[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

Article Types

Countries / Regions

remove_circle_outline
remove_circle_outline
remove_circle_outline
remove_circle_outline
remove_circle_outline

Search Results (801)

Search Parameters:
Keywords = COX-2 inhibitor

Order results
Result details
Results per page
Select all
Export citation of selected articles as:
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
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 ">
17 pages, 4485 KiB  
Article
Tumor Intrinsic Immunogenicity Suppressor SETDB1 Worsens the Prognosis of Patients with Hepatocellular Carcinoma
by Chang-Qing Yin and Chun-Qing Song
Cells 2024, 13(24), 2102; https://doi.org/10.3390/cells13242102 - 19 Dec 2024
Viewed by 217
Abstract
Hepatocellular carcinoma (HCC) is clinically distinguished by its covert onset, rapid progression, high recurrence rate, and poor prognosis. Studies have revealed that SETDB1 (SET Domain Bifurcated 1) is a histone H3 methyltransferase located on chromosome 1 and plays a crucial role in carcinogenesis. [...] Read more.
Hepatocellular carcinoma (HCC) is clinically distinguished by its covert onset, rapid progression, high recurrence rate, and poor prognosis. Studies have revealed that SETDB1 (SET Domain Bifurcated 1) is a histone H3 methyltransferase located on chromosome 1 and plays a crucial role in carcinogenesis. Therefore, we aimed to evaluate the clinical significance of SETDB1 expression in HCC. In patients with HCC, elevated levels of SETDB1 correlated with a poorer overall survival (OS) rate, marking it as an independent prognostic factor for HCC, as revealed by both univariate and multivariate Cox analyses. Furthermore, we utilized the SangerBox and TISIDB databases to profile the tumor immune microenvironment in HCC, including scoring the tumor microenvironment and assessing immune cell infiltration. The TIDE algorithm was employed to examine the association between SETDB1 expression and immune responses. Our findings indicated that SETDB1 expression negatively correlated with the majority of immune cells, a wide range of immune cell marker genes, and numerous immune pathways, thereby leading to the reduced effectiveness of immune checkpoint inhibitors. Lastly, both in vivo and ex vivo experiments were conducted to substantiate the role of SETDB1 in HCC tumorigenesis. In conclusion, the upregulation of SETDB1 is associated with a poorer prognosis in HCC patients and inversely correlates with immune cell infiltration, potentially serving as a predictive marker for immunotherapy response. Full article
Show Figures

Graphical abstract

Graphical abstract
Full article ">Figure 1
<p>SETDB1 was upregulated in pan-cancers. (<b>A</b>) The expression of the SETDB1 gene in pan-cancers and their corresponding normal tissues from TCGA data. (<b>B</b>) The non-paired expression of SETDB1 between normal and tumor tissues. (<b>C</b>) The paired expression of SETDB1 between normal and tumor tissues. The mRNA expression of SETDB1 was higher in HCC than in normal liver tissue in the GSE-14529-GPL3921 (<b>D</b>) and GSE14520-GPL571 (<b>E</b>) datasets. (<b>F</b>) The protein levels of SETDB1 were higher in primary tumor tissues than in normal tissues in CPTAC samples. * <span class="html-italic">p</span> &lt; 0.05; ** <span class="html-italic">p</span> &lt; 0.01; *** <span class="html-italic">p</span> &lt; 0.001.</p>
Full article ">Figure 2
<p>Correlation between SETDB1 expression and the prognosis of HCC patients. (<b>A</b>) The protein expression of SETDB1 was obtained from the Human Protein Atlas. (<b>B</b>) The subcellular localization of SETDB1 in mouse primary hepatocytes and Hepa1-6 cells. The correlation of SETDB1 expression with OS (<b>C</b>) and DFS (<b>D</b>) in HCC patients. Scale bar = 20 μm.</p>
Full article ">Figure 3
<p>In vivo and ex vivo validation of the SETDB1 function in HCC. (<b>A</b>) HCC model design. FVB mice were injected with normal saline or sgP53/c-Myc/SB plasmids, respectively. (<b>B</b>) Representative liver tissues of normal (<b>left</b>) and HCC mice (<b>right</b>). (<b>C</b>) The protein expression of SETDB1 and MYC in the normal and HCC tissues. (<b>D</b>) The IHC staining of SETDB1 and MYC in HCC liver tissues. (<b>E</b>) The SETDB1 was successfully knocked out in Hepa1-6 cells. (<b>F</b>) The SETDB1 knockout inhibited colony formation in Hepa1-6 cells. (<b>G</b>) The SETDB1 knockout suppressed tumor formation in the subcutaneous xenograft tumor model.</p>
Full article ">Figure 4
<p>SETDB1 Interaction Network. PPI network of SETDB1 in the GeneMANIA (<b>A</b>) and STRING (<b>B</b>) tools. (<b>C</b>) Metascape analysis of SETDB1.</p>
Full article ">Figure 5
<p>Enrichment analysis of SETDB1 expression-correlated DEGs in HCC. (<b>A</b>) Clustering analysis heatmap of SETDB1 expression-correlated DEGs. (<b>B</b>) Volcano plot of DEGs between samples with high SETDB1 expression and low SETDB1 expression. KEGG analysis (<b>C</b>) and GO analysis (biological process) (<b>D</b>) in SETDB1 expression-correlated upregulated DEGs. KEGG analysis (<b>E</b>) and GO analysis (biological process) (<b>F</b>) in SETDB1 expression-correlated downregulated DEGs.</p>
Full article ">Figure 6
<p>Correlation analysis between SETDB1 and immune microenvironment. Correlation of SETDB1 with Immune score (<b>A</b>), Stromal score (<b>B</b>), and ESTIMATE score (<b>C</b>) in HCC. (<b>D</b>) The landscape of the relationship between SETDB1 expression and TILs in multiple types of cancers (red denotes positive correlation, and blue denotes negative correlation). (<b>E</b>) SETDB1 expression was significantly negatively associated with infiltrating levels of act_CD8, Th1, NK, and Monocyte in HCC.</p>
Full article ">Figure 7
<p>(<b>A</b>) SETDB1 expression was negatively related with CD96, CD244, CD274, CSF1R, HAVCR2, LGALS9, PDCD1LG2, and TIGIT in HCC. (<b>B</b>) Differences in the T-cell dysfunction score between the low and high expression of SETDB1 in HCC. (<b>C</b>) Differences in the T-cell rejection score between the low and high expression of SETDB1 in HCC. (<b>D</b>) Differences in the tumor immune dysfunction and exclusion (TIDE) score between the low and high expression of SETDB1 in HCC. *** <span class="html-italic">p</span> &lt; 0.001.</p>
Full article ">
17 pages, 1966 KiB  
Article
Anti-Inflammatory and Anticancer Effects of Kaurenoic Acid in Overcoming Radioresistance in Breast Cancer Radiotherapy
by Tae Woo Kim and Seong-Gyu Ko
Nutrients 2024, 16(24), 4320; https://doi.org/10.3390/nu16244320 - 14 Dec 2024
Viewed by 433
Abstract
Background/Objectives: Peroxisome proliferator–activated receptor γ (PPARγ) plays a key role in mediating anti-inflammatory and anticancer effects in the tumor microenvironment. Kaurenoic acid (KA), a diterpene compound isolated from Sphagneticola trilobata (L.) Pruski, has been demonstrated to exert anti-inflammatory, anticancer, and antihuman immunodeficiency [...] Read more.
Background/Objectives: Peroxisome proliferator–activated receptor γ (PPARγ) plays a key role in mediating anti-inflammatory and anticancer effects in the tumor microenvironment. Kaurenoic acid (KA), a diterpene compound isolated from Sphagneticola trilobata (L.) Pruski, has been demonstrated to exert anti-inflammatory, anticancer, and antihuman immunodeficiency virus effects. Methods: In this study, we identified KA as a novel activator of PPARγ with potent anti-inflammatory and antitumor effects both in vitro and in vivo. Given the potential of PPARγ regulators in overcoming radioresistance and chemoresistance in cancer therapies, we hypothesized that KA may enhance the efficacy of breast cancer radiotherapy. Results: In a lipopolysaccharide (LPS)-induced mouse inflammation model, KA treatment reduced the levels of pro-inflammatory cytokines, including COX-2, IL-6, IL-1β, and TNFα. In a xenograft mouse mode of breast cancer, KA treatment inhibited tumor growth. Specifically, KA treatment enhanced caspase-3 activity and cytotoxicity against MDA-MB-231 and MCF-7 breast cancer cells. When KA was co-treated with a caspase inhibitor, Z-VAD-FMK, caspase-dependent apoptosis was suppressed in these cells. KA was found to induce the generation of cytosolic calcium ions (Ca2+) and reactive oxygen species (ROS), triggering endoplasmic reticulum (ER) stress via the PERK-ATF4-CHOP axis. Hence, the ER stressor thapsigargin (TG) synergized with KA treatment to enhance apoptosis in these cells, while the loss of the PERK or CHOP function inhibited this phenomenon. KA treatment was shown to induce oxidative stress via the NADPH oxidase 4 (NOX4) and stimulate ROS production. Specifically, NOX4 knockdown (KD) and antioxidant treatment (N-acetyl cysteine or diphenyleneiodonium) suppressed such ER stress–mediated apoptosis by inhibiting KA-enhanced caspase-3 activity, cytotoxicity, and intracellular ROS production in the treated cells. In radioresistant MDA-MB-231R and MCF-7R cells, KA combined with 2 Gy radiation overcame radioresistance by upregulating PPARγ and modulating epithelial–mesenchymal transition (EMT) markers, such as E-cadherin, N-cadherin, and vimentin. In PPARγ KD MDA-MB-231R and MCF-7R cells, this phenomenon was inhibited due to reduced PPARγ and NOX4 expression. Conclusions: In conclusion, these findings demonstrated KA as a novel PPARγ regulator with promising potential to enhance the efficacy of breast cancer radiotherapy. Full article
(This article belongs to the Section Nutritional Immunology)
Show Figures

Figure 1

Figure 1
<p>KA-induced enhancement in PPARγ activity of kaurenoic acid (KA). (<b>A</b>) The chemical structure of KA. (<b>B</b>,<b>C</b>) Luciferase activity corresponding to the PPAR response element reporter gene. The mRNA and protein levels in 3T3-L1, MCF-7, MDA-MB-231, and SK-BR-3 cells treated with 100 µM KA, 10 µM ciglitazone, or 20 µM rosiglitazone. *, <span class="html-italic">p</span> &lt; 0.05.</p>
Full article ">Figure 2
<p>The impact of KA on the mRNA and protein levels of inflammatory cytokines in LPS-treated macrophages. (<b>A</b>) C57BL/6 mice were administered with LPS (20 mg/kg) via i.p. injection. The treatment group received KA (100 mg/kg) via i.p. injection. The survival rate of all groups (<span class="html-italic">n</span> = 10) was analyzed daily for 12 days following LPS injection. (<b>B</b>–<b>E</b>) The protein levels of IL-1β, IL-6, and TNF-α in the serum, lungs, liver, and kidneys of the treated mice, as assessed by ELISA and Western blot. (<b>F</b>–<b>H</b>) The mRNA and protein levels of IL-1β, IL-6, and TNF-α in LPS (1 µg/mL)-treated Raw264.7 and J774.1 cells in the presence or absence of KA (0, 25, 50, and 100 µM; 24 h), as assessed by ELISA, Western blot, and qRT-PCR. β-actin was used to normalize the relative mRNA and protein levels. *, <span class="html-italic">p</span> &lt; 0.05. All experiments were conducted three times.</p>
Full article ">Figure 3
<p>The in vitro and in vivo anticancer effects of KA. (<b>A</b>,<b>B</b>) LDH and WST-1 assays were conducted for cells (MCF-10A, HT-20, SK-BR-3, T-47D, HCC1419, MDA-MB-231, and MCF-7) treated with varying doses of KA (0, 10, 50, 100, 200, and 300 µM; 24 h). (<b>C</b>,<b>D</b>) The MDA-MB-231 tumor model was established by injecting 1 × 10<sup>7</sup> cells into the right dorsal flank of nude mice (<span class="html-italic">n</span> = 10 per group). KA (100 and 200 mg/kg) was administered (i.p. injection) twice weekly. The body weights of the treated mice were measured twice weekly. (<b>E</b>–<b>H</b>) MDA-MB-231 and MCF-7 cells were treated with KA for varying durations (0, 8, 16, and 24 h; 100 µM) and subjected to caspase-3, LDH cytotoxicity, and WST-1 assays. Western blot analysis of cleaved caspase-9 and -3 was also conducted following KA treatment for the indicated durations; *, <span class="html-italic">p</span> &lt; 0.05. β-actin was used as the loading control. (<b>I</b>–<b>L</b>) MCF-7 and MDA-MB-231 cells were pretreated with Z-VAD-FMK (50 μM) for 4 h before KA treatment (100 µM, 24 h). Caspase-3 activity, LDH cytotoxicity, and WST-1 assays were conducted; *, <span class="html-italic">p</span> &lt; 0.05. n.s, no significance. Western blot analysis was performed to assess the level of cleaved caspase-3. β-actin was used as the loading control.</p>
Full article ">Figure 4
<p>KA promotes the production of intracellular Ca<sup>2+</sup> and apoptosis. (<b>A</b>) MCF-7 and MDA-MB-231 cells were treated with KA for varying durations (0, 8, 16, and 24 h; 100 µM) and subjected to intracellular Ca<sup>2+</sup> assay; *, <span class="html-italic">p</span> &lt; 0.05. (<b>B</b>) The mRNA levels of CHOP, ATF4, and GRP78 were assessed by qRT-PCR. β-actin was used as the loading control. (<b>C</b>) MDA-MB-231 and MCF-7 cells were treated with KA for varying durations (0, 8, 16, and 24 h; 100 µM). Western blot analysis was conducted for the proteins associated with the ER stress signaling pathway: CHOP, ATF4, GRP78, p-eIF2α, and p-PERK. β-actin was used as the loading control. (<b>D</b>–<b>F</b>) MCF-7 and MDA-MB-231 cells were treated with 3 μM TG and 100 µM KA for 24 h. LDH cytotoxicity, intracellular Ca<sup>2+</sup>, and cell viability assays were conducted; *, <span class="html-italic">p</span> &lt; 0.05. (<b>G</b>,<b>H</b>) The mRNA levels of CHOP and ATF4 were assessed by qRT-PCR. Western blot analysis was conducted to examine the protein levels of p-eIF2α, p-PERK, CHOP, and ATF4 in MDA-MB-231 and MCF-7 cells treated with 100 µM KA and 3 µM TG for 24 h. β-actin was used as the loading control.</p>
Full article ">Figure 5
<p>PERK silencing inhibits KA-induced apoptosis in breast cancer cells. (<b>A</b>–<b>E</b>) PERK siRNA was transfected into MDA-MB-231 and MCF-7 cells. The cells were then treated with 100 µM KA for 24 h. Intracellular Ca<sup>2+</sup>, caspase-3 activity, LDH cytotoxicity, and WST-1 assays were conducted; *, <span class="html-italic">p</span> &lt; 0.05. N.S, no significance. Western blot analysis was conducted to assess the protein levels of cleaved caspase-3, CHOP, ATF4, p-eIF2α, and p-PERK in MDA-MB-231 and MCF-7 cells treated with 100 µM KA for 24 h. β-actin was used as the loading control. (<b>F</b>–<b>J</b>) CHOP siRNA was transfected into MDA-MB-231 and MCF-7 cells. The cells were then treated with 100 µM KA for 24 h. Intracellular Ca<sup>2+</sup>, caspase-3 activity, LDH cytotoxicity, and WST-1 assays were conducted; *, <span class="html-italic">p</span> &lt; 0.05. Western blot analysis was conducted to assess the protein levels of cleaved caspase-3, CHOP, and DR5 in MDA-MB-231 and MCF-7 cells treated with 100 µM KA for 24 h. β-actin was used as the loading control.</p>
Full article ">Figure 6
<p>NOX4 silencing inhibits ROS-induced ER stress and apoptosis in KA-treated breast cancer. (<b>A</b>) MDA-MB-231 and MCF-7 cells were treated with 100 µM KA for the indicated durations and subjected to intracellular ROS assay DCFDA; *, <span class="html-italic">p</span> &lt; 0.05. (<b>B</b>–<b>E</b>) MDA-MB-231 and MCF-7 cells were treated with 100 µM KA, 1 µM DPI, and 100 µM NAC for 24 h. Caspase-3 activity, LDH cytotoxicity, intracellular ROS, and WST-1 assays were conducted; *, <span class="html-italic">p</span> &lt; 0.05. N.S, no significance. (<b>F</b>–<b>I</b>) NOX4 siRNA was transfected into MDA-MB-231 and MCF-7 cells and then treated with 100 µM KA for 24 h. Intracellular ROS, WST-1, and LDH cytotoxicity assays were conducted; *, <span class="html-italic">p</span> &lt; 0.05. N.S, no significance. Western blot analysis was conducted to assess the protein levels of cleaved caspase-3, CHOP, NOX4, PERK, and p-PERK in MDA-MB-231 and MCF-7 cells treated with 100 µM KA for 24 h. β-actin was used as the loading control.</p>
Full article ">Figure 7
<p>Radiation combined with KA overcomes radioresistance in breast cancer cells. (<b>A</b>) Colony formation analysis was performed for MDA-MB-231, MCF-7, MDA-MB-231R, and MCF-7R cells following radiation at varying intensities (0, 2, 4, and 6 Gy) and/or KA treatment. The cell survival rate was quantified; *, <span class="html-italic">p</span> &lt; 0.05. (<b>B</b>–<b>D</b>) MDA-MB-231, MCF-7, MDA-MB-231R, and MCF-7R cells treated with 100 µM KA and 2 Gy radiation for 24 h were subjected to caspase-3 activity, WST-1, and LDH cytotoxicity assays; *, <span class="html-italic">p</span> &lt; 0.05. n.s, no significance. (<b>E</b>) The mRNA levels of vimentin, N-cadherin, and E-cadherin were assessed by qRT-PCR in MDA-MB-231, MCF-7, MDA-MB-231R, and MCF-7R cells treated with 100 µM KA and 2 Gy radiation for 24 h; *, <span class="html-italic">p</span> &lt; 0.05. n.s, no significance. β-actin was used the loading control.</p>
Full article ">Figure 8
<p>PPARγ silencing inhibits KA + radiation-induced apoptosis in MDA-MB-231R and MCF-7R cells. (<b>A</b>–<b>F</b>) MDA-MB-231R and MCF-7R cells were treated with PPAR<b>γ</b> shRNA particles for gene knockdown. These cells were treated with 100 µM KA and 2 Gy radiation for 24 h and subjected to LDH cytotoxicity, intracellular Ca<sup>2+</sup> and ROS, caspase-3 activity, and WST-1 assays. Western blot analysis was conducted to examine the protein levels of cleaved caspase-3, PERK, CHOP, PPARγ, NOX4, and p-PERK; * <span class="html-italic">p</span> &lt; 0.05. N.S, no significance. β-actin was used as the loading control.</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 394
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 ">
19 pages, 7612 KiB  
Article
Identification of a Potential PGK1 Inhibitor with the Suppression of Breast Cancer Cells Using Virtual Screening and Molecular Docking
by Xianghui Chen, Zanwen Zuo, Xianbin Li, Qizhang Li and Lei Zhang
Pharmaceuticals 2024, 17(12), 1636; https://doi.org/10.3390/ph17121636 - 5 Dec 2024
Viewed by 454
Abstract
Background/Objectives: Breast cancer is the second most common malignancy worldwide and poses a significant threat to women’s health. However, the prognostic biomarkers and therapeutic targets of breast cancer are unclear. A prognostic model can help in identifying biomarkers and targets for breast cancer. [...] Read more.
Background/Objectives: Breast cancer is the second most common malignancy worldwide and poses a significant threat to women’s health. However, the prognostic biomarkers and therapeutic targets of breast cancer are unclear. A prognostic model can help in identifying biomarkers and targets for breast cancer. In this study, a novel prognostic model was developed to optimize treatment, improve clinical prognosis, and screen potential phosphoglycerate kinase 1 (PGK1) inhibitors for breast cancer treatment. Methods: Using data from the Gene Expression Omnibus (GEO) database, differentially expressed genes (DEGs) were identified in normal individuals and breast cancer patients. The biological functions of the DEGs were examined using bioinformatics analysis. A novel prognostic model was then constructed using the DEGs through LASSO and multivariate Cox regression analyses. The relationship between the prognostic model, survival, and immunity was also evaluated. In addition, virtual screening was conducted based on the risk genes to identify novel small molecule inhibitors of PGK1 from Chemdiv and Targetmol libraries. The effects of the potential inhibitors were confirmed through cell experiments. Results: A total of 230 up- and 325 down-regulated DEGs were identified in HER2, LumA, LumB, and TN breast cancer subtypes. A new prognostic model was constructed using ten risk genes. The analysis from The Cancer Genome Atlas (TCGA) indicated that the prognosis was poorer in the high-risk group compared to the low-risk group. The accuracy of the model was confirmed using the ROC curve. Furthermore, functional enrichment analyses indicated that the DEGs between low- and high-risk groups were linked to the immune response. The risk score was also correlated with tumor immune infiltrates. Moreover, four compounds with the highest score and the lowest affinity energy were identified. Notably, D231-0058 showed better inhibitory activity against breast cancer cells. Conclusions: Ten genes (ACSS2, C2CD2, CXCL9, KRT15, MRPL13, NR3C2, PGK1, PIGR, RBP4, and SORBS1) were identified as prognostic signatures for breast cancer. Additionally, results showed that D231-0058 (2-((((4-(2-methyl-1H-indol-3-yl)-1,3-thiazol-2-yl)carbamoyl)methyl)sulfanyl)acetic acid) may be a novel candidate for treating breast cancer. Full article
(This article belongs to the Section Pharmacology)
Show Figures

Figure 1

Figure 1
<p>Identification and functional enrichment analysis of DEGs. (<b>A</b>–<b>D</b>) Top 20 up-regulated and down-regulated genes in HER2 (<b>A</b>), LumA (<b>B</b>), LumB (<b>C</b>), and TN (<b>D</b>) subtype tumor samples from GSE29431, GSE38959, GSE45827, GSE65194, and GSE115275 datasets. The red color represents up-regulated genes, while green indicates down-regulated genes. The numbers shown in the figure represent the log fold change (logFC) of genes in each dataset. The cutoff criteria are <span class="html-italic">p</span> &lt; 0.05 and |logFC| &gt; 0.5. (<b>E</b>) The Venn diagram of DEGs of HER2, LumA, LumB, and TN subtype tumor samples from GSE29431, GSE38959, GSE45827, GSE65194, and GSE115275 datasets. (<b>F</b>) The bar plot of GO functional enrichment analysis. The top 10 terms of biological process, cellular component, and molecular function are shown. (<b>G</b>) The bar plot illustrates the results of KEGG functional enrichment analysis.</p>
Full article ">Figure 2
<p>Analysis of the prognostic model in BC. (<b>A</b>) Forest plot of the signature risk model. (<b>B</b>) Lasso model for screening the key genes. (<b>C</b>) Multivariate Cox analysis confirming hub genes for risk model. (<b>D</b>) The expression levels of ten hub genes in breast cancer tissues compared to normal tissues. (<b>E</b>,<b>G</b>,<b>I</b>) Kaplan–Meier analysis of survival differences between high-risk and low-risk groups in training (<b>E</b>), test (<b>G</b>), and entire (<b>I</b>) sets. (<b>F</b>,<b>H</b>,<b>J</b>) Receiver operating characteristic (ROC) curve analysis on the ten model gene signatures in the training (<b>F</b>), test (<b>H</b>), and entire (<b>J</b>) sets. AUC, the area under the curve. These curves are performed by R package survival ROC. (<b>K</b>) Univariate Cox analysis of risk score and clinicopathological features in the entire set. (<b>L</b>) Multivariate Cox analysis of clinicopathological features and risk score in the entire set. (<b>M</b>) The ROC curve of the risk score and clinical characteristics. (<b>N</b>) The ROC curve and AUC values for the predictive signature at 1-year, 3-year, and 5-year survival rates.</p>
Full article ">Figure 3
<p>Analysis of the relationships between risk score and clinical characteristics of breast cancer in the TCGA cohort. (<b>A</b>) Heat map of ten model genes and clinical characteristics in the high- and low-risk groups. *, <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. (<b>B</b>) Analysis of overall survival in TCGA-BC patients based on clinical stratification, focusing on high- and low-risk groups by age, clinical stage, N stage, and T stage.</p>
Full article ">Figure 4
<p>The nomogram in predicting overall survival of breast cancer. (<b>A</b>) The nomogram predicts 1-, 3-, and 5-year overall survival. (<b>B</b>) Calibration maps were utilized to predict survival rates at 1, 3, and 5 years.</p>
Full article ">Figure 5
<p>Analysis of functional enrichment across different risk groups. (<b>A</b>) Volcano chart of differentially expressed genes; (<b>B</b>) GO analysis explored the potential function in terms of biological process (BP), cellular component (CC), and molecular function (MF); (<b>C</b>) KEGG analysis showed the potential pathway enrichment; (<b>D</b>) GSEA analysis demonstrated the potential activated and suppressed pathway enrichment in the high-risk group compared with the low-risk group.</p>
Full article ">Figure 6
<p>Immune features analysis in risk groups. (<b>A</b>,<b>B</b>) ssGSEA (single-sample gene set enrichment analysis) scores for immune cells (<b>A</b>) and immune function (<b>B</b>) in TCGA cohort. (<b>C</b>) The expression of immune checkpoint-related genes and the correlation between risk scores. aDCs, activated dendritic cells; APC, antigen-presenting cell; CCR, chemokine receptor; HLA, human leukocyte antigen; iDCs, immature dendritic cells; IFN, interferon; MHC, major histocompatibility complex; NK, natural killer; pDCs, plasmacytoid dendritic cells; Tfh, T follicular helper; Th, T helper cell; TIL, tumor-infiltrating lymphocyte; Treg, T regulatory cell. * <span class="html-italic">p</span> &lt; 0.05; ** <span class="html-italic">p</span> &lt; 0.01; *** <span class="html-italic">p</span> &lt; 0.001; ns, non-significant.</p>
Full article ">Figure 7
<p>Three-dimensional interaction between PGK1 (2X13) and D715-2871 (<b>A</b>), Y040-8304 (<b>B</b>), D715-0344 (<b>C</b>), and D231-0058 (<b>D</b>). Yellow dotted lines represent hydrogen bonds, pinkish-red dotted lines represent salt bridges, and green balls depict magnesium ions.</p>
Full article ">Figure 8
<p>Inhibitory activity of D715-2871, Y040-8304, D715-0344, and D231-0058 against breast cancer cells T-47D and MCF-7. (<b>A</b>) CCK8 assay for cell viability. Cancer cells were treated with D715-2871, Y040-8304, D715-0344, or D231-0058 (0, 0.1, 1, 10, and 100 μg/mL) for 24 h. Data were presented as mean ± SD (<span class="html-italic">n</span> = 6). (<b>B</b>) CCK8 assay for cell viability. Cancer cells were treated with D231-0058 (0, 1, 3, 10, 30, and 100 μg/mL) for 24 and 48 h. Data were presented as mean ± SD (<span class="html-italic">n</span> = 6). (<b>C</b>) Microscopic observation of the cells treated with D231-0058 (0, 1, 3, 10, 30, and 100 μg/mL) for 24 h.</p>
Full article ">
16 pages, 3149 KiB  
Article
GW0742 as a Potential TRα and TRβ Antagonist Reduces the Viability and Metabolic Activity of an Adult Granulosa Tumour Cell Line and Simultaneously Upregulates TRβ Expression
by Justyna Gogola-Mruk, Izabela Kumor, Gabriela Wojtaszek, Karolina Kulig and Anna Ptak
Cancers 2024, 16(23), 4069; https://doi.org/10.3390/cancers16234069 - 5 Dec 2024
Viewed by 450
Abstract
Background/Objectives: Clinical studies have demonstrated a correlation between alterations in the expression level of TRα and TRβ receptors in ovarian cancer cells and overall survival. Celecoxib and GW0742, commonly known as a COX-2 inhibitor and a PPARβ/δ agonist, are novel thyroid hormone receptor [...] Read more.
Background/Objectives: Clinical studies have demonstrated a correlation between alterations in the expression level of TRα and TRβ receptors in ovarian cancer cells and overall survival. Celecoxib and GW0742, commonly known as a COX-2 inhibitor and a PPARβ/δ agonist, are novel thyroid hormone receptor antagonists that bind to TRβ or both TRα and TRβ. Methods: The study was conducted on a non-luteinized ovarian granulosa cell line (HGrC1) and two rare ovarian cancer cell lines (COV434 and KGN). The expression of TRα and TRβ at the gene and protein levels was examined by real-time PCR and Western blot, respectively. The impact of GW0742 and celecoxib on the cell viability of the HGrC1, COV434 and KGN lines was evaluated using the PrestoBlue™ Cell Viability Reagent. The metabolic activity of the cells was analysed using the Seahorse XFp Analyzer. Results: Initially, we observed that the gene and protein expression levels of TRα and TRβ were higher in COV434 and KGN cells than in HGrC1 cells. Subsequently, it was demonstrated that T3 enhances the viability of HGrC1, COV434 and KGN cells. Furthermore, autoregulatory feedback loops were not observed during TRα or TRβ signalling in ovarian cancer cells, in contrast to the findings in healthy granulosa cells. Finally, we demonstrated that GW0742 reduced the viability and metabolic activity of granulosa cell tumours (GCTs). Simultaneously, we observed that GW0742 upregulated the expression of TRβ in GCT. Conclusions: These findings suggest that GW0742 may be a novel adjuvant therapy for GCTs expressing TRα and TRβ. Full article
(This article belongs to the Special Issue Advances in Ovarian Cancer Research and Treatment)
Show Figures

Figure 1

Figure 1
<p>Basal expression profile of TRα and TRβ. Basal mRNA and protein expression of TRα (<span class="html-italic">THRA</span>) (<b>A</b>,<b>B</b>) and TRβ (<span class="html-italic">THRB)</span> (<b>C</b>,<b>D</b>) in HGrC1, COV434 and KGN cells. The expression level of <span class="html-italic">THRA</span> and <span class="html-italic">THRB</span> in HGrC1 cells was set to 1.0 RQ. Each bar represents the mean ± SD of three independent experiments. Statistically significant differences are denoted by mean values not sharing letters (<span class="html-italic">p</span> ≤ 0.05). The uncropped bolts are shown in <a href="#app1-cancers-16-04069" class="html-app">Supplementary Materials</a>.</p>
Full article ">Figure 2
<p>Effect of triiodothyronine (T3) on the viability and metabolic activity of non-luteinised ovarian granulosa cells. Dose-dependent effects of T3 (0.1, 1, 10 and 100 nM) on the viability of HGrC1 cells after 24 h (<b>A</b>) and 48 h (<b>B</b>) of treatment. C, control (0.01% DMSO, dimethyl sulfoxide). RFU, relative fluorescence units. Effect of T3 (1 and 10 nM) on the expression of TRα and TRβ mRNA in HGrC1 cells (<b>C</b>) after 24 h. mRNA expression by vehicle-treated cells was set to 1.0. RQ, relative quantity. Total ATP production rates in HGrC1 cells after stimulation of T3 (10 nM) (<b>D</b>). ATP production rates from glycolysis respiration (glycoATP) (<b>E</b>) and mitochondrial (mitoATP) (<b>F</b>) in HGrC1 cells after treatment with T3 (10 nM). Oxygen consumption rates (OCRs) (<b>G</b>) and extra cellular acidification rates (ECARs) (<b>H</b>) during the challenge with T3 (10 nM) in HGrC1 cells. Data represent the mean ± SD of three independent experiments. * <span class="html-italic">p</span> &lt; 0.05, ** <span class="html-italic">p</span> &lt; 0.01.</p>
Full article ">Figure 3
<p>The effect of Triiodothyronine (T3) on viability and metabolic activity rare types of ovarian cancer cells. Dose-dependent effects of T3 (0.1, 1, 10 and 100 nM) on the viability of COV434 and KGN cells after 24 h (<b>A</b>,<b>D</b>) and 48 h (<b>B</b>,<b>E</b>) of treatment. C, control (0.01% DMSO; dimethyl sulfoxide). RFU, relative fluorescence units. Effect of T3 (1 and 10 nM) on the expression of TRα and TRβ mRNA in COV434 cells (<b>C</b>) and KGN cells (<b>F</b>) after 24 h. mRNA expression by vehicle-treated cells was set to 1.0. RQ, relative quantity. Total ATP production rates in KGN after stimulation with T3 (10 nM) (<b>G</b>). ATP production rates from glycolysis respiration (glycoATP) (<b>H</b>) and mitochondrial (mitoATP) (<b>I</b>) in KGN cells after treatment with T3 (10 nM). Oxygen consumption rates (OCRs) (<b>J</b>) and extra cellular acidification rates (ECAR) (<b>K</b>) during challenge with T3 (10 nM) in KGN cells. Data represent the mean ± SD of three independent experiments. * <span class="html-italic">p</span> &lt; 0.05, ** <span class="html-italic">p</span> &lt; 0.01.</p>
Full article ">Figure 4
<p>The effect of GW0724 and celecoxib on the viability and/or metabolic activity of non-luteinised ovarian granulosa cells. Dose-dependent effects of celecoxib (0.1–15 µM) (<b>A</b>) and GW0742 (0.001–25 µM) (<b>B</b>) on the viability of HGrC1 cells after 24 and 48 h. C, control (0.01% DMSO; dimethyl sulfoxide). RFU, relative fluorescence units. Effects of celecoxib (10 µM) (<b>C</b>) and GW0742 (10 µM) (<b>D</b>) on mRNA expression of TRα, TRβ in HGrC1 cells after 24 h. mRNA expression in vehicle-treated cells was set to 1.0. RQ, relative quantity. ATP production rates of HGrC1 after stimulation with GW0742 (10 µM) (<b>E</b>). ATP production rates from glycolysis respiration (glycoATP) (<b>F</b>) and mitochondrial respiration (mitoATP) (<b>G</b>) in HGrC1 cells after treatment with GW0742 (10 µM). Oxygen consumption rates (OCRs) (<b>H</b>) and extra cellular acidification rates (ECARs) (<b>I</b>) during the challenge with GW0742 (10 µM) in HGrC1 cells. Data represent the mean ± SD of three independent experiments. * <span class="html-italic">p</span> &lt; 0.05, ** <span class="html-italic">p</span> &lt; 0.01 and *** <span class="html-italic">p</span> &lt; 0.001.</p>
Full article ">Figure 5
<p>The effect of GW0724 and celecoxib on the viability and/or metabolic activity of a granulosa cell tumour (GCT). Dose-dependent effects of celecoxib (0.1–15 µM) (<b>A</b>,<b>D</b>) and GW0742 (0.001–25 µM) (<b>B</b>,<b>E</b>) after 24 and 48 h treatment on the viability of COV434 and KGN cells, respectively. C, control (0.01% DMSO; dimethyl sulfoxide). RFU, relative fluorescence units. Effects of GW0742 (10 µM) (<b>C</b>,<b>F</b>) on the mRNA expression of TRα and TRβ in COV434 and KGN cells after 24 h, respectively. mRNA expression in vehicle-treated cells was set to 1.0. RQ, relative quantity. ATP production rates of KGN cells after stimulation with GW0742 (10 µM) (<b>G</b>). ATP production rates from glycolytic respiration (glycoATP) (<b>H</b>) and mitochondrial (mitoATP) (<b>I</b>) in KGN cells after treatment with GW0742 (10 µM). Oxygen consumption rates (OCRs) (<b>J</b>) and extra cellular acidification rates (ECARs) (<b>K</b>) during the challenge with GW0742 (10 µM) in KGN cells. Data represent the mean ± SD of three independent experiments. * <span class="html-italic">p</span> &lt; 0.05, ** <span class="html-italic">p</span> &lt; 0.01 and *** <span class="html-italic">p</span> &lt; 0.001.</p>
Full article ">Figure 6
<p>The effect of GW074 and celecoxib on the viability and metabolic activity in HGrC1, COV434 and KGN stimulated by T<sub>3</sub>. The effects of the addition of triiodothyronine (T<sub>3</sub>) (10 nM) in combination with pretreatment with the TRα and TRβ antagonists (GW0742) (10 µM) or a TRβ antagonist (celecoxib) (10 µM) for 48 h on the viability of HGrC1 (<b>A</b>), COV434 (<b>B</b>) and KGN cells (<b>C</b>). C, control (0.01% DMSO; dimethyl sulfoxide); RFU, relative fluorescence units. Synergism quotient (SQ) values were calculated to determine if there was any synergistic activity in the treatment with GW0742 or celecoxib with T<sub>3</sub>. ATP production rates of KGN cells after stimulation with T<sub>3</sub> (10 nM), GW0742 (10 µM) and with T<sub>3</sub> (10 nM) in combination with the pretreatment with GW0742 (10 µM) (<b>D</b>). ATP production rates from glycolysis respiration (glycoATP) (<b>E</b>) and mitochondrial respiration (mitoATP) (<b>F</b>) in KGN cells after treatment with T<sub>3</sub> (10 nM), GW0742 (10 µM) and with T<sub>3</sub> (10 nM) in combination with the pretreatment with GW0742 (10 µM). Oxygen consumption rates (OCRs) (<b>G</b>) and extra cellular acidification rates (ECARs) (<b>H</b>) during the challenge with T<sub>3</sub> (10 nM), GW0742 (10 µM) and with T<sub>3</sub> (10 nM) in combination with the pretreatment with GW0742 (10 µM) in KGN cells. Data represent the mean ± SD of three independent experiments. * <span class="html-italic">p</span> &lt; 0.05, ** <span class="html-italic">p</span> &lt; 0.01 and *** <span class="html-italic">p</span> &lt; 0.001.</p>
Full article ">
18 pages, 8859 KiB  
Article
Cinnamaldehyde-Mediated Suppression of MMP-13, COX-2, and IL-6 Through MAPK and NF-κB Signaling Inhibition in Chondrocytes and Synoviocytes Under Inflammatory Conditions
by Jaishree Sankaranarayanan, Seok Cheol Lee, Hyung Keun Kim, Ju Yeon Kang, Sree Samanvitha Kuppa and Jong Keun Seon
Int. J. Mol. Sci. 2024, 25(23), 12914; https://doi.org/10.3390/ijms252312914 - 30 Nov 2024
Viewed by 654
Abstract
Inflammatory disorders encompass a range of conditions, including osteoarthritis (OA), characterized by the body’s heightened immune response to diverse stimuli. OA is a prevalent degenerative joint disease characterized by the progressive deterioration of joint cartilage and subchondral bone, leading to pain, limited mobility, [...] Read more.
Inflammatory disorders encompass a range of conditions, including osteoarthritis (OA), characterized by the body’s heightened immune response to diverse stimuli. OA is a prevalent degenerative joint disease characterized by the progressive deterioration of joint cartilage and subchondral bone, leading to pain, limited mobility, and physical disability. Synovitis, the inflammation of the synovial membrane, is increasingly recognized as a critical factor in OA pathogenesis and progression. This study evaluates the therapeutic potential of cinnamaldehyde (CA), a bioactive compound derived from cinnamon, on synovial and articular inflammation in OA. Given CA’s established anti-inflammatory, antioxidant, and antibacterial properties, this research explores its specific impact on OA and synovitis. The cytotoxicity of CA was assessed using a CCK-8 assay in human IL-1β pretreated chondrocytes and synoviocytes, which serve as in vitro models of OA and synovitis. The study further examined the effects of CA on the expression of proinflammatory cytokines, including IL-6, COX-2, and TNF-α, utilizing multiple analytical techniques. Additionally, the production of matrix metalloproteinases (MMP-3 and MMP-13) and the activation of the NF-κB signaling pathway, particularly the phosphorylation of p65 (pp65), were investigated. The role of the NF-κB inhibitor 5HPP-33 and its downstream effects on gene expression, including COX-2 and IL-6, as well as the MAPK pathway components (p38, ERK, and JNK), were also explored. An MEK inhibitor (U0126) was employed to assess its downstream impact on COX-2 and IL-6 expressions. The results demonstrated that CA significantly inhibited the expression of proinflammatory cytokines and suppressed NF-κB activation in IL-1β pretreated chondrocytes and synoviocytes. These findings suggest that CA, in a dose-dependent manner, may serve as an effective therapeutic agent for preventing OA and synovitis, offering valuable insights into its potential role in managing synovial inflammation and OA. Full article
Show Figures

Figure 1

Figure 1
<p>(<b>A</b>) Cell cytotoxicity of chondrocytes and (<b>B</b>) cell cytotoxicity of synoviocytes with the treatment of cinnamaldehyde in a dose-dependent manner (0.5, 5, 50, and 500 μM) for 24 h. (<b>C</b>) Cell viability of chondrocytes and (<b>D</b>) cell viability of synoviocytes with the treatment IL-1β prior to cinnamaldehyde in a dose-dependent manner (0.5, 5, 50, and 500 μM) for 24 h. Data are presented as the mean ± standard deviation (<span class="html-italic">n</span> = 3). ns = non-significant, * <span class="html-italic">p</span> &lt; 0.01, *** <span class="html-italic">p</span> &lt; 0.001, and **** <span class="html-italic">p</span> &lt; 0.0001 compared with the control group.</p>
Full article ">Figure 2
<p>(<b>A</b>) Cell microscopy of chondrocytes followed by confocal microscopy of COL-II surface marker expression, and (<b>B</b>) cell microscopy of synoviocytes followed by confocal microscopy of CD 86 surface marker expression at 40X magnification (scale bar = 100 μM). (DAPI (blue)-Nucleus, Phalloidin (red)-F-actin, and COL-II and CD-86 (green)).</p>
Full article ">Figure 3
<p>(<b>A</b>–<b>C</b>) Protein expression of MAPK pathways (p38, pERK, and pJNK) in chondrocytes followed by qualitative analysis and confocal microscopy at 40× magnification in a dose-dependent manner (DAPI (blue)-Nucleus, Phalloidin (red)-F-actin, and p38, pERK, and pJNK (green)). (<b>D</b>–<b>F</b>) Protein expression of MAPK pathways (p38, pERK, and pJNK) in synoviocytes followed by qualitative analysis and confocal microscopy at 40× magnification in a dose-dependent manner (DAPI (blue)-Nucleus, Phalloidin (red)-F-actin, and p38, pERK, and pJNK (green)) (scale bar = 100 μM). Data are presented as the mean ± standard deviation (<span class="html-italic">n</span> = 3). ns = non-significant, * <span class="html-italic">p</span> &lt; 0.05, ** <span class="html-italic">p</span> &lt; 0.01, *** <span class="html-italic">p</span> &lt; 0.001, and **** <span class="html-italic">p</span> &lt; 0.0001 compared with the control group.</p>
Full article ">Figure 3 Cont.
<p>(<b>A</b>–<b>C</b>) Protein expression of MAPK pathways (p38, pERK, and pJNK) in chondrocytes followed by qualitative analysis and confocal microscopy at 40× magnification in a dose-dependent manner (DAPI (blue)-Nucleus, Phalloidin (red)-F-actin, and p38, pERK, and pJNK (green)). (<b>D</b>–<b>F</b>) Protein expression of MAPK pathways (p38, pERK, and pJNK) in synoviocytes followed by qualitative analysis and confocal microscopy at 40× magnification in a dose-dependent manner (DAPI (blue)-Nucleus, Phalloidin (red)-F-actin, and p38, pERK, and pJNK (green)) (scale bar = 100 μM). Data are presented as the mean ± standard deviation (<span class="html-italic">n</span> = 3). ns = non-significant, * <span class="html-italic">p</span> &lt; 0.05, ** <span class="html-italic">p</span> &lt; 0.01, *** <span class="html-italic">p</span> &lt; 0.001, and **** <span class="html-italic">p</span> &lt; 0.0001 compared with the control group.</p>
Full article ">Figure 4
<p>(<b>A</b>–<b>D</b>) Gene expression and protein expression of inflammatory mediators in chondrocytes, followed by qualitative analysis. (<b>E</b>–<b>H</b>) Gene expression and protein expression of inflammatory mediators in synoviocytes, followed by qualitative analysis. Data are presented as the mean ± standard deviation (<span class="html-italic">n</span> = 3). ns = non-significant, * <span class="html-italic">p</span> &lt; 0.05, ** <span class="html-italic">p</span> &lt; 0.01, *** <span class="html-italic">p</span> &lt; 0.001, and **** <span class="html-italic">p</span> &lt; 0.0001 compared with the control group.</p>
Full article ">Figure 4 Cont.
<p>(<b>A</b>–<b>D</b>) Gene expression and protein expression of inflammatory mediators in chondrocytes, followed by qualitative analysis. (<b>E</b>–<b>H</b>) Gene expression and protein expression of inflammatory mediators in synoviocytes, followed by qualitative analysis. Data are presented as the mean ± standard deviation (<span class="html-italic">n</span> = 3). ns = non-significant, * <span class="html-italic">p</span> &lt; 0.05, ** <span class="html-italic">p</span> &lt; 0.01, *** <span class="html-italic">p</span> &lt; 0.001, and **** <span class="html-italic">p</span> &lt; 0.0001 compared with the control group.</p>
Full article ">Figure 5
<p>(<b>A</b>) Protein expression of NF-κB pathways (pp65) in chondrocytes in a dose-dependent manner, followed by qualitative analysis. (<b>B</b>) Protein expression of NF-κB pathways (pp65) in the presence and absence of NF-κB inhibitor (5HPP-33) in chondrocytes. (<b>C</b>) Comparison of the downstream mediators of inflammation (COX-2, IL-6) in the presence of an inhibitor and 50 μM of CA in chondrocytes, followed by qualitative analysis. (<b>D</b>) Confocal microscopy (pp65) at 40× magnification in the presence and absence of NF-κB inhibitor (5HPP-33) in chondrocytes (Figure) (DAPI (blue)-Nucleus, Phalloidin (red)-F-actin, and pp65 (green)). (<b>E</b>) Protein expression of NF-κB pathways (pp65) in synoviocytes in a dose-dependent manner, followed by qualitative analysis. (<b>F</b>) Protein expression of NF-κB pathways (pp65) in the presence and absence of NF-κB inhibitor (5HPP-33) in synoviocytes. (<b>G</b>) Comparison of the downstream mediators of inflammation (COX-2, IL-6) in the presence of NF-κB inhibitor and 50 μM of CA in synoviocytes, followed by qualitative analysis. (<b>H</b>) Confocal microscopy (pp65) at 40× magnification in the presence and absence of NF-κB inhibitor (5HPP-33) in synoviocytes. (DAPI (blue)-Nucleus, Phalloidin (red)-F-actin, and pp65 (green)). (Scale bar = 100 μM). Data are presented as the mean ± standard deviation (<span class="html-italic">n</span> = 3). ns = non-significant, ** <span class="html-italic">p</span> &lt; 0.01, *** <span class="html-italic">p</span> &lt; 0.001, and **** <span class="html-italic">p</span> &lt; 0.0001 compared with the control group.</p>
Full article ">Figure 6
<p>(<b>A</b>) Protein expression of MAPK pathways (pERK phosphorylation) in the presence of MEK1/2 inhibitor (U0126) in chondrocytes, followed by qualitative analysis. (<b>B</b>) Comparison of the downstream mediators of inflammation (COX-2, IL-6) in the presence of MEK 1/2 inhibitor and 50 μM of CA in synoviocytes, followed by qualitative analysis. (<b>C</b>) Protein expression pERK phosphorylation in the presence of MEK1/2 inhibitor (U0126) in synoviocytes, followed by qualitative analysis. (<b>D</b>) Comparison of the downstream mediators of inflammation (COX-2, IL-6) in the presence of MEK 1/2 inhibitor and 50 μM of CA in synoviocytes, followed by qualitative analysis. The figure represents the activity of CA and Inhibitor role in the ERK pathway inhibition. Data are presented as the mean ± standard deviation (<span class="html-italic">n</span> = 3). ns = non-significant, * <span class="html-italic">p</span> &lt; 0.05, *** <span class="html-italic">p</span> &lt; 0.001, and **** <span class="html-italic">p</span> &lt; 0.0001 compared with the control group.</p>
Full article ">Figure 7
<p>A schematic explanation of the mechanism by which cinnamaldehyde (CA) upregulates <span class="html-italic">p</span>-NF-κB (<span class="html-italic">p</span>-p65) expression in chondrocytes and synoviocytes because of IL-1β stimulation. IL-1β alters the transcriptional activity of MAPK by decreasing the expression of downstream mediators such as IL-6, MMP13, and COX 2. The effect of CA is exerted through a reduction in pNF-κB (<span class="html-italic">p</span>-p65) expression and the pMAPK (pp38, pERK, pJNK) expression in both cell types. (→ denotes the process of activation and ⊥ denotes the inhibition of the activity).</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 572
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 528
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 567
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 ">
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 541
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 ">
13 pages, 2077 KiB  
Article
Prognostic Role of Inflammatory and Nutritional Biomarkers in Non-Small-Cell Lung Cancer Patients Treated with Immune Checkpoint Inhibitors Alone or in Combination with Chemotherapy as First-Line
by Antonello Veccia, Mariachiara Dipasquale, Stefania Kinspergher and Orazio Caffo
Cancers 2024, 16(22), 3871; https://doi.org/10.3390/cancers16223871 - 19 Nov 2024
Viewed by 675
Abstract
Introduction: In recent years, several inflammation-related factors and nutritional parameters have been evaluated to develop prognostic scores as potential biomarkers in non-small-cell lung cancer (NSCLC) patients receiving immune checkpoint inhibitors (ICIs). The aim of this study was to retrospectively investigate the prognostic role [...] Read more.
Introduction: In recent years, several inflammation-related factors and nutritional parameters have been evaluated to develop prognostic scores as potential biomarkers in non-small-cell lung cancer (NSCLC) patients receiving immune checkpoint inhibitors (ICIs). The aim of this study was to retrospectively investigate the prognostic role of the advanced lung cancer inflammation (ALI) index, lung immune prognostic index (LIPI), prognostic nutritional index (PNI) and systemic inflammation score (SIS) in metastatic NSCLC patients receiving ICI alone or in combination with chemotherapy. Methods and patients: We retrospectively included 191 patients with advanced NSCLC who received first-line ICI with or without chemotherapy from 2017 to 2024. The association between pretreatment ALI, LIPI, PNI, and SIS and overall survival (OS) was evaluated using the Kaplan–Meier method and Cox regression models. Results: After a median follow-up of 27.7 months, significantly longer OS was associated with an ALI score > 18 vs. ≤18 (18.0 vs. 7.3 months; p = 0.00111), LIPI score 0 vs. 1 and 2 [18.9 vs. 8.2 and 4.2 months; (p = 0.001)], PNI ≥ 45 vs. <45 (22.7 vs. 9.6 months; p = 0.002), and SIS score 0 vs. 1 and 2 (27.4 vs. 7.1 and 8.6 months, respectively; p < 0.001). The OS benefit was independent of treatment (ICI vs. ICI + chemotherapy). At multivariate analysis, pretreatment albumin was positively associated with OS, while ECOG PS 1 and liver metastases were negatively associated with OS. Conclusions: Inflammatory and nutritional biomarkers such as the ALI, LIPI, PNI, and SIS represent useful tools to prognosticate survival in metastatic lung cancer patients treated with ICI alone or in combination with chemotherapy as first-line. Full article
(This article belongs to the Special Issue Feature Papers in Section "Cancer Biomarkers" in 2023–2024)
Show Figures

Figure 1

Figure 1
<p>OS according to ALI score.</p>
Full article ">Figure 2
<p>OS according to LIPI score.</p>
Full article ">Figure 3
<p>OS according to PNI score.</p>
Full article ">Figure 4
<p>OS according to SIS score.</p>
Full article ">
18 pages, 9624 KiB  
Article
Galangin Triggers Eryptosis and Hemolysis Through Ca2+ Nucleation and Metabolic Collapse Mediated by PKC/CK1α/COX/p38/Rac1 Signaling Axis
by Mohammad A. Alfhili, Sumiah A. Alghareeb, Ghada A. Alotaibi and Jawaher Alsughayyir
Int. J. Mol. Sci. 2024, 25(22), 12267; https://doi.org/10.3390/ijms252212267 - 15 Nov 2024
Viewed by 602
Abstract
Anticancer drugs cause anemia in patients through eryptosis and hemolysis. We thus studied the in vitro toxicity of galangin (GAL) in red blood cells (RBCs). RBCs were exposed to 50–500 μM of GAL and analyzed for markers of eryptosis and hemolysis. Ca2+ [...] Read more.
Anticancer drugs cause anemia in patients through eryptosis and hemolysis. We thus studied the in vitro toxicity of galangin (GAL) in red blood cells (RBCs). RBCs were exposed to 50–500 μM of GAL and analyzed for markers of eryptosis and hemolysis. Ca2+ nucleation, phosphatidylserine (PS) externalization, oxidative stress, and cell size were detected via fluorescence-activated cell sorting using Fluo4/AM, annexin-V-FITC, 2′,7′-dichlorodihydrofluorescein diacetate, and forward scatter (FSC), respectively. Acetylcholinesterase (AChE) activity was measured via Ellman’s assay and ultrastructural morphology was examined via scanning electron microscopy. Membrane rupture and extracellular hemoglobin, aspartate transaminase (AST), and lactate dehydrogenase (LDH) were assessed via colorimetric methods. Distinct experiments were carried out to identify protective agents and signaling pathways using small-molecule inhibitors. GAL triggered sucrose-sensitive hemolysis with AST and LDH leakage, increased annexin-V-FITC and Fluo4 fluorescence, and decreased FSC and AChE activity which was associated with the formation of granulated echinocytes. Ca2+ omission and energy replenishment with glucose, adenine, and guanosine blunted PS externalization and preserved cellular volume. Moreover, caffeine, Trolox, heparin, and uric acid had similar ameliorative effects. Hemolysis was abrogated via caffeine, Trolox, heparin, mannitol, lactate, melatonin, and PEG 8000. Notably, co-treatment of cells with GAL and staurosporin, D4476, or acetylsalicylic acid prevented PS externalization whereas only the presence of SB203580 and NSC23766 rescued the cells from GAL-induced hemolysis. Ca2+ nucleation and metabolic collapse mediated by PKC/CK1α/COX/p38/Rac1 drive GAL-induced eryptosis and hemolysis. These novel findings carry ramifications for the clinical prospects of GAL in anticancer therapy. Full article
(This article belongs to the Special Issue Erythrocyte Cell Death: Molecular Insights)
Show Figures

Figure 1

Figure 1
<p>Experimental design. Prepared with BioRender.</p>
Full article ">Figure 2
<p>Eryptotic and hemolytic activities of GAL. (<b>a</b>) Chemical structure of GAL. (<b>b</b>) Original histograms of annexin-V-FITC fluorescence. (<b>c</b>) Percentage of eryptotic cells. (<b>d</b>) Percentage of hemolytic cells. (<b>e</b>) AST activity. (<b>f</b>) LDH activity. (<b>g</b>) CK activity. (<b>h</b>) K<sup>+</sup> levels. (<b>i</b>) Correlation between eryptosis and hemolysis. (<b>j</b>) Osmotic fragility curves. (<b>k</b>) AChE activity. (<b>l</b>) B<sub>12</sub> levels. (<b>m</b>) ESR. Graphs show means ± SD. * (<span class="html-italic">p</span> &lt; 0.05), ** (<span class="html-italic">p</span> &lt; 0.01), *** (<span class="html-italic">p</span> &lt; 0.001), and **** (<span class="html-italic">p</span> &lt; 0.0001).</p>
Full article ">Figure 3
<p>GAL causes loss of cellular volume and Ca<sup>2+</sup> nucleation. (<b>a</b>) Original histograms of FSC signals. (<b>b</b>) Original histograms of Fluo4 fluorescence. (<b>c</b>) Percentage of cell shrinkage. (<b>d</b>) Percentage with increased Ca<sup>2+</sup>. (<b>e</b>) Original histograms of annexin-V-FITC with and without Ca<sup>2+</sup>. (<b>f</b>) Original histograms of FSC with and without Ca<sup>2+</sup>. (<b>g</b>) Percentage of eryptotic cells. (<b>h</b>) Percentage of cell shrinkage. (<b>i</b>) Percentage of hemolyzed cells. (<b>j</b>) Original histograms of annexin-V-FITC in 5 and 125 mM KCl. (<b>k</b>) Original histograms of FSC in 5 and 125 mM KCl. (<b>l</b>) Percentage of eryptotic cells. (<b>m</b>) Percentage of cell shrinkage. (<b>n</b>) Percentage of hemolyzed cells. Graphs show means ± SD. No significance is indicated by ns whereas * (<span class="html-italic">p</span> &lt; 0.05), ** (<span class="html-italic">p</span> &lt; 0.01), *** (<span class="html-italic">p</span> &lt; 0.001), and **** (<span class="html-italic">p</span> &lt; 0.0001).</p>
Full article ">Figure 4
<p>Ultrastructural morphology of RBCs. GAL induces the formation of granulated echinocytes. Magnification: ×5000. Scale bar: 1 μm.</p>
Full article ">Figure 5
<p>Energy replenishment reverses GAL-induced cytotoxicity. (<b>a</b>) Original annexin-V-FITC histograms in 5 and 50 mM glucose. (<b>b</b>) Original FSC histograms in 5 and 50 mM glucose. (<b>c</b>) Percentage of eryptotic cells. (<b>d</b>) Percentage of cell shrinkage. (<b>e</b>) Percentage of hemolyzed cells. (<b>f</b>) Original annexin-V-FITC histograms with and without lactate. (<b>g</b>) Original FSC histograms with and without lactate. (<b>h</b>) Percentage of eryptotic cells. (<b>i</b>) Percentage of cell shrinkage. (<b>j</b>) Percentage of hemolyzed cells. (<b>k</b>) Original annexin-V-FITC histograms with and without adenine. (<b>l</b>) Original FSC histograms with and without adenine. (<b>m</b>) Percentage of eryptotic cells. (<b>n</b>) Percentage of cell shrinkage. (<b>o</b>) Percentage of hemolyzed cells. (<b>p</b>) Original annexin-V-FITC histograms with and without guanosine. (<b>q</b>) Original FSC histograms with and without guanosine. (<b>r</b>) Percentage of eryptotic cells. (<b>s</b>) Percentage of cell shrinkage. (<b>t</b>) Percentage of hemolyzed cells. Graphs show means ± SD. No significance is indicated by ns whereas * (<span class="html-italic">p</span> &lt; 0.05), *** (<span class="html-italic">p</span> &lt; 0.001) and **** (<span class="html-italic">p</span> &lt; 0.0001).</p>
Full article ">Figure 6
<p>Antieryptotic and antihemolytic effects of Trolox, uric acid, and ASA. (<b>a</b>) Original annexin-V-FITC histograms with and without Trolox. (<b>b</b>) Original FSC histograms with and without Trolox. (<b>c</b>) Percentage of eryptotic cells. (<b>d</b>) Percentage of cell shrinkage. (<b>e</b>) Percentage of hemolyzed cells. (<b>f</b>) Original annexin-V-FITC histograms with and without uric acid. (<b>g</b>) Original FSC histograms with and without uric acid. (<b>h</b>) Percentage of eryptotic cells. (<b>i</b>) Percentage of cell shrinkage. (<b>j</b>) Percentage of hemolyzed cells. (<b>k</b>) Original annexin-V-FITC histograms with and without ASA. (<b>l</b>) Original FSC histograms with and without ASA. (<b>m</b>) Percentage of eryptotic cells. (<b>n</b>) Percentage of cell shrinkage. (<b>o</b>) Percentage of hemolyzed cells. Graphs show means ± SD. No significance is indicated by ns whereas ** (<span class="html-italic">p</span> &lt; 0.01) and **** (<span class="html-italic">p</span> &lt; 0.0001).</p>
Full article ">Figure 7
<p>Ameliorative effects of caffeine and heparin. (<b>a</b>) Original annexin-V-FITC histograms with and without caffeine. (<b>b</b>) Original FSC histograms with and without caffeine. (<b>c</b>) Percentage of eryptotic cells. (<b>d</b>) Percentage of cell shrinkage. (<b>e</b>) Percentage of hemolyzed cells. (<b>f</b>) Original annexin-V-FITC histograms with and without heparin. (<b>g</b>) Original FSC histograms with and without heparin. (<b>h</b>) Percentage of eryptotic cells. (<b>i</b>) Percentage of cell shrinkage. (<b>j</b>) Percentage of hemolyzed cells. Graphs show means ± SD **** (<span class="html-italic">p</span> &lt; 0.0001).</p>
Full article ">Figure 8
<p>Inhibitors of GAL-induced eryptosis. (<b>a</b>) Original annexin-V-FITC histograms with and without staurosporin. (<b>b</b>) Original FSC histograms with and without staurosporin. (<b>c</b>) Percentage of eryptotic cells. (<b>d</b>) Percentage of cell shrinkage. (<b>e</b>) Percentage of hemolyzed cells. (<b>f</b>) Original annexin-V-FITC histograms with and without D4476. (<b>g</b>) Original FSC histograms with and without D4476. (<b>h</b>) Percentage of eryptotic cells. (<b>i</b>) Percentage of cell shrinkage. (<b>j</b>) Percentage of hemolyzed cells. Graphs show means ± SD. No significance is indicated by ns whereas * (<span class="html-italic">p</span> &lt; 0.05), ** (<span class="html-italic">p</span> &lt; 0.01), *** (<span class="html-italic">p</span> &lt; 0.001), and **** (<span class="html-italic">p</span> &lt; 0.0001).</p>
Full article ">Figure 9
<p>Inhibitors of GAL-induced hemolysis. (<b>a</b>) Percentage of eryptotic cells, (<b>b</b>) shrinkage, and (<b>c</b>) hemolysis with and without SB203580. (<b>d</b>) Percentage of eryptotic cells, (<b>e</b>) shrinkage, and (<b>f</b>) hemolysis with and without NSC23766. (<b>g</b>) Percentage of eryptotic cells, (<b>h</b>) shrinkage, and (<b>i</b>) hemolysis with and without MTN. (<b>j</b>) Effect of GAL on hemolysis with and without PEG. Graphs show means ± SD. No significance is indicated by ns whereas * (<span class="html-italic">p</span> &lt; 0.05), *** (<span class="html-italic">p</span> &lt; 0.001), and **** (<span class="html-italic">p</span> &lt; 0.0001).</p>
Full article ">Figure 10
<p>Analysis of antioxidants and signaling pathways. Hemolytic rates in the presence and absence of (<b>a</b>) L-NAME, (<b>b</b>) vitamin C, (<b>c</b>) GSH, (<b>d</b>) Z-VAD-FMK, (<b>e</b>) myriocin, (<b>f</b>) BAPTA-AM, (<b>g</b>) necrostatin-2, (<b>h</b>) NSA, and (<b>i</b>) ATP. Graphs show means ± SD. No significance is indicated by ns whereas **** (<span class="html-italic">p</span> &lt; 0.0001).</p>
Full article ">Figure 11
<p>GAL toxicity under hyperosmotic stress. (<b>a</b>) Original annexin-V-FITC histograms with and without urea. (<b>b</b>) Original FSC histograms with and without urea. (<b>c</b>) Percentage of eryptotic cells. (<b>d</b>) Percentage of cell shrinkage. (<b>e</b>) Percentage of hemolyzed cells. (<b>f</b>) Original annexin-V-FITC histograms with and without mannitol. (<b>g</b>) Original FSC histograms with and without mannitol. (<b>h</b>) Percentage of eryptotic cells. (<b>i</b>) Percentage of cell shrinkage. (<b>j</b>) Percentage of hemolyzed cells. (<b>k</b>) Original annexin-V-FITC histograms with and without sucrose. (<b>l</b>) Original FSC histograms with and without sucrose. (<b>m</b>) Percentage of eryptotic cells. (<b>n</b>) Percentage of cell shrinkage. (<b>o</b>) Percentage of hemolyzed cells. Graphs show means ± SD. No significance is indicated by ns whereas * (<span class="html-italic">p</span> &lt; 0.05), *** (<span class="html-italic">p</span> &lt; 0.001), and **** (<span class="html-italic">p</span> &lt; 0.0001).</p>
Full article ">Figure 12
<p>A working model of GAL-induced RBC death. Prepared with BioRender.</p>
Full article ">
17 pages, 1611 KiB  
Article
Sertraline as a Multi-Target Modulator of AChE, COX-2, BACE-1, and GSK-3β: Computational and In Vivo Studies
by Minhajul Arfeen and Vasudevan Mani
Molecules 2024, 29(22), 5354; https://doi.org/10.3390/molecules29225354 - 14 Nov 2024
Viewed by 841
Abstract
Alzheimer’s disease (AD) is a neurodegenerative disorder associated with the dysregulation of several key enzymes, including acetylcholinesterase (AChE), cyclooxygenase-2 (COX-2), glycogen synthase kinase 3β (GSK-3β), β-site amyloid precursor protein cleaving enzyme 1 (BACE-1), and caspase-3. In this study, machine learning algorithms such as [...] Read more.
Alzheimer’s disease (AD) is a neurodegenerative disorder associated with the dysregulation of several key enzymes, including acetylcholinesterase (AChE), cyclooxygenase-2 (COX-2), glycogen synthase kinase 3β (GSK-3β), β-site amyloid precursor protein cleaving enzyme 1 (BACE-1), and caspase-3. In this study, machine learning algorithms such as Random Forest (RF), Gradient Boost (GB), and Extreme Gradient Boost (XGB) were employed to screen US-FDA approved drugs from the ZINC15 database to identify potential dual inhibitors of COX-2 and AChE. The models were trained using molecules obtained from the ChEMBL database, with 5039 molecules for AChE and 3689 molecules for COX-2. Specifically, 1248 and 3791 molecules were classified as active and inactive for AChE, respectively, while 858 and 2831 molecules were classified as active and inactive for COX-2. The three machine learning models achieved prediction accuracies ranging from 92% to 95% for both AChE and COX-2. Virtual screening of US-FDA drugs from the ZINC15 database identified sertraline (SETL) as a potential dual inhibitor of AChE and COX-2. Further docking studies of SETL in the active sites of AChE and COX-2, as well as BACE-1, GSK-3β, and caspase-3, revealed strong binding affinities for all five proteins. In vivo validation was conducted using a lipopolysaccharide (LPS)-induced rat model pretreated with SETL for 30 days. The results demonstrated a significant decrease in the levels of AChE (p < 0.001), BACE-1 (p < 0.01), GSK-3β (p < 0.05), and COX-2 (p < 0.05). Additionally, the downstream effects were evaluated, showing significant decreases in the apoptosis marker caspase-3 (p < 0.05) and the oxidative stress marker malondialdehyde (MDA) (p < 0.001), indicating that SETL is clinically localized in its effectiveness, mitigating both enzymatic activity and the associated pathological changes of cognitive impairment and AD. Full article
Show Figures

Figure 1

Figure 1
<p>Workflow utilized in this report.</p>
Full article ">Figure 2
<p>Chemical space analysis of training and test sets. The molecular weight and MlogP define the chemical space considered in this study.</p>
Full article ">Figure 3
<p>Hydrogen (green) and hydrophobic interactions (pink and purple) of SETL in the active sites of AChE, COX-2, GSK-3β, BACE-1, and caspase-3.</p>
Full article ">Figure 4
<p>Effect of sertraline (SETL) on (<b>A</b>) AChE, (<b>B</b>) BACE-1, (<b>C</b>) GSK-3β, (<b>D</b>) COX-2, (<b>E</b>) caspase-3, and (<b>F</b>) MDA in LPS-induced neurotoxicity rats. The results are expressed as mean ± SEM, with a sample size of n = 6. One-way ANOVA followed by Tukey–Kramer multiple comparisons test. * <span class="html-italic">p</span> &lt; 0.05 and *** <span class="html-italic">p</span> &lt; 0.001 as compared to the control group; <sup>#</sup> <span class="html-italic">p</span> &lt; 0.05, <sup>##</sup> <span class="html-italic">p</span> &lt; 0.01, and <sup>###</sup> <span class="html-italic">p</span> &lt; 0.001 as compared to the LPS group.</p>
Full article ">
21 pages, 2568 KiB  
Review
Exploring COX-Independent Pathways: A Novel Approach for Meloxicam and Other NSAIDs in Cancer and Cardiovascular Disease Treatment
by Lixia Cheng, Zhenghui Hu, Jiawei Gu, Qian Li, Jiahao Liu, Meiling Liu, Jie Li and Xiaowen Bi
Pharmaceuticals 2024, 17(11), 1488; https://doi.org/10.3390/ph17111488 - 6 Nov 2024
Viewed by 845
Abstract
As a fundamental process of innate immunity, inflammation is associated with the pathologic process of various diseases and constitutes a prevalent risk factor for both cancer and cardiovascular disease (CVD). Studies have indicated that several non-steroidal anti-inflammatory drugs (NSAIDs), including Meloxicam, may prevent [...] Read more.
As a fundamental process of innate immunity, inflammation is associated with the pathologic process of various diseases and constitutes a prevalent risk factor for both cancer and cardiovascular disease (CVD). Studies have indicated that several non-steroidal anti-inflammatory drugs (NSAIDs), including Meloxicam, may prevent tumorigenesis, reduce the risk of carcinogenesis, improve the efficacy of anticancer therapies, and reduce the risk of CVD, in addition to controlling the body’s inflammatory imbalances. Traditionally, most NSAIDs work by inhibiting cyclooxygenase (COX) activity, thereby blocking the synthesis of prostaglandins (PGs), which play a role in inflammation, cancer, and various cardiovascular conditions. However, long-term COX inhibition and reduced PGs synthesis can result in serious side effects. Recent studies have increasingly shown that some selective COX-2 inhibitors and NSAIDs, such as Meloxicam, may exert effects beyond COX inhibition. This emerging understanding prompts a re-evaluation of the mechanisms by which NSAIDs operate, suggesting that their benefits in cancer and CVD treatment may not solely depend on COX targeting. In this review, we will explore the potential COX-independent mechanisms of Meloxicam and other NSAIDs in addressing oncology and cardiovascular health. Full article
(This article belongs to the Section Pharmacology)
Show Figures

Figure 1

Figure 1
<p>Overview of NSAIDs targeting COX in response to tumor and CVD pathways. As a substrate for the action of cyclooxygenase, ARA is catalyzed by phospholipases from membrane phospholipids. COX is a key enzyme in the metabolism of ARA, with two isoforms, structural (COX-1) and inducible (COX-2), and NSAIDs block prostaglandin synthesis, which is involved in cancer and CVD, through inhibition of the enzymatic activity of COX. Abbreviations: Phospholipases A<sub>2</sub>, PLA<sub>2</sub>; Prostaglandin H2 synthase, PGH2; Thromboxane A<sub>2,</sub> TXA<sub>2</sub>; Prostaglandins (respective receptors): prostaglandins E<sub>2</sub> (PGE<sub>2</sub>), prostaglandins F<sub>2</sub> (PGF<sub>2</sub>), prostaglandins D<sub>2</sub> (PGD<sub>2</sub>), and prostaglandins I<sub>2</sub> (PGI<sub>2</sub>).</p>
Full article ">Figure 2
<p>Protein targets of Meloxicam and other NSAIDs against cancer. (A) NSAIDs inhibit the phosphorylation of AXL. (B) NSAIDs promote ubiquitinated degradation of AXL. (C) NSAIDs inhibit the deacetylase activity of SIRT1. (D,E) The protein expression and phosphorylation of STAT3 were inhibited by NSAIDs. (F,G) NSAIDs inhibit the phosphorylation of mTOR, in part by activating the AMPK pathway. (H) NSAIDs inhibit the enzymatic activity of Neu-1.</p>
Full article ">Figure 3
<p>Meloxicam and other NSAIDs mediate cell behavior. Drugs marked in red font represent that the drug is a facilitator of a cellular behavior; while drugs marked in green font express an inhibitory effect on a cellular behavior.</p>
Full article ">Figure 4
<p>Effects of Meloxicam and other NSAIDs on activation and transduction of NF-κB, MAPKs, and Wnt/β-Catenin signaling pathways.</p>
Full article ">
Back to TopTop