[go: up one dir, main page]
More Web Proxy on the site http://driver.im/
You seem to have javascript disabled. Please note that many of the page functionalities won't work as expected without javascript enabled.
 
 
Sign in to use this feature.

Years

Between: -

Subjects

remove_circle_outline
remove_circle_outline
remove_circle_outline
remove_circle_outline
remove_circle_outline
remove_circle_outline
remove_circle_outline
remove_circle_outline
remove_circle_outline

Journals

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

Article Types

Countries / Regions

remove_circle_outline
remove_circle_outline
remove_circle_outline
remove_circle_outline

Search Results (1,857)

Search Parameters:
Keywords = cyclooxygenase-2

Order results
Result details
Results per page
Select all
Export citation of selected articles as:
18 pages, 3444 KiB  
Review
Search for New Compounds with Anti-Inflammatory Activity Among 1,2,4-Triazole Derivatives
by Teresa Glomb, Julia Minta, Michalina Nowosadko, Julia Radzikowska and Piotr Świątek
Molecules 2024, 29(24), 6036; https://doi.org/10.3390/molecules29246036 - 21 Dec 2024
Viewed by 275
Abstract
Compounds containing the 1,2,4-triazole moiety in their structure exhibit broad biological activities. Many of these compounds demonstrate anti-inflammatory activity in vitro through various mechanisms, such as inhibiting COX-1/COX-2 and LOX, modulating pro-inflammatory cytokine levels, or having effects on other specific enzymes. Some also [...] Read more.
Compounds containing the 1,2,4-triazole moiety in their structure exhibit broad biological activities. Many of these compounds demonstrate anti-inflammatory activity in vitro through various mechanisms, such as inhibiting COX-1/COX-2 and LOX, modulating pro-inflammatory cytokine levels, or having effects on other specific enzymes. Some also display activities in vivo. In many publications, the activities of new 1,2,4-triazole-based compounds exceed those of the reference drugs, suggesting their promising potential as new therapeutic agents. This review of active 1,2,4-triazole derivatives with anti-inflammatory activity is based on literature published from 2015–2024. Full article
Show Figures

Figure 1

Figure 1
<p>Examples of known drugs containing a 1,2,4-triazole ring.</p>
Full article ">Figure 2
<p>1,2,4-triazole derivatives as COX inhibitors (part 1).</p>
Full article ">Figure 3
<p>1,2,4-triazole derivatives as COX inhibitors (part 2).</p>
Full article ">Figure 4
<p>1,2,4-triazole derivatives as COX inhibitors (part 3).</p>
Full article ">Figure 5
<p>1,2,4-triazole derivatives as COX inhibitors with sulfamoylphenyl or methylsulfonylphenyl substituents.</p>
Full article ">Figure 6
<p>1,2,4-triazole derivatives as 15-LOX inhibitors (part 1).</p>
Full article ">Figure 7
<p>1,2,4-triazole derivatives as 15-LOX inhibitors (part 2).</p>
Full article ">Figure 8
<p>1,2,4-triazole derivatives as FLAP inhibitors.</p>
Full article ">Figure 9
<p>1,2,4-triazole derivatives as COX-2/5-LOX inhibitors (part 1).</p>
Full article ">Figure 10
<p>1,2,4-triazole derivatives as COX-2/5-LOX inhibitors (part 2).</p>
Full article ">Figure 11
<p>1,2,4-triazole derivatives with confirmed anti-inflammatory activity in vivo (part 1).</p>
Full article ">Figure 12
<p>1,2,4-triazole derivatives with confirmed anti-inflammatory activity in vivo (part 2).</p>
Full article ">Figure 13
<p>1,2,4-triazole derivatives with confirmed anti-inflammatory activity in vivo (part 3).</p>
Full article ">Figure 14
<p>1,2,4-triazole derivatives with miscellaneous mechanisms for anti-inflammatory activity (part 1).</p>
Full article ">Figure 15
<p>1,2,4-triazole derivatives with miscellaneous mechanisms for anti-inflammatory activity (part 2).</p>
Full article ">
14 pages, 5656 KiB  
Article
Celecoxib Combined with Tocilizumab Has Anti-Inflammatory Effects and Promotes the Recovery of Damaged Cartilage via the Nrf2/HO-1 Pathway In Vitro
by Miyako Shimasaki, Shusuke Ueda, Masaru Sakurai, Norio Kawahara, Yoshimichi Ueda and Toru Ichiseki
Biomolecules 2024, 14(12), 1636; https://doi.org/10.3390/biom14121636 - 20 Dec 2024
Viewed by 291
Abstract
Inflammation and oxidative stress are crucial for osteoarthritis (OA) pathogenesis. Despite the potential of pharmacological pretreatment of chondrocytes in preventing OA, its efficacy in preventing the progression of cartilage damage and promoting its recovery has not been examined. In this study, an H [...] Read more.
Inflammation and oxidative stress are crucial for osteoarthritis (OA) pathogenesis. Despite the potential of pharmacological pretreatment of chondrocytes in preventing OA, its efficacy in preventing the progression of cartilage damage and promoting its recovery has not been examined. In this study, an H2O2-induced human OA-like chondrocyte cell model was created using H1467 primary human chondrocytes to evaluate the efficacy of interleukin (IL)-6 and cyclooxygenase (COX)-2 inhibitors (tocilizumab and celecoxib, respectively) in the prevention and treatment of cartilage damage. H2O2 significantly elevated the IL-6, COX-2, and matrix metalloproteinase (MMP)-13 levels. Although monotherapy decreased the levels, nuclear shrinkage and altered cell morphology, similar to those in the H2O2 group, were observed. The expression of these factors was significantly lower in the combination therapy group, and the cell morphology was maintained. Moreover, the nuclear factor erythroid 2-related factor 2 (Nrf2) pathway was activated, and levels of the antioxidant protein heme oxygenase-1 (HO-1) were increased, especially in the combination group, indicating an anti-inflammatory effect. The treatment groups, particularly the combination group, demonstrated increased cell viability. Overall, the drug combination exhibited superior efficacy in preventing the progression of cartilage damage and promoted its recovery compared with the monotherapy. Given that the drugs herein are already in clinical use, they are suitable candidates for OA treatment. Full article
(This article belongs to the Section Cellular Biochemistry)
Show Figures

Figure 1

Figure 1
<p>Schematic of the experimental platform. (<b>a</b>) Assessment of the preventive potential of mono- and combination therapy on H<sub>2</sub>O<sub>2</sub>-induced chondrocyte damage. (<b>b</b>) Assessment of therapeutic potential of mono- and combination therapy on H<sub>2</sub>O<sub>2</sub>-induced chondrocyte damage. (For details, see <a href="#sec2-biomolecules-14-01636" class="html-sec">Section 2</a>).</p>
Full article ">Figure 2
<p>Fluorescence immunostaining of inflammatory (IL-6, COX-2, and MMP-13) and anti-inflammatory (Nrf2 and HO-1) factors in the human osteoarthritis-like chondrocyte model (human chondrocytes stimulated with H<sub>2</sub>O<sub>2</sub> for 30 min or 2 h). Representative images showing IL-6, COX-2, and MMP-13 expression in red, Nrf2, and HO-1 expression in green, and DAPI-stained nuclei in blue. Scale bar, 200 μm. After stimulation of chondrocytes with H<sub>2</sub>O<sub>2</sub> for 2 h, the expression of IL-6, Cox-2, and MMP-13 was increased; Nrf2 was expressed in the cytoplasm in controls and was translocated into the nucleus after 2-h H<sub>2</sub>O<sub>2</sub> stimulation. HO-1 expression was decreased in the treated groups. (<b>a</b>) IL-6, Cox-2, MMP-13, HO-1; Low magnification. (<b>b</b>) Quantification of IL-6, COX-2, MMP-13, and HO-1 in chondrocytes was conducted by calculating the ratio of the number of cells positive for the protein expression to the total number of cells. Expression of IL-6, Cox-2, and MMP-13 were both significantly increased after 30 min and 2 h stimulation with H<sub>2</sub>O<sub>2</sub> compared to control (*** <span class="html-italic">p</span> &lt; 0.001). Compared with the 30 min stimulation with H<sub>2</sub>O<sub>2</sub>, there was a significant increase at 2 h (IL-6, COX-2; ** <span class="html-italic">p</span> &lt; 0.01; MMP-13, HO-1; *** <span class="html-italic">p</span> &lt; 0.001). (<b>c</b>) Nrf2; representative images in low magnification (Scale bar, 200 μm) and high magnification (Scale bar, 20 μm) are shown. (<b>d</b>) Quantification of the number of cells demonstrating nuclear migration of Nrf2. Nuclear translocation of Nrf2 was significantly increased at both 30 min and 2 h stimulated with H<sub>2</sub>O<sub>2</sub> compared to control (*** <span class="html-italic">p</span> &lt; 0.001). Comparing 30 min to 2 h showed a significant increase at 2 h (* <span class="html-italic">p</span> &lt; 0.05).</p>
Full article ">Figure 3
<p>Fluorescence immunostaining of inflammatory factors (IL-6 and COX-2) in the human osteoarthritis chondrocyte injury model (human chondrocytes stimulated with H<sub>2</sub>O<sub>2</sub> for 2 h) after pretreatment with celecoxib or tocilizumab alone or in combination for 24 h. Control: control group. Celecoxib+/H<sub>2</sub>O<sub>2</sub>-, Tocilizumab+/H<sub>2</sub>O<sub>2</sub>-, Celecoxib+/Tocilizumab+/H<sub>2</sub>O<sub>2</sub>-; Human chondrocytes pretreated with inhibitors alone or in combination for 24 h. Inhibitor-/H<sub>2</sub>O<sub>2</sub>+; Inflammation group, in which human chondrocytes were stimulated with H<sub>2</sub>O<sub>2</sub> for 2 h. Celecoxib+/H<sub>2</sub>O<sub>2</sub>+, Tocilizumab+/H<sub>2</sub>O<sub>2</sub>+, Celecoxib+/Tocilizumab+/H<sub>2</sub>O<sub>2</sub>+; Prevention group, in which human chondrocytes were treated with inhibitors alone or in combination for 24 h, followed by stimulation with H<sub>2</sub>O<sub>2</sub> for 2 h. (<b>a</b>) Representative images showing IL-6 and COX-2 expression in red and DAPI-stained nuclei in blue. Scale bar, 200 µm (Low magnification). (<b>b</b>) Quantification of IL-6 and COX-2 in chondrocytes was conducted by calculating the ratio of the number of cells positive for the protein expression to the total number of cells. The expression of IL-6 and COX-2 was upregulated and downregulated, respectively, in the inflammation and prevention groups. The expression of IL-6 and COX-2 was significantly diminished in all prevention groups in comparison to the inflammation group (*** <span class="html-italic">p</span> &lt; 0.001). Comparing inhibitors alone to the combination showed a significant reduction in the combined (IL-6; *** <span class="html-italic">p</span> &lt; 0.001 for C or T vs. C+T; COX-2; * <span class="html-italic">p</span> &lt; 0.05 for C vs. C+T, *** <span class="html-italic">p</span> &lt; 0.001 for T vs. C+T).</p>
Full article ">Figure 4
<p>Fluorescence immunostaining and gene expression analysis of inflammatory factors (IL-6, COX-2, and MMP-13) in the human osteoarthritis chondrocyte injury model treated with celecoxib or tocilizumab alone or in combination. Control: control group. Inhibitor-/H<sub>2</sub>O<sub>2</sub>+: Inflammation group, in which human chondrocytes were stimulated with H<sub>2</sub>O<sub>2</sub> for 2 h. H<sub>2</sub>O<sub>2</sub>+/Celecoxib+, H<sub>2</sub>O<sub>2</sub>+/Tocilizumab+, H<sub>2</sub>O<sub>2</sub>+/Celecoxib+/Tocilizumab+; Inhibitor-/H<sub>2</sub>O<sub>2</sub>+/Celecoxib+/Tocilizumab+; Inflammation group subjected to monotherapy or combination therapy with the inhibitors for 24 h. Representative images showing IL-6, COX-2, and MMP-13 expression in red and DAPI-stained nuclei in blue. (<b>a</b>) Low magnification (Scale bar, 200 µm). (<b>b</b>) Quantification of IL-6, COX-2, and MMP-13 in chondrocytes was conducted by calculating the ratio of the number of cells positive for the protein expression to the total number of cells. The expression of IL-6, COX-2, and MMP-13 was observed to be significantly decreased in all treatment groups that included inhibitors in comparison to the inflammation group that was stimulated with H<sub>2</sub>O<sub>2</sub> (*** <span class="html-italic">p</span> &lt; 0.001). A significant reduction was observed with the combination when compared to the inhibitors alone, with MMP-13 exhibiting the most notable decline (*** <span class="html-italic">p</span> &lt; 0.001). (<b>c</b>) High magnification (Scale bar, 20 µm). (<b>d</b>) Gene expression analysis in inflammation and treatment groups. <span class="html-italic">MMP-13</span> expression was upregulated in the inflammation group and was significantly suppressed in the combination treatment with celecoxib (C) and tocilizumab (T): * <span class="html-italic">p</span> &lt; 0.05 for H<sub>2</sub>O<sub>2</sub> vs. C+T.</p>
Full article ">Figure 5
<p>Fluorescence immunostaining and gene expression analysis of anti-inflammatory factors (Nrf2 and HO-1) in the human osteoarthritis chondrocyte injury model treated with celecoxib or tocilizumab alone or in combination. Control: control group. Inhibitor-/H<sub>2</sub>O<sub>2</sub>+: Inflammation group, in which human chondrocytes were stimulated with H<sub>2</sub>O<sub>2</sub> for 2 h. H<sub>2</sub>O<sub>2</sub>+/Celecoxib+, H<sub>2</sub>O<sub>2</sub>+/Tocilizumab+, H<sub>2</sub>O<sub>2</sub>+/Celecoxib+/Tocilizumab+; Inhibitor-/H<sub>2</sub>O<sub>2</sub>+/Celecoxib+/Tocilizumab+; Inflammation group subjected to monotherapy or combination therapy with the inhibitors for 24 h. Representative fluorescence immunostaining images showing Nrf2 and HO-1 expression in green and DAPI-stained nuclei in blue. (<b>a</b>) Low magnification (Scale bar, 200 µm). (<b>b</b>) Quantification of HO-1 in chondrocytes was conducted by calculating the ratio of the number of cells positive for the protein to the total number of cells. Quantification of the number of cells demonstrating nuclear migration of Nrf2. The nuclear translocation of Nrf2 was significantly enhanced in the combined treatment group relative to the H<sub>2</sub>O<sub>2</sub> stimulation group (*** <span class="html-italic">p</span> &lt; 0.001). Nuclear translocation was also significantly greater with the combination than the inhibitors alone (*** <span class="html-italic">p</span> &lt; 0.001). The expression of HO-1 was observed to be significantly increased in all treatment groups that included inhibitors in comparison to the inflammation group that was stimulated with H<sub>2</sub>O<sub>2</sub> (*** <span class="html-italic">p</span> &lt; 0.001). A significant increase was observed with the combination when compared to the inhibitors alone (*** <span class="html-italic">p</span> &lt; 0.001). (<b>c</b>) High magnification (Scale bar, 20 µm). (<b>d</b>) Gene expression analysis in inflammation and treatment groups. <span class="html-italic">Nrf2</span> showed increased nuclear translocation in the treatment group compared with that in the inflammation group. In particular, an increasing trend was observed in treatment with a combination of celecoxib (C) and tocilizumab (T); <span class="html-italic">p</span> = 0.0623 for H<sub>2</sub>O<sub>2</sub> vs. C+T.</p>
Full article ">Figure 6
<p>Analysis of cell viability in the human osteoarthritic cartilage injury model. Compared to the inflammation group, there was an increase in cell proliferation in the treatment group at 24 and 48 h after treatment, which was particularly significant in the combination group: H<sub>2</sub>O<sub>2</sub> vs. C+T (24 h: ***, <span class="html-italic">p</span> &lt; 0.001; 48 h: **, <span class="html-italic">p</span> &lt; 0.01).</p>
Full article ">
26 pages, 10330 KiB  
Article
Anti-Inflammatory Properties of Novel 1,2-Benzothiazine Derivatives and Their Interaction with Phospholipid Model Membranes
by Berenika M. Szczęśniak-Sięga, Jadwiga Maniewska, Benita Wiatrak, Tomasz Janek, Paulina Nowotarska and Żaneta Czyżnikowska
Membranes 2024, 14(12), 274; https://doi.org/10.3390/membranes14120274 - 18 Dec 2024
Viewed by 394
Abstract
The design of novel anti-inflammatory drugs remains a critical area of research in the development of effective treatments for inflammatory diseases. In this study, a series of 1,2-benzothiazine was evaluated through a multifaceted approach. In particular, we investigated the potential interactions of the [...] Read more.
The design of novel anti-inflammatory drugs remains a critical area of research in the development of effective treatments for inflammatory diseases. In this study, a series of 1,2-benzothiazine was evaluated through a multifaceted approach. In particular, we investigated the potential interactions of the potential drugs with lipid bilayers, an important consideration for membrane permeability and overall pharmacokinetics. In addition, we evaluated their ability to inhibit cyclooxygenase 1 and cyclooxygenase 2 activity and selectivity using both a cyclooxygenase inhibition assay and molecular docking simulations. To evaluate their therapeutic potential, we performed in vitro assays to measure cytokine mRNA expression in inflamed cells. The antioxidant activity was evaluated using both in vitro assays, such as 2,2-diphenyl-1-picrylhydrazyl and 2,2-azino-bis-3-ethylbenzothiazoline-6-sulphonic acid scavenging, to determine the compounds’ capacity to neutralize free radicals and reduce oxidative stress. Theoretical calculations, including density functional theory, were used to predict the reactivity profiles of the compounds. Full article
Show Figures

Figure 1

Figure 1
<p>Structure of the model drug—meloxicam.</p>
Full article ">Figure 2
<p>Structures of a non-ulcerogenic 1,2-benzothiazine (<b>a</b>) and new designed arylpiperazine derivatives of 1,2-benzothiazine (<b>b</b>–<b>d</b>).</p>
Full article ">Figure 3
<p>The exemplary thermograms obtained for DPPC mixed with compounds: (<b>A</b>) <b>BS23</b> and (<b>B</b>) <b>BS24</b> as well as for pure lipid (the first curve from the top—black color): DPPC molar ratios from the top to the bottom: 0 (pure lipid), 0.06, 0.08, 0.10, and 0.12; the direction of the endothermic reaction is upwards.</p>
Full article ">Figure 4
<p>Influence of all studied compounds on the DPPC main transition temperature (<b>A</b>); influence of two studied compounds (<b>BS24</b> and <b>BS27</b>) and meloxicam on the DPPC main transition temperature (<b>B</b>).</p>
Full article ">Figure 5
<p>Influence of studied compounds on the DPPC peak half-width.</p>
Full article ">Figure 6
<p>Relative main transition temperature (T<sub>m</sub>/T<sub>m</sub><sup>0</sup>) for DPPC and DMPC in the presence of studied compounds at compound: phospholipid molar ratio 1:10; T<sub>m</sub><sup>0</sup>—the main transition temperature for pure lipid; T<sub>m</sub>—the main transition temperature in the presence of studied compounds at compound: phospholipid molar ratio 1:10; full symbols—DPPC, open symbols—DMPC.</p>
Full article ">Figure 7
<p>Anti-inflammatory effect after incubation of NHDF cells with lipopolysaccharide (LPS) and then the tested compounds. The black column presents the <b>BS23</b> compound and the white column the <b>BS28</b> compound. (<b>A</b>) MTT assay; (<b>B</b>) DCF-DA assay. Data presented as a mean and SD; * <span class="html-italic">p</span> &lt; 0.05—significant difference compared to the control (cells with only medium); # <span class="html-italic">p</span> &lt; 0.05 significant difference compared to the positive control (cells treated with only LPS).</p>
Full article ">Figure 8
<p>Relative expression of TNF-α (<b>left panel</b>) and IL-6 (<b>right panel</b>) in NHDF cells. (A) Control—only NHDF cells, (B) NHDF cells incubated with lipopolysaccharide (LPS), (C) NHDF cells incubated with lipopolysaccharide (LPS) treated with 10 µM of <b>BS23,</b> and (D) NHDF cells incubated with lipopolysaccharide (LPS) treated with 10 µM of <b>BS28</b>. The values represent mean ± SD from three experiments. * <span class="html-italic">p</span> &lt; 0.05—significant difference compared to the control (cells with only medium); # <span class="html-italic">p</span> &lt; 0.05 significant difference compared to the positive control (cells treated with only LPS).</p>
Full article ">Figure 9
<p>The binding mode of meloxicam (blue and top) and compound <b>BS23</b> (pink and bottom) in the active center of COX-2.</p>
Full article ">Scheme 1
<p>Synthesis of new 1,2-benzothiazine derivatives <b>BS23</b>–<b>BS30</b>.</p>
Full article ">
12 pages, 3011 KiB  
Article
Chondroprotective Effect of Campylaephora hypnaeoides Extract in Primary Chondrocytes and Rat OA Model
by Ji Yun Jang, Seul Ah Lee, Do Kyung Kim, Sook-Young Lee and Chun Sung Kim
Int. J. Mol. Sci. 2024, 25(24), 13391; https://doi.org/10.3390/ijms252413391 - 13 Dec 2024
Viewed by 340
Abstract
Campylaephora hypnaeoides (C. hypnaeoides) was extracted using fermented ethanol. The effect of fermented ethanol extract of C. hypnaeoides (FeCH) on chondrocyte viability was analyzed by 3-(4,5-dimethylthiazol-2-yl)-2,5-iphenyltetrazolium bromide assay, which showed no cytotoxicity at 2 mg/mL. FeCH pretreatment in IL-1β-stimulated chondrocytes significantly [...] Read more.
Campylaephora hypnaeoides (C. hypnaeoides) was extracted using fermented ethanol. The effect of fermented ethanol extract of C. hypnaeoides (FeCH) on chondrocyte viability was analyzed by 3-(4,5-dimethylthiazol-2-yl)-2,5-iphenyltetrazolium bromide assay, which showed no cytotoxicity at 2 mg/mL. FeCH pretreatment in IL-1β-stimulated chondrocytes significantly inhibited the accumulation of nitric oxide and prostaglandin E2, which was analyzed using the ELISA assay. In addition, protein expression levels of inflammatory-related factors, such as inducible nitric oxide synthase, cyclooxygenase-2, interleukin-6, tumor necrosis factor-alpha, and cartilage-degrading-related enzymes, such as matrix metalloproteinases-1, -3, and -13, and a disintegrin and metalloproteinase with thrombospondin motifs-4 and -5 were significantly decreased in IL-1β-stimulated chondrocytes pretreated with FeCH, which were analyzed using western blot analysis. In addition, as a result of analyzing the content of collagen type II (Col II) and proteoglycan through western blot analysis and alcian blue staining, FeCH pretreatment prevented the degradation of Col II and proteoglycan. It was analyzed through western blot analysis that the chondroprotective effect of FeCH may be mediated through MAPKs and NF-κB-signaling mechanisms. In an in vivo study, an osteoarthritis experimental animal model with damaged medial meniscus (DMM) was utilized and orally administered daily for 8 weeks after surgery. At the study end point, knee joints were harvested and subjected to histological analysis with safranin O staining. As a result, articular cartilage was significantly protected in the FeCH group compared to the DMM group. These results suggest FeCH as a candidate material for the development of pharmaceutical materials for the treatment or prevention of degenerative arthritis. Full article
(This article belongs to the Section Bioactives and Nutraceuticals)
Show Figures

Figure 1

Figure 1
<p>Effects of fermentation ethanol of <span class="html-italic">Campylaephora hypnaeoides</span> (FeCH) on primary rat chondrocyte viability. (<b>A</b>) primary rat chondrocytes were treated with FeCH (0.25, 0.5, 1, and 2 mg/mL) for 24 h, and viability was determined by MTT assay. (<b>B</b>) ICC staining of chondrocytes using collagen type II antibody after 3 days of culture (×40). Data are represented as mean ± SD of three independent experiments.</p>
Full article ">Figure 2
<p>Inhibitory effects of FeCH on IL-1β-induced nitrite, PGE<sub>2</sub>, TNF-α, IL-6, and COX-2 in primary rat chondrocytes. Primary rat chondrocytes were pretreated with FeCH (0.125, 0.25, and 0.5 mg/mL) for 1 h, followed by IL-1β (5 ng/mL) for 24 h. Nitrite production (<b>A</b>), PGE₂ production (<b>B</b>), and TNF-α production (<b>C</b>) in the cell culture medium were determined by ELISA kit. (<b>D</b>) Expression of iNOS, COX-2, TNF-α, IL-6, and COX-1 expression was determined by western blot. (<b>E</b>) Quantitative data of (<b>D</b>) were analyzed using Image J software (version 8). Data are represented as mean ± SD of three independent experiments. α-tubulin served as an internal control. ## <span class="html-italic">p</span> &lt; 0.01 vs. control; * <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 vs. IL-1β-treated group.</p>
Full article ">Figure 3
<p>Inhibitory effects of FeCH on IL-1β-induced ADAMTS-5 and -4 and MMP-1, -3, and -13 in primary rat chondrocytes. Primary rat chondrocytes were pretreated with FeCH (0.125, 0.25, and 0.5 mg/mL) for 1 h, followed by IL-1β (5 ng/mL) for 24 h. (<b>A</b>) Protein levels of matrix-degrading enzymes (ADAMTS-5 and -4 and MMP-1, -3, and -13) was determined by western blot. (<b>B</b>) Quantitative data of (<b>A</b>) were analyzed using Image J software. Data are represented as mean ± SD of three independent experiments. α-tubulin served as an internal control. ## <span class="html-italic">p</span> &lt; 0.01 vs. control; ** <span class="html-italic">p</span> &lt; 0.01, and *** <span class="html-italic">p</span> &lt; 0.001 vs. IL-1β-treated group.</p>
Full article ">Figure 4
<p>Effect of FeCH on collagen type II and proteoglycan in IL-1β-stimulated primary rat chondrocytes. Primary rat chondrocytes were pretreated with FeCH (0.125, 0.25, and 0.5 mg/mL) for 1 h, followed by IL-1β (5 ng/mL) for 24 h. (<b>A</b>) Protein levels of collagen type II degradation were determined by western blot. (<b>B</b>) Quantitative data of (<b>A</b>) were analyzed using Image J software. (<b>C</b>) Proteoglycan degradation was determined by alcian blue stain. (<b>D</b>) After dissolving alcian blue dye in 6M guanidine-HCL, the absorbance at 620 nm was measured. Data are represented as mean ± SD of three independent experiments. α-tubulin served as an internal control. <span class="html-italic">## p</span> &lt; 0.01 vs. control; <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 vs. IL-1β-treated group.</p>
Full article ">Figure 5
<p>Effects of FeCH on phosphorylation of MAPKs and activation of NF-κB in IL-1β-stimulated primary rat chondrocytes. Primary rat chondrocytes were pretreated with FeCH (0.125, 0.25, and 0.5 mg/mL) for 1 h, followed by IL-1β (5 ng/mL) for 24 h. (<b>A</b>) Protein levels of phosphorylation of MAPKs were determined by western blot. (<b>B</b>) Quantitative data of (<b>A</b>) were analyzed using Image J software. (<b>C</b>) Phosphorylation levels of NF-κB p65 were determined by western blot. (<b>D</b>) Quantitative data of (<b>C</b>) were analyzed using Image J software. Data are represented as mean ± SD of three independent experiments. α-tubulin and Lamin B1 served as internal controls. <span class="html-italic">## p</span> &lt; 0.01 vs. control; <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 vs. IL-1β-treated group.</p>
Full article ">Figure 6
<p>Histological evaluation of cartilage-protective effect of FeCH against cartilage degradation in a DMM model. (<b>A</b>) In vivo schemic (×100). White box indicates the enlarged area. (<b>B</b>) Rats underwent surgical destabilization of the medial meniscus (DMM). The day after the DMM surgery, rats were orally administered with FeCH (25 and 50 mg/kg) or celecoxib (2 mg/kg) daily for 8 weeks. Histological analysis of cartilage destruction was evaluated by safranin O/fast green staining.</p>
Full article ">
12 pages, 5281 KiB  
Article
PAC1 Agonist Maxadilan Reduces Atherosclerotic Lesions in Hypercholesterolemic ApoE-Deficient Mice
by Lilli Mey, Gabriel A. Bonaterra, Joy Hoffmann, Hans Schwarzbach, Anja Schwarz, Lee E. Eiden, Eberhard Weihe and Ralf Kinscherf
Int. J. Mol. Sci. 2024, 25(24), 13245; https://doi.org/10.3390/ijms252413245 - 10 Dec 2024
Viewed by 322
Abstract
A possible involvement of immune- and vasoregulatory PACAP signaling at the PAC1 receptor in atherogenesis and plaque-associated vascular inflammation has been suggested. Therefore, we tested the PAC1 receptor agonist Maxadilan and the PAC1 selective antagonist M65 on plaque development and lumen stenosis in [...] Read more.
A possible involvement of immune- and vasoregulatory PACAP signaling at the PAC1 receptor in atherogenesis and plaque-associated vascular inflammation has been suggested. Therefore, we tested the PAC1 receptor agonist Maxadilan and the PAC1 selective antagonist M65 on plaque development and lumen stenosis in the ApoE−/− atherosclerosis model for possible effects on atherogenesis. Adult male ApoE−/− mice were fed a cholesterol-enriched diet (CED) or standard chow (SC) treated with Maxadilan, M65 or Sham. Effects of treatment on atherosclerotic plaques, lumen stenosis, apoptosis and pro-inflammatory signatures were analyzed in the brachiocephalic trunk (BT). The percentage of Maxadilan treated mice exhibiting plaques under SC and CED was lower than that of Sham or M65 treatment indicating opposite effects of Maxadilan and M65. Maxadilan application inhibited lumen stenosis in SC and CED mice compared to the Sham mice. In spite of increased cholesterol levels, lumen stenosis of Maxadilan-treated mice was similar under CED and SC. In contrast, M65 under SC or CED did not reveal a significant influence on lumen stenosis. Maxadilan significantly reduced the TNF-α-immunoreactive (TNF-α+) area in the plaques under CED, but not under SC. In contrast, the IL-1β+ area was reduced after Maxadilan treatment in SC mice but remained unchanged in CED mice compared to Sham mice. Maxadilan reduced caspase-3 immunoreactive (caspase-3+) in the tunica media under both, SC and CED without affecting lipid content in plaques. Despite persistent hypercholesterolemia, Maxadilan reduces lumen stenosis, apoptosis and TNF-α driven inflammation. Our data suggest that Maxadilan provides atheroprotection by acting downstream of hypercholesterolemia-induced vascular inflammation. This implicates the potential of PAC1-specific agonist drugs against atherosclerosis even beyond statins and PCSK9 (proprotein convertase subtilisin/kexin type 9) inhibitors. Full article
(This article belongs to the Special Issue Atherosclerosis: From Molecular Basis to Therapy)
Show Figures

Figure 1

Figure 1
<p>Percentage of mice exhibiting macroscopically visible atherosclerotic plaques in the ascending aorta (aao), descending aorta (dao), aortic arch (aa), and branches of Sham and Maxadilan-treated ApoE<sup>−/−</sup> mice under standard chow (SC) or cholesterol-enriched diet (CED): Brachiocephalic trunk (BT), right/left carotid artery (rca/lca), right/left subclavian artery (rsca/lsca). Intraluminal contrast staining with methylene blue enhances the visibility of white atherosclerosis plaques, indicated by black arrows. For statistical significance, see <span class="html-italic">p</span> values in the individual micrographs. The numbers in parentheses denote the numbers of animals.</p>
Full article ">Figure 2
<p>Lumen stenosis and plaque development in the brachiocephalic trunk (BT)of ApoE<sup>−/−</sup> mice at the age of 20 weeks, fed with standard chow (SC) or cholesterol-enriched diet (CED). Treated mice received PAC1 agonist Maxadilan (20 nmol/kg) or PAC1 antagonist M65 (20 nmol/kg), dissolved in physiological saline solution i.p., Sham groups received physiological saline solution. Cross sections of areas with maximum plaque sizes were prepared and stained with hematoxylin/eosin. Lumen stenosis was measured in %; data are provided as mean ± SEM. For statistical significance, see <span class="html-italic">p</span> values in the individual micrographs.</p>
Full article ">Figure 3
<p>Immunohistochemistry and quantification of the inflammatory markers TNF-α and IL-1β in the plaque and tunica media of BT in ApoE<sup>−/−</sup> mice under SC and CED, after Maxadilan treatment or Sham (<b>A</b>). The percentages of immunohistochemically positive-stained areas per plaque or media were evaluated. Black arrowheads indicate immunoreactive positive cells. Data (<b>B</b>) TNF-α and (<b>C</b>) IL-1β are provided as mean + SEM (n = 7–9). * <span class="html-italic">p</span> ≤ 0.05 vs. Sham under SC or CED. Scale bar: 100 µm.</p>
Full article ">Figure 4
<p>Immunohistochemistry and quantification of immunoreactive areas (in %) in relation to the whole plaque or media of cross sections of BT of ApoE<sup>−/−</sup> mice under SC or CED after Maxadilan treatment or Sham (<b>A</b>). Antibodies were directed against cyclooxygenase-2 (COX-2) and cleaved caspase-3 (cl. casp.3, apoptosis). Black arrowheads indicate immunoreactive positive cells. Data (<b>B</b>) COX-2 and (<b>C</b>) cleaved caspase-3 are provided as mean + SEM (n = 7–9). ** <span class="html-italic">p</span> ≤ 0.01 vs. Sham under SC; <sup>#</sup> <span class="html-italic">p</span> ≤ 0.05 <sup>###</sup> <span class="html-italic">p</span> ≤ 0.001 vs. Sham under CED. Scale bar: 100 µm.</p>
Full article ">Figure 5
<p>ORO (on the left) and CD68 (on the right) (immuno)histochemistry of atherosclerotic BT cross-sections of ApoE<sup>−/−</sup> mice under Maxadilan treatment or Sham SC or CED. Oil red positive lipid deposits are shown in red. (<b>A</b>,<b>B</b>) Lipid-loaden MΦ in the plaque-forming foam cells (black arrowheads) are distinguishable from intimal cells, smooth muscle cells of the tunica media with small lipid droplets (arrows) and cells of perivascular fat tissue showing distinct fat vacuoles. (<b>C</b>) CD68<sup>+</sup> Mo and MΦ (right side, white arrowheads) stained with HRP-DAB in brown. Data are provided as mean + SEM (n = 7–9). Diagrams show % of lipid or Mo/MΦ positive area in relation to the whole plaque area. Intima (i), media (m) and adventitia (a). Scale bar: 100 µm.</p>
Full article ">
17 pages, 4993 KiB  
Article
A Novel Platform Featuring Nanomagnetic Ligand Fishing Based on Fixed-Orientation Immobilized Magnetic Beads for Screening Potential Cyclooxygenase-2 Inhibitors from Panax notoginseng Leaves
by Fan Zhang, Fan Sun, Lequan Yu, Fei Li, Lixia Liu, Xiaoyan Cao, Yi Zhang and Lijie Wu
Molecules 2024, 29(23), 5801; https://doi.org/10.3390/molecules29235801 - 9 Dec 2024
Viewed by 425
Abstract
A novel screening platform based on an Fe3O4@C@PDA-Ni2+@COX-2 ligand fishing combination with high-performance liquid chromatography–mass spectrometry was first designed, synthesized, and employed to screen and identify COX-2 inhibitors from Panax notoginseng leaves. The obtained magnetic nanoparticles exhibit [...] Read more.
A novel screening platform based on an Fe3O4@C@PDA-Ni2+@COX-2 ligand fishing combination with high-performance liquid chromatography–mass spectrometry was first designed, synthesized, and employed to screen and identify COX-2 inhibitors from Panax notoginseng leaves. The obtained magnetic nanoparticles exhibit outstanding preconcentration ability that allows for controlling the enzyme orientation to avoid enzyme active site blocking, conformational changes, or denaturing during immobilization. The as-prepared Fe3O4@C@PDA-Ni2+@COX-2 composite was carefully characterized by scanning electron microscopy (SEM), transmission electron microscopy (TEM), Fourier-transform infrared spectrometry (FT-IR), Xray powder diffraction (XRD), thermal gravimetric analyzer (TGA), vibrating sample magnetometer (VSM), and Zeta potential analysis. The analytical parameters influencing the magnetic solid-phase fishing efficiency were optimized by univariate and multivariate methods (Box–Behnken design) by testing a positive control and celecoxib with active and inactive COX-2. Under the optimized ligand fishing conditions, twelve potential COX-2 inhibitors were screened and characterized in Panax notoginseng leaves. The results indicate that the proposed method provides a simple, feasible, selective, and effective platform for the efficient screening and identification of active compounds from Chinese herbal medicine. It has guiding significance for the synthesis and development of novel anti-inflammatory drugs, and provides a reference for the efficient discovery of anti-inflammatory drugs or lead compounds from the complex system of Chinese herbal medicine. Full article
Show Figures

Graphical abstract

Graphical abstract
Full article ">Figure 1
<p>SEM image of (<b>A</b>) Fe<sub>3</sub>O<sub>4</sub>@C, (<b>B</b>) Fe<sub>3</sub>O<sub>4</sub>@C@PDA, and (<b>C</b>) Fe<sub>3</sub>O<sub>4</sub>@C@PDA-Ni<sup>2+</sup>; TEM image of (<b>D</b>) Fe<sub>3</sub>O<sub>4</sub>@C@PDA-Ni<sup>2+</sup>.</p>
Full article ">Figure 2
<p>(<b>A</b>) FT-IR spectra; (<b>B</b>) TGA thermograms; (<b>C</b>) Nitrogen adsorption–desorption isotherms with pore diameter distribution (inset); (<b>D</b>) VSM magnetization curves; (<b>E</b>) particle size distribution curve; and (<b>F</b>) Zeta potential distribution diagram of Fe<sub>3</sub>O<sub>4</sub>@PDA-Ni<sup>2+</sup>@COX-2 NPs.</p>
Full article ">Figure 3
<p>Confocal laser scanning images of immobilized COX-2 using MNPs in (<b>A</b>) bright-field, (<b>B</b>) dark-field, and (<b>C</b>) merged bright-darkfield. The bar is 50 μm.</p>
Full article ">Figure 4
<p>Effects of (<b>A</b>) concentration of Ni<sup>2+</sup>, (<b>B</b>) volume of COX-2, (<b>C</b>) content of methanol, (<b>D</b>) volume of elution solvent, and (<b>E</b>) desorption time. The data shown are the mean ± SD, <span class="html-italic">n</span> = 3. One-way analysis of variance (ANOVA) was used, <span class="html-italic">p</span> &lt; 0.05 was considered as statistically significant (**** <span class="html-italic">p</span> &lt; 0.0001, *** <span class="html-italic">p</span> &lt; 0.001,** <span class="html-italic">p</span> &lt; 0.01, * <span class="html-italic">p</span> &lt; 0.05, ns <span class="html-italic">p</span> &gt; 0.05).</p>
Full article ">Figure 5
<p>Three-dimensional response surface variables for extraction of celecoxib: (<b>A</b>) extraction time and extraction temperature; (<b>B</b>) extraction time and adsorbent mass; and (<b>C</b>) extraction temperature and adsorbent mass.</p>
Full article ">Figure 6
<p>(<b>A</b>) The pseudo-first-order kinetic model and (<b>B</b>) the pseudo-second-order kinetic model fitting diagram; (<b>C</b>) the Langmuir isothermal adsorption model; and (<b>D</b>) Freundlich isothermal adsorption model fitting diagram.</p>
Full article ">Figure 7
<p>Schematic of preparation of Fe<sub>3</sub>O<sub>4</sub>@C@PDA-Ni<sup>2+</sup>@COX-2.</p>
Full article ">Figure 8
<p>MSPF procedure. Peaks 1–12 are fishing compounds that are potential COX-2 inhibitors.</p>
Full article ">
13 pages, 1225 KiB  
Review
Low Prostaglandin E2 but High Prostaglandin D2, a Paradoxical Dissociation in Arachidonic Acid Metabolism in Aspirin-Exacerbated Airway Disease: Role of Airway Epithelium
by César Picado, Liliana Machado-Carvalho and Jordi Roca-Ferrer
J. Clin. Med. 2024, 13(23), 7416; https://doi.org/10.3390/jcm13237416 - 5 Dec 2024
Viewed by 489
Abstract
In patients with aspirin-exacerbated respiratory disease (AERD), there is disparate regulation of prostaglandin E2 (PGE2) and prostaglandin D2 (PGD2). Both prostanoids are synthesised by cyclooxygenase 1 (COX-1) and cyclooxygenase 2 (COX-2). However, while the basal synthesis of PGE [...] Read more.
In patients with aspirin-exacerbated respiratory disease (AERD), there is disparate regulation of prostaglandin E2 (PGE2) and prostaglandin D2 (PGD2). Both prostanoids are synthesised by cyclooxygenase 1 (COX-1) and cyclooxygenase 2 (COX-2). However, while the basal synthesis of PGE2 tends to decrease, that of PGD2 increases in patients with AERD. Furthermore, both behave differently in response to the inhibitory action of NSAIDs on COX-1: PGE2 levels decrease while PGD2 increases. Increased PGD2 release correlates with nasal, bronchial, and extra-pulmonary symptoms caused by aspirin in AERD. The proposed hypothesis establishes that the answer to this paradoxical dissociation can be found in the airway epithelium. This is based on the observation that reduced COX-2 mRNA and/or protein expression is associated with reduced PGE2 synthesis in cultured fibroblast and epithelial cells from AERD compared to patients with asthma who are aspirin-tolerant and healthy subjects. The low production of PGE2 by the airway epithelium in AERD results in an excessive release of alarmins (TSLP, IL-33), which in turn contributes to activating group 2 innate lymphoid cells (ILC2s) and PGD2 synthesis by mast cells and eosinophils. Aspirin, by further increasing the diminished PGE2 regulation capacity in AERD, leads to respiratory reactions associated with the surge in PGD2 from mast cells and eosinophils. In summary, the downregulation of COX-2 and the subsequent low production of PGE2 by airway cells account for the apparently paradoxical increased production of PGD2 by mast cells and eosinophils at the baseline and after aspirin provocation in patients with AERD. A better understanding of the role of the airway epithelium would contribute to elucidating the mechanism of AERD. Full article
(This article belongs to the Section Pulmonology)
Show Figures

Figure 1

Figure 1
<p>Arachidonic acid metabolism pathways.</p>
Full article ">Figure 2
<p>Response of airway epithelium to environmental triggers in subjects who are healthy or have AERD at the baseline and after aspirin exposure. In healthy airway epithelium, PGE2 modulates alarmin release, which is maintained within homeostatic limits. In AERD, decreased COX-2 expression and reduced PGE2 production by the disrupted epithelium facilitate excessive alarmin release, which in turn increases PGD2 production by mast cells. In patients with AERD, aspirin exposure further decreases PGE2 production, precipitating an acute increase in PGD2 by mast cells, contributing to the development of bronchoconstriction, nasal obstruction, and extra thoracic symptoms (urticaria, abdominal cramps).</p>
Full article ">
16 pages, 8973 KiB  
Article
Broccoli Sprout Extract Suppresses Particulate-Matter-Induced Matrix-Metalloproteinase (MMP)-1 and Cyclooxygenase (COX)-2 Expression in Human Keratinocytes by Direct Targeting of p38 MAP Kinase
by Jaehyeok Yun and Jong-Eun Kim
Nutrients 2024, 16(23), 4156; https://doi.org/10.3390/nu16234156 - 30 Nov 2024
Viewed by 601
Abstract
Background/Objectives: Particulate matter (PM) is an environmental pollutant that negatively affects human health, particularly skin health. In this study, we investigated the inhibitory effects of broccoli sprout extract (BSE) on PM-induced skin aging and inflammation in human keratinocytes. Methods: HaCaT keratinocytes were pretreated [...] Read more.
Background/Objectives: Particulate matter (PM) is an environmental pollutant that negatively affects human health, particularly skin health. In this study, we investigated the inhibitory effects of broccoli sprout extract (BSE) on PM-induced skin aging and inflammation in human keratinocytes. Methods: HaCaT keratinocytes were pretreated with BSE before exposure to PM. Cell viability was assessed using the MTT assay. The expression of skin aging and inflammation markers (MMP-1, COX-2, IL-6) was measured using Western blot, ELISA, and qRT-PCR. Reactive oxygen species levels were determined using the DCF-DA assay. Kinase assays and pull-down assays were conducted to investigate the interaction between BSE and p38α MAPK. Results: Our findings demonstrate that BSE effectively suppressed the expression of MMP-1, COX-2, and IL-6—critical skin aging and inflammation markers—by inhibiting p38 MAPK activity. BSE binds directly to p38α without competing with ATP, thereby selectively inhibiting its activity and downstream signaling pathways, including MSK1/2, AP-1, and NF-κB. Conclusions: These results suggest that BSE is a potential functional ingredient in skincare products to mitigate PM-induced skin damage. Full article
(This article belongs to the Special Issue Dietary Phytochemicals: Implications for Health and Disease)
Show Figures

Figure 1

Figure 1
<p>Effects of BSE on PM-induced MMP-1 production and expression. (<b>A</b>) Influence of BSE on MMP-1 production in HaCaT cells. HaCaT cells were pretreated with varying concentrations of BSE for 1 h prior to PM exposure (40 μg/mL). After 24 h, culture media was collected, and MMP-1 levels were quantified using ELISA kits following the protocol in the Materials and Methods section. (<b>B</b>) Influence of BSE on MMP-1 expression in HaCaT cells. Cells were pretreated with specified concentrations of BSE for 1 h before PM exposure (40 μg/mL). Following a 24 h incubation, cells were lysed as outlined in Materials and Methods, and MMP-1 protein levels were assessed via Western blotting. β-Actin was used as a loading control. Images were analyzed using the FUSION Solo S system (VILBER Lourmat, Paris, France). (<b>C</b>) Effect of BSE on MMP-1 mRNA levels in HaCaT cells. Quantitative real-time PCR was used to assess MMP-1 mRNA expression following 1 h pretreatment with BSE at specified concentrations and 6 h PM exposure (40 μg/mL). Cells were lysed according to the Materials and Methods. Data (<span class="html-italic">n</span> = 3) are presented as mean ± SD. (<b>D</b>) Impact of BSE on HaCaT cell viability. Cell viability was evaluated using the MTT assay 24 h after pretreatment with BSE at indicated concentrations and PM exposure (40 μg/mL). Data (<span class="html-italic">n</span> = 5) are shown as mean ± SD. Different letters (a–d) on the graph indicate statistically significant differences (<span class="html-italic">p</span> &lt; 0.05) determined by one-way ANOVA, with Duncan’s Multiple Range test as a post hoc test.</p>
Full article ">Figure 2
<p>Effects of BSE on PM-induced COX-2 and IL-6 production and expression. (<b>A</b>) Influence of BSE on PGE2 production in HaCaT cells. HaCaT cells were pretreated with various concentrations of BSE for 1 h before PM exposure (40 μg/mL). Following 24 h, culture media was collected, and PGE2 production was measured using enzyme immunoassay kits as detailed in the Materials and Methods. (<b>B</b>,<b>C</b>) Influence of BSE on COX-2 expression in HaCaT cells. Cells were pretreated with the indicated concentrations of BSE for 1 h prior to PM exposure (40 μg/mL). After 24 h, cells were lysed as per the protocol, and COX-2 protein levels were analyzed by Western blotting. β-Actin was used as a loading control. The FUSION Solo S imaging system (VILBER Lourmat, Paris, France) was used for measurements. (<b>D</b>) Effect of BSE on COX-2 mRNA levels in HaCaT cells. Quantitative real-time PCR was employed to assess COX-2 mRNA levels following 1 h BSE pretreatment and 6 h PM exposure (40 μg/mL). Cells were lysed as described in the Materials and Methods. (<b>E</b>) Effect of BSE on IL-6 production in HaCaT cells. Cells were pretreated with BSE for 1 h before exposure to PM (40 μg/mL), and after 24 h, IL-6 levels in the culture media were quantified using ELISA kits as per the protocol. (<b>F</b>) Influence of BSE on IL-6 mRNA expression in HaCaT cells. Quantitative real-time PCR was used to assess IL-6 mRNA levels after 1 h BSE pretreatment and 6 h PM exposure (40 μg/mL). Data (<span class="html-italic">n</span> = 3) are shown as mean ± SD. Means with different letters (a–d) indicate significant differences (<span class="html-italic">p</span> &lt; 0.05), determined by one-way ANOVA, with Duncan’s Multiple Range test as the post hoc test.</p>
Full article ">Figure 3
<p>Effect of BSE on PM-induced AP-1 and NF-κB activity in transactivated cells. The transactivation of (<b>A</b>) AP-1 and (<b>B</b>) NF-κB was assessed using a GFP reporter gene assay. HaCaT cells transduced with AP-1 and NF-κB were pretreated with BSE at specified concentrations for 1 h prior to PM exposure (40 μg/mL). After 24 h, the transactivation was measured at an excitation wavelength of 469 nm and an emission wavelength of 525 nm (GFP). Data (<span class="html-italic">n</span> = 3) are presented as the mean ± SD. The means labeled with different letters (a–c) on the graph indicate significant differences (<span class="html-italic">p</span> &lt; 0.05) according to one-way ANOVA, with Duncan’s Multiple Range Test used for post hoc analysis.</p>
Full article ">Figure 4
<p>Effect of SBE on PM-induced MSK1 phosphorylation in HaCaT cells. Effects of BSE on PM-induced phosphorylation of (<b>A</b>,<b>B</b>) MSK1/2, (<b>C</b>,<b>D</b>) p38, (<b>C</b>,<b>E</b>) ERK1/2, and (<b>C</b>,<b>F</b>) JNK1/2. HaCaT cells were pretreated with BSE at the indicated concentrations for 1 h before exposure to PM (40 μg/mL). After 2 h, the HaCaT cells were lysed as described in the Materials and Methods. Phosphorylated and total protein levels were analyzed by Western blotting. Images were measured with FUSION Solo S (VILVER Lourmat, Paris, France). Data (<span class="html-italic">n</span> = 3) are presented as the mean ± SD. The means with letters (a,b) on the graph indicate a significant difference (<span class="html-italic">p</span> &lt; 0.05) according to one-way ANOVA, and Duncan’s Multiple Range test was used as a post hoc test.</p>
Full article ">Figure 5
<p>Effects of BSE on p38 MAPK activity via direct binding. (<b>A</b>) BSE inhibits p38α activity. (<b>B</b>) BSE does not inhibit ERK1 activity. Kinase activity was measured as outlined in the Materials and Methods section. (<b>C</b>) Interaction between BSE and active p38α. Binding of BSE to p38α was confirmed by Western blot analysis using an anti-p38α antibody: Lane 1 (input control) contains 10 ng of p38α, serving as a protein standard. Lane 2 (control) contains Sepharose 4B beads. Lane 3 shows the p38α protein pulled down by BSE conjugated to Sepharose 4B beads. (<b>D</b>) Binding of BSE to p38 in HaCaT cell lysate. Western blotting, using an anti-p38 antibody, confirmed the binding: Lane 1 (input control) contains 0.5 µg of HaCaT cell lysate as a protein standard. Lane 2 (control) contains Sepharose 4B beads. Lane 3 indicates that p38 was pulled down by BSE–Sepharose 4B beads. (<b>E</b>) BSE does not compete with ATP for binding to p38α. Lane 1 (input control) includes 10 ng of p38α, used as a protein standard. Lane 2 (control) contains Sepharose 4B beads. Lane 3 shows p38α pulled down by BSE–Sepharose 4B beads, while Lanes 4-6 demonstrate active p38α pulled down by BSE–Sepharose 4B beads in the presence of ATP at concentrations of 10, 100, and 1000 µM. The means with letters (a,b) on the graph indicate a significant difference (<span class="html-italic">p</span> &lt; 0.05) according to one-way ANOVA, and Duncan’s Multiple Range test was used as a post hoc test.</p>
Full article ">Figure 6
<p>Effect of BSE on PM-induced ROS production. (<b>A</b>,<b>B</b>) Images and quantification of PM-induced ROS reduction by BSE were determined using a DCF-DA fluorescence assay. Data (<span class="html-italic">n</span> = 3) are expressed as the mean ± SD. Different letters (a–c) on the graph indicate significant differences (<span class="html-italic">p</span> &lt; 0.05), as determined by one-way ANOVA followed by Duncan’s Multiple Range Test for post hoc analysis.</p>
Full article ">
14 pages, 4925 KiB  
Article
The Effect of Meloxicam on Inflammatory Response and Oxidative Stress Induced by Klebsiella pneumoniae in Bovine Mammary Epithelial Cells
by Kangjun Liu, Shangfei Qiu, Li Fang, Luying Cui, Junsheng Dong, Long Guo, Xia Meng, Jianji Li and Heng Wang
Vet. Sci. 2024, 11(12), 607; https://doi.org/10.3390/vetsci11120607 - 29 Nov 2024
Viewed by 556
Abstract
Klebsiella pneumoniae (K. pneumoniae) is a significant pathogen associated with clinical mastitis in cattle. Anti-inflammatory drugs are necessary to alleviate pain and inflammation during clinical mastitis. Among many drugs, meloxicam (MEL) has been widely used in clinical mastitis because of its [...] Read more.
Klebsiella pneumoniae (K. pneumoniae) is a significant pathogen associated with clinical mastitis in cattle. Anti-inflammatory drugs are necessary to alleviate pain and inflammation during clinical mastitis. Among many drugs, meloxicam (MEL) has been widely used in clinical mastitis because of its excellent inhibitory effect on the cyclooxygenase-2 (COX-2) enzyme. However, the effectiveness of MEL on the inflammatory response and oxidative stress induced by K. pneumoniae are unclear. In the present study, primary BMECs were infected with K. pneumoniae in the presence or absence of plasma maintenance concentration of MEL (0.5 and 5 μM). Following 1 or 3 h of combined treatment with K. pneumoniae and MEL, BMECs were gathered to assess the related indicators. The results showed that MEL at plasma maintenance concentrations exerted no influence on the viability of uninfected BMECs and also had no impact on bacterial load in BMECs. At these concentrations, MEL was able to inhibit the mRNA expression of COX-2, Interleukin (IL)-1β, Tumor necrosis factor α (TNF-α), and IL-6 while simultaneously elevating the mRNA levels of IL-8 in K. pneumoniae-infected BMECs. MEL had clear effects on relieving oxidative stress by increasing the activity of superoxide dismutase (SOD) and catalase (CAT) and the level of total antioxidant capacity (T-AOC). The mechanisms by which MEL mitigated the inflammatory response and oxidative stress were partially attributed to inhibition of the nuclear transcription factor-kappa B (NF-κB) signaling pathway and improvement of the activation of the nuclear factor erythroid 2-related factors (Nrf2) signaling pathway. To conclude, the results manifested that MEL at plasma maintenance concentrations protected BMECs from inflammatory and oxidative damage induced by K. pneumoniae. Full article
(This article belongs to the Special Issue Ruminant Mastitis: Therapies and Control)
Show Figures

Figure 1

Figure 1
<p>The effects of MEL on cell viability. BMECs were treated with different concentrations (0, 0.5, 5, 10, 20, and 40 μΜ) of MEL for 12 h. The cell viability was evaluated with the CCK-8 method.</p>
Full article ">Figure 2
<p>The effects of MEL on bacterial load. BMECs were infected with <span class="html-italic">K. pneumoniae</span> for 3 h in the presence or absence of MEL. Each experiment was repeated 4 times.</p>
Full article ">Figure 3
<p>The effects of MEL on the mRNA expression (<b>A</b>) and protein levels (<b>B</b>,<b>C</b>) of COX-2 in BMECs. BMECs were infected with <span class="html-italic">K. pneumoniae</span> for 3 h in the presence or absence of MEL. <span class="html-italic"># p</span> &lt; 0.05 and ## <span class="html-italic">p</span> &lt; 0.01 compared with the control group. * <span class="html-italic">p</span> &lt; 0.05 and ** <span class="html-italic">p</span> &lt; 0.01 compared with the group infected with <span class="html-italic">K. pneumoniae</span> alone (see <a href="#app1-vetsci-11-00607" class="html-app">Supplementary Materials</a>).</p>
Full article ">Figure 4
<p>The effects of MEL on the mRNA expression of IL-1β (<b>A</b>), IL-6 (<b>B</b>), IL-8 (<b>C</b>), and TNF-α (<b>D</b>) in BMECs. BMECs were infected with <span class="html-italic">K. pneumoniae</span> for 3 h in the presence or absence of MEL. ## <span class="html-italic">p</span> &lt; 0.01 compared with the control group. * <span class="html-italic">p</span> &lt; 0.05 and ** <span class="html-italic">p</span> &lt; 0.01 compared with the group infected with <span class="html-italic">K. pneumoniae</span> alone.</p>
Full article ">Figure 5
<p>The effects of MEL on the NF-κB signaling pathway. BMECs were infected with <span class="html-italic">K. pneumoniae</span> for 1 h in the presence or absence of MEL. (<b>A</b>) The protein expression of p-p65, p65, p-IκBα, and IκBα in BMECs (see <a href="#app1-vetsci-11-00607" class="html-app">Supplementary Materials</a>). (<b>B</b>) Changes in the phosphorylation level of p65. (<b>C</b>) Changes in the phosphorylation level of IκBα. ## <span class="html-italic">p</span> &lt; 0.01 compared with the control group. * <span class="html-italic">p</span> &lt; 0.05 and ** <span class="html-italic">p</span> &lt; 0.01 compared with the group infected with <span class="html-italic">K. pneumoniae</span> alone.</p>
Full article ">Figure 6
<p>The effect of MEL on the nuclear accumulation of the p65 protein in BMECs. BMECs were infected with <span class="html-italic">K. pneumoniae</span> for 1 h in the presence or absence of MEL.</p>
Full article ">Figure 7
<p>The effects of MEL on the oxidative state of BMECs. (<b>A</b>,<b>B</b>) Changes in the level of ROS. (<b>C</b>) Changes in the level of MDA. BMECs were infected with <span class="html-italic">K. pneumoniae</span> for 3 h in the presence or absence of MEL. ## <span class="html-italic">p</span> &lt; 0.01 compared with the control group. ** <span class="html-italic">p</span> &lt; 0.01 compared with the group infected with <span class="html-italic">K. pneumoniae</span> alone.</p>
Full article ">Figure 8
<p>The effects of MEL on the antioxidant capacity of BMECs. BMECs were infected with <span class="html-italic">K. pneumoniae</span> for 3 h in the presence or absence of MEL. The activity of SOD (<b>A</b>) and CAT (<b>B</b>) and the level of T-AOC (<b>C</b>) were detected using commercial kits. ## <span class="html-italic">p</span> &lt; 0.01 compared with the control group. * <span class="html-italic">p</span> &lt; 0.05 and ** <span class="html-italic">p</span> &lt; 0.01 compared with the group infected with <span class="html-italic">K. pneumoniae</span> alone.</p>
Full article ">Figure 9
<p>The effects of MEL on the Nrf2 signaling pathway in BMECs. (<b>A</b>) The key proteins expressed in the Nrf2 signaling pathway were detected by Western blot (see <a href="#app1-vetsci-11-00607" class="html-app">Supplementary Materials</a>). (<b>B</b>) Changes in Nrf2 nuclear accumulation. Nuclear protein was extracted from cells. The protein expression levels of Nrf2 (<b>C</b>), Keap1 (<b>D</b>), HO-1 (<b>E</b>), and NQO1 (<b>F</b>) were detected using total protein. BMECs were infected with <span class="html-italic">K. pneumoniae</span> for 3 h in the presence or absence of MEL. <span class="html-italic"># p</span> &lt; 0.05 and ## <span class="html-italic">p</span> &lt; 0.01 compared with the control group. * <span class="html-italic">p</span> &lt; 0.05 and ** <span class="html-italic">p</span> &lt; 0.01 compared with the group infected with <span class="html-italic">K. pneumoniae</span> alone.</p>
Full article ">
12 pages, 2345 KiB  
Article
Photoaging Protective Effects of Quercitrin Isolated from ‘Green Ball’ Apple Peel
by Eun-Ho Lee, Junhyo Cho and In-Kyu Kang
Horticulturae 2024, 10(12), 1258; https://doi.org/10.3390/horticulturae10121258 - 27 Nov 2024
Viewed by 586
Abstract
Premature skin aging, also known as photoaging, refers to the changes in the structure and function of the skin caused by chronic sun exposure. The ultraviolet radiation in sunlight is one of the key factors that cause photoaging. Thus, matrix metalloproteinases (MMPs), transforming [...] Read more.
Premature skin aging, also known as photoaging, refers to the changes in the structure and function of the skin caused by chronic sun exposure. The ultraviolet radiation in sunlight is one of the key factors that cause photoaging. Thus, matrix metalloproteinases (MMPs), transforming growth factor beta-1 (TGFB1), and nuclear factor kappa B (NF-κB) signaling can be an effective therapeutic strategy for ultraviolet B (UVB) exposure. In this study, we used human dermal fibroblast and mouse macrophage cells to identify the mediators of skin photoaging. Quercitrin isolated from ‘Green Ball’ apple peel was treated to UVB-irradiated fibroblast cells and lipopolysaccharide (LPS)-induced macrophages to identify the photoaging prevention effect of quercitrin. Genes that are associated with photoaging were determined by using enzyme-linked immunosorbent assay (ELISA), Western blot, and quantitative polymerase chain reaction (qPCR). Quercitrin increased the collagen biosynthesis in UVB-irradiated fibroblast cells via regulating MMPs, TIMP metallopeptidase inhibitor 1 (TIMP-1), TGFB1, hyaluronan synthase 2 (HAS2), and collagen type I alpha 1 chain (COL1A2). In addition, quercitrin regulated p-65, inducible nitric oxide synthase (iNOS), and cyclooxygenase-2 (COX-2), and its mediators (prostaglandin E2 and nitric oxide), in the NF-κB signaling process, and it inhibited the production of cytokines in LPS-induced macrophages. These results indicate that quercitrin can improve photoaging damaged skin by regulating MMPs, TGFB1, and NF-κB signaling pathway modulators. Full article
Show Figures

Figure 1

Figure 1
<p>Flow chart of the separation and purification of quercitrin isolated from ‘Green Ball’ apple peel.</p>
Full article ">Figure 2
<p>The cell cytotoxicity (<b>A</b>) and cell numbers (<b>B</b>,<b>C</b>) of quercitrin (5–100 μM) in CCD-986sk cells and raw 264.7 cells. (<b>A</b>) The cell cytotoxicity was assessed using an MTT reduction assay and the results are expressed as the percentage of surviving cells compared with that in the negative control group (no addition of quercitrin). (<b>B</b>,<b>C</b>) The cell numbers were counted in the cell culture plates after incubation with quercitrin for 0, 12, and 24 h or 0, 24, and 48 h. Control groups were obtained in the absence of sample. Data are presented the means ± SD (<span class="html-italic">n</span> = 3, collected from 3 independent experiments). Data are considered significant when <span class="html-italic">p</span> &lt; 0.05 compared to control group (**: &lt;0.01).</p>
Full article ">Figure 3
<p>Effect of quercitrin on MMP-1, MMP-9, and TIMP-1 protein expression (<b>A</b>), protein expression rate (<b>B</b>), MMP-1, MMP-9, and TIMP-1 mRNA expression (<b>C</b>) in CCD-986sk cells with UVB irradiation. CCD-986sk cells were treated with UVB (30 mJ/cm<sup>2</sup>) and various concentrations of quercitrin (5, 10, and 25 μM) were added to the cells and incubated for 48 h. Control groups were obtained with only UVB irradiation without treatment. Data are presented the means ± SD (<span class="html-italic">n</span> = 3, collected from 3 independent experiments). Data with different letters were considered significant compared to the control group (<span class="html-italic">p</span> &lt; 0.05).</p>
Full article ">Figure 4
<p>Effect of quercitrin on COL1A2 protein expression (<b>A</b>), protein expression rate (<b>B</b>), TGFB1, COL1A2, and HAS2 mRNA expression (<b>C</b>) in CCD-986sk cells with UVB irradiation. CCD-986sk cells were treated with UVB (30 mJ/cm<sup>2</sup>) and various concentrations of quercitrin (5, 10, and 25 μM) were added and the cells were further incubated for 48 h. Control groups were obtained with only UVB irradiation without treatment. Data are presented the means ± SD (<span class="html-italic">n</span> = 3, collected from 3 independent experiments). Data with different letters were considered significant compared to the control group (<span class="html-italic">p</span> &lt; 0.05).</p>
Full article ">Figure 5
<p>Effect of quercitrin on p-p65, iNOS, and COX-2 protein expression (<b>A</b>), protein expression rate (<b>B</b>), PTGES2 mRNA expression (<b>C</b>), and nitrite production (<b>D</b>) in raw 264.7 cells with LPS-induced. Raw 264.7 cells were treated with LPS (1 µg/mL) and various concentrations of quercitrin (5, 10, 25, and 50 μM) were added to the cells and incubated for 24 h. Control groups were obtained with only LPS stimulation without treatment. Data are presented the means ± SD (<span class="html-italic">n</span> = 3, collected from 3 independent experiments). Data with different letters were considered significant compared to the control group (<span class="html-italic">p</span> &lt; 0.05).</p>
Full article ">Figure 6
<p>Effect of quercitrin on IL-1β, IL-6, TNF-α, and MCP-1 mRNA expression in raw 264.7 cells with LPS-induced. Raw 264.7 cells were treated with LPS (1 µg/mL) and various concentrations of quercitrin (10, 25, and 50 μM) were added to raw cells and incubated for 24 h. Control groups were obtained with only LPS stimulation without treatment. Data are presented the means ± SD (<span class="html-italic">n</span> = 3, collected from 3 independent experiments). Data with different letters were considered significant compared to the control group (<span class="html-italic">p</span> &lt; 0.05).</p>
Full article ">Figure 7
<p>Summary of the activity of quercitrin on photoaging.</p>
Full article ">
23 pages, 8373 KiB  
Article
Cyclodextrin-Nanosponge-Loaded Cyclo-Oxygenase-2 Inhibitor-Based Topical Gel for Treatment of Psoriatic Arthritis: Formulation Design, Development, and In vitro Evaluations
by Umme Hani, Sharanya Paramshetti, Mohit Angolkar, Wajan Khalid Alqathanin, Reema Saeed Alghaseb, Saja Mohammed Al Asmari, Alhanouf A. Alsaab, Farhat Fatima, Riyaz Ali M. Osmani and Ravi Gundawar
Pharmaceuticals 2024, 17(12), 1598; https://doi.org/10.3390/ph17121598 - 27 Nov 2024
Viewed by 559
Abstract
Background: Psoriatic arthritis (PsA), a chronic inflammatory disease, mainly affects the joints, with approximately 30% of psoriasis patients eventually developing PsA. Characterized by both innate and adaptive immune responses, PsA poses significant challenges for effective treatment. Recent advances in drug delivery systems have [...] Read more.
Background: Psoriatic arthritis (PsA), a chronic inflammatory disease, mainly affects the joints, with approximately 30% of psoriasis patients eventually developing PsA. Characterized by both innate and adaptive immune responses, PsA poses significant challenges for effective treatment. Recent advances in drug delivery systems have sparked interest in developing novel formulations to improve therapeutic outcomes. The current research focuses on the development and evaluation of a nanosponge-loaded, cyclo-oxygenase-2 (COX-2) inhibitor-based topical gel for the treatment of PsA. Methods: Nanosponges (NSs) were prepared by using beta-cyclodextrin as a polymer and dimethyl carbonate (DMC) as a crosslinker by melting, and gels were prepared by employing carbopol and badam gum as polymers. Results: Solubility studies confirmed that the prepared nanosponges were highly soluble. FT-IR studies confirmed the formation of hydrogen bonds between lumiracoxib and beta-cyclodextrin. SEM confirmed that the prepared formulations were roughly spherical and porous in nature. The average particle size was 190.5 ± 0.02 nm, with a zeta potential of −18.9 mv. XRD studies showed that the crystallinity of lumiracoxib decreased after encapsulation, which helped to increase its solubility. The optimized nanosponges (NS2) were incorporated in an optimized gel (FG10) to formulate a nanosponge-loaded topical gel. The optimized gel formulation exhibited a homogeneous consistency, with a pH of 6.8 and a viscosity of 1.15 PaS, indicating its suitability for topical application and stability. The in vitro diffusion studies for the topical gel showed drug release of 82.32% in 24 h. The optimized formulation demonstrated significant antipsoriatic activity, as confirmed through cytotoxicity studies conducted on HaCaT cells. Conclusions: On the basis of the findings, it can be concluded that the prepared nanosponge-loaded topical gel formulation presents a promising solution for the effective management of PsA, offering enhanced drug solubility, sustained release, and improved therapeutic potential. Full article
Show Figures

Figure 1

Figure 1
<p>FT-IR spectra of (<b>A</b>) lumiracoxib, (<b>B</b>) β-cyclodextrin, and (<b>C</b>) lumiracoxib-loaded β-CD NSs (NS2).</p>
Full article ">Figure 2
<p>DSC thermograms of (<b>A</b>) lumiracoxib, (<b>B</b>) β-cyclodextrin, and (<b>C</b>) lumiracoxib-loaded β-CD NSs (NS2).</p>
Full article ">Figure 3
<p>X-ray diffractograms: (<b>A</b>) lumiracoxib, (<b>B</b>) β-cyclodextrin, and (<b>C</b>) lumiracoxib loaded β-CD NSs (NS2).</p>
Full article ">Figure 4
<p>Particle size distribution for NS2 formulation.</p>
Full article ">Figure 5
<p>Surface morphology of optimized nanosponge (NS2).</p>
Full article ">Figure 6
<p>Graph of <span class="html-italic">in vitro</span> diffusion studies for formulations NS1 to NS3. Each value is expressed as the mean ± SD, with experiments conducted in triplicate (n = 3).</p>
Full article ">Figure 7
<p>Three-dimensional (3D) response surface plots of (<b>A</b>) viscosity (PsA) and (<b>B</b>) spreadability (cm). (<b>C</b>) Overlay plot of optimized gel formulation.</p>
Full article ">Figure 8
<p>Viscosity vs. shear rate graph.</p>
Full article ">Figure 9
<p><span class="html-italic">In vitro</span> diffusion studies for topical gel and pure drug. Each value is expressed as the mean ± SD, with experiments conducted in triplicate (n = 3).</p>
Full article ">Figure 10
<p>Cytotoxicity study performed using MTT assay for (<b>A</b>) L929 and (<b>B</b>) HaCaT cell lines. Each value is expressed as the mean ± SD, with experiments conducted in triplicate (n = 3).</p>
Full article ">Figure 11
<p>Representative images of test animal for skin irritation at (<b>A</b>) 0 h and (<b>B</b>) 72 h.</p>
Full article ">Figure 12
<p>Graphical representation of <span class="html-italic">Ex vivo</span> permeation studies.</p>
Full article ">
20 pages, 4640 KiB  
Article
In Vivo and Computational Studies on Sitagliptin’s Neuroprotective Role in Type 2 Diabetes Mellitus: Implications for Alzheimer’s Disease
by Vasudevan Mani and Minhajul Arfeen
Brain Sci. 2024, 14(12), 1191; https://doi.org/10.3390/brainsci14121191 - 26 Nov 2024
Viewed by 585
Abstract
Background/Objectives: Diabetes mellitus (DM), a widespread endocrine disorder characterized by chronic hyperglycemia, can cause nerve damage and increase the risk of neurodegenerative diseases such as Alzheimer’s disease (AD). Effective blood glucose management is essential, and sitagliptin (SITG), a dipeptidyl peptidase-4 (DPP-4) [...] Read more.
Background/Objectives: Diabetes mellitus (DM), a widespread endocrine disorder characterized by chronic hyperglycemia, can cause nerve damage and increase the risk of neurodegenerative diseases such as Alzheimer’s disease (AD). Effective blood glucose management is essential, and sitagliptin (SITG), a dipeptidyl peptidase-4 (DPP-4) inhibitor, may offer neuroprotective benefits in type 2 diabetes mellitus (T2DM). Methods: T2DM was induced in rats using nicotinamide (NICO) and streptozotocin (STZ), and biomarkers of AD and DM-linked enzymes, inflammation, oxidative stress, and apoptosis were evaluated in the brain. Computational studies supported the in vivo findings. Results: SITG significantly reduced the brain enzyme levels of acetylcholinesterase (AChE), beta-secretase-1 (BACE-1), DPP-4, and glycogen synthase kinase-3β (GSK-3β) in T2DM-induced rats. It also reduced inflammation by lowering cyclooxygenase-2 (COX-2), prostaglandin E2 (PGE2), tumor necrosis factor-α (TNF-α), and nuclear factor-κB (NF-κB). Additionally, SITG improved oxidative stress markers by reducing malondialdehyde (MDA) and enhancing glutathione (GSH). It increased anti-apoptotic B-cell lymphoma protein-2 (Bcl-2) while reducing pro-apoptotic markers such as Bcl-2-associated X (BAX) and Caspace-3. SITG also lowered blood glucose levels and improved plasma insulin levels. To explore potential molecular level mechanisms, docking was performed on AChE, COX-2, GSK-3β, BACE-1, and Caspace-3. The potential binding affinity of SITG for the above-mentioned target enzymes were 10.8, 8.0, 9.7, 7.7, and 7.9 kcal/mol, respectively, comparable to co-crystallized ligands. Further binding mode analysis of the lowest energy conformation revealed interactions with the critical residues. Conclusions: These findings highlight SITG’s neuroprotective molecular targets in T2DM-associated neurodegeneration and its potential as a therapeutic approach for AD, warranting further clinical investigations. Full article
Show Figures

Figure 1

Figure 1
<p>The timeline of the drug treatment and the experiment schedule.</p>
Full article ">Figure 2
<p>Effect of diabetes and sitagliptin on body weight in rats over a 30-day treatment period (<span class="html-italic">n</span> = 6). Data are presented as mean ± SEM. ** <span class="html-italic">p</span> &lt; 0.01 and *** <span class="html-italic">p</span> &lt; 0.001 vs. Day-1 in Control; ## <span class="html-italic">p</span> &lt; 0.01 and ### <span class="html-italic">p</span> &lt; 0.001 vs. Day 1 in SITG10; <span>$</span> <span class="html-italic">p</span> &lt; 0.05 vs. Day 1 in T2DM + SITG30.</p>
Full article ">Figure 3
<p>Effect of sitagliptin on blood glucose levels in diabetes-induced rats (<span class="html-italic">n</span> = 6). Data are presented as mean ± SEM. ** <span class="html-italic">p</span> &lt; 0.01 and *** <span class="html-italic">p</span> &lt; 0.001 vs. Day 1 in T2DM + SITG10; ### <span class="html-italic">p</span> &lt; 0.001 vs. Day 1 in T2DM + SITG30.</p>
Full article ">Figure 4
<p>Effect of sitagliptin on plasma insulin levels in diabetes-induced rats (<span class="html-italic">n</span> = 6). Data are presented as mean ± SEM. *** <span class="html-italic">p</span> &lt; 0.001 vs. Control; ### <span class="html-italic">p</span> &lt; 0.001 vs. T2DM.</p>
Full article ">Figure 5
<p>Effect of sitagliptin on enzyme activity in the brains of diabetes-induced rats (<span class="html-italic">n</span> = 6): (<b>A</b>) <span class="html-italic">AChE</span>, (<b>B</b>) <span class="html-italic">BACE-1</span>, (<b>C</b>) <span class="html-italic">DPP-4</span>, and (<b>D</b>) <span class="html-italic">GSK-3β</span>. Data are presented as mean ± SEM. *** <span class="html-italic">p</span> &lt; 0.001 vs. Control; ## <span class="html-italic">p</span> &lt; 0.01 and ### <span class="html-italic">p</span> &lt; 0.001 vs. T2DM.</p>
Full article ">Figure 6
<p>Effect of sitagliptin on inflammatory markers in the brains of diabetes-induced rats (<span class="html-italic">n</span> = 6): (<b>A</b>) <span class="html-italic">COX-2</span>, (<b>B</b>) PGE2, (<b>C</b>) TNF-α, and (<b>D</b>) NF-κB. Data are presented as mean ± SEM. * <span class="html-italic">p</span> &lt; 0.05 and *** <span class="html-italic">p</span> &lt; 0.001 vs. Control; # <span class="html-italic">p</span> &lt; 0.05 and ## <span class="html-italic">p</span> &lt; 0.01 vs. T2DM.</p>
Full article ">Figure 7
<p>Effect of sitagliptin on oxidative and antioxidant markers in the brains of diabetes-induced rats (<span class="html-italic">n</span> = 6): (<b>A</b>) MDA, (<b>B</b>) GSH, and (<b>C</b>) Catalase. Data are presented as mean ± SEM. ** <span class="html-italic">p</span> &lt; 0.01 and *** <span class="html-italic">p</span> &lt; 0.001 vs. Control; # <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 vs. T2DM; <span>$</span><span>$</span> <span class="html-italic">p</span> &lt; 0.01 vs. T2DM + SITG10.</p>
Full article ">Figure 8
<p>Effect of sitagliptin on apoptotic proteins in the brains of diabetes-induced rats (<span class="html-italic">n</span> = 6): (<b>A</b>) Bcl-2, (<b>B</b>) BAX, and (<b>C</b>) Caspace-3. Data are presented as mean ± SEM. * <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 vs. Control; # <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 vs. T2DM.</p>
Full article ">Figure 9
<p>Binding mode of SITG in the active site of <span class="html-italic">AChE</span>, <span class="html-italic">COX-2</span>, <span class="html-italic">GSK-3β</span>, <span class="html-italic">BACE-1</span>, and Caspace-3.</p>
Full article ">
13 pages, 5119 KiB  
Article
Sorbus commixta Fruit Extract Suppresses Lipopolysaccharide-Induced Neuroinflammation in BV-2 Microglia Cells via the MAPK and NF-κB Signaling Pathways
by Yon-Suk Kim, Jin-Hwa Jung and Ki-Tae Kim
Molecules 2024, 29(23), 5592; https://doi.org/10.3390/molecules29235592 - 26 Nov 2024
Viewed by 449
Abstract
Sorbus commixta Hedl. is a traditional medicinal plant in Korea, China, and Japan with known antioxidative, anti-inflammatory, anti-atherogenic, and anti-melanin activities. However, its anti-neuroinflammatory effects remain largely unknown. In this study, we investigated the inhibitory effects of S. commixta fruit extracts on lipopolysaccharide-stimulated [...] Read more.
Sorbus commixta Hedl. is a traditional medicinal plant in Korea, China, and Japan with known antioxidative, anti-inflammatory, anti-atherogenic, and anti-melanin activities. However, its anti-neuroinflammatory effects remain largely unknown. In this study, we investigated the inhibitory effects of S. commixta fruit extracts on lipopolysaccharide-stimulated pro-inflammatory factors in BV-2 microglia. We compared the anti-neuroinflammatory activity of S. commixta fruit water extract (SFW) and 70% ethanol extract using a nitric oxide assay. Our data indicated that the SFW (25–100 μg/mL) treatment significantly inhibited excessive nitric oxide production in lipopolysaccharide-stimulated BV-2 microglia compared to the 70% ethanol extract. It also attenuated the expression of inducible nitric oxide synthase, cyclooxygenase-2, and pro-inflammatory cytokines such as interleukin-6 and tumor necrosis factor α. Moreover, SFW exhibited its anti-inflammatory properties by downregulating the expression of factors involved in the extracellular signal-regulated kinase, c-Jun N-terminal kinase, and p38 mitogen-activated protein kinase pathways and by suppressing nuclear factor kappa B. Caffeic acid was identified as a primary component of SFW showing anti-neuroinflammatory activity. These findings suggest that SFW may offer substantial therapeutic potential for the treatment of neurodegenerative diseases involving microglia activation. Full article
Show Figures

Figure 1

Figure 1
<p>Effects of SFW and SFE on NO production (<b>A</b>) and cell viabilities (<b>B</b>) in LPS-induced BV-2 cells. BV-2 cells (2 × 10<sup>5</sup> cells/mL) were treated with various concentrations of SFW (12.5–200 μg/mL) for 1 h and treated with LPS (200 ng/mL) for 24 h. Controls were samples without LPS or SFW treatment. (<b>A</b>) NO production in the culture medium was determined using the Griess reagent and a standard curve using NaNO<sub>2</sub>. (<b>B</b>) Cell viability was assessed by the MTT assay, and the results are expressed as the percentage of control cells. Data are presented as mean ± SD from three independent experiments. <sup>###</sup> <span class="html-italic">p</span> &lt; 0.001 versus control. *** <span class="html-italic">p</span> &lt; 0.001 versus LPS.</p>
Full article ">Figure 2
<p>Effects of pro-inflammatory cytokines IL-6 and TNF-<span class="html-italic">α</span> in LPS-induced BV-2 cells (<b>A</b>,<b>B</b>). BV-2 cells were pretreated with SFW for 1 h and stimulated with LPS (200 ng/mL) for 18 h (<b>A</b>) or 6 h (<b>B</b>), and then IL-6 and TNF-<span class="html-italic">α</span> levels were determined using ELISA. Data are presented as mean ± SD from three independent experiments. <sup>###</sup> <span class="html-italic">p</span> &lt; 0.001 versus control. *** <span class="html-italic">p</span> &lt; 0.001 versus LPS.</p>
Full article ">Figure 3
<p>Effect of SFW on LPS-induced iNOS and COX-2 protein expressions in BV-2 cells (<b>A</b>,<b>B</b>). BV-2 cells were pretreated with SFW for 1 h, stimulated with LPS (200 ng/mL) for 18 h, and examined by Western blotting. Data are presented as mean ± SD from three independent experiments. β-actin was used as an internal control. <sup>###</sup> <span class="html-italic">p</span> &lt; 0.001 versus control. *** <span class="html-italic">p</span> &lt; 0.001 versus LPS.</p>
Full article ">Figure 4
<p>Effect of SFW on LPS-induced MAPK activation in BV-2 cells (<b>A</b>–<b>D</b>). BV-2 cells were pretreated with SFW (50–100 μg/mL) for 1 h prior to stimulation with LPS (200 ng/mL) for 30 min. Total protein (15 μg) was subjected to 12% SDS-PAGE, followed by Western blotting using anti-ERK, anti-p38, and anti-JNK. Results are representative of those obtained from three independent experiments. <sup>###</sup> <span class="html-italic">p</span> &lt; 0.001 versus control. *** <span class="html-italic">p</span> &lt; 0.001 versus LPS.</p>
Full article ">Figure 5
<p>Effects of SFW on nuclear translocation of NF-<span class="html-italic">κ</span>B p65 through suppression of I<span class="html-italic">κ</span>Bα phosphorylation in LPS-stimulated BV-2 microglia. BV-2 cells were pretreated with SFW (200 μg/mL) for 1 h prior to stimulation with LPS (200 ng/mL) for 30 min. Cell lysates were prepared and analyzed by Western blotting with anti-phospho-I<span class="html-italic">κ</span>Bα antibody (<b>A</b>,<b>B</b>). Results are representative of those obtained from three independent experiments. <sup>###</sup> <span class="html-italic">p</span> &lt; 0.001 versus control. *** <span class="html-italic">p</span> &lt; 0.001 versus LPS. Localization of NF-<span class="html-italic">κ</span>B p65 was visualized with fluorescence microscopy after immunofluorescence staining with NF-<span class="html-italic">κ</span>B p65 antibody (red fluorescence) (<b>C</b>). Cells were stained with DAPI for visualization of nuclei (blue fluorescence). Scale bar: 50 μm. Arrows: highlight areas of cells with a translocation of NF-<span class="html-italic">κ</span>B p65.</p>
Full article ">Figure 6
<p>Chemical structure of gallic acid (<b>1</b>), chlorogenic acid (<b>2</b>), caffeic acid (<b>3</b>), <span class="html-italic">p</span>-coumaric acid (<b>4</b>), and rutin (<b>5</b>).</p>
Full article ">Figure 7
<p>HPLC chromatogram of caffeic acid (<b>A</b>) and SFW (<b>B</b>). <span class="html-italic">X</span>-axis, retention time; <span class="html-italic">Y</span>-axis, absorbance unit. The monitoring wavelength was set at 320 nm.</p>
Full article ">Figure 8
<p>Effect of caffeic acid on NO production (<b>A</b>), cell viability (<b>B</b>) and iNOS and COX-2 protein expression (<b>C</b>,<b>D</b>) in LPS-induced BV-2 microglia. The BV-2 cells were pretreated with caffeic acid for 1 h, stimulated with LPS (200 ng/mL) for 18 h, and examined by Western blotting. Data are presented as mean ± SD from three independent experiments. β-actin was used as an internal control. <sup>###</sup> <span class="html-italic">p</span> &lt; 0.001 versus control. *** <span class="html-italic">p</span> &lt; 0.001 versus LPS.</p>
Full article ">
14 pages, 1605 KiB  
Article
Hydroethanolic Extract of Polygonum aviculare L. Mediates the Anti-Inflammatory Activity in RAW 264.7 Murine Macrophages Through Induction of Heme Oxygenase-1 and Inhibition of Inducible Nitric Oxide Synthase
by Chan Ho Jang, You Chul Chung, Ami Lee and Youn-Hwan Hwang
Plants 2024, 13(23), 3314; https://doi.org/10.3390/plants13233314 - 26 Nov 2024
Viewed by 580
Abstract
Polygonum aviculare L. (PAL), commonly known as knotgrass, has been utilized as a traditional folk medicine across Asian, African, Latin American and Middle Eastern countries to treat various inflammatory diseases, including arthritis and airway inflammation. Numerous medicinal herbs exert anti-inflammatory and antioxidative effects [...] Read more.
Polygonum aviculare L. (PAL), commonly known as knotgrass, has been utilized as a traditional folk medicine across Asian, African, Latin American and Middle Eastern countries to treat various inflammatory diseases, including arthritis and airway inflammation. Numerous medicinal herbs exert anti-inflammatory and antioxidative effects that are mediated through the activation of nuclear factor-erythroid 2-related factor 2 (Nrf2) and the inhibition of nuclear factor kappa B (NF-κB). However, the underlying molecular mechanisms linking the antioxidative and anti-inflammatory effects remain poorly understood. Heme oxygenase-1 (HO-1) is an antioxidant enzyme that catalyzes heme degradation, ultimately leading to the production of carbon monoxide (CO). Elevated levels of CO have been correlated with the decreased level of inducible nitric oxide synthase (iNOS). In this study, we examined whether HO-1 plays a key role in the relationship between the antioxidative and anti-inflammatory properties of PAL. The anti-inflammatory and antioxidative activities of PAL in an in vitro system were evaluated by determining NF-κB activity, antioxidant response element (ARE) activity, pro-inflammatory cytokine and protein levels, as well as antioxidant protein levels. To examine whether HO-1 inhibition interfered with the anti-inflammatory effect of PAL, we measured nitrite, reactive oxygen species, iNOS, and HO-1 levels in RAW 264.7 murine macrophages pre-treated with Tin protoporphyrin (SnPP, an HO-1 inhibitor). Our results demonstrated that PAL increased ARE activity and the Nrf2-regulated HO-1 level, exerting antioxidative activities in RAW 264.7 macrophages. Additionally, PAL reduced cyclooxygenase-2 (COX-2) and iNOS protein levels by inactivating NF-κB in lipopolysaccharide (LPS)-activated RAW 264.7 macrophages. Further investigation using the HO-1 inhibitor revealed that HO-1 inhibition promoted iNOS expression, subsequently elevating nitric oxide (NO) generation in LPS-activated RAW 264.7 macrophages treated with PAL compared to those in the macrophages without the HO-1 inhibitor. Overall, our findings suggest that HO-1 induction by PAL may exert anti-inflammatory effects through the reduction of the iNOS protein level. Hence, this study paves the way for further investigation to understand molecular mechanisms underlying the antioxidative and anti-inflammatory activities of medicinal herbs. Full article
Show Figures

Graphical abstract

Graphical abstract
Full article ">Figure 1
<p>Inhibition of NF-κB activation by PAL hydroethanolic extract. Cell viability of PAL hydroethanolic extract in (<b>A</b>) NF-κB Luciferase Reporter-RAW 264.7 cells, and (<b>B</b>) RAW 264.7 macrophages was quantified using a CCK-8 assay. (<b>C</b>) Concentration-dependent inhibition of NF-κB luciferase activity by PAL hydroethanolic extract in LPS-activated NF-κB Luciferase Reporter-RAW 264.7 cells. (<b>D</b>) Expression level of nuclear NF-κB in LPS-activated RAW 264.7 macrophages were quantitatively analyzed. Data are presented as mean  ±  standard error of the mean (SEM) from three independent experiments (<span class="html-italic">N</span>  =  3). A statistical significance compared with LPS alone treatment at <span class="html-italic">p</span> &lt; 0.05 and <span class="html-italic">p</span> &lt; 0.01 was marked by an asterisk (*) and double asterisk (**), respectively. LPS, lipopolysaccharide; DEX, dexamethasone; PAL, <span class="html-italic">Polygonum aviculare</span> L.; NS, not significant.</p>
Full article ">Figure 2
<p>Anti-inflammatory effects of PAL hydroethanolic extract in LPS-activated RAW 264.7 macrophages. Expression levels of (<b>A</b>) COX-2 and (<b>B</b>) iNOS in LPS-activated RAW 264.7 macrophages were quantitatively analyzed. The levels of extracellular (<b>C</b>) PGE<sub>2</sub> and (<b>D</b>) NO were analyzed in LPS-activated RAW 264.7 macrophages. Data are presented as the mean  ±  SEM (<span class="html-italic">N</span>  =  3). A statistical significance compared with LPS alone treatment at <span class="html-italic">p</span> &lt; 0.05 and <span class="html-italic">p</span> &lt; 0.01 was marked by an asterisk (*) and double asterisk (**), respectively.</p>
Full article ">Figure 3
<p>Effects of PAL hydroethanolic extract on the production of pro-inflammatory cytokines in LPS-activated RAW 264.7 macrophages. Cellular inflammatory response was provoked by LPS in RAW 264.7 macrophages. Pro-inflammatory cytokines, including (<b>A</b>) IL-1β, (<b>B</b>) IL-6, and (<b>C</b>) TNF-α were analyzed in LPS-activated RAW 264.7 cells. Data are presented as the mean  ±  SEM (<span class="html-italic">N</span>  =  3). A statistical significance compared with LPS alone treatment at <span class="html-italic">p</span> &lt; 0.05 and <span class="html-italic">p</span> &lt; 0.01 was marked by an asterisk (*) and double asterisk (**), respectively.</p>
Full article ">Figure 4
<p>Antioxidant effects of PAL hydroethanolic extract by activation of Nrf2 signaling pathway in RAW 264.7 macrophages. Protein levels of (<b>A</b>) nuclear Nrf2 and (<b>B</b>) cytoplasmic HO-1 in RAW 264.7 macrophages were quantitatively analyzed. (<b>C</b>) ARE activity by PAL hydroethanolic extract in HepG2-ARE cells. (<b>D</b>) Intracellular ROS level by PAL hydroethanolic extract in RAW 264.7 macrophages. Data are presented as the mean  ±  SEM (<span class="html-italic">N</span>  =  3). A statistical significance compared with control group at <span class="html-italic">p</span> &lt; 0.05 and <span class="html-italic">p</span> &lt; 0.01 was marked by an asterisk (*) and double asterisk (**), respectively. SFN, sulforaphane; tBHP, tert-butyl hydroperoxide; tBHQ, tertiary-butylhydroquinone.</p>
Full article ">Figure 5
<p>HO-1 inhibition nullified anti-inflammatory effects by PAL hydroethanolic extract. (<b>A</b>) Cytoplasmic iNOS protein level was quantitatively analyzed in RAW 264.7 cells pre-treated with or without SnPP. (<b>B</b>) Extracellular NO level by PAL hydroethanolic extract in RAW 264.7 cells pre-treated with or without SnPP. (<b>C</b>) Cytoplasmic HO-1 protein level was quantitatively analyzed in RAW 264.7 cells pre-treated with or without SnPP. (<b>D</b>) Intracellular ROS level was analyzed in LPS-activated RAW 264.7 cells pre-treated with and without SnPP. Values are mean  ±  SEM (<span class="html-italic">N</span>  =  3). A significance difference compared with LPS alone or control group at <span class="html-italic">p</span> &lt; 0.05 and <span class="html-italic">p</span> &lt; 0.01 was indicated by an asterisk (*) and double asterisk (**), respectively. A hash (# <span class="html-italic">p</span> &lt; 0.05) and double hash (## <span class="html-italic">p</span> &lt; 0.01) indicate a significant difference between groups. SnPP, Tin Protoporphyrin IX dichloride.</p>
Full article ">Figure 6
<p>HO-1 inhibition abrogates anti-inflammatory effects by KAE and QUE abundant in PAL hydroethanolic extract. NF-κB luciferase activity by (<b>A</b>) KAE and (<b>B</b>) QUE in LPS-activated NF-κB Luciferase Reporter-RAW 264.7 cells pre-treated with or without SnPP. Extracellular NO level by (<b>C</b>) KAE and (<b>D</b>) QUE in LPS-activated RAW 264.7 macrophages pre-treated with and without SnPP. Data are presented as the mean  ±  SEM (<span class="html-italic">N</span>  =  3). A statistical significance compared with LPS alone group at <span class="html-italic">p</span> &lt; 0.05 and <span class="html-italic">p</span> &lt; 0.01 was marked by an asterisk (*) and double asterisk (**), respectively. A hash (# <span class="html-italic">p</span> &lt; 0.05) and double hash (## <span class="html-italic">p</span> &lt; 0.01) indicate a statistical significance between groups. NS, not significant.</p>
Full article ">
16 pages, 4320 KiB  
Article
Bondarzewia dickinsii Against Colitis-Associated Cancer Through the Suppression of the PI3K/AKT/COX-2 Pathway and Inhibition of PGE2 Production in Mice
by Junliang Chen, Shuai Liu, Xin Zhang, Xiaojing Dai, Yu Li, Yonglin Han and Lanzhou Li
Nutrients 2024, 16(23), 4048; https://doi.org/10.3390/nu16234048 - 26 Nov 2024
Viewed by 542
Abstract
Background: Bondarzewia dickinsii (BD) is a newly discovered edible mushroom with rich nutritional components. This study presents a thorough analysis of the components of BD, examining its inhibitory effects and the underlying mechanisms by which BD influences colitis-associated cancer (CAC). Methods: AOM/DSS-induced CAC [...] Read more.
Background: Bondarzewia dickinsii (BD) is a newly discovered edible mushroom with rich nutritional components. This study presents a thorough analysis of the components of BD, examining its inhibitory effects and the underlying mechanisms by which BD influences colitis-associated cancer (CAC). Methods: AOM/DSS-induced CAC mice (male C57BL/6) were used, and a histopathological analysis, intestinal microbiota assessment, and metabolomics profiling were carried out, as well as an evaluation of relevant proteins and factors, to investigate the CAC-inhibitory effects of BD. Results: BD is rich in nutritional components, including a total sugar content of 37.29% and total protein content of 24.9%. BD significantly diminished colon inflammation, as well as the size and quantity of tumors. In addition, BD modified the diversity of intestinal microbiota and changed the levels of 19 serum metabolites, including arachidonic acid. BD significantly reduced prostaglandin E2 (PGE2) and cyclooxygenase-2 (COX-2) in colon tissue. Furthermore, it was found to inhibit the phosphatidylinositol 3-kinase (PI3K)/protein kinase B (AKT)/COX-2 signaling pathway. Conclusions: In general, BD inhibited the onset and progression of CAC by modulating the composition of intestinal microbiota and metabolite levels, suppressing the PI3K/AKT/COX-2 pathway, and decreasing PGE2 expression. This study provides a significant reference for the development of BD as a dietary supplement and pharmaceutical agent in the treatment of CAC. Full article
(This article belongs to the Special Issue Functional Evaluation of Edible Mushrooms and Their Active Materials)
Show Figures

Figure 1

Figure 1
<p>Effect of BD on AOM/DSS-induced CAC mice. (<b>A</b>) Flow chart of experimental protocol. (<b>B</b>) Bodyweight changes in mice (<span class="html-italic">n</span> = 8). (<b>C</b>) BD inhibits tumor development in CAC mice (<span class="html-italic">n</span> = 5).</p>
Full article ">Figure 2
<p>BD alleviates colorectal status in CAC mice. (<b>A</b>) The number of tumors. (<b>B</b>) The length of the colorectal. (<b>C</b>) Colon coefficient. (<b>D</b>) Histopathological observation of colorectal tumors in CAC mice (<span class="html-italic">n</span> = 3) (40× scale bar: 500 μm; 200× scale bar: 100 μm). ### <span class="html-italic">p</span> &lt; 0.001 vs. Ctrl group; * <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 vs. model group.</p>
Full article ">Figure 3
<p>The impact of BD on the intestinal microbiota in CAC mice. (<span class="html-italic">n</span> = 4) (<b>A</b>) Venn diagram. (<b>B</b>) PCoA based on weighted UniFrac distances. (<b>C</b>) Heatmap illustrating the composition of the top 20 most advantageous genera.</p>
Full article ">Figure 4
<p>BD modulates serum metabolite concentrations and colorectal cytokine levels in CAC mice. (<b>A</b>) OPLS-DA score plot. (<b>B</b>) Venn diagram. (<b>C</b>) Heatmap of 37 significantly altered metabolites in CAC mice.</p>
Full article ">Figure 5
<p>(<b>A</b>) The associated heatmap of significantly altered metabolites and genus microbiota. (<b>B</b>) PGE2, (<b>C</b>) COX-2, (<b>D</b>) IL-6 in CAC mice colon. ### <span class="html-italic">p</span> &lt; 0.001 vs. Ctrl group; * <span class="html-italic">p</span> &lt; 0.05 vs. model group.</p>
Full article ">Figure 6
<p>BD regulated protein expression in the colorectal tissue of CAC mice. (<b>A</b>) BD reduced the phosphorylation of PI3K and AKT and decreased the levels of COX-2, PGE2 and other proteins. Quantification of (<b>B</b>) P-PI3K, (<b>C</b>) P-AKT, (<b>D</b>) COX-2, (<b>E</b>) PGES2, (<b>F</b>) PKA, (<b>G</b>) CRTC1, (<b>H</b>) AREG, and (<b>I</b>) IL-6. # <span class="html-italic">p</span> &lt; 0.05, ## <span class="html-italic">p</span> &lt; 0.01 vs. Ctrl group; * <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 vs. model group.</p>
Full article ">
Back to TopTop