[go: up one dir, main page]

 
 
Sign in to use this feature.

Years

Between: -

Subjects

remove_circle_outline
remove_circle_outline
remove_circle_outline
remove_circle_outline
remove_circle_outline
remove_circle_outline
remove_circle_outline
remove_circle_outline
remove_circle_outline

Journals

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

Article Types

Countries / Regions

remove_circle_outline
remove_circle_outline
remove_circle_outline
remove_circle_outline
remove_circle_outline
remove_circle_outline
remove_circle_outline
remove_circle_outline
remove_circle_outline

Search Results (2,443)

Search Parameters:
Keywords = curcumin

Order results
Result details
Results per page
Select all
Export citation of selected articles as:
22 pages, 19388 KiB  
Article
Network Pharmacology Approaches Used to Identify Therapeutic Molecules for Chronic Venous Disease Based on Potential miRNA Biomarkers
by Oscar Salvador Barrera-Vázquez, Juan Luis Escobar-Ramírez and Gil Alfonso Magos-Guerrero
J. Xenobiot. 2024, 14(4), 1519-1540; https://doi.org/10.3390/jox14040083 - 15 Oct 2024
Viewed by 349
Abstract
Chronic venous disease (CVD) is a prevalent condition in adults, significantly affecting the global elderly population, with a higher incidence in women than in men. The modulation of gene expression through microRNA (miRNA) partly regulated the development of cardiovascular disease (CVD). Previous research [...] Read more.
Chronic venous disease (CVD) is a prevalent condition in adults, significantly affecting the global elderly population, with a higher incidence in women than in men. The modulation of gene expression through microRNA (miRNA) partly regulated the development of cardiovascular disease (CVD). Previous research identified a functional analysis of seven genes (CDS2, HDAC5, PPP6R2, PRRC2B, TBC1D22A, WNK1, and PABPC3) as targets of miRNAs related to CVD. In this context, miRNAs emerge as essential candidates for CVD diagnosis, representing novel molecular and biological knowledge. This work aims to identify, by network analysis, the miRNAs involved in CVD as potential biomarkers, either by interacting with small molecules such as toxins and pollutants or by searching for new drugs. Our study shows an updated landscape of the signaling pathways involving miRNAs in CVD pathology. This latest research includes data found through experimental tests and uses predictions to propose both miRNAs and genes as potential biomarkers to develop diagnostic and therapeutic methods for the early detection of CVD in the clinical setting. In addition, our pharmacological network analysis has, for the first time, shown how to use these potential biomarkers to find small molecules that may regulate them. Between the small molecules in this research, toxins, pollutants, and drugs showed outstanding interactions with these miRNAs. One of them, hesperidin, a widely prescribed drug for treating CVD and modulating the gene expression associated with CVD, was used as a reference for searching for new molecules that may interact with miRNAs involved in CVD. Among the drugs that exhibit the same miRNA expression profile as hesperidin, potential candidates include desoximetasone, curcumin, flurandrenolide, trifluridine, fludrocortisone, diflorasone, gemcitabine, floxuridine, and reversine. Further investigation of these drugs is essential to improve the treatment of cardiovascular disease. Additionally, supporting the clinical use of miRNAs as biomarkers for diagnosing and predicting CVD is crucial. Full article
Show Figures

Figure 1

Figure 1
<p>Network analysis of CVD-associated miRNAs with their expression, sources, countries of origin, and detection methods. The network depicts the interconnected structure of miRNAs derived from CVD patients, organized by their expression, sources, countries of origin, and detection methods. In the network, miRNAs are denoted as blue (upregulated) or red (downregulated) nodes, green nodes represent countries, yellow nodes represent sources, and gray nodes represent detection methods. The connections between the nodes signify the frequency of independent study reports. The three most outstanding sources were the proximal part of the significant saphenous vein tissue, vein tissues, and peripheral blood mononuclear cells. China was the country with the most available finds from miRNAs. Microarrays and RT-PCR are the most effective methods for diagnosing CVD. At least two tissues are expected to contain five specific miRNAs: miR-34a, miR-34c, miR-202-3, miR-1202, and miR-130a. The network, constructed using Cytoscape software (v.3.10.2), comprises 78 nodes and 193 edges, with a diameter and a network density of 6 and 0.106, respectively. Please refer to the <a href="#app1-jox-14-00083" class="html-app">Supplementary Materials (Figure S1)</a> for a better image resolution.</p>
Full article ">Figure 2
<p>Network analysis of miRNAs associated with CVD and their predicted targets. (<b>A</b>) The bar plot visually presents the number of targetable genes in the miRNA curated dataset. In the plot, gray bars represent upregulated genes, blue bars denote downregulated genes, and orange bars indicate genes with undetermined expression (ND). (<b>B</b>) We constructed a structural network using reported and predicted interactions between miRNAs and their targeted genes. This network consists of 1882 nodes and 5267 edges, with a diameter and a network density of 12 and 0.001. The network was created using Cytoscape software (v.3.10.2). Please refer to the <a href="#app1-jox-14-00083" class="html-app">Supplementary Materials (Figure S2)</a> for a better image resolution.</p>
Full article ">Figure 3
<p>The structural network represents the ten most connected nodes, including miRNAs and targets. Nodes are color-coded from orange to yellow based on their degree of connection, representing the most connected genes and miRNAs in the network. The most relevant nodes in this network are WNK-1, hsa-miR-106b-3p, IL1NR, hsa-miR-92a-3p, PPP6R2, hsa-miR-454-3p, PRRC2B, hsa-miR-548ac, hsa-miR-128-3p, and ADIPOQ. The network consists of 921 nodes and 1256 edges, with a diameter and a network density of 7 and 0.003, respectively. This network was created using Cytoscape software (v.3.10.2). Please refer to the <a href="#app1-jox-14-00083" class="html-app">Supplementary Materials (Figure S3)</a> for a better image resolution.</p>
Full article ">Figure 4
<p>The structural network of small molecules, phlebotonic, and miRNAs. The structural network depicts the targeted miRNAs (green nodes) between small molecules (yellow nodes) and the phlebotonic hesperidin (pink node). It was observed that curcumin exhibited the highest connection of miRNAs with the reported phlebotonic hesperidin. The network involves 246 nodes and 1153 edges, with a diameter and a network density of 6 and 0.038, respectively. The network construction utilized Cytoscape software (v.3.10.2). Please refer to the <a href="#app1-jox-14-00083" class="html-app">Supplementary Materials (Figure S4)</a> for a better image resolution.</p>
Full article ">Figure 5
<p>Structural network of hesperidin-specific miRNAs that are also altered by small molecules. This structural network depicts the miRNA profiles shared by both miRNAs that are specifically upregulated or downregulated by the reference compound hesperidin and the small molecules studied in this work. These shared miRNA profiles between hesperidin and small molecules allow the selection of potential candidates for CVD treatment. Downregulated miRNAs are shown in red, upregulated in blue, and those shared with small molecules in orange. The network involves 93 nodes and 320 edges, with a diameter and a density of 4 and 0.075, respectively. This network was created using Cytoscape software (v.3.10.2). Please refer to the <a href="#app1-jox-14-00083" class="html-app">Supplementary Materials (Figure S5)</a> for a better image resolution.</p>
Full article ">Scheme 1
<p>Flow diagram of the bibliographical screening performed for this research.</p>
Full article ">
14 pages, 1891 KiB  
Article
Impact of κ-Carrageenan on the Freshwater Mussel (Solenaia oleivora) Protein Emulsion Gels: Gel Formation, Stability, and Curcumin Delivery
by Wanwen Chen, Wu Jin, Xueyan Ma, Haibo Wen, Gangchun Xu, Pao Xu and Hao Cheng
Gels 2024, 10(10), 659; https://doi.org/10.3390/gels10100659 - 14 Oct 2024
Viewed by 218
Abstract
Protein-based emulsion gels are an ideal delivery system due to their unique structure, remarkable encapsulation efficiency, and tunable digestive behavior. Freshwater mussel (Solenaia oleivora) protein isolate (SoPI), an emerging sustainable protein with high nutritional value, possesses unique value in the development [...] Read more.
Protein-based emulsion gels are an ideal delivery system due to their unique structure, remarkable encapsulation efficiency, and tunable digestive behavior. Freshwater mussel (Solenaia oleivora) protein isolate (SoPI), an emerging sustainable protein with high nutritional value, possesses unique value in the development of functional foods. Herein, composite emulsion gels were fabricated with SoPI and κ-carrageenan (κ-CG) for the delivery of curcumin. SoPI/κ-CG stabilized emulsions possessed a high encapsulation efficiency of curcumin with a value of around 95%. The addition of κ-CG above 0.50% facilitated the emulsion gel formation and significantly improved the gel strength with 1326 g. Furthermore, the storage and digestive stability of curcumin were significantly improved as the κ-CG concentration increased. At 1.50% κ-CG, around 80% and 90% curcumin remained after 21-day storage at 45 °C and the 6 h in vitro gastrointestinal digestion, respectively. The addition of 0.50% κ-CG obtained the highest bioaccessibility of curcumin (~60%). This study illustrated the potential of SoPI emulsion gels as a carrier for stabilizing and delivering hydrophobic polyphenols. Full article
(This article belongs to the Section Gel Applications)
Show Figures

Figure 1

Figure 1
<p>Effect of κ-carrageenan concentration on the size distribution (<b>A</b>) and ζ-potential (<b>B</b>) of SoPI-stabilized emulsions.</p>
Full article ">Figure 2
<p>Encapsulation efficiency of curcumin in SoPI-stabilized emulsions at various κ-carrageenan concentrations.</p>
Full article ">Figure 3
<p>Appearance image of curcumin-loaded SoPI emulsion gels with different κ-carrageenan concentrations.</p>
Full article ">Figure 4
<p>Effects of κ-carrageenan concentration on the water-holding capacity of SoPI emulsion gels. Different letters indicate the statistically significant difference (<span class="html-italic">p</span> &lt; 0.05).</p>
Full article ">Figure 5
<p>Retention of curcumin encapsulated within SoPI emulsion gels at various κ-carrageenan concentrations during storage for 21 days at 20 °C (<b>A</b>) and 45 °C (<b>B</b>).</p>
Full article ">Figure 6
<p>The release of free fatty acid from SoPI emulsion gels at various concentrations of κ-carrageenan during in vitro intestinal digestion.</p>
Full article ">Figure 7
<p>Retention (<b>A</b>) and bioaccessibility (<b>B</b>) of curcumin in SoPI emulsion gels at various concentrations of κ-carrageenan during in vitro digestion. Different letters indicate the statistically significant difference (<span class="html-italic">p</span> &lt; 0.05).</p>
Full article ">
32 pages, 11462 KiB  
Article
Selective Inhibition of Deamidated Triosephosphate Isomerase by Disulfiram, Curcumin, and Sodium Dichloroacetate: Synergistic Therapeutic Strategies for T-Cell Acute Lymphoblastic Leukemia in Jurkat Cells
by Luis A. Flores-López, Ignacio De la Mora-De la Mora, Claudia M. Malagón-Reyes, Itzhel García-Torres, Yoalli Martínez-Pérez, Gabriela López-Herrera, Gloria Hernández-Alcántara, Gloria León-Avila, Gabriel López-Velázquez, Alberto Olaya-Vargas, Saúl Gómez-Manzo and Sergio Enríquez-Flores
Biomolecules 2024, 14(10), 1295; https://doi.org/10.3390/biom14101295 - 13 Oct 2024
Viewed by 436
Abstract
T-cell acute lymphoblastic leukemia (T-ALL) is a challenging childhood cancer to treat, with limited therapeutic options and high relapse rates. This study explores deamidated triosephosphate isomerase (dTPI) as a novel therapeutic target. We hypothesized that selectively inhibiting dTPI could reduce T-ALL cell viability [...] Read more.
T-cell acute lymphoblastic leukemia (T-ALL) is a challenging childhood cancer to treat, with limited therapeutic options and high relapse rates. This study explores deamidated triosephosphate isomerase (dTPI) as a novel therapeutic target. We hypothesized that selectively inhibiting dTPI could reduce T-ALL cell viability without affecting normal T lymphocytes. Computational modeling and recombinant enzyme assays revealed that disulfiram (DS) and curcumin (CU) selectively bind and inhibit dTPI activity without affecting the non-deamidated enzyme. At the cellular level, treatment with DS and CU significantly reduced Jurkat T-ALL cell viability and endogenous TPI enzymatic activity, with no effect on normal T lymphocytes, whereas the combination of sodium dichloroacetate (DCA) with DS or CU showed synergistic effects. Furthermore, we demonstrated that dTPI was present and accumulated only in Jurkat cells, confirming our hypothesis. Finally, flow cytometry confirmed apoptosis in Jurkat cells after treatment with DS and CU or their combination with DCA. These findings strongly suggest that targeting dTPI represents a promising and selective target for T-ALL therapy. Full article
(This article belongs to the Section Enzymology)
Show Figures

Figure 1

Figure 1
<p>Molecular docking analysis of n-dTPI and dTPI crystallographic structures. Conformational representations of DS docked at the interface of n-dTPI (<b>a</b>) and dTPI (<b>b</b>) are shown. The proximity of Cys residues near the interface is highlighted, suggesting potential interaction sites for DS. Conformational representations of CU docked at the interface of n-dTPI (<b>c</b>) and dTPI (<b>d</b>) are shown; the hydrophobic surfaces of proteins are highlighted in red. A deeper penetration of both DS and CU into the interface of the deamidated enzyme (dTPI) compared to the non-deamidated form (n-dTPI) is observed. Figures modeled with PyMOL version 2.5.0 (Schrödinger Inc., New York, NY, USA).</p>
Full article ">Figure 2
<p>Inactivation assays of recombinant n-dTPI and dTPI enzymes. Both enzymes (0.2 mg/mL) were incubated for 2 h at 37 °C with gradually increasing concentrations of the respective compounds: (<b>a</b>) DS (0 to 1000 μM) and (<b>b</b>) CU (0 to 1500 μM). Following incubation, aliquots were taken for enzyme activity measurement as described in the Methods section. The enzymatic activity was normalized, and 100% corresponds to the assays in the absence of the compound. Filled black squares represent n-dTPI activity, while filled red squares represent dTPI activity. The results represent the average of three independent experiments, with error bars indicating the variation observed in the experiments.</p>
Full article ">Figure 3
<p>Extrinsic fluorescence spectra of TPIs after incubation with DS and CU. Enzymes (0.2 mg/mL) were incubated without or with 250 μM DS and 1500 μM CU for 2 h at 37 °C. Following incubation, excess compounds were removed, and extrinsic fluorescence was measured in the presence of 100 μM ANS with excitation at 395 nm. (<b>a</b>) shows the spectra in the absence of any compounds (control). (<b>b</b>,<b>c</b>) show the spectra after incubation with DS and CU, respectively. n-dTPI (black line) and dTPI (red line). The results represent the mean of three independent experiments.</p>
Full article ">Figure 4
<p>Effects of DS and CU on cell viability and TPI activity in normal T-cell lymphocytes and Jurkat cells. Cells (1 × 10<sup>5</sup> per well) were incubated with increasing concentrations of DS and CU. Following incubation, cell viability was assessed using MTT assays and endogenous TPI activity was determined by enzymatic activity assays. (<b>a</b>,<b>c</b>) normal T lymphocytes, (<b>b</b>,<b>d</b>) Jurkat cells. Results are expressed as percentages relative to the untreated control set to 100%. The results represent the average of three independent experiments, with error bars indicating the variation observed in the experiments. Statistical differences were analyzed using one-way ANOVA with Tukey’s post-hoc test, with a significance level set at <span class="html-italic">p</span> = 0.01 (**).</p>
Full article ">Figure 5
<p>Effects of combined DCA and CU treatment on cell viability and TPI activity in normal T lymphocytes and Jurkat cells. Cells (1 × 10<sup>5</sup> per well) were pre-treated with 12 mM DCA for 24 h at 37 °C, followed by exposure to increasing concentrations of DS or CU for an additional 24 h. MTT and enzyme activity assays were then performed to assess cell viability and endogenous TPI activity, respectively. (<b>a</b>,<b>c</b>) normal T lymphocytes and (<b>b</b>,<b>d</b>) Jurkat cells. Results are expressed as percentages relative to the untreated control group (set to 100%). The results represent the average of three independent experiments, with error bars indicating the variation observed in the experiments. Statistical differences were analyzed using one-way ANOVA with Tukey’s post-hoc test, with a significance level set at <span class="html-italic">p</span> = 0.01 (**).</p>
Full article ">Figure 6
<p>Western blot analysis of TPI isoforms in normal T lymphocytes and Jurkat cells. In (<b>a</b>) and (<b>b</b>), lanes 1, 2, and 3, each loaded with 1 μg of protein, serve as migration standards for recombinant n-dTPI, dTPI, and ddTPI, respectively. Lanes 4–6 in (<b>a</b>) contain total protein extracts from normal T lymphocytes, while lanes 7–9 in (<b>a</b>) and lanes 4–10 in (<b>b</b>) correspond to protein extracts from Jurkat cells, with 100 μg of protein loaded per lane. For normal T lymphocytes, lane 4 is the control (untreated), lane 5 is treated with 250 μM DS, and lane 6 is treated with 1500 μM CU. In Jurkat cells, lanes 7, 8, and 9 in (<b>a</b>) represent the control (untreated), treatment with 250 μM DS, and treatment with 1500 μM CU, respectively. In (<b>b</b>), lane 4 is the control condition (untreated), while lanes 6 and 7 were treated with 100 and 250 μM DS, respectively. Finally, lanes 8, 9, and 10 were treated with 500, 1000, and 1500 μM of CU, respectively. The positive and negative poles of the gel are indicated on the left side of each panel. Full-length (uncropped) blots for panels (<b>a</b>,<b>b</b>) are presented in <a href="#app1-biomolecules-14-01295" class="html-app">Supplementary Figure S5</a>.</p>
Full article ">Figure 7
<p>Western blot analysis of TPI isoforms in normal T-cell lymphocytes and Jurkat cells with combined DCA and CU treatment. In both panels, lanes 1–3 represent recombinant n-dTPI, dTPI, and ddTPI enzymes loaded at 1 μg protein per lane, serving as migration standards. In both panels, lanes 4–9 with 100 μg of protein loaded per lane. In a and b, lane 4 shows control Jurkat cells (untreated); lane 5 shows 12 mM DCA. In (<b>a</b>) lane 6 and 7 are Jurkat cells pretreated with 12 mM DCA followed by incubation with 100 and 250 μM DS, respectively. In (<b>b</b>) lane 6 and 7 are Jurkat cells pretreated with 12 mM DCA followed by incubation with 1000 μM and 1500 μM CU, respectively. Finally, in both, a and b, lanes 8–9 show normal T lymphocytes pretreated with 12 mM DCA followed by incubation with 250 μM DS (<b>a</b>) or 1500 μM CU (<b>b</b>). The polarity of the gel is indicated on the left side of the panel. Full-length blots (uncropped blots) are shown in <a href="#app1-biomolecules-14-01295" class="html-app">Supplementary Figure S6</a>.</p>
Full article ">Figure 8
<p>MGO and AGEs levels in Jurkat cells following DCA, DS, and CU treatment. Quantification of MGO (<b>a</b>,<b>c</b>) and AGEs (<b>b</b>,<b>d</b>) in Jurkat cells following treatment with DCA, DS, CU, or their combination (*). As observed, increasing concentrations of DS and CU lead to a dose-dependent rise in both MGO and AGEs. Notably, pretreatment with DCA before DS or CU administration results in a further significant increase in MGO and AGEs production. The results represent the average of three independent experiments, with error bars indicating the variation observed in the experiments.</p>
Full article ">Figure 9
<p>Western blot analysis of apoptosis-related proteins in Jurkat cells. (<b>a</b>) Expression of ERK1/2 and its phosphorylated form following treatment with DCA, CU, or their combination. The corresponding bar graphs quantify the decrease in total and phosphorylated ERK1/2 levels relative to the control. (<b>b</b>) Expression of Bcl-2 and Bax proteins under treatment with DCA, CU, or their combination. The bar graphs represent the relative expression levels of these proteins. (<b>c</b>) Procaspase-7 and cleaved caspase-7 levels following treatment with DCA, CU, or their combination. The bar graphs quantify the relative abundance of procaspase-7 and cleaved caspase-7 compared to the control. β-Actin was used as a loading control for all Western blots. Each lane was loaded with 100 μg of total protein extract. Statistical analysis was performed using a one-way ANOVA followed by the Tukey-Kramer test. Significance levels are indicated as follows: * <span class="html-italic">p</span> ≤ 0.01 compared to the control, ¥ <span class="html-italic">p</span> ≤ 0.01 compared to 12 mM DCA treatment, &amp; <span class="html-italic">p</span> ≤ 0.01 compared to 0.5 mM CU treatment, <span>$</span> <span class="html-italic">p</span> ≤ 0.01 compared to 12 mM DCA + 0.5 mM CU treatment. Full-length blots (uncropped blots) are shown in <a href="#app1-biomolecules-14-01295" class="html-app">Supplementary Figure S7b–d</a> for Total ERK 1/2, pERK, and β-Actin, respectively; <a href="#app1-biomolecules-14-01295" class="html-app">Supplementary Figure S8b–d</a> for Bcl2, Bax, and β-Actin, respectively, and <a href="#app1-biomolecules-14-01295" class="html-app">Supplementary Figure S9b,c</a> for Procaspase-7, cleaved caspase-7, and β-Actin, respectively.</p>
Full article ">Figure 10
<p>Detection of cell death in normal T lymphocytes and Jurkat cells after treatment with DCA, DS, CU, or their combinations (DCA + DS or DCA + CU). Cells were incubated with 12 mM DCA, 100 µM DS, or 250 µM CU for 24 h at 37 °C. For combination treatments, cells were first exposed to 12 mM DCA for 24 h, followed by an additional 24-h incubation with 100 µM DS or 250 µM CU. At the end of the incubation period, cells were washed, resuspended at a density of 1 × 10<sup>6</sup>/mL, and stained with Annexin V and Propidium Iodide, for flow cytometry analysis. (<b>a</b>) shows untreated control cells. (<b>b</b>–<b>d</b>) represent cells treated with DCA, DS, and CU, respectively. (<b>e</b>,<b>f</b>) display cells pretreated with DCA followed by DS and CU, respectively. Quadrants indicate distinct cell populations: Q1 (necrotic cells), Q2 (late apoptotic cells), and Q4 (early apoptotic cells). The data presented in the figure are representative of 100,000 cells analyzed across two independent experiments.</p>
Full article ">Figure 10 Cont.
<p>Detection of cell death in normal T lymphocytes and Jurkat cells after treatment with DCA, DS, CU, or their combinations (DCA + DS or DCA + CU). Cells were incubated with 12 mM DCA, 100 µM DS, or 250 µM CU for 24 h at 37 °C. For combination treatments, cells were first exposed to 12 mM DCA for 24 h, followed by an additional 24-h incubation with 100 µM DS or 250 µM CU. At the end of the incubation period, cells were washed, resuspended at a density of 1 × 10<sup>6</sup>/mL, and stained with Annexin V and Propidium Iodide, for flow cytometry analysis. (<b>a</b>) shows untreated control cells. (<b>b</b>–<b>d</b>) represent cells treated with DCA, DS, and CU, respectively. (<b>e</b>,<b>f</b>) display cells pretreated with DCA followed by DS and CU, respectively. Quadrants indicate distinct cell populations: Q1 (necrotic cells), Q2 (late apoptotic cells), and Q4 (early apoptotic cells). The data presented in the figure are representative of 100,000 cells analyzed across two independent experiments.</p>
Full article ">
34 pages, 1167 KiB  
Review
Potential Applications of the Anti-Inflammatory, Antithrombotic and Antioxidant Health-Promoting Properties of Curcumin: A Critical Review
by Elli Rapti, Theodora Adamantidi, Pavlos Efthymiopoulos, George Z. Kyzas and Alexandros Tsoupras
Nutraceuticals 2024, 4(4), 562-595; https://doi.org/10.3390/nutraceuticals4040031 - 11 Oct 2024
Viewed by 1043
Abstract
The major constituent of turmeric, curcumin, is a bioactive phenolic compound that has been studied for its potential health benefits and therapeutic properties. Within this article, the anti-inflammatory, antioxidant and antithrombotic properties and mechanisms of action of curcumin are thoroughly reviewed and the [...] Read more.
The major constituent of turmeric, curcumin, is a bioactive phenolic compound that has been studied for its potential health benefits and therapeutic properties. Within this article, the anti-inflammatory, antioxidant and antithrombotic properties and mechanisms of action of curcumin are thoroughly reviewed and the main focus is shifted to its associated health-promoting effects against inflammation-related chronic disorders. An overview of the cardio-protective, anti-tumor, anti-diabetic, anti-obesity, anti-microbial and neuro–protective health-promoting properties of curcumin are thoroughly reviewed, while relative outcomes obtained from clinical trials are also presented. Emphasis is given to the wound-healing properties of curcumin, as presented by several studies and clinical trials, which further promote the application of curcumin as a bioactive ingredient in several functional products, including functional foods, nutraceuticals, cosmetics and drugs. Limitations and future perspectives of such uses of curcumin as a bio-functional ingredient are also discussed. Full article
Show Figures

Graphical abstract

Graphical abstract
Full article ">Figure 1
<p>A glimpse of the multifaceted activities of curcumin.</p>
Full article ">Figure 2
<p>Global curcumin market by application.</p>
Full article ">Figure 3
<p>Curcumin utilized forms, applications and health-promoting properties.</p>
Full article ">Figure 4
<p>Curcumin’s antioxidant, anti-inflammatory and anticancer properties.</p>
Full article ">
15 pages, 16627 KiB  
Article
Vesicle-Transported Multidrug Resistance as a Possible Therapeutic Target of Natural Compounds
by Salvatrice Rigogliuso, Alessandra Cusimano, Lucia Condorelli, Manuela Labbozzetta, Gabriella Schiera, Paola Poma and Monica Notarbartolo
Pharmaceuticals 2024, 17(10), 1358; https://doi.org/10.3390/ph17101358 - 11 Oct 2024
Viewed by 379
Abstract
Background/Objectives: A key role of extracellular vesicles (EVs) is mediating both cell–cell and cell–stroma communication in pathological/physiological conditions. EVs from resistant tumor cells can transport different molecules like P-glycoprotein (P-gp), acting as a shuttle between donor and recipient cells, resulting in a phenotypic [...] Read more.
Background/Objectives: A key role of extracellular vesicles (EVs) is mediating both cell–cell and cell–stroma communication in pathological/physiological conditions. EVs from resistant tumor cells can transport different molecules like P-glycoprotein (P-gp), acting as a shuttle between donor and recipient cells, resulting in a phenotypic change. The aim of our work was to isolate, characterize, and inhibit the release of EVs in two multidrug resistance (MDR) cancer models: MCF-7R (breast cancer cell line) and HL-60R (acute myeloid leukemia cell line). Methods: The existence of P-gp in EVs from MDR cells was confirmed by Western blotting assays. The characterization of EVs was carried out by evaluating the size using NTA and the presence of specific markers such as CD63, Hsp70 and Syntenin. The ability of HL-60R and MCF-7R to perform horizontal transfer of P-gp via EVs to sensitive cells was assessed using three different methods. The acquisition of resistance and its inhibition in recipient cells was confirmed by MTS 3-(4,5-dimethylthiazol-2-yl)-5-(3-carboxymethoxyphenyl)-2-(4-sulfophenyl)-2H-tetrazolium (MTS) assay. Results: Our data showed that cell lines (MDR) release P-gp-loaded EVs, unlike sensitive cells. The acquisition of resistance determined by the incorporation of P-gp into the membrane of sensitive cells was confirmed by the reduced cytotoxic activity of doxorubicin. Natural compounds such as curcumin, lupeol, and heptacosane can block vesicular transfer and restore the sensitivity of HL-60 and MCF-7 cells. Conclusions: Our study demonstrates that natural inhibitors able to reverse this mechanism may represent a new therapeutic strategy to limit the propagation of the resistant phenotype. Full article
(This article belongs to the Section Natural Products)
Show Figures

Graphical abstract

Graphical abstract
Full article ">Figure 1
<p>Evaluation of P-gp expression in EVs isolated from resistant cell lines. (<b>A</b>) Immunofluorescence analysis of HL-60, HL-60R, MCF-7, and MCF-7R cells. Nuclei (DAPI Blue); P-gp (green). (<b>B</b>) Western blotting analysis of cell extract and 38,000 rpm fractions (38K) of EVs isolated, respectively, from HL-60R and MCF-7R cell lines. (<b>C</b>) Comparison of the amount of protein content in EVs released from both sensitive cell lines compared to resistant cell lines. Data were normalized to the same number of cells: HL-60R vs. HL-60 * <span class="html-italic">p</span> &lt; 0.05; MCF-7R vs. MCF-7 * <span class="html-italic">p</span> &lt; 0.01 (Tukey’s test).</p>
Full article ">Figure 2
<p>Analysis of cells producing EVs (38K) with AO staining: (<b>A</b>) fluorescence microscopy images; (<b>B</b>) Trypan Blue exclusion assay; (<b>C</b>) spectrofluorometric analysis. Fluorescence emission of the cells was compared to positive control (apoptosis induced by doxorubicin treatment; RFU—Relative Fluorescence Units).</p>
Full article ">Figure 3
<p>Characterization analysis of EVs isolated from resistant cell lines. (<b>A</b>) Western blot analysis of the EV markers, Hsp-70, CD63, and Syntenin. (<b>B</b>) Size distributions of EV fractions isolated, respectively, from HL-60R and MCF-7R cell lines measured by nanoparticle tracking analysis (NTA). (<b>C</b>) Separation of heterogeneous EV populations by differential ultracentrifugation and relative %; Western blotting analysis for P-gp and Syntenin expression in a 35k EV fraction.</p>
Full article ">Figure 4
<p>Analysis of the horizontal transfer of resistance. Representative image of evaluation of P-gp expression after following treatment of the sensitive HL-60 and MCF-7 cell lines: EVs isolated from the respective resistant cell lines, with the medium conditioned by the cells of the respective resistant lines or by the co-culture growth of the sensitive and resistant line in a Transwell plate. Nuclei (DAPI Blue); P-gp (green); magnification 20× and scale bar was 25 µm.</p>
Full article ">Figure 5
<p>(<b>A</b>) Cells were exposed to Doxorubicin (1.8 μM) for 16 h after co-culture. Cell viability was assessed by an MTS assay. Letters indicate significant differences (Tukey’s test) in cell viability among the concentrations of each cell line (Cytochalasin, Curcumin, and Heptacosane vs. co-culture + Doxorubicin <span class="html-italic">p</span> &lt; 0.005; Cytochalasin vs. Lupeol <span class="html-italic">p</span> &lt; 0.001; Curcumin vs. Lupeol and Heptacosane <span class="html-italic">p</span> &lt; 0.05). Treatments are likened to the control: * <span class="html-italic">p</span> &lt; 0.005; ** <span class="html-italic">p</span> &lt; 0.001. (<b>B</b>) Cells were exposed to Doxorubicin (9.2 μM) for 24 h after co-culture. Cell viability was assessed by cell counting. Letters indicate significant differences (Tukey test) in cell viability among the concentrations (Cytochalasin and Heptacosane vs. co-culture + Doxorubicin <span class="html-italic">p</span> &lt; 0.05; Cytochalasin vs. Lupeol and Curcumin <span class="html-italic">p</span> &lt; 0.05). Treatments are likened to the control: * <span class="html-italic">p</span> &lt; 0.05.</p>
Full article ">Figure 6
<p>Co-culture of MCF-7 and MCF-7R cell spheroids. MCF-7 and MCF-7R cell spheroids in co-culture for 24 h, labeled with DAPI (blue nuclei) and anti-P-gp mAb (green fluorescence). (<b>A</b>) Immunofluorescence at time zero and after 24 h of co-culture with an anti-P-gp antibody; magnification 60× and scale bar was 25 µm. (<b>B</b>) Confocal microscopy with Z-stack imaging; magnification 40× and scale bar was 50 μm.</p>
Full article ">
24 pages, 1821 KiB  
Review
Unraveling Mitochondrial Reactive Oxygen Species Involvement in Psoriasis: The Promise of Antioxidant Therapies
by Hajar Ahmad Jamil and Norwahidah Abdul Karim
Antioxidants 2024, 13(10), 1222; https://doi.org/10.3390/antiox13101222 - 11 Oct 2024
Viewed by 488
Abstract
Psoriasis is a chronic inflammatory skin disorder characterized by immune dysregulation and aberrant keratinocyte proliferation. Despite tremendous advances in understanding its etiology, effective therapies that target its fundamental mechanisms remain necessary. Recent research highlights the role of reactive oxygen species dysregulation and mitochondrial [...] Read more.
Psoriasis is a chronic inflammatory skin disorder characterized by immune dysregulation and aberrant keratinocyte proliferation. Despite tremendous advances in understanding its etiology, effective therapies that target its fundamental mechanisms remain necessary. Recent research highlights the role of reactive oxygen species dysregulation and mitochondrial dysfunction in psoriasis pathogenesis. Mitochondrial reactive oxygen species mediate cellular signaling pathways involved in psoriasis, such as proliferation, apoptosis, and inflammation, leading to oxidative stress, exacerbating inflammation and tissue damage if dysregulated. This review explores oxidative stress biomarkers and parameters in psoriasis, including myeloperoxidase, paraoxonase, sirtuins, superoxide dismutase, catalase, malondialdehyde, oxidative stress index, total oxidant status, and total antioxidant status. These markers provide insights into disease mechanisms and potential diagnostic and therapeutic targets. Modulating mitochondrial reactive oxygen species levels and enhancing antioxidant defenses can alleviate inflammation and oxidative damage, improving patient outcomes. Natural antioxidants like quercetin, curcumin, gingerol, resveratrol, and other antioxidants show promise as complementary treatments targeting oxidative stress and mitochondrial dysfunction. This review aims to guide the development of personalized therapeutic methods and diagnostic techniques, emphasizing the importance of comprehensive clinical studies to validate the efficacy and safety of these interventions, paving the way for more effective and holistic psoriasis care. Full article
(This article belongs to the Special Issue Role of Mitochondria and ROS in Health and Disease)
Show Figures

Figure 1

Figure 1
<p>Summary of oxidative stress biomarkers in psoriasis inflammation. This figure illustrates key oxidative stress biomarkers involved in the inflammatory processes of psoriasis. The biomarkers depicted, including myeloperoxidase (MPO), paraoxonase (PON), sirtuins (SIRTs), superoxide dismutase (SOD), and catalase (CAT), are emerging as critical indicators of psoriasis pathogenesis. The red arrows indicate decreasing, while green arrow represent increasing.</p>
Full article ">Figure 2
<p>Overview of oxidative stress parameters in psoriasis inflammation. This figure provides a summary of key oxidative stress parameters associated with psoriasis. It depicts the roles of malondialdehyde (MDA), total antioxidant status (TAS), oxidative stress index (OSI), and total oxidant status (TOS) in evaluating oxidative stress and antioxidant defense mechanisms in psoriasis patients.</p>
Full article ">
50 pages, 8706 KiB  
Review
Metabolic-Associated Fatty Liver Disease: The Influence of Oxidative Stress, Inflammation, Mitochondrial Dysfunctions, and the Role of Polyphenols
by Raissa Bulaty Tauil, Paula Takano Golono, Enzo Pereira de Lima, Ricardo de Alvares Goulart, Elen Landgraf Guiguer, Marcelo Dib Bechara, Claudia C. T. Nicolau, José Luiz Yanaguizawa Junior, Adriana M. R. Fiorini, Nahum Méndez-Sánchez, Ludovico Abenavoli, Rosa Direito, Vitor Engrácia Valente, Lucas Fornari Laurindo and Sandra Maria Barbalho
Pharmaceuticals 2024, 17(10), 1354; https://doi.org/10.3390/ph17101354 - 10 Oct 2024
Viewed by 513
Abstract
Metabolic-Associated Fatty Liver Disease (MAFLD) is a clinical–pathological scenario that occurs due to the accumulation of triglycerides in hepatocytes which is considered a significant cause of liver conditions and contributes to an increased risk of death worldwide. Even though the possible causes of [...] Read more.
Metabolic-Associated Fatty Liver Disease (MAFLD) is a clinical–pathological scenario that occurs due to the accumulation of triglycerides in hepatocytes which is considered a significant cause of liver conditions and contributes to an increased risk of death worldwide. Even though the possible causes of MAFLD can involve the interaction of genetics, hormones, and nutrition, lifestyle (diet and sedentary lifestyle) is the most influential factor in developing this condition. Polyphenols comprise many natural chemical compounds that can be helpful in managing metabolic diseases. Therefore, the aim of this review was to investigate the impact of oxidative stress, inflammation, mitochondrial dysfunction, and the role of polyphenols in managing MAFLD. Some polyphenols can reverse part of the liver damage related to inflammation, oxidative stress, or mitochondrial dysfunction, and among them are anthocyanin, baicalin, catechin, curcumin, chlorogenic acid, didymin, epigallocatechin-3-gallate, luteolin, mangiferin, puerarin, punicalagin, resveratrol, and silymarin. These compounds have actions in reducing plasma liver enzymes, body mass index, waist circumference, adipose visceral indices, lipids, glycated hemoglobin, insulin resistance, and the HOMA index. They also reduce nuclear factor-KB (NF-KB), interleukin (IL)-1β, IL-6, tumor necrosis factor-α (TNF-α), blood pressure, liver fat content, steatosis index, and fibrosis. On the other hand, they can improve HDL-c, adiponectin levels, and fibrogenesis markers. These results show that polyphenols are promising in the prevention and treatment of MAFLD. Full article
Show Figures

Figure 1

Figure 1
<p>Factors related to the occurrence of Metabolic-Associated Fatty Liver Disease (MAFLD) and the possibility of the inhibition of this condition. An unhealthy diet, sedentary lifestyle, obesity, insulin resistance/diabetes, dyslipidemia, genetics, and excessive drug consumption are related to the pathogenesis of MAFLD and its progression to fibrosis, cirrhosis, and cancer. A healthy diet, physical exercise, and weight loss can improve metabolic conditions and can prevent or reduce MAFLD.</p>
Full article ">Figure 2
<p>The liver in the context of MAFLD. Lifestyle and metabolic alterations lead to an increased lipolysis of visceral adipose tissue, stimulating de novo lipogenesis, and an increase in FFA and VLDL (and a consequent efflux of this lipoprotein). Increased glucose intake results in increased pyruvate and Acetyl-CoA production, leading to increased TCA activity. Furthermore, there is augmented β-oxidation resulting in mitochondrial dysfunction. The consequences are mitochondrial dysfunction, altered mtDNA, an imbalance in respiration (reduction in ATP production), and RE stress. All these events are related to increased inflammation and ROS, which results in apoptosis and liver damage. Systemic inflammation occurs due to Kupffer cell activation. DNL: de novo lipogenesis; FFA: free fatty acid; IL: interleukin; JNK: c-Jun N-terminal kinase; M2: macrophage; mtDNA: mitochondrial DNA; NF-KB: nuclear factor-KB, NO: nitric acid; NLRP3: NLR family pyrin domain-containing 3; ROS: reactive oxygen species; VLDL: very-low-density lipoprotein; TG: triglyceride; TNF-α: tumor necrosis factor-α; TCA: tricarboxylic acid cycle.</p>
Full article ">Figure 3
<p>The activation of DNL and an increase in FFAs lead to mitochondrial alterations and an increase in oxidative stress and inflammation. The stimulation of the mitochondrial membrane permeability transition pore is also observed by mitochondrial alterations and the deposit of fatty acids. There is stimulation in the activity of inner membrane proteins, leading to a reduction in ATP production. Mitochondrial gene mutation (mt-DNA) also activates uncoupling proteins. AMPK: AMP-activated protein kinase; CoQ: coenzyme Q; Cyt C: cytochrome C; DNL: de novo lipogenesis; FAO: fatty acid oxidation; FFA: free fatty acid; PGC1α: peroxisome proliferator-activated receptor-γ coactivator 1-α; JNK: c-Jun N-terminal kinase; NF-KB: nuclear factor kappa B; SIRT3: sirtuin 3; TCA: tricarboxylic acid cycle; TNF-α: tumor necrosis factor-α; UCP: uncoupling protein.</p>
Full article ">Figure 4
<p>Polyphenols: classification and origin. Polyphenols are found in many fruits and vegetables and can be separated into phenolic acids, flavonoids, and non-flavonoids. Phenolic acids can be found in onion, tea, and coffee; flavonoids in grapes, pepper, broccoli, green tea, lemon, and soy; and non-flavonoids in grapes, peanut skin, and <span class="html-italic">Curcuma longa</span>. These compounds can protect the liver since they can reduce the risks for MAFLD, such as oxidative stress, inflammation, and lipid deposits. IL: interleukin; JNK: c-Jun N-terminal kinase; MAFLD: Metabolic-Associated Fatty Liver Disease; NF-KB: nuclear factor kappa B; Nrf2: nuclear factor erythroid 2-related factor 2, PKC: protein kinase C; ROS: reactive oxygen species; SREBP-1c: Sterol regulatory element-binding protein 1c.</p>
Full article ">Figure 5
<p>The main mechanisms of action promoted by phenols in MAFLD. A salubrious diet with an increased consumption of fruits and vegetables elevates the intake of polyphenols. These phytochemicals can inhibit liver cellular damage associated with MAFLD through varied mechanisms that may include a decrease in de novo lipogenesis due to the downregulation of SREBP-1c, elevating β-fatty acid oxidation through PPAR α upregulation, ameliorating insulin sensitivity, and reducing oxidative stress and inflammation processes. This scenario is related to a reduction in liver damage and systemic inflammation. JNK: c-Jun N-terminal kinase; NF-KB: nuclear factor kappa B; Nrf2: nuclear factor erythroid 2-related factor 2, PKC: protein kinase C; PPAR-α: peroxisome proliferator-activated receptor gamma; SREBP-1c: Sterol regulatory element-binding protein 1c; TCA: tricarboxylic acid cycle; TAG: triglyceride.</p>
Full article ">
30 pages, 2414 KiB  
Review
Promising Phytogenic Feed Additives Used as Anti-Mycotoxin Solutions in Animal Nutrition
by Sergio Quesada-Vázquez, Raquel Codina Moreno, Antonella Della Badia, Oscar Castro and Insaf Riahi
Toxins 2024, 16(10), 434; https://doi.org/10.3390/toxins16100434 - 10 Oct 2024
Viewed by 899
Abstract
Mycotoxins are a major threat to animal and human health, as well as to the global feed supply chain. Among them, aflatoxins, fumonisins, zearalenone, T-2 toxins, deoxynivalenol, and Alternaria toxins are the most common mycotoxins found in animal feed, with genotoxic, cytotoxic, carcinogenic, [...] Read more.
Mycotoxins are a major threat to animal and human health, as well as to the global feed supply chain. Among them, aflatoxins, fumonisins, zearalenone, T-2 toxins, deoxynivalenol, and Alternaria toxins are the most common mycotoxins found in animal feed, with genotoxic, cytotoxic, carcinogenic, and mutagenic effects that concern the animal industry. The chronic negative effects of mycotoxins on animal health and production and the negative economic impact on the livestock industry make it crucial to develop and implement solutions to mitigate mycotoxins. In this review, we summarize the current knowledge of the mycotoxicosis effect in livestock animals as a result of their contaminated diet. In addition, we discuss the potential of five promising phytogenics (curcumin, silymarin, grape pomace, olive pomace, and orange peel extracts) with demonstrated positive effects on animal performance and health, to present them as potential anti-mycotoxin solutions. We describe the composition and the main promising characteristics of these bioactive compounds that can exert beneficial effects on animal health and performance, and how these phytogenic feed additives can help to alleviate mycotoxins’ deleterious effects. Full article
(This article belongs to the Special Issue Mitigation and Detoxification Strategies of Mycotoxins)
Show Figures

Figure 1

Figure 1
<p>Synthesis of curcumin by the general method proposed by Pabon [<a href="#B79-toxins-16-00434" class="html-bibr">79</a>] (image adapted from Zerazion et al., 2016) [<a href="#B86-toxins-16-00434" class="html-bibr">86</a>]. <b>1.</b> A suspension of B<sub>2</sub>O<sub>3</sub> (3 mmol) and acetylacetone (2 mmol) in 1.5 mL of DMF was stirred for 30 min at 80 °C. <b>2.</b> To this was added 4-hydroxy-3-methoxybenzaldehyde (5.4 mmol), followed by the slow addition of n-butylamine solution, and this was stirred at 80 °C for 4 h. <b>3.</b> The solution was acidified with 0.5 M HCl at 80 °C for 1 h. <b>4.</b> Curcumin was dried and recrystallized.</p>
Full article ">Figure 2
<p>Chemical structure of main flavonolignans contained in silymarin complex.</p>
Full article ">Figure 3
<p>Composition and principal bioactive components of grape pomace extract.</p>
Full article ">Figure 4
<p>Composition and principal bioactive components of olive pomace extract.</p>
Full article ">Figure 5
<p>Composition and principal bioactive components of orange peel extract.</p>
Full article ">
26 pages, 10057 KiB  
Article
EF24, a Curcumin Analog, Reverses Interleukin-18-Induced miR-30a or miR-342-Dependent TRAF3IP2 Expression, RECK Suppression, and the Proinflammatory Phenotype of Human Aortic Smooth Muscle Cells
by Yusuke Higashi, Ryan Dashek, Patrice Delafontaine, Randy Scott Rector and Bysani Chandrasekar
Cells 2024, 13(20), 1673; https://doi.org/10.3390/cells13201673 - 10 Oct 2024
Viewed by 511
Abstract
Curcumin, a polyphenolic compound derived from the widely used spice Curcuma longa, has shown anti-atherosclerotic effects in animal models and cultured vascular cells. Inflammation is a major contributor to atherosclerosis development and progression. We previously reported that the induction of the proinflammatory molecule [...] Read more.
Curcumin, a polyphenolic compound derived from the widely used spice Curcuma longa, has shown anti-atherosclerotic effects in animal models and cultured vascular cells. Inflammation is a major contributor to atherosclerosis development and progression. We previously reported that the induction of the proinflammatory molecule TRAF3IP2 (TRAF3 Interacting Protein 2) or inhibition of the matrix metallopeptidase (MMP) regulator RECK (REversion Inducing Cysteine Rich Protein with Kazal Motifs) contributes to pro-oxidant, proinflammatory, pro-mitogenic and pro-migratory effects in response to external stimuli in vascular smooth muscle cells. Here we hypothesized that EF24, a curcumin analog with a better bioavailability and bioactivity profile, reverses interleukin (IL)-18-induced TRAF3IP2 induction, RECK suppression and the proinflammatory phenotype of primary human aortic smooth muscle cells (ASMC). The exposure of ASMC to functionally active recombinant human IL-18 (10 ng/mL) upregulated TRAF3IP2 mRNA and protein expression, but markedly suppressed RECK in a time-dependent manner. Further investigations revealed that IL-18 inhibited both miR-30a and miR-342 in a p38 MAPK- and JNK-dependent manner, and while miR-30a mimic blunted IL-18-induced TRAF3IP2 expression, miR-342 mimic restored RECK expression. Further, IL-18 induced ASMC migration, proliferation and proinflammatory phenotype switching, and these effects were attenuated by TRAF3IP2 silencing, and the forced expression of RECK or EF24. Together, these results suggest that the curcumin analog EF24, either alone or as an adjunctive therapy, has the potential to delay the development and progression of atherosclerosis and other vascular inflammatory and proliferative diseases by differentially regulating TRAF3IP2 and RECK expression in ASMC. Full article
Show Figures

Figure 1

Figure 1
<p>Interleukin-18 (IL-18) upregulates TRAF3IP2 but suppresses RECK expression in primary human aortic smooth muscle cells (ASMC). (<b>A</b>–<b>C</b>) IL-18 upregulates TRAF3IP2 expression. ASMCs were grown in complete media, and at 70–80% confluency, made quiescent for 48 h, and then incubated with rhIL-18 (10 ng/mL) for up to 12 h (experimental design in (<b>A</b>)). TRAF3IP2 mRNA expression was analyzed by RT-qPCR using a TaqMan™ probe (<b>B</b>) and its protein levels by Western blotting (<b>C</b>), with GAPDH and Tubulin serving as loading controls, respectively. * <span class="html-italic">p</span> &lt; 0.05, ** at least <span class="html-italic">p</span> &lt; 0.01 vs. untreated controls (<span class="html-italic">n</span> = 4). (<b>D</b>,<b>E</b>) Quiescent ASMCs were incubated with IL-18 as in (A) and were analyzed for RECK mRNA expression by RT-qPCR using a TaqMan™ probe (<b>D</b>) and protein levels by Western blotting (<b>E</b>). * <span class="html-italic">p</span> &lt; 0.05, ** at least <span class="html-italic">p</span> &lt; 0.01 vs. untreated controls (<span class="html-italic">n</span> = 4). (<b>F</b>–<b>H</b>) Specificity of IL-18 on TRAF3IP2 induction (<b>G</b>) and RECK suppression (<b>H</b>) was verified by incubating with neutralizing IL-18R1 antibody or IL-18BP-Fc chimera for 1 h prior to IL-18 addition for 2 (<b>G</b>) or 6 h (<b>H</b>), with normal goat IgG or Fc serving as controls. mRNA expressions of TRAF3IP2 and RECK were analyzed by RT-qPCR. (<b>C</b>,<b>E</b>) While a representative immunoblot is shown, the intensities of immunoreactive bands from 4 independent experiments were semiquantified by densitometry and are summarized on the right. (<b>G</b>) * <span class="html-italic">p</span> &lt; 0.05, ** <span class="html-italic">p</span> &lt; 0.01 vs. untreated controls, † <span class="html-italic">p</span> &lt; 0.01 versus IL-18 (n = 6); (<b>H</b>) * <span class="html-italic">p</span> &lt; 0.05, ** <span class="html-italic">p</span> &lt; 0.01 vs. untreated controls, † <span class="html-italic">p</span> &lt; 0.01 versus IL-18 (<span class="html-italic">n</span> = 4).</p>
Full article ">Figure 2
<p>Targeting TRAF3IP2 or RECK overexpression blunts IL-18-induced ASMC proliferation and migration. (<b>A</b>–<b>E</b>) Silencing TRAF3IP2 inhibits IL-18-induced ASMC proliferation (<b>B</b>) and migration (<b>C</b>). ASMC were transduced with adenovirus-expressing shRNA targeting human TRAF3IP2 (moi 10 for 48 h), made quiescent, and then exposed to IL-18 (10 ng/mL; experimental design in (<b>A</b>)). ASMC proliferation was assessed after 48 h of IL-18 addition using the CyQUANT Cell proliferation assay (<b>B</b>), and migration after 18 h using Boyden chamber assay (<b>C</b>). ASMCs migrating to the lower surface of the membrane were counted in 10 different fields and summarized as mean ± SEM. (<b>B</b>,<b>C</b>) * <span class="html-italic">p</span> &lt; at least 0.01 vs. Untreated; † <span class="html-italic">p</span> &lt; 0.01 vs. IL-18 or IL-18+GFP (n = 6). Knockdown of TRAF3IP2 was confirmed by RT-qPCR using a TaqMan™ probe (<b>D</b>) and Western blotting (<b>E</b>). (<b>D</b>,<b>E</b>) * <span class="html-italic">p</span> &lt; 0.01 vs. untreated (n = 3). (<b>F</b>,<b>G</b>) Dose-dependent effects of Ad.RECK on RECK expression (experimental design in (<b>F</b>)). Induction of RECK following adenoviral transduction was confirmed by Western blotting with tubulin serving as an internal control (<b>G</b>). (<b>H</b>–<b>K</b>) Forced expression of RECK inhibits IL-18-stimulated ASMC proliferation and migration. ASMCs were transduced with adenovirus-expressing human RECK cDNA (moi 10 for 24 h), made quiescent, and then treated with IL-18 (experimental design in (<b>H</b>)) and analyzed for proliferation (<b>I</b>) and migration (<b>J</b>) as in (<b>B</b>,<b>C</b>). (<b>C</b>,<b>J</b>) The insets show representative images of Matrigel™ Transwell invasion. Scale bar: 20 μM. (<b>E</b>,<b>G</b>,<b>K</b>) While a representative immunoblot is shown, the intensities of immunoreactive bands from three (<b>E</b>), four (<b>G</b>) and three (<b>K</b>) independent experiments were semiquantified by densitometry and are summarized on the right. (<b>I</b>) * <span class="html-italic">p</span> &lt; at least 0.01 vs. Untreated; † <span class="html-italic">p</span> &lt; 0.01 vs. IL-18 or IL-18+GFP (n = 6); (<b>J</b>) * <span class="html-italic">p</span> &lt; at least 0.01 vs. Untreated; † <span class="html-italic">p</span> &lt; 0.01 vs. IL-18 or IL-18+eGFP (n = 4); (<b>K</b>) * <span class="html-italic">p</span> &lt; 0.01 vs. Untreated; † <span class="html-italic">p</span> &lt; 0.05 vs. IL-18 or IL-18+eGFP (n = 3).</p>
Full article ">Figure 3
<p>TRAF3IP2 knockdown restores SMC marker expression and inhibits ASMC proinflammatory phenotype without affecting cell viability. (<b>A</b>–<b>G</b>) Silencing TRAF3IP2 restores IL-18-mediated suppression in SMC markers, but inhibits the expression of proinflammatory phenotype markers, without significantly modulating cell viability. ASMCs were transduced with adenoviral TRAF3IP2 shRNA (moi10 for 48 h), made quiescent and then treated with IL-18 (10 ng/mL for 48 h; experimental design in (<b>A</b>)). Expressions of the SMC markers ACTA2 (<b>B</b>,<b>C</b>) and MYH11 (<b>D</b>,<b>E</b>) were analyzed by both RT-qPCR (<b>B</b>,<b>D</b>) and Western blotting (<b>C</b>,<b>E</b>). The proinflammatory phenotype markers Galectin 3, Olr1, VCAM, CCL2, IL-6, IL-8, and TNF-α were analyzed by RT-qPCR using TaqMan™ probes (<b>F</b>). Cell viability was assessed by analyzing cleaved caspase-3 levels using a commercially available Caspase-3 (Cleaved) Human ELISA (<b>G</b>). H<sub>2</sub>O<sub>2</sub> (100 μM for 18 h) served as a positive control and induced a significant increase in cleaved capase-3 levels. (<b>C</b>,<b>E</b>) While a representative immunoblot is shown, the intensities of immunoreactive bands from three independent experiments were semiquantified by densitometry and are summarized on the right. (<b>B</b>,<b>D</b>,<b>F</b>,<b>G</b>) * <span class="html-italic">p</span> &lt; at least 0.01 vs. Untreated; † <span class="html-italic">p</span> &lt; 0.01 vs. IL-18 or IL-18+GFP (n = 6 or 7). (<b>C</b>,<b>E</b>) * <span class="html-italic">p</span> &lt; 0.05 vs. Untreated; † <span class="html-italic">p</span> &lt; 0.05 vs. IL-18 or IL-18+GFP (n = 3).</p>
Full article ">Figure 4
<p>Ectopic expression of RECK blunts IL-18-induced inhibition in the expression of SMC markers and the induction of proinflammatory phenotype markers. (<b>A</b>–<b>F</b>) Forced expression of RECK restores IL-18-induced suppression of SMC markers and inhibits the induction of proinflammatory phenotype markers. ASMCs were transduced with adenovirus-expressing human RECK cDNA (moi10 for 24 h), made quiescent, and then treated with IL-18 at 10 ng/mL for 48 h (experimental design in (<b>A</b>)). The expression levels of SMC markers ACTA2 (<b>B</b>,<b>C</b>) and MYH11 (<b>D</b>,<b>E</b>) were analyzed by RT-qPCR (<b>B</b>,<b>D</b>) and Western blotting (<b>C</b>,<b>E</b>). The proinflammatory phenotype markers Galectin 3, Olr1, VCAM, CCL2, IL-6, IL-8, and TNF-α were analyzed by RT-qPCR using TaqMan™ probes (<b>F</b>). (<b>C</b>,<b>E</b>) While a representative immunoblot is shown, the intensities of immunoreactive bands from three independent experiments were semiquantified by densitometry and are summarized on the right. (<b>B</b>,<b>D</b>,<b>F</b>) * <span class="html-italic">p</span> &lt; at least 0.01 vs. Untreated; † <span class="html-italic">p</span> &lt; 0.01 vs. IL-18 or IL-18+eGFP (n = 5). (<b>C</b>,<b>E</b>) * <span class="html-italic">p</span> &lt; 0.05 vs. Untreated; † <span class="html-italic">p</span> &lt; 0.05 vs. IL-18 or IL-18+eGFP (n = 3).</p>
Full article ">Figure 5
<p>IL-18 inhibits miR-30a and miR-342 expression via stress-activated kinases. (<b>A</b>–<b>C</b>) IL-18 inhibits miR-30a and miR-342 expression via stress-activated kinases. Quiescent ASMCs were treated with inhibitors of either p38 MAPK (SB239063, 10 μM in DMSO for 1 h), ERK1/2 (SCH772984 10 μM in DMSO for 1 h) or JNK (SP600125, 20 μM in DMSO for 1 h) prior to IL-18 addition at 10 ng/mL for 30 min (experimental design in (<b>A</b>)). (<b>B</b>,<b>C</b>) Fresh DMSO (0.1%) served as a solvent control. miR-30a and miR-342 expressions were analyzed by TaqMan<span class="html-italic">®</span> Advanced miRNA assays, with U6 serving as a loading control. (<b>B</b>) * <span class="html-italic">p</span> &lt; 0.001 vs. untreated, † <span class="html-italic">p</span> &lt; at least 0.01 vs. IL-18 or IL-18+DMSO (n = 11), (<b>C</b>) * <span class="html-italic">p</span> &lt; 0.01 vs. untreated, † <span class="html-italic">p</span> &lt; at least 0.05 vs. IL-18 or IL-18+DMSO (n = 5). (<b>D</b>–<b>F</b>) miR-30a mimic inhibits IL-18-induced TRAF3IP2 expression. ASMC were transfected with miR-30a mimic (80 nM), made quiescent and then exposed to IL-18 at 10 ng/mL for 2 h (experimental design in (<b>D</b>)). TRAF3IP2 mRNA expression was analyzed by RT-qPCR (E) and its protein levels by Western blotting (<b>F</b>). (<b>G</b>–<b>I</b>) miR-342 mimic restores IL-18-induced RECK suppression. ASMCs were transfected with miR-342 mimic (80 nM), made quiescent and then exposed to IL-18 at 10 ng/mL for 6 h (experimental design in (<b>G</b>)). RECK mRNA expression was analyzed by RT-qPCR (<b>H</b>) and its protein levels by Western blotting (<b>I</b>). (<b>F</b>,<b>I</b>) While a representative immunoblot is shown, the intensities of immunoreactive bands from 4 independent experiments were semiquantified by densitometry and are summarized as mean ± SEM on the right. (<b>E</b>,<b>F</b>,<b>H</b>,<b>I</b>) * <span class="html-italic">p</span> &lt; 0.05, † at least <span class="html-italic">p</span> &lt; 0.01 vs. untreated controls (n = 4).</p>
Full article ">Figure 6
<p>EF24 inhibits IL-18-induced ASMC proliferation and migration, without affecting cell viability. (<b>A</b>) Chemical structure of EF24, a curcumin analog. (<b>B</b>–<b>F</b>) EF24 is not cytotoxic to ASMC at the concentrations used. Quiescent ASMCs were exposed to EF24 at the indicated concentrations for 48 h (experimental design in (<b>B</b>)). Cell viability was analyzed by trypan blue dye exclusion (<b>C</b>), cleaved caspase-3 levels by ELISA (<b>D</b>) with H<sub>2</sub>O<sub>2</sub> (100 μM) serving as a positive control, ELISA of mono-oligonucleosomal fragmented DNA (<b>E</b>) with H<sub>2</sub>O<sub>2</sub> (100 μM) serving as a positive control, and LDH release with LDH release in response to 0.2% Triton-X100 being considered as 100% (<b>F</b>). DMSO (0.1%) alone served as a solvent control (depicted as “0”). Cells without any treatment served as a control (<b>C</b>). (<b>D</b>–<b>F</b>) * <span class="html-italic">p</span> &lt; 0.01 vs. untreated controls or treated with DMSO alone. (<b>G</b>–<b>H</b>) EF24 inhibits IL-18-induced ASMC proliferation and migration. Quiescent ASMCs were incubated with EF24 at various concentrations ranging from 1 to 10 μM in DMSO for 1 h, followed by the addition of IL-18 at 10 ng/mL for 48 h (experimental design in (<b>G</b>)). Cell proliferation was analyzed by the CyQUANT Cell proliferation assay (<b>H</b>). Cell migration was analyzed by Boyden chamber assay after 18 h ((<b>I</b>) The insets show representative images of Matrigel™ transwell invasion). The combination of IL-18 and EF24 did not affect cell viability (<b>J</b>). (<b>C</b>,<b>I</b>) Scale bars, 20 μM. (<b>H</b>,<b>I</b>) * <span class="html-italic">p</span> &lt; at least 0.05 vs. untreated controls, † <span class="html-italic">p</span> &lt; 0.05 vs. IL-18 without EF24 (n = 6). (<b>J</b>) * <span class="html-italic">p</span> &lt; 0.01 vs. untreated controls (n = 12).</p>
Full article ">Figure 7
<p>EF24 reverses IL-18-induced upregulation in TRAF3IP2 expression and RECK suppression. (<b>A</b>–<b>C</b>) EF24 blunts IL-18-induced TRAF3IP2 expression. Quiescent ASMCs were treated with EF24 (2.5 μM in DMSO for 1 h) prior to IL-18 addition at 10 ng/mL for 3 h (experimental design in (<b>A</b>)). DMSO alone (0.025%) served as a solvent control. TRAF3IP2 mRNA expression was analyzed by RT-qPCR (<b>B</b>) and its protein levels by Western blotting (<b>C</b>). (<b>D</b>–<b>F</b>) EF24 restores IL-18-induced RECK suppression. Quiescent ASMCs were treated with EF24 (2.5 μM in DMSO for 1 h) prior to IL-18 addition at 10 ng/mL for 6 h (experimental design in (<b>D</b>)). RECK mRNA expression was analyzed by RT-qPCR (<b>E</b>) and its protein levels by Western blotting (<b>F</b>). (<b>C</b>,<b>F</b>) While a representative immunoblot is shown, the intensities of immunoreactive bands from 3–4 independent experiments were semiquantified by densitometry and are summarized as mean ± SEM on the right. (<b>B</b>,<b>E</b>) * <span class="html-italic">p</span> &lt; at least 0.01 vs. Untreated; † <span class="html-italic">p</span> &lt; at least 0.05 vs. IL-18 or IL-18+DMSO (n = 3–4), (<b>C</b>,<b>F</b>) * <span class="html-italic">p</span> &lt; 0.05 vs. Untreated; † <span class="html-italic">p</span> &lt; 0.05 vs. IL-18 or IL-18+DMSO.</p>
Full article ">Figure 8
<p>EF24 reverses ASMC proinflammatory phenotype without affecting cell viability. (<b>A</b>–<b>F</b>) Pretreatment with EF24 restores the IL-18-induced suppression of SMC markers and inhibits the expression of proinflammatory phenotype markers, without significantly modulating cell viability. Quiescent ASMCs were treated with EF24 (2.5 μM for 1 h in DMSO) prior to the addition of IL-18 at 10 ng/mL for 48 h (experimental design in (<b>A</b>)). The expression levels of SMC markers ACTA2 ((<b>B</b>) mRNA, (<b>C</b>) protein levels) and MYH11 ((<b>D</b>) mRNA, (<b>E</b>) protein levels) were analyzed by RT-qPCR and Western blotting, and those of proinflammatory phenotype markers Galectin 3, Olr1, VCAM, CCL2, IL-6, IL-8, and TNF-α (<b>F</b>) were analyzed by RT-qPCR using TaqMan™ probes. Cell viability was assessed by analyzing cleaved caspase-3 levels using a Caspase-3 (Cleaved) Human ELISA kit (<b>G</b>). H<sub>2</sub>O<sub>2</sub> (100 μM) for 24 h served as a positive control and induced a significant increase in cleaved capase-3 levels. (<b>C</b>,<b>E</b>) While a representative immunoblot is shown, the intensities of immunoreactive bands from three independent experiments were semiquantified by densitometry and are summarized on the right. (<b>B</b>,<b>D</b>,<b>F</b>,<b>G</b>) * <span class="html-italic">p</span> &lt; at least 0.01 vs. Untreated; † <span class="html-italic">p</span> &lt; at least 0.05 vs. IL-18 or IL-18+DMSO (n = 5–6), (<b>C</b>–<b>E</b>) * <span class="html-italic">p</span> &lt; 0.05 vs. Untreated; † <span class="html-italic">p</span> &lt; 0.05 vs. IL-18 or IL-18+DMSO (n = 3).</p>
Full article ">Figure 9
<p>EF24 inhibits IL-18-induced MMP2 and MMP9 expressions. (<b>A</b>–<b>D</b>) Silencing MMPs 2 and 9 inhibits IL-18-induced ASMC migration. ASMCs were transduced with adenoviral vector-expressing human MMP2 or MMP9 shRNA (moi10 for 48 h), made quiescent in basal medium containing ITS-G 1X supplement and then treated with IL-18 at 10 ng/mL for 18 h (experimental design in (<b>A</b>)). Ad.siGFP at moi 10 for 48 h served as a control. Cell migration was analyzed by the Boyden chamber assay (<b>B</b>). (<b>C</b>,<b>D</b>) Knockdown of MMPs 2 and 9 was analyzed by RT-qPCR and their protein levels by Western blotting (insets). (<b>C</b>,<b>D</b>) While a representative immunoblot is shown, the intensities of immunoreactive bands from 3 independent experiments were semiquantified by densitometry and are summarized as mean ± SEM on the right. (<b>E</b>–<b>G</b>) EF24 inhibits IL-18-induced MMP expression. SMCs made quiescent in basal medium containing ITS-G 1X supplement for 48 h were incubated with EF24 (2.5 μM) for 1 h followed by IL-18 at 10 ng/mL for 2 h (experimental design in (<b>E</b>)). MMP2 (<b>F</b>) and MMP9 (<b>G</b>) mRNA expressions were analyzed by RT-qPCR. (<b>B</b>,<b>F</b>,<b>G</b>) * <span class="html-italic">p</span> &lt; at least 0.01 vs. Untreated; † <span class="html-italic">p</span> &lt; at least 0.05 vs. IL-18 or IL-18+siGFP (n = 3–6); ((<b>C</b>,<b>D</b>) left) * <span class="html-italic">p</span> &lt; 0.001 vs. Untreated; ((<b>C</b>,<b>D</b>) right) * <span class="html-italic">p</span> &lt; 0.05 vs. Untreated.</p>
Full article ">Figure 10
<p>Schematic showing that EF24, a curcumin analog with better bioavailability and biologic activity, inhibits the proinflammatory IL-18-induced primary human aortic smooth muscle cells’ (ASMCs) proliferation, migration, and proinflammatory phenotype changes. EF24 inhibits the stress-activated kinase-dependent miR-30a and miR-342 inhibition, TRAF3IP2 upregulation and RECK suppression. While miR-30a mimic reverses IL-18-induced TRAF3IP2 upregulation, the miR-342 mimic restores IL-18-mediated RECK suppression, potentially via reduced DNMT1 expression and promoter demethylation (dashed purple box). While IL-18 promotes ASMC migration and proliferation, these effects were reversed by TRAF3IP2 knockdown or the ectopic expression of RECK. Further, while IL-18 induced ASMC migration in part via induction of the gelatinases MMP2 and MMP9, these effects were inhibited by EF24. Moreover, EF24 restores IL-18-mediated suppression in SMC markers and blunts the expression of the proinflammatory phenotype markers. These results suggest that the curcumin analog EF24 reverses IL-18-induced ASMC proliferation and migration and proinflammatory phenotypic changes by targeting TRAF3IP2 and restoring RECK expression. These results suggest the therapeutic potential of EF24 in vascular inflammatory and proliferative diseases, including atherosclerosis.</p>
Full article ">
19 pages, 2473 KiB  
Systematic Review
The Role of Curcumin in Modulating Vascular Function and Structure during Menopause: A Systematic Review
by Amanina Athirah Mad Azli, Norizam Salamt, Amilia Aminuddin, Nur Aishah Che Roos, Mohd Helmy Mokhtar, Jaya Kumar, Adila A. Hamid and Azizah Ugusman
Biomedicines 2024, 12(10), 2281; https://doi.org/10.3390/biomedicines12102281 - 8 Oct 2024
Viewed by 472
Abstract
The risk of developing cardiovascular disease (CVD) escalates in women during menopause, which is associated with increased vascular endothelial dysfunction, arterial stiffness, and vascular remodeling. Meanwhile, curcumin has been demonstrated to enhance vascular function and structure in various studies. Therefore, this study systematically [...] Read more.
The risk of developing cardiovascular disease (CVD) escalates in women during menopause, which is associated with increased vascular endothelial dysfunction, arterial stiffness, and vascular remodeling. Meanwhile, curcumin has been demonstrated to enhance vascular function and structure in various studies. Therefore, this study systematically reviewed the recent literature regarding the potential role of curcumin in modulating vascular function and structure during menopause. The Ovid MEDLINE, PubMed, Scopus, and Web of Science electronic databases were searched to identify relevant articles. Clinical and preclinical studies involving menopausal women and postmenopausal animal models with outcomes related to vascular function or structure were included. After thorough screening, seven articles were selected for data extraction, comprising three animal studies and four clinical trials. The findings from this review suggested that curcumin has beneficial effects on vascular function and structure during menopause by addressing endothelial function, arterial compliance, hemodynamic parameters, and the formation of atherosclerotic lesions. Therefore, curcumin has the potential to be utilized as a supplement to enhance vascular health in menopausal women. However, larger-scale clinical trials employing gold-standard techniques to evaluate vascular health in menopausal women are necessary to validate the preliminary results obtained from small-scale randomized clinical trials involving curcumin supplementation (INPLASY, INPLASY202430043). Full article
(This article belongs to the Special Issue Compounds from Natural Products as Sources for Drug Discovery)
Show Figures

Figure 1

Figure 1
<p>The chemical structure of curcumin.</p>
Full article ">Figure 2
<p>Preferred Reporting Items for Systematic Reviews and Meta-Analyses (PRISMA) 2020 flow diagram for systematic reviews.</p>
Full article ">Figure 3
<p>Systematic Review Center for Laboratory Animal Experimentation (SYRCLE) risk of bias summary: the reviewers’ assessments of the risk of bias for each item in every animal study included in the review, namely Abd. Aziz et al. 2012 [<a href="#B26-biomedicines-12-02281" class="html-bibr">26</a>], Jusoh et al. 2013 [<a href="#B27-biomedicines-12-02281" class="html-bibr">27</a>] and Morrone et al. 2016 [<a href="#B25-biomedicines-12-02281" class="html-bibr">25</a>].</p>
Full article ">Figure 4
<p>The Cochrane risk of bias tool for every randomized controlled trial included in the review, namely Akazawa et al. 2012 [<a href="#B28-biomedicines-12-02281" class="html-bibr">28</a>], Akazawa et al. 2013 [<a href="#B29-biomedicines-12-02281" class="html-bibr">29</a>], Sugawara et al. 2012 [<a href="#B30-biomedicines-12-02281" class="html-bibr">30</a>] and Farshbaf et al. 2022 [<a href="#B20-biomedicines-12-02281" class="html-bibr">20</a>].</p>
Full article ">Figure 5
<p>Meta-analysis of the effects of curcumin versus placebo on brachial blood pressure based on the findings of Akazawa et al. 2012 [<a href="#B28-biomedicines-12-02281" class="html-bibr">28</a>], Akazawa et al. 2013 [<a href="#B29-biomedicines-12-02281" class="html-bibr">29</a>] and Sugawara et al. 2012 [<a href="#B30-biomedicines-12-02281" class="html-bibr">30</a>]. The green square with horizontal line indicates individual study effect size (mean difference) together with its 95% confidence interval (CI), whilst the black diamond indicates the summary effect size together with its 95% CI.</p>
Full article ">
20 pages, 9133 KiB  
Article
Utilizing an Ex Vivo Skin Penetration Analysis Model for Predicting Ocular Drug Penetration: A Feasibility Study with Curcumin Formulations
by Christian Raab, Stefan Brugger, Jara-Sophie Lechner, Geisa Nascimento Barbalho, Taís Gratieri, Priyanka Agarwal, Ilva D. Rupenthal and Cornelia M. Keck
Pharmaceutics 2024, 16(10), 1302; https://doi.org/10.3390/pharmaceutics16101302 - 6 Oct 2024
Viewed by 611
Abstract
Objective: This study aimed to investigate the feasibility of using the digital image processing technique, developed to semi-quantitatively study dermal penetration, to study corneal penetration in an ex vivo porcine eye model. Here, we investigated various formulation strategies intended to enhance dermal and [...] Read more.
Objective: This study aimed to investigate the feasibility of using the digital image processing technique, developed to semi-quantitatively study dermal penetration, to study corneal penetration in an ex vivo porcine eye model. Here, we investigated various formulation strategies intended to enhance dermal and corneal bioavailability of the model hydrophobic drug, curcumin. Methods: Several formulation principles were explored, including oily solutions, oily suspensions, aqueous nanosuspension, micelles, liposomes and cyclodextrins. The dermal penetration efficacy was tested using an ex vivo porcine ear model previously developed at Philipps-Universität Marburg with subsequent digital image processing. This image analysis method was further applied to study corneal penetration using an ex vivo porcine whole-eye model. Results: For dermal penetration, oily solutions, oily suspensions and nanosuspensions exhibited the least penetration, whereas liposomes and cyclodextrins showed enhanced penetration. Corneal curcumin penetration correlated with dermal penetration, with curcumin loaded into cyclodextrins penetrating the deepest. Conclusions: Overall, our study suggests that the image analysis method previously developed for ex vivo skin penetration can easily be repurposed to study corneal penetration of hydrophobic drugs. Full article
(This article belongs to the Special Issue Curcumin in Biomedical Applications, 2nd Edition)
Show Figures

Figure 1

Figure 1
<p>Scheme of dermal penetration testing with the ‘Marburg skin penetration’ model. (<b>A</b>) Fresh pig ear with areas for treatment. (<b>B</b>,<b>C</b>) Cross-sectional images of skin biopsies taken by inverted epifluorescence microscopy. (<b>D</b>,<b>E</b>) Images after digital image processing, where an automated threshold was applied to subtract the autofluorescence of the skin. The remaining pixels represent the penetrated API. SCT—stratum corneum thickness; MPD = mean penetration depth. Scale bar = 50 µm.</p>
Full article ">Figure 2
<p>Microscopic images of curcumin formulations.</p>
Full article ">Figure 3
<p>Particle size analysis (LD data) of curcumin formulations.</p>
Full article ">Figure 4
<p>Fluorescence images of tissue cross-sections showing untreated skin and skin treated with the different curcumin formulations. Scale bar = 50 µm.</p>
Full article ">Figure 5
<p>Dermal penetration efficacy [MGV/px] and relative (rel.) penetration efficacy [%] of curcumin from different formulations. *: <span class="html-italic">p</span>-value &lt; 0.05; ***: <span class="html-italic">p</span>-value &lt; 0.001.</p>
Full article ">Figure 6
<p>Mean dermal penetration depth [µm] and relative (rel.) penetration depth [%] of curcumin from different formulations. *: <span class="html-italic">p</span>-value &lt; 0.05; ***: <span class="html-italic">p</span>-value &lt; 0.001.</p>
Full article ">Figure 7
<p>Influence of unloaded vehicles and curcumin-loaded formulations on SCT. *: <span class="html-italic">p</span>-value &lt; 0.05; **: <span class="html-italic">p</span>-value &lt; 0.01, ***: <span class="html-italic">p</span>-value &lt; 0.001, n.s.: non-significant.</p>
Full article ">Figure 8
<p>Fluorescence images of untreated cornea and cornea treated with different curcumin-containing formulations. Scale bar = 50 µm.</p>
Full article ">Figure 9
<p>Fluorescence images after RGB threshold of untreated cornea and cornea treated with different curcumin-containing formulations. Scale bar = 50 µm.</p>
Full article ">Figure 10
<p>Influence of formulation principle on corneal penetration efficacy of curcumin—data analysis from images obtained after automated RGB threshold. (<b>A</b>) Corneal penetration efficacy [MGV/px] and rel. penetration efficacy [%] of curcumin from different formulations. (<b>B</b>) Mean penetration depth [µm] and rel. penetration depth [%] of curcumin from different formulations. *: <span class="html-italic">p</span>-value &lt; 0.05, ***: <span class="html-italic">p</span>-value &lt; 0.001, n.s.: non-significant.</p>
Full article ">Figure 11
<p>Influence of formulation principle on corneal penetration efficacy of curcumin—data analysis from original images obtained from inverted epifluorescence microscopy. (<b>A</b>) Corneal penetration efficacy [MGV/px] and rel. penetration efficacy [%] of curcumin from different formulations. (<b>B</b>) Mean penetration depth [µm] and rel. penetration depth [%] of curcumin from different formulations. *: <span class="html-italic">p</span>-value &lt; 0.05; **: <span class="html-italic">p</span>-value &lt; 0.01, ***: <span class="html-italic">p</span>-value &lt; 0.001, n.s.: non-significant.</p>
Full article ">
13 pages, 1854 KiB  
Article
Drug Combination Studies of Isoquinolinone AM12 with Curcumin or Quercetin: A New Combination Strategy to Synergistically Inhibit 20S Proteasome
by Carla Di Chio, Santo Previti, Josè Starvaggi, Fabiola De Luca, Maria Luisa Calabrò, Maria Zappalà and Roberta Ettari
Int. J. Mol. Sci. 2024, 25(19), 10708; https://doi.org/10.3390/ijms251910708 - 4 Oct 2024
Viewed by 521
Abstract
In the eukaryotic cells, the ubiquitin–proteasome system (UPS) plays a crucial role in the intracellular protein turnover. It is involved in several cellular functions such as the control of the regular cell cycle progression, the immune surveillance, and the homeostasis. Within the 20S [...] Read more.
In the eukaryotic cells, the ubiquitin–proteasome system (UPS) plays a crucial role in the intracellular protein turnover. It is involved in several cellular functions such as the control of the regular cell cycle progression, the immune surveillance, and the homeostasis. Within the 20S proteasome barrel-like structure, the catalytic subunits, β1, β2 and β5, are responsible for different proteolytic activities: caspase-like (C-L), trypsin-like (T-L) and chymotrypsin-like (ChT-L), respectively. The β5 subunit is particularly targeted for its role in antitumor activity: the synthesis of β5 subunit inhibitors could be a promising strategy for the treatment of solid and hematologic tumors. In the present work, we performed two combination studies of AM12, a recently developed synthetic proteasome inhibitor, with curcumin and quercetin, two nutraceuticals endowed of many pharmacological properties. We measured the combination index (CI), applying the Chou and Talalay method, comparing the two studies, from 50% to 90% of proteasome inhibition. In the case of the combination AM12 + curcumin, an increasing synergism was observed from 50% to 90% of proteasome inhibition, while in the case of the combination AM12 + quercetin an additive effect was observed only from 50% to 70% of β5 subunit inhibition. These results suggest that combining AM12 with curcumin is a more promising strategy than combining it with quercetin for potential therapeutic applications, especially in treating tumors. Full article
Show Figures

Figure 1

Figure 1
<p>Chemical structures of <b>AM12</b>, curcumin and quercetin.</p>
Full article ">Figure 2
<p>Dose–response curves for β5 inhibition by <b>AM12</b> (<b>A</b>), curcumin (<b>B</b>) and <b>AM12</b> + curcumin in combination (<b>C</b>). Each experiment was performed two times, each in duplicate.</p>
Full article ">Figure 3
<p>Median Effect Plot for <b>AM12</b> (<b>A</b>), curcumin (<b>B</b>) and <b>AM12</b> + curcumin in combination (molar ratio 4.95:1) (<b>C</b>). D is the dose, and fa and fu the affected and the unaffected fraction of β5 activity, respectively, by the dose D.</p>
Full article ">Figure 4
<p>Dose–response curves for β5 inhibition by <b>AM12</b> (<b>A</b>), quercetin (<b>B</b>) and <b>AM12</b> + quercetin in combination (<b>C</b>). Each experiment was performed two times, each in duplicate.</p>
Full article ">Figure 5
<p>Median Effect Plot for <b>AM12</b> (<b>A</b>), quercetin (<b>B</b>) and <b>AM12</b> + quercetin in combination (molar ratio 4.95:1) (<b>C</b>). D is the dose, and fa and fu the affected and the unaffected fraction of β5 activity, respectively, by the dose D.</p>
Full article ">Figure 6
<p>Computer-generated graphical presentation of the combination index (CI) vs. the fraction affected (fa), i.e., the effect of reduction in proteasome activity exerted by <b>AM12</b> + curcumin (<b>A</b>) and by <b>AM12</b> + quercetin (<b>B</b>). The circles indicate the trend of the combination index.</p>
Full article ">Scheme 1
<p>Reagents and conditions: (<span class="html-italic">i</span>) NaH, N<sub>2</sub>, dry DMF, 0 °C, <b>1</b>, 1 h, and then methyl 4-bromocrotonate, 0 °C to room temperature (rt), 12 h; (<span class="html-italic">ii</span>) MeOH, LiOH, 0° C to rt, TLC monitoring; and (<span class="html-italic">iii</span>) dry DMF, resulting acid from <b>2</b>, HATU, 0 °C, 10 min and then DIPEA, iso-pentyl amine, 0 °C to rt, 12 h.</p>
Full article ">
31 pages, 3629 KiB  
Article
Biocompatible Poly(ε-Caprolactone) Nanocapsules Enhance the Bioavailability, Antibacterial, and Immunomodulatory Activities of Curcumin
by Floriana D’Angeli, Giuseppe Granata, Ivana Roberta Romano, Alfio Distefano, Debora Lo Furno, Antonella Spila, Mariantonietta Leo, Chiara Miele, Dania Ramadan, Patrizia Ferroni, Giovanni Li Volti, Paolo Accardo, Corrada Geraci, Fiorella Guadagni and Carlo Genovese
Int. J. Mol. Sci. 2024, 25(19), 10692; https://doi.org/10.3390/ijms251910692 - 4 Oct 2024
Viewed by 688
Abstract
Curcumin (Cur), the primary curcuminoid found in Curcuma longa L., has garnered significant attention for its potential anti-inflammatory and antibacterial properties. However, its hydrophobic nature significantly limits its bioavailability. Additionally, adipose-derived stem cells (ADSCs) possess immunomodulatory properties, making them useful for treating inflammatory [...] Read more.
Curcumin (Cur), the primary curcuminoid found in Curcuma longa L., has garnered significant attention for its potential anti-inflammatory and antibacterial properties. However, its hydrophobic nature significantly limits its bioavailability. Additionally, adipose-derived stem cells (ADSCs) possess immunomodulatory properties, making them useful for treating inflammatory and autoimmune conditions. This study aims to verify the efficacy of poly(ε-caprolactone) nanocapsules (NCs) in improving Cur’s bioavailability, antibacterial, and immunomodulatory activities. The Cur-loaded nanocapsules (Cur-NCs) were characterized for their physicochemical properties (particle size, polydispersity index, Zeta potential, and encapsulation efficiency) and stability over time. A digestion test simulated the behavior of Cur-NCs in the gastrointestinal tract. Micellar phase analyses evaluated the Cur-NCs’ bioaccessibility. The antibacterial activity of free Cur, NCs, and Cur-NCs against various Gram-positive and Gram-negative strains was determined using the microdilution method. ADSC viability, treated with Cur-NCs and Cur-NCs in the presence or absence of lipopolysaccharide, was analyzed using the 3-[4,5-dimethylthiazol-2-yl]-2,5-diphenyl tetrazolium bromide assay. Additionally, ADSC survival was assessed through the Muse apoptotic assay. The expression of both pro-inflammatory (interleukin-1β and tumor necrosis factor-α) and anti-inflammatory (IL-10 and transforming growth factor-β) cytokines on ADSCs was evaluated by real-time polymerase chain reaction. The results demonstrated high stability post-gastric digestion of Cur-NCs and elevated bioaccessibility of Cur post-intestinal digestion. Moreover, Cur-NCs exhibited antibacterial activity against Escherichia coli without affecting Lactobacillus growth. No significant changes in the viability and survival of ADSCs were observed under the experimental conditions. Finally, Cur-NCs modulated the expression of both pro- and anti-inflammatory cytokines in ADSCs exposed to inflammatory stimuli. Collectively, these findings highlight the potential of Cur-NCs to enhance Cur’s bioavailability and therapeutic efficacy, particularly in cell-based treatments for inflammatory diseases and intestinal dysbiosis. Full article
(This article belongs to the Special Issue New Perspective on Inflammatory Diseases: Role of Natural Compounds)
Show Figures

Figure 1

Figure 1
<p>Ultraviolet–visible (UV−Vis) spectrum of curcumin-nanocapsules (Cur-NCs) in water. To perform the UV−Vis spectrum, 10 μL of the Cur-NCs nanosuspension was diluted with 3 mL of pure water.</p>
Full article ">Figure 2
<p>Fluorescence spectra of curcumin-nanocapsules (Cur-NCs) in water and after curcumin (Cur) release in acetonitrile. To perform the fluorescence spectra, 10 μL of the Cur-NCs nanosuspension was diluted with 3 mL of pure water or acetonitrile.</p>
Full article ">Figure 3
<p>Intensity-weighted distribution of the hydrodynamic diameter (<span class="html-italic">D<sub>H</sub></span>) of freshly prepared curcumin-nanocapsules, after 30 days of storage at 25 °C and 40 °C.</p>
Full article ">Figure 4
<p>Curcumin (Cur) retention percentage in curcumin-nanocapsules (Cur-NCs) after 30 days of storage at 25 °C and 40 °C.</p>
Full article ">Figure 5
<p>Dose and time-dependent effects of curcumin (Cur), empty nanocapsules (NCs), and curcumin-loaded nanocapsules (Cur-NCs) on primary human adipose-derived stem cells (ADSCs) viability. Cells were grown in a culture medium (CTRL) or exposed to increasing concentrations (0.06–1 µg/mL) of Cur or Cur-NCs or different dilutions of NCs for 24 h (<b>A</b>,<b>A’</b>) and 48 h (<b>B</b>,<b>B’</b>). Results are expressed as a percent of control. The bars represent means ± SD from three independent experiments performed in triplicate (SD = standard deviation). Statistically significant differences, determined by one-way analysis of variance ANOVA and the Tukey post-test, are indicated: * <span class="html-italic">p</span> &lt; 0.05 vs. CTRL at the same incubation time. WS: working standard: diluted solution derived from the stock solution.</p>
Full article ">Figure 6
<p>Effect of curcumin (Cur), empty nanocapsules (NCs), curcumin-loaded nanocapsules (Cur-NCs), and dexamethasone (Dexa) on primary human adipose-derived stem cells (ADSCs) cell viability, at the steady state, and under inflammatory stimuli. Cells were grown alone (control: CTRL) or in the presence of 0.125 µg/mL of Cur or Cur-NCs, NCs diluted in the 1:8 ratio, 10 nM Dexa, with or without 1 µg/mL lipopolysaccharide (LPS), for 24 h and 48 h. Histograms showed ADSCs cell viability, at 24 h (<b>A</b>) and 48 h (<b>B</b>). Results are expressed as a percent of control. The bars represent means ± SD from three independent experiments performed in triplicate (SD = standard deviation). Statistically significant differences, determined by one-way analysis of variance ANOVA and the Tukey post-test, are indicated: * <span class="html-italic">p</span> &lt; 0.05 vs. CTRL; § <span class="html-italic">p</span> &lt; 0.05 vs. Cur; # <span class="html-italic">p</span> &lt; 0.05 vs. LPS, at the same incubation time.</p>
Full article ">Figure 7
<p>Evaluation of apoptotic cell death on human adipose-derived stem cells (ADSCs) exposed to curcumin (Cur), empty nanocapsules (NCs), curcumin-loaded nanocapsules (Cur-NCs), dexamethasone (Dexa), and lipopolysaccharide (LPS) in the presence or absence of inflammatory stimuli. A–B Scatter plots of ADSCs grown, for 24 h (<b>A</b>) and 48 h (<b>B</b>), in normal culture medium (control; CTRL) or the presence of 0.125 µg/mL Cur or Cur-NCs, NCs diluted in the 1:8 ratio, 10 nM Dexa, with or without 1 µg/mL LPS. Each plot reports four squares in which cells are distributed based on their staining. (<b>A’</b>,<b>B’</b>) Histograms showed the rate of vital cells (Alive), early apoptotic cells (EA), late apoptotic (LA)/dead cells, and debris for each experimental condition, at 24 h (<b>A’</b>) and 48 h (<b>B’</b>). The bars represent means ± SD of three independent experiments (SD = standard deviation). Statistically significant differences, determined by two-way analysis of variance ANOVA and the Tukey post-test, are indicated: <b>*</b> <span class="html-italic">p</span> &lt; 0.05 vs. CTRL; <b>#</b> <span class="html-italic">p</span> &lt; 0.05 vs. LPS, at the same incubation time.</p>
Full article ">Figure 7 Cont.
<p>Evaluation of apoptotic cell death on human adipose-derived stem cells (ADSCs) exposed to curcumin (Cur), empty nanocapsules (NCs), curcumin-loaded nanocapsules (Cur-NCs), dexamethasone (Dexa), and lipopolysaccharide (LPS) in the presence or absence of inflammatory stimuli. A–B Scatter plots of ADSCs grown, for 24 h (<b>A</b>) and 48 h (<b>B</b>), in normal culture medium (control; CTRL) or the presence of 0.125 µg/mL Cur or Cur-NCs, NCs diluted in the 1:8 ratio, 10 nM Dexa, with or without 1 µg/mL LPS. Each plot reports four squares in which cells are distributed based on their staining. (<b>A’</b>,<b>B’</b>) Histograms showed the rate of vital cells (Alive), early apoptotic cells (EA), late apoptotic (LA)/dead cells, and debris for each experimental condition, at 24 h (<b>A’</b>) and 48 h (<b>B’</b>). The bars represent means ± SD of three independent experiments (SD = standard deviation). Statistically significant differences, determined by two-way analysis of variance ANOVA and the Tukey post-test, are indicated: <b>*</b> <span class="html-italic">p</span> &lt; 0.05 vs. CTRL; <b>#</b> <span class="html-italic">p</span> &lt; 0.05 vs. LPS, at the same incubation time.</p>
Full article ">Figure 8
<p>Expression levels of inflammatory cytokines IL-1β, TNF-α, IL-10, and TGF-β on human adipose-derived stem cells (ADSCs) exposed to curcumin (Cur), empty nanocapsules (NCs), curcumin-loaded nanocapsules (Cur-NCs), dexamethasone (Dexa), with or without lipopolysaccharide (LPS), for 24 h and 48 h. ADSCs were cultured in normal culture medium (control; CTRL) or the presence of 0.125 µg/mL Cur or Cur-NCs, NCs diluted in the 1:8 ratio, 10 nM Dexa, with or without 1 µg/mL LPS, for 24 h (<b>A</b>–<b>D</b>) and 48 h (<b>A’</b>–<b>D’</b>). Histograms showed IL-1β, TNF-α, IL-10, and TGF-β mRNA expression levels on ADSCs grown in our experimental conditions for 24 h (<b>A</b>–<b>D</b>) and 48 h (<b>A’</b>–<b>D’</b>). The bars represent means ± SD of three independent experiments (SD = standard deviation). Statistically significant differences, determined by one-way analysis of variance ANOVA and the Tukey post-test, are indicated: * <span class="html-italic">p</span> &lt; 0.05 vs. CTRL; # <span class="html-italic">p</span> &lt; 0.05 vs. LPS, at the same incubation time.</p>
Full article ">
12 pages, 3090 KiB  
Article
Ultrasound-Activated Multifunctional Bioactive Calcium Phosphate Composites for Enhanced Osteosarcoma Treatment
by Mingjie Wang, Dong Xu, Chunfeng Xu, Menghong Li, Chang Du and Yuelian Liu
Coatings 2024, 14(10), 1267; https://doi.org/10.3390/coatings14101267 - 2 Oct 2024
Viewed by 682
Abstract
Bone defects caused by surgical interventions and the challenges of tumor recurrence and metastasis due to residual cancer cells significantly complicate the treatment of osteosarcoma (OS). To address these complex clinical challenges, we propose an innovative therapeutic strategy that centers on an ultrasound-activated [...] Read more.
Bone defects caused by surgical interventions and the challenges of tumor recurrence and metastasis due to residual cancer cells significantly complicate the treatment of osteosarcoma (OS). To address these complex clinical challenges, we propose an innovative therapeutic strategy that centers on an ultrasound-activated multifunctional bioactive calcium phosphate (BioCaP) composite. A modified curcumin (mcur)-mediated wet biomimetic mineralization process was used to develop an anticancer-drug-integrated multifunctional BioCaP (mcur@BioCaP), exploring its potential biological effects for OS treatment activated by ultrasound (US). The mcur@BioCaP demonstrates a drug dose-dependent, tailorable alteration in its micro/nanostructure. The US stimulus significantly enhanced this composite to generate reactive oxygen species (ROS) in cancer cells. The results show that the OS cell viability of the mcur@BioCaP with US is 62.2% ± 6.3%, the migration distance is 63.9% ± 6.6%, and the invaded OS cell number is only 57.0 ± 3.7 OS cells per version, which were all significantly lower than US or mcur@BioCaP alone, suggesting that the anticancer, anti-migratory and anti-invasive effects of mcur@BioCaP on OS 143B cells were amplified by ultrasonic stimulation. This amplification can be attributed to the US-activated ROS production from the drug molecules, which regulates the wet biomimetic mineralization of the multifunctional composite. Furthermore, mcur@BioCaP with US increased calcium nodule formation by 1.8-fold, which was significantly higher than mcur@BioCaP or US group, indicating its potential in promoting bone regeneration. The anticancer and osteogenic potentials of mcur@BioCaP were found to be consistent with the mcur concentration in the multifunctional composite. Our research provides a novel therapeutic approach that leverages a multifunctional biomimetic mineral and ultrasonic activation, highlighting its potential applications in OS therapy. Full article
(This article belongs to the Special Issue Latest Trends in Coatings of Medical Implants)
Show Figures

Figure 1

Figure 1
<p>Surface morphology of BioCaP and different mcur@BioCaP crystals. (<b>A</b>) SEM images of mcur@BioCaP coating synthesized in the mineralization solutions containing different mcur concentration, scale bar = 5 µm. (<b>B</b>) Crystal morphology of the mcur@BioCaP, scale bar = 1 µm.</p>
Full article ">Figure 2
<p>Ultrasound enhances drug penetration of mcur@BioCaP. (<b>A</b>) The change in fluorescence intensity observed by CLSM, scale bar = 100 µm. (<b>B</b>) Quantitative analysis of the fluorescence intensity change. Statistical difference ** <span class="html-italic">p</span> &lt; 0.01.</p>
Full article ">Figure 3
<p>Ultrasound-induced ROS production in OS cells. (<b>A</b>) ROS production in OS cells observed by CLSM, scale bar = 100 µm. (<b>B</b>) Quantitative analysis of intracellular ROS after US treatment. Statistical difference ** <span class="html-italic">p</span> &lt; 0.01.</p>
Full article ">Figure 4
<p>US-enhanced anticancer activities of mcur@BioCaP. (<b>A</b>) Cell viability of OB cells after treatment with US, <span class="html-italic">n</span> = 6. (<b>B</b>) Cell viability of OS cells after treatment with US, <span class="html-italic">n</span> = 6. (<b>C</b>) Cell viability of OB cells following treatment with mcur@BioCaP and 1 min of US exposure, <span class="html-italic">n</span> = 6. (<b>D</b>) Cell viability of OS cells following treatment with mcur@BioCaP and 1 min US, <span class="html-italic">n</span> = 6. Statistical difference ** <span class="html-italic">p</span> &lt; 0.01.</p>
Full article ">Figure 5
<p>US-mediated retardation of migration and invasion in OS cells. (<b>A</b>) Migration of OS cells under the influence of US treatment, scare bar = 100 µm. (<b>B</b>) Quantitative analysis of OS cell migration following US treatment. (<b>C</b>) Invasion of OS cells facilitated by US treatment, scare bar = 100 µm. (<b>D</b>) Quantitative analysis of OS cell invasion post US treatment. Statistical difference <span class="html-italic">* p</span> &lt; 0.05, <span class="html-italic">** p</span> &lt; 0.01.</p>
Full article ">Figure 6
<p>Osteogenic differentiation of OB cells treated by different stimulus. (<b>A</b>) ALP staining of OB cell samples treated with mcur@BioCaP-16 on day 7, bar = 500 µm. (<b>B</b>) ALP activity of OB cells, <span class="html-italic">n</span> = 6. (<b>C</b>) ARS staining of OB cells samples treated with mcur@BioCaP-16 after US treatment at day 14. (<b>D</b>) RAS activity of OB cells, <span class="html-italic">n</span> = 6. Statistical difference <span class="html-italic">** p</span> &lt; 0.01.</p>
Full article ">
19 pages, 4738 KiB  
Article
Eco-Friendly Microwave Synthesis of Sodium Alginate-Chitosan Hydrogels for Effective Curcumin Delivery and Controlled Release
by Ivan Ristić, Ljubiša Nikolić, Suzana Cakić, Vesna Nikolić, Jelena Tanasić, Jelena Zvezdanović and Marija Krstić
Gels 2024, 10(10), 637; https://doi.org/10.3390/gels10100637 - 2 Oct 2024
Viewed by 605
Abstract
In this study, we developed sodium alginate-chitosan hydrogels using a microwave-assisted synthesis method, aligning with green chemistry principles for enhanced sustainability. This eco-friendly approach minimizes chemical use and waste while boosting efficiency. A curcumin:2-hydroxypropyl-β-cyclodextrin complex was incorporated into the hydrogels, significantly increasing the [...] Read more.
In this study, we developed sodium alginate-chitosan hydrogels using a microwave-assisted synthesis method, aligning with green chemistry principles for enhanced sustainability. This eco-friendly approach minimizes chemical use and waste while boosting efficiency. A curcumin:2-hydroxypropyl-β-cyclodextrin complex was incorporated into the hydrogels, significantly increasing the solubility and bioavailability of curcumin. Fourier Transform Infrared Spectroscopy (FTIR) analysis confirmed the structure and successful incorporation of curcumin, in both its pure and complexed forms, into the polymer matrix. Differential scanning calorimetry revealed distinct thermal transitions influenced by the hydrogel composition and physical cross-linking. Hydrogels with higher alginate content had higher swelling ratios (338%), while those with more chitosan showed the lowest swelling ratios (254%). Scanning Electron Microscopy (SEM) micrographs showed a porous structure as well as successful incorporation of curcumin or its complex. Curcumin release studies indicated varying releasing rates between its pure and complexed forms. The chitosan-dominant hydrogel exhibited the slowest release rate of pure curcumin, while the alginate-dominant hydrogel exhibited the fastest. Conversely, for curcumin from the inclusion complex, a higher chitosan proportion led to the fastest release rate, while a higher alginate proportion resulted in the slowest. This study demonstrates that the form of curcumin incorporation and gel matrix composition critically influence the release profile. Our findings offer valuable insights for designing effective curcumin delivery systems, representing a significant advancement in biodegradable and sustainable drug delivery technologies. Full article
(This article belongs to the Special Issue Designing Hydrogels for Sustained Delivery of Therapeutic Agents)
Show Figures

Graphical abstract

Graphical abstract
Full article ">Figure 1
<p>FTIR spectra of hydrogels 4A_1CH; 1A_1CH; and 1A_4CH.</p>
Full article ">Figure 2
<p>FTIR spectra of pure hydrogels 4A_1CH; with incorporated curcumin; and complex.</p>
Full article ">Figure 3
<p>FTIR spectra of pure hydrogels 1A_1CH; with incorporated curcumin; and with complex.</p>
Full article ">Figure 4
<p>FTIR spectra of pure hydrogels 1A_4CH; with incorporated curcumin; and with complex.</p>
Full article ">Figure 5
<p>DSC thermographs of Alg-Chi hydrogels.</p>
Full article ">Figure 6
<p>DSC thermographs of Alg-Chi hydrogels with incorporated (<b>a</b>) curcumin and (<b>b</b>) curcumin-2-hydroxypropyl-β-cyclodextrin complex.</p>
Full article ">Figure 6 Cont.
<p>DSC thermographs of Alg-Chi hydrogels with incorporated (<b>a</b>) curcumin and (<b>b</b>) curcumin-2-hydroxypropyl-β-cyclodextrin complex.</p>
Full article ">Figure 7
<p>Swelling of Alg-Chi hydrogels at pH 7.4.</p>
Full article ">Figure 8
<p>SEM micrographs of samples: (<b>a</b>) 1A_1CH; (<b>b</b>) 1A_4CH; (<b>c</b>) 4A_1CH.</p>
Full article ">Figure 9
<p>SEM micrographs of the xerogels with incorporated curcumin: (<b>a</b>) 1A_1CH_Cu; (<b>b</b>) 1A_4CH_Cu; (<b>c</b>) 4A_1CH_Cu.</p>
Full article ">Figure 10
<p>SEM micrographs of the xerogels with incorporated complex curcumin:2-hydroxypropyl-β-cyclodextrin: (<b>a</b>) 1A_1CH_Com; (<b>b</b>) 1A_4CH_Com; (<b>c</b>) 4A_1CH_Com.</p>
Full article ">Figure 11
<p>Profile of curcumin release, depending on time (t) in hours (h), from: (<b>a</b>) 1A_4CH_Cu, (<b>b</b>) 1A_4CH_Com, (<b>c</b>) 1A_1CH_Cu, (<b>d</b>) 1A_1CH_Com, (<b>e</b>) 4A_1CH_Cu, and (<b>f</b>) 4A_1CH_Com.</p>
Full article ">
Back to TopTop