 components and screening of components relevant to NAFLD intervention. Venn diagram illustrating the potential targets of CQWF30 in NAFLD intervention (<b>A</b>). The protein-protein interaction (PPI) network of 53 potential targets modulated by 13 flavonoids in CQWF30 in the context of NAFLD (<b>B</b>). Gene Ontology (GO) enrichment analysis of biological processes for the potential targets (<b>C</b>). GO enrichment analysis of cellular components for the potential targets (<b>D</b>). GO enrichment analysis of molecular functions for the potential targets (<b>E</b>). KEGG pathway enrichment analysis for the potential targets (<b>F</b>).</p>
Full article ">Figure 2
<p>Effects of isorhamnetin on lipid accumulation and cell viability. Effects of kaempferol (<b>A</b>), quercetin (<b>B</b>), and isorhamnetin (<b>C</b>) on the viability of HepG2 cells and effects of kaempferol (<b>D</b>), quercetin (<b>E</b>), and isorhamnetin (<b>F</b>) on the viability of BEL-7402 cells. Oil red O staining and Nile red fluorescence staining showing the effect of 10 μM kaempferol, quercetin, and isorhamnetin on lipid droplets in FFA-induced (at a volume ratio of OA to PA 2:1, 1 mM for HepG2) HepG2 cells (<b>G</b>–<b>I</b>,<b>M</b>–<b>O</b>). Oil red O staining and Nile red fluorescence staining showing the effect of 10 μM kaempferol, quercetin, and isorhamnetin on lipid droplets in FFA-induced (0.5 mM for BEL-7402) BEL-7402 cells (<b>J</b>–<b>L</b>,<b>P</b>–<b>R</b>). The scale bar for Oil red O staining is 50 μm and 20 μm for Nile red fluorescence staining. T-CHO and TG content in HepG2 cells (<b>S</b>,<b>T</b>). T-CHO and TG content in BEL-7402 cells (<b>U</b>,<b>V</b>). The images below are enlarged views of the red or yellow frames in the upper image. * <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 statistically significant.</p>
Full article ">Figure 3
<p>Isorhamnetin reduces lipid accumulation in NAFLD mice. Schematic representation of the animal experiment design and grouping (<b>A</b>). Representative morphological images of mice from each group (<b>B</b>). Growth curve of body weight (<b>C</b>). Body weight gain of mice in each group (<b>D</b>). Average daily food intake (<b>E</b>). Representative images of epididymal and inguinal white adipose tissues (<b>F</b>). Weight of inguinal white adipose tissue (<b>G</b>). Weight of epididymal white adipose tissue (<b>H</b>). In images (<b>G</b>) and (<b>H</b>), red indicates the Control group, blue indicates the HFD group, yellow indicates the HFD + Isorhamnetin group, and green indicates the HFD + Simvastatin group. ** <span class="html-italic">p</span> &lt; 0.01, *** <span class="html-italic">p</span> &lt; 0.001.</p>
Full article ">Figure 4
<p>Isorhamnetin alleviates liver morphology and injury and regulates lipid levels in the liver and serum of mice. Representative histological images of liver tissues (<b>A</b>). Liver weight (<b>B</b>). Liver index (liver-to-body weight ratio) (<b>C</b>). H&amp;E staining of liver tissues (<b>D</b>). Serum T-CHO (<b>E</b>), TG (<b>F</b>), HDL-C (<b>G</b>), LDL-C (<b>H</b>), AST (<b>I</b>), and ALT (<b>J</b>) levels, and AST/ALT ratio (<b>K</b>) in serum. Liver T-CHO (<b>L</b>), TG (<b>M</b>), HDL-C (<b>N</b>), LDL-C (<b>O</b>), AST (<b>P</b>), and ALT (<b>Q</b>) levels, and AST/ALT ratio (<b>R</b>) in liver tissues. TG content in feces (<b>S</b>). * <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.</p>
Full article ">Figure 5
<p>Isorhamnetin mitigates NAFLD by reducing bile acid stasis. Effect of isoquercitrin on TBA in serum (<b>A</b>), liver (<b>B</b>), and feces (<b>C</b>) of C57BL/6N mice. Two-dimensional principal component analysis–discriminant analysis (PCA-DA) score plot (<b>D</b>). Heatmap of 20 BAs with a significant contribution to differentiation, where green indicates low Z-scores and red indicates high Z-scores (<b>E</b>). Concentrations of various BAs such as Tβ-MCA (<b>M</b>), CDCA (<b>Q</b>), and UDCA (<b>W</b>) in the liver (<b>F</b>–<b>Y</b>). * <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.</p>
Full article ">Figure 6
<p>Isorhamnetin relieves NAFLD by increasing FXR protein expression and regulating downstream proteins. Expression of FXR, BSEP, and SLCO1B3 proteins in HepG2 cells (<b>A</b>) and FXR, BSEP, SLCO1B3 proteins expressed in BEL-7402 cells (<b>B</b>). Expression of FXR, BSEP, and SLCO1B3 proteins in mouse liver (<b>C</b>). Quantification of protein expression for FXR (<b>D</b>), BSEP (<b>E</b>), and SLCO1B3 (<b>F</b>) * <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.</p>
Full article ">
19 pages, 4115 KiB  
Article
The Antioxidant and Anti-Fatigue Effects of Rare Ginsenosides and γ-Aminobutyric Acid in Fermented Ginseng and Germinated Brown Rice Puree
by Shiwen Feng, Tao Li, Xinrui Wei, Yifei Zheng, Yumeng Zhang, Gao Li and Yuqing Zhao
Int. J. Mol. Sci. 2024, 25(19), 10359; https://doi.org/10.3390/ijms251910359 - 26 Sep 2024
Viewed by 403
Abstract
γ-aminobutyric acid (GABA) and rare ginsenosides are good antioxidant and anti-fatigue active components that can be enriched via probiotic fermentation. In this study, ginseng and germinated brown rice were used as raw materials to produce six fermented purees using fermentation and non-fermentation technology. [...] Read more.
γ-aminobutyric acid (GABA) and rare ginsenosides are good antioxidant and anti-fatigue active components that can be enriched via probiotic fermentation. In this study, ginseng and germinated brown rice were used as raw materials to produce six fermented purees using fermentation and non-fermentation technology. We tested the chemical composition of the purees and found that the content of GABA and rare ginsenoside (Rh4, Rg3, and CK) in the puree made of ginseng and germinated brown rice (FGB) increased significantly after fermentation. The antioxidant activity of the six purees was determined using cell-free experiments, and it was found that FGB had better ferric-ion-reducing antioxidant power (FRAP) and 1,1-diphenyl-2-picryl-hydrazyl (DPPH) free radical scavenging rates, exhibiting better antioxidant effects. We then evaluated the antioxidant effect of FGB in HepG2 cells induced by H2O2 and found that FGB can reduce the generation of reactive oxygen species (ROS) in HepG2 cells and increase the membrane potential level, thereby improving oxidative damage in these cells. In vivo experiments also showed that FGB has good antioxidant and anti-fatigue activities, which can prolong the exhaustive swimming time of mice and reduce the accumulation of metabolites, and is accompanied by a corresponding increase in liver glycogen and muscle glycogen levels as well as superoxide dismutase and lactate dehydrogenase activities. Finally, we believe that the substances with good antioxidant and anti-fatigue activity found in FGB are derived from co-fermented enriched GABA and rare ginsenosides. Full article
(This article belongs to the Special Issue Effects of Functional Food Components in Health and Disease)
Show Figures

Figure 1

Figure 1
<p>Transformation pathway of ginsenoside in FGB.</p>
Full article ">Figure 2
<p>Inhibitory activity of FGB and H<sub>2</sub>O<sub>2</sub> on proliferation of HepG<sub>2</sub> cells (24 h). (<b>a</b>). determination of activity of FGB on HepG2 cells; (<b>b</b>). H<sub>2</sub>0<sub>2</sub> activity determination of HepG2 cells.</p>
Full article ">Figure 3
<p>Effects of FGB and H<sub>2</sub>O<sub>2</sub> on ROS release in HepG<sub>2</sub> cells. Data are presented as mean ± SD (<span class="html-italic">n</span> = 3). Different letters on the bar are significantly different (<span class="html-italic">p</span> &lt; 0.05).</p>
Full article ">Figure 4
<p>Effects of FGB and H<sub>2</sub>O<sub>2</sub> on mitochondrial membrane potential of HepG<sub>2</sub> cells. Data are presented as mean ± SD (<span class="html-italic">n</span> = 3). Different letters on the bar are significantly different (<span class="html-italic">p</span> &lt; 0.05).</p>
Full article ">Figure 5
<p>Effects of FGB on H<sub>2</sub>O<sub>2</sub>-induced acridine orange in HepG<sub>2</sub> cells (×200).</p>
Full article ">Figure 6
<p>Effects of different purees on body weight (<b>A</b>) and food intake (<b>B</b>) of mice. Data are presented as mean ± SD (<span class="html-italic">n</span> = 10). <span class="html-italic">p</span> &lt; 0.05 indicates significant difference.</p>
Full article ">Figure 7
<p>Effects of different purees on the weight-loaded forced swimming time in mice. Data are presented as mean ± SD (<span class="html-italic">n</span> = 10). Different letters on the bar are significantly different (<span class="html-italic">p</span> &lt; 0.05).</p>
Full article ">Figure 8
<p>Effects of different purees on blood urea nitrogen (<b>A</b>) and creatinine (<b>B</b>) in mice. Data are presented as mean ± SD (<span class="html-italic">n</span> = 10). Different letters on the bar are significantly different (<span class="html-italic">p</span> &lt; 0.05).</p>
Full article ">Figure 9
<p>Effects of different purees on blood lactic acid (<b>A</b>) and lactic dehydrogenase (<b>B</b>) in mice. Data are presented as mean ± SD (<span class="html-italic">n</span> = 10). Different letters on the bar are significantly different (<span class="html-italic">p</span> &lt; 0.05).</p>
Full article ">Figure 10
<p>Effect of different purees on glycogen accumulation of liver (<b>A</b>) and muscle (<b>B</b>) in exhausted mice. Data are presented as mean ± SD (<span class="html-italic">n</span> = 10). Different letters on the bar are significantly different (<span class="html-italic">p</span> &lt; 0.05).</p>
Full article ">Figure 11
<p>Effect of different purees on malondialdehyde (<b>A</b>) and superoxide dismutase (<b>B</b>) in mice. Data are presented as mean ± SD (<span class="html-italic">n</span> = 10). Different letters on the bar are significantly different (<span class="html-italic">p</span> &lt; 0.05).</p>
Full article ">Figure 12
<p>Schematic representation of puree processing and fermentation.</p>
Full article ">
18 pages, 5942 KiB  
Article
An Efficient Fabrication Approach for Multi-Cancer Responsive Chemoimmuno Co-Delivery Nanoparticles
by Jianxi Huang, Yu-Ting Chien, Qingxin Mu and Miqin Zhang
Pharmaceutics 2024, 16(10), 1246; https://doi.org/10.3390/pharmaceutics16101246 - 25 Sep 2024
Viewed by 557
Abstract
Background/Objectives: Cancer remains one of the leading causes of death, with breast, liver, and pancreatic cancers significantly contributing to this burden. Traditional treatments face issues including dose-limiting toxicity, drug resistance, and limited efficacy. Combining therapeutic agents can enhance effectiveness and reduce toxicity, but [...] Read more.
Background/Objectives: Cancer remains one of the leading causes of death, with breast, liver, and pancreatic cancers significantly contributing to this burden. Traditional treatments face issues including dose-limiting toxicity, drug resistance, and limited efficacy. Combining therapeutic agents can enhance effectiveness and reduce toxicity, but separate administration often leads to inefficiencies due to differing pharmacokinetics and biodistribution. Co-formulating hydrophobic chemotherapeutics such as paclitaxel (PTX) and hydrophilic immunologic agents such as polyinosinic-polycytidylic acid (Poly IC) is particularly challenging due to their distinct physicochemical properties. This study presents a novel and efficient approach for the co-delivery of PTX and Poly IC using chitosan-based nanoparticles. Method: Chitosan-PEG (CP) nanoparticles were developed to encapsulate both PTX and Poly IC, overcoming their differing physicochemical properties and enhancing therapeutic efficacy. Results: With an average size of ~100 nm, these nanoparticles facilitate efficient cellular uptake and stability. In vitro results showed that CP-PTX-Poly IC nanoparticles significantly reduced cancer cell viability in breast (4T1), liver (HepG2), and pancreatic (Pan02) cancer types, while also enhancing dendritic cell (DC) maturation. Conclusions: This dual-modal delivery system effectively combines chemotherapy and immunotherapy, offering a promising solution for more effective cancer treatment and improved outcomes. Full article
(This article belongs to the Special Issue Combination Therapeutic Delivery Systems)
Show Figures

Figure 1

Figure 1
<p>CP-PTX-Poly IC nanoparticle that co-delivers PTX and Poly IC to cancer cells and dendritic cells for combined chemo-immunotherapy. (<b>a</b>) Schematic illustration of nanoparticle synthesis. PTX was covalently conjugated on CP polymers that contain chitosan and PEG, forming CP-PTX (intermediate (1)). The CP-PTX was incubated with Poly IC (intermediate (2)) before albumin stabilization into CP-PTX-Poly IC nanoparticles (product (3)). (<b>b</b>) Potential mechanisms of combined chemo-immunotherapy for the co-delivery of PTX and Poly IC for direct tumor killing and immune-triggered indirect tumor killing in vivo. The latter includes the activation of the host’s immune system by NPs through dendritic cell (DC) maturation and the subsequent activation of anti-cancer innate and adaptive (activation of cytotoxic T cells) immune responses.</p>
Full article ">Figure 2
<p>Synthesis of CP-PTX. (<b>a</b>) Schematic representation of CP-PTX synthesis from CP and carboxylated PTX via EDC/NHS chemistry. (<b>b</b>) NMR spectra of CP, CP-PTX, and PTX.</p>
Full article ">Figure 3
<p>Physicochemical properties of CP-PTX-Poly IC nanoparticles. (<b>a</b>) Zeta potentials of CP-PTX (intermediate (1) in <a href="#pharmaceutics-16-01246-f001" class="html-fig">Figure 1</a>a) and a series of intermediate CP-PTX-Poly IC complexes (intermediate (2) in <a href="#pharmaceutics-16-01246-f001" class="html-fig">Figure 1</a>a) at different CP-PTX: Poly IC ratios (wt: wt% = 46:1, 18:1, 11:1, 6:1) at room temperature. (<b>b</b>) TEM micrographs of final CP-PTX-poly IC nanoparticles. Scale bar is 200 nm. (<b>c</b>) Size distribution of CP-PTX-Poly IC nanoparticles analyzed from TEM images. (<b>d</b>) Hydrodynamic size and (<b>e</b>) zeta potential of albumin-stabilized CP-PTX and CP-PTX-Poly IC nanoparticles (step (3) in <a href="#pharmaceutics-16-01246-f001" class="html-fig">Figure 1</a>a). Measurements were conducted at room temperature in water (pH ~7). (<b>f</b>) Hydrodynamic size of CP-PTX-Poly IC at 37 °C in water over 7 days.</p>
Full article ">Figure 4
<p>Cellular uptake of CP-PTX-Poly IC NPs in three cancer types. Confocal images of 4T1, HepG2, and Pan02 cells incubated for 2 h with Cy5-labeled CP-PTX-Poly IC. Untreated cells were used as a reference. Blue color represents cell nucleus and green color represents cell membrane. Red color is the signal of CP-PTX-Poly IC. Scale bar is 20 μm.</p>
Full article ">Figure 5
<p>Cancer killing effect of CP-PTX-Poly IC. (<b>a</b>) Brightfield images and (<b>b</b>) Relative ATP level of 4T1 (Breast), HepG2 (Liver), and Pan02 (Pancreatic) cells treated with Poly IC, CP, PTX, CP-PTX, and CP-PTX-Poly IC for 48 h. Scale bar is 50 μm. The dose of each treatment was the same amount of the corresponding components in CP-PTX-Poly IC at 2 µM PTX (Poly IC = 425 μg/mL). * <span class="html-italic">p</span> &lt; 0.05, **** <span class="html-italic">p</span> &lt; 0.0001, ns = statistically insignificant.</p>
Full article ">Figure 6
<p>Flow cytometry study of DC maturation. (<b>a</b>) CD80 and (<b>b</b>) CD86 expressions in BMDCs were evaluated after treatment of Poly IC, CP-PTX, or CP-PTX-Poly IC for 24 h. The percentages of CD80 positive or CD86 positive populations are shown in each subfigure. The mean fluorescence intensity of (<b>c</b>) CD80 and (<b>d</b>) CD86 signals in BMDC were calculated from the treatment groups in (<b>a</b>,<b>b</b>). The dose of each treatment was the same amount of the corresponding components in CP-PTX-Poly IC at 10 μg/mL Poly IC (PTX = 850 μM).</p>
Full article ">Figure 7
<p>Biocompatibility of Poly IC, CP-PTX, and CP-PTX-Poly IC on BMDCs. (<b>a</b>) Brightfield images and (<b>b</b>) Relative ATP levels of untreated BMDCs or BMDCs treated by Poly IC, CP-PTX, and CP-PTX-Poly IC for 24 h. The dose of each treatment had the same amount of the corresponding components in CP-PTX-Poly IC at 10 μg/mL Poly IC (PTX = 850 μM). Scale bar is 50 μm.</p>
Full article ">
Back to TopTop