[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

Article Types

Countries / Regions

remove_circle_outline
remove_circle_outline
remove_circle_outline
remove_circle_outline

Search Results (217)

Search Parameters:
Keywords = chrysin

Order results
Result details
Results per page
Select all
Export citation of selected articles as:
15 pages, 3143 KiB  
Article
Differential Interactions of Flavonoids with the Aryl Hydrocarbon Receptor In Silico and Their Impact on Receptor Activity In Vitro
by Monique Reis de Santana, Ylanna Bonfim dos Santos, Késsia Souza Santos, Manoelito Coelho Santos Junior, Mauricio Moraes Victor, Gabriel dos Santos Ramos, Ravena Pereira do Nascimento and Silvia Lima Costa
Pharmaceuticals 2024, 17(8), 980; https://doi.org/10.3390/ph17080980 - 24 Jul 2024
Viewed by 557
Abstract
The molecular mechanisms underlying the observed anticancer effects of flavonoids remain unclear. Increasing evidence shows that the aryl hydrocarbon receptor (AHR) plays a crucial role in neoplastic disease progression, establishing it as a potential drug target. This study evaluated the potential of hydroxy [...] Read more.
The molecular mechanisms underlying the observed anticancer effects of flavonoids remain unclear. Increasing evidence shows that the aryl hydrocarbon receptor (AHR) plays a crucial role in neoplastic disease progression, establishing it as a potential drug target. This study evaluated the potential of hydroxy flavonoids, known for their anticancer properties, to interact with AHR, both in silico and in vitro, aiming to understand the mechanisms of action and identify selective AHR modulators. A PAS-B domain homology model was constructed to evaluate in silico interactions of chrysin, naringenin, quercetin apigenin and agathisflavone. The EROD activity assay measured the effects of flavonoids on AHR’s activity in human breast cancer cells (MCF7). Simulations showed that chrysin, apigenin, naringenin, and quercetin have the highest AHR binding affinity scores (−13.14 to −15.31), while agathisflavone showed low scores (−0.57 and −5.14). All tested flavonoids had the potential to inhibit AHR activity in a dose-dependent manner in the presence of an agonist (TCDD) in vitro. This study elucidates the distinct modulatory effects of flavonoids on AHR, emphasizing naringenin’s newly described antagonistic potential. It underscores the importance of understanding flavonoid’s molecular mechanisms, which is crucial for developing novel cancer therapies based on these molecules. Full article
(This article belongs to the Special Issue Therapeutic Agents for the Treatment of Tumors in the CNS)
Show Figures

Graphical abstract

Graphical abstract
Full article ">Figure 1
<p>AHR PAS-B ligand-binding domain homology model. (<b>A</b>) Schematic structure of human AHR domains. Numbers at the domain boundaries refer to the amino acids of human proteins. (<b>B</b>) Three-dimensional representation of the PAS-B domain of human AHR. The model reveals the presence of four alpha helices along with a region characterized by antiparallel beta sheets. (<b>C</b>) Three-dimensional representation of 3 potential PAS-B ligand domain predicted binding sites. Amino acids of the binding pockets are highlighted in red. The binding regions were named based on which docking system they originated from. AHR PAS-B site 1: described in the work of Leclair et al., (2020) [<a href="#B23-pharmaceuticals-17-00980" class="html-bibr">23</a>]; AHR PAS-B site 2: described in the work of SZÖLLÖSI et al. (2016) [<a href="#B28-pharmaceuticals-17-00980" class="html-bibr">28</a>]; 3: calculated by CASTp (TIAN et al., 2018) [<a href="#B29-pharmaceuticals-17-00980" class="html-bibr">29</a>] using the search radius of 1.4 Å. Three-dimensional models were built on the SWISS-MODEL platform.</p>
Full article ">Figure 2
<p>Flavonoids interact differently with the AHR binding domain (PAS-B). (<b>a</b>) Heat map showing the flavonoids chrysin (CHRY), apigenin (API), naringenin (NAR), quercetin (QUER), and agathisflavone (FAB) and their putative interactions with amino acids of the PAS-B domain (site 2). (<b>b</b>) Heat map showing the flavonoids CHRY, API, NAR, QUER, and FAB and their putative interactions with amino acids of the PAS- B domain (site 3). (<b>c</b>) Two-dimensional diagram of flavonoid structures and proposed hydroxyl group positions of CHRY, API, NAR, QUER, and FAB hydrogen bond interaction with AHR residues in the AHR binding site model (PAS-B). Orange-colored areas indicate hydrogen bonds with AHR residues at AHR binding sites (PAS-B). Residue names colored in orange indicate interactions with residues at model AHR binding site 2 (PAS-B), and names colored in blue indicate interactions with residues at AHR binding site 3 (PAS-B).</p>
Full article ">Figure 3
<p>Flavonoids modulate the canonical AHR activity induced by TCDD. MCF7 cells were pretreated with the flavonoids (<b>A</b>) chrysin (1, 5, 10, 50 µM), (<b>B</b>) apigenin (1, 5, 10, 50 µM), (<b>C</b>) naringenin (5, 10, 20, 30 µM), (<b>D</b>) quercetin (1, 5, 10, 20 µM), and (<b>E</b>) agathisflavone (10, 20, 30 µM) for 2 h and exposed to agonist (TCDD 5 nM) for 6 h to measure the induction of CYP1A1 activity using an EROD activity assay. The CYP1A1 activity was analyzed considering the positive control condition (TCDD 5 nM). The results were compared to the control (100%) <span class="html-italic">n</span> = 3. The significance was evaluated by a one-way ANOVA test followed by the Tukey test; * <span class="html-italic">p</span> &lt; 0.05, ** <span class="html-italic">p</span> &lt; 0.01, *** <span class="html-italic">p</span> &lt; 0.001, **** <span class="html-italic">p</span> &lt; 0.0001.</p>
Full article ">
13 pages, 1516 KiB  
Article
Development of an Antioxidant, Anti-Aging, and Photoprotective Phytocosmetic from Discarded Agave sisalana Perrine Roots
by Guilherme dos Santos Mazo, Julia Amanda Rodrigues Fracasso, Luísa Taynara Silvério da Costa, Valdecir Farias Ximenes, Natália Alves Zoppe, Amanda Martins Viel, Lucas Pires Guarnier, Beatriz de Castro Silva, Luan Victor Coelho de Almeida and Lucinéia dos Santos
Cosmetics 2024, 11(3), 104; https://doi.org/10.3390/cosmetics11030104 - 20 Jun 2024
Viewed by 1269
Abstract
The primary source of hard fiber globally is Agave sisalana Perrine, also known as sisal. In areas where sisal is grown, the roots of the plant are usually left in the field after it has stopped producing, which leads to soil degradation and [...] Read more.
The primary source of hard fiber globally is Agave sisalana Perrine, also known as sisal. In areas where sisal is grown, the roots of the plant are usually left in the field after it has stopped producing, which leads to soil degradation and decreased sisal productivity. It is, therefore, critical to find alternatives to reuse this waste. This study explores the potential use of sisal waste in the cosmetic industry by incorporating a hydroethanolic extract (HER) into a cream–gel formulation, taking advantage of the plant’s recognized ethnopharmacological value. The study involves analyzing the extract’s phytochemical composition (flavonoids) and evaluating its cytotoxicity. Subsequently, the antioxidant and antiglycation activities of the extract and cream–gel are evaluated, as well as ex vivo ocular toxicity, photoprotective activity, and preliminary stability analyses. The HER extract showed a flavonoid composition (catechin, kaempferol, isorhamnetin, and chrysin) and maintained cell viability above 70% throughout all time points analyzed in the MTT assay. Furthermore, the extract and the formulation demonstrated proven antioxidant and antiglycation activities. The cream–gel’s UVB and UVA protection effectiveness with the HER was comparable to that of synthetic UVB/UVA sunscreens, with the samples proving nonirritating and stable. In conclusion, the extract has a significant presence of flavonoids, and the cream–gel developed with it did not present cytotoxicity and met the stability requirements, indicating phytocosmetic potential with antioxidant, antiglycation, and photoprotective properties. Full article
(This article belongs to the Special Issue Natural Sources for Cosmetic Ingredients: Challenges and Innovations)
Show Figures

Figure 1

Figure 1
<p>Chromatograms of (<b>A</b>) reference flavonoids and (<b>B</b>) HER fingerprint obtained by high-performance liquid chromatography. (1) rutin; (2) morin; (3) kaempferol; (4) isorhamnetin; (5) fisetin; (7) chrysin; and (8) catechin.</p>
Full article ">Figure 2
<p>Mean ± SD of cell viability of the NC—negative control (physiologic solution 0.9%), PC—positive control (2% Tween 80%), and the HER (100 µg/mL, 200 µg/mL, 400 µg/mL, 800 µg/mL, and 1600 µg/mL) by the MTT method. A one-way ANOVA followed by Tukey’s post hoc test was performed. An asterisk (*) indicates if there is a significant difference (<span class="html-italic">p</span> &lt; 0.05) from the negative control.</p>
Full article ">Figure 3
<p>Mean ± SD of values in % of antioxidant activity in the (<b>A</b>) DPPH and (<b>B</b>) Lipoperoxidation tests after the following treatments: NC—negative control (physiologic solution 0.9%), PC—positive control (quercetin solution 300 µg/mL), HER (50 µg/mL, 100 µg/mL, 200 µg/mL, 400 µg/mL, and 600 µg/mL), and the cream–gel with HER (C1—200 µg/mL, C2—400 µg/mL, and C3—600 µg/mL). A one-way ANOVA followed by Tukey’s post hoc test was performed. An asterisk (*) indicates if there is a significant difference (<span class="html-italic">p</span> &lt; 0.05) from the negative control.</p>
Full article ">Figure 4
<p>Mean ± SD of the values in % of anti-aging activity through (<b>A</b>) BSA/GLU and (<b>B</b>) BSA/MGO after the following treatments: NC—negative control (physiologic solution 0.9%), PC—positive control (quercetin 300 μg/mL), HER (50 µg/mL, 100 µg/mL, 200 µg/mL, 400 µg/mL, and 600 µg/mL), and the cream–gel with HER (C1—200 µg/mL, C2—400 µg/mL, and C3—600 µg/mL). A one-way ANOVA followed by Tukey’s post hoc test was performed. An asterisk (*) indicates if there is a significant difference (<span class="html-italic">p</span> &lt; 0.05) from the negative control.</p>
Full article ">
15 pages, 2301 KiB  
Article
The Neuroprotective Effects of Oroxylum indicum Extract in SHSY-5Y Neuronal Cells by Upregulating BDNF Gene Expression under LPS Induced Inflammation
by Shareena Sreedharan, Alpana Pande, Anurag Pande, Muhammed Majeed and Luis Cisneros-Zevallos
Nutrients 2024, 16(12), 1887; https://doi.org/10.3390/nu16121887 - 14 Jun 2024
Viewed by 1042
Abstract
The brain-derived neurotrophic factor (BDNF) plays a crucial role during neuronal development as well as during differentiation and synaptogenesis. They are important proteins present in the brain that support neuronal health and protect the neurons from detrimental signals. The results from the present [...] Read more.
The brain-derived neurotrophic factor (BDNF) plays a crucial role during neuronal development as well as during differentiation and synaptogenesis. They are important proteins present in the brain that support neuronal health and protect the neurons from detrimental signals. The results from the present study suggest BDNF expression can be increase up to ~8-fold by treating the neuroblastoma cells SHSY-5Y with an herbal extract of Oroxylum indicum (50 μg/mL) and ~5.5-fold under lipopolysaccharides (LPS)-induced inflammation conditions. The Oroxylum indicum extract (Sabroxy) was standardized to 10% oroxylin A, 6% chrysin, and 15% baicalein. In addition, Sabroxy has shown to possess antioxidant activity that could decrease the damage caused by the exacerbation of radicals during neurodegeneration. A mode of action of over expression of BDNF with and without inflammation is proposed for the Oroxylum indicum extract, where the three major hydroxyflavones exert their effects through additive or synergistic effects via five possible targets including GABA, Adenoside A2A and estrogen receptor bindings, anti-inflammatory effects, and reduced mitochondrial ROS production. Full article
(This article belongs to the Special Issue Preventive and Therapeutic Nutraceuticals)
Show Figures

Graphical abstract

Graphical abstract
Full article ">Figure 1
<p>HPLC chromatogram of a standardized <span class="html-italic">Oroxylum indicum</span> methanolic extract (Sabroxy).</p>
Full article ">Figure 2
<p>Mass spectrum analysis of a standardized <span class="html-italic">Oroxylum indicum</span> extract (Sabroxy). Mass spectrum of (<b>A</b>,<b>B</b>) peak 1—baicalein; (<b>C</b>,<b>D</b>) peak 2—chrysin; and (<b>E</b>,<b>F</b>) peak 3—Oroxylin A in ESI positive and ESI negative, respectively.</p>
Full article ">Figure 3
<p>Cell viability of SHSY-5Y neural cells after treatment with a standardized <span class="html-italic">Oroxylum indicum</span> extract (Sabroxy containing 10% OA, 6% CH, 15% BA) was determined by the MTS assay. Assays were performed on 3 replicates for each treatment. Viability measured by MTS is expressed as the percentage of controls at 24 h treatment. Data are expressed as mean ± standard errors (SE). One-way analysis of variance (ANOVA) followed by Tukey HSD showing values with a different letter(s) are statistically different (<span class="html-italic">p</span> ≤ 0.05).</p>
Full article ">Figure 4
<p>The effect of a standardized <span class="html-italic">Oroxylum indicum</span> extract (Sabroxy) on BDNF gene expression in SHSY-5Y neural cell with and without inflammation conditions. The gene expression analysis was performed by RT-PCR, using differentiated SHSY-5Y neural cells. BDNF expression increases with 10, 20, and 50 µg/mL Sabroxy in the presence and absence of inflammation induced by LPS (10 µg/mL). The gene transcripts were normalized using β-actin as a control. Data, obtained from n = 6 repeats at least, are shown as mean ± SE. One-way analysis of variance (ANOVA) followed by Tukey HSD showing values with a different letter(s) are statistically different (<span class="html-italic">p</span> &lt; 0.05). * t-student analysis showed significant difference between controls and LPS treated cells at <span class="html-italic">p</span> = 0.001.</p>
Full article ">Figure 5
<p>Effect of a standardized <span class="html-italic">Oroxylum indicum</span> extract (Sabroxy) in LPS-induced ROS expression status in SH-SY5Y neural cells after 19 h of incubation, using DCFHDA as a reporter. Data represent the mean ± SE of six independent experiments in triplicate. One-way analysis of variance (ANOVA) followed by Tukey HSD showing values with a different letter(s) are statistically different (<span class="html-italic">p</span> ≤ 0.05).</p>
Full article ">Figure 6
<p>Cellular Antioxidant Activity of different concentrations of a standardized <span class="html-italic">Oroxylum indicum</span> extract (Sabroxy) in differentiated SH-SY5Y neural cells. Cells were exposed to DCFCH-DA and Sabroxy extracts, then the ROS generator AAPH was added, and the reaction was monitored for 1.5 h. Data represent the mean ± SE of six independent experiments in triplicate. One-way analysis of variance (ANOVA) followed by Tukey HSD showing values with a different letter(s) are statistically different (<span class="html-italic">p</span> ≤ 0.05).</p>
Full article ">Figure 7
<p>Proposed mechanisms of action of a standardized <span class="html-italic">Oroxylum indicum</span> extract (Sabroxy) in SH-SY5Y neural cells with or without inflammation. Active components of the extract like oroxylin A (OA), chrysin (CH), or baicalein (BA) elicit upregulation of BDNF through five possible targets. Namely, OA binds to the GABA receptor and blocks its expression resulting in activation of the NMDA receptor that turns out in a sustained increase of intracellular calcium levels that will lead to the activation of two pathways. One related to Calcium/calmodulin-dependent protein kinase IV (CaMKIV) and the other related to the Adenosine A2A receptor and the cAMP/PAK-MAPK pathway. Both routes will lead to the phosphorylation of cyclic AMP response element binding protein (CREB) and the subsequent BDNF expression. CH and BA may exert positive allosteric modulation of GABA receptor enhancing the OA antagonist effect. The activation of the A2A receptor will also transactivate TrkB, which initiates the TrkB-Akt pathway for neuronal survival. In addition, CH may activate estrogen receptors, while BA may attenuate this effect. Furthermore, all three compounds in the <span class="html-italic">Oroxylum indicum</span> extract exhibit free-radical scavenging activities that could reduce the reactive oxygen species (ROS) produced by the mitochondria as well as anti-inflammatory properties by attenuating NF-kβ and reducing pro-inflammatory cytokine levels. In the proposed model LPS induces inflammation through activation of pNF-kβ, where cytokine IL-1β suppresses BDNF expression. A + signal indicates positive allosteric modulation of receptor. Adapted from: [<a href="#B12-nutrients-16-01887" class="html-bibr">12</a>,<a href="#B31-nutrients-16-01887" class="html-bibr">31</a>,<a href="#B32-nutrients-16-01887" class="html-bibr">32</a>,<a href="#B33-nutrients-16-01887" class="html-bibr">33</a>,<a href="#B36-nutrients-16-01887" class="html-bibr">36</a>,<a href="#B37-nutrients-16-01887" class="html-bibr">37</a>,<a href="#B38-nutrients-16-01887" class="html-bibr">38</a>,<a href="#B39-nutrients-16-01887" class="html-bibr">39</a>].</p>
Full article ">
26 pages, 1684 KiB  
Article
Exploring Bioactive Components and Assessing Antioxidant and Antibacterial Activities in Five Seaweed Extracts from the Northeastern Coast of Algeria
by Nawal Bouzenad, Nesrine Ammouchi, Nadjla Chaib, Mohammed Messaoudi, Walid Bousabaa, Chawki Bensouici, Barbara Sawicka, Maria Atanassova, Sheikh F. Ahmad and Wafa Zahnit
Mar. Drugs 2024, 22(6), 273; https://doi.org/10.3390/md22060273 - 12 Jun 2024
Cited by 1 | Viewed by 1322
Abstract
The main goal of this study was to assess the bioactive and polysaccharide compositions, along with the antioxidant and antibacterial potentials, of five seaweeds collected from the northeastern coast of Algeria. Through Fourier transform infrared spectroscopy analysis and X-ray fluorescence spectroscopy, the study [...] Read more.
The main goal of this study was to assess the bioactive and polysaccharide compositions, along with the antioxidant and antibacterial potentials, of five seaweeds collected from the northeastern coast of Algeria. Through Fourier transform infrared spectroscopy analysis and X-ray fluorescence spectroscopy, the study investigated the elemental composition of these seaweeds and their chemical structure. In addition, this study compared and identified the biochemical makeup of the collected seaweed by using cutting-edge methods like tandem mass spectrometry and ultra-high-performance liquid chromatography, and it searched for new sources of nutritionally valuable compounds. According to the study’s findings, Sargassum muticum contains the highest levels of extractable bioactive compounds, showing a phenolic compound content of 235.67 ± 1.13 µg GAE·mg−1 and a total sugar content of 46.43 ± 0.12% DW. Both S. muticum and Dictyota dichotoma have high concentrations of good polyphenols, such as vanillin and chrysin. Another characteristic that sets brown algae apart is their composition. It showed that Cladophora laetevirens has an extracted bioactive compound content of 12.07% and a high capacity to scavenge ABTS+ radicals with a value of 78.65 ± 0.96 µg·mL−1, indicating high antioxidant activity. In terms of antibacterial activity, S. muticum seaweed showed excellent growth inhibition. In conclusion, all five species of seaweed under investigation exhibited unique strengths, highlighting the variety of advantageous characteristics of these seaweeds, especially S. muticum. Full article
Show Figures

Figure 1

Figure 1
<p>FT-IR spectra of the five studied seaweeds that were collected from the northeastern coast of Algeria.</p>
Full article ">Figure 2
<p>Chemical structure of identified seaweed bioactive compounds.</p>
Full article ">Figure 3
<p>Overview of sampling locations.</p>
Full article ">
23 pages, 1405 KiB  
Article
Multi-Endpoint Toxicological Assessment of Chrysin Loaded Oil-in-Water Emulsion System in Different Biological Models
by Pornsiri Pitchakarn, Pisamai Ting, Pensiri Buacheen, Jirarat Karinchai, Woorawee Inthachat, Boonrat Chantong, Uthaiwan Suttisansanee, Onanong Nuchuchua and Piya Temviriyanukul
Nanomaterials 2024, 14(12), 1001; https://doi.org/10.3390/nano14121001 - 8 Jun 2024
Cited by 1 | Viewed by 919
Abstract
Chrysin is hypothesized to possess the ability to prevent different illnesses, such as diabetes, cancer, and neurodegenerative disorders. Nonetheless, chrysin has a low solubility under physiological conditions, resulting in limited bioavailability. In a previous study, we utilized an oil-in-water emulsion system (chrysin-ES or [...] Read more.
Chrysin is hypothesized to possess the ability to prevent different illnesses, such as diabetes, cancer, and neurodegenerative disorders. Nonetheless, chrysin has a low solubility under physiological conditions, resulting in limited bioavailability. In a previous study, we utilized an oil-in-water emulsion system (chrysin-ES or chrysin-NE) to encapsulate chrysin, thereby increasing its bioaccessibility and preserving its antioxidant and anti-Alzheimer’s properties. To promote the chrysin-ES as a supplementary and functional food, it was obligatory to carry out a safety assessment. Cytotoxicity testing showed that chrysin-ES was harmless, with no killing effect on 3T3-L1 (adipocytes), RAW 264.7 (macrophages), HEK293 (kidney cells), and LX-2 (hepatic stellate cells). The acute toxicity evaluation demonstrated that the 50% lethal dose (LD50) for chrysin-ES was greater than 2000 mg/kg BW. Genotoxicity assessments found that chrysin-ES did not induce DNA mutations in vitro or in vivo. Furthermore, chrysin and chrysin-ES exhibited anti-mutagenic properties against PhIP-induced and IQ-induced mutagenesis in the Ames test, while they inhibited urethane-, ethyl methanesulfonate-, mitomycin C-, and N-nitrosomethylurea-mediated mutations in Drosophila. The present study illustrates the safety and anti-genotoxicity properties of chrysin-ES, allowing for the further development of chrysin-based food supplements and nutraceuticals. Full article
(This article belongs to the Special Issue Advances in Toxicity of Nanoparticles in Organisms (2nd Edition))
Show Figures

Figure 1

Figure 1
<p>(<b>A</b>) Appearance of chrysin-ES. This chrysin-ES was obtained from the previous study (data from Ting et al., 2021 [<a href="#B11-nanomaterials-14-01001" class="html-bibr">11</a>]). (<b>B</b>) Morphological study of chrysin-ES upon Nile red staining under confocal fluorescence microscope (scale bar = 30 µm). When using Nile red staining, chrysin-ES revealed oil droplets dispersed in an aqueous solution. The oil droplets could be the micelle formation of chrysin/MCT oil surrounded by surfactant and co-surfactant used in the formulation. (<b>C</b>) Size distribution by volume of chrysin-ES. Chrysin-ES contained heterogeneous sizes of nano-colloids about 154.4 ± 64.62 nm in diameter (88.9% volume) and microparticles about 5498 ± 626.4 nm in diameter (9.8% volume). (<b>D</b>) Zeta potential distribution of chrysin-ES used in the present study.</p>
Full article ">Figure 2
<p>Food (<b>A</b>) and drink (<b>B</b>) consumptions of rats after administration of chrysin-ES at 300 and 2000 mg/kg BW. Values are presented as mean ± SD; N = 3.</p>
Full article ">
22 pages, 3260 KiB  
Article
Preliminary Investigation of Astragalus arpilobus subsp. hauarensis: LC-MS/MS Chemical Profiling, In Vitro Evaluation of Antioxidant, Anti-Inflammatory Properties, Cytotoxicity, and In Silico Analysis against COX-2
by Sabrina Lekmine, Ouided Benslama, Kenza Kadi, Abir Brik, Ouidad Djeffali, Manar Ounissi, Meriem Slimani, Mohammad Shamsul Ola, Omayma A. Eldahshan, Antonio Ignacio Martín-García and Ahmad Ali
Antioxidants 2024, 13(6), 654; https://doi.org/10.3390/antiox13060654 - 27 May 2024
Cited by 1 | Viewed by 1780
Abstract
The search results offer comprehensive insights into the phenolic compounds, antioxidant, anti-inflammatory, cytotoxic effects, LC-MS/MS analysis, molecular docking, and MD simulation of the identified phenolic compounds in the Astragalus arpilobus subsp. hauarensis extract (AAH). The analysis revealed substantial levels of total phenolic content [...] Read more.
The search results offer comprehensive insights into the phenolic compounds, antioxidant, anti-inflammatory, cytotoxic effects, LC-MS/MS analysis, molecular docking, and MD simulation of the identified phenolic compounds in the Astragalus arpilobus subsp. hauarensis extract (AAH). The analysis revealed substantial levels of total phenolic content (TPC), with a measured value of 191 ± 0.03 mg GAE/g DM. This high TPC was primarily attributed to two key phenolic compounds: total flavonoid content (TFC) and total tannin content (TTC), quantified at 80.82 ± 0.02 mg QE/g DM and 51.91 ± 0.01 mg CE/g DM, respectively. LC-MS/MS analysis identified 28 phenolic compounds, with gallic acid, protocatechuic acid, catechin, and others. In the DPPH scavenging assay, the IC50 value for the extract was determined to be 19.44 ± 0.04 μg/mL, comparable to standard antioxidants like BHA, BHT, ascorbic acid, and α-tocopherol. Regarding anti-inflammatory activity, the extract demonstrated a notably lower IC50 value compared to both diclofenac and ketoprofen, with values of 35.73 µg/mL, 63.78 µg/mL, and 164.79 µg/mL, respectively. Cytotoxicity analysis revealed significant cytotoxicity of the A. arpilobus extract, with an LC50 value of 28.84 µg/mL, which exceeded that of potassium dichromate (15.73 µg/mL), indicating its potential as a safer alternative for various applications. Molecular docking studies have highlighted chrysin as a promising COX-2 inhibitor, with favorable binding energies and interactions. Molecular dynamic simulations further support chrysin’s potential, showing stable interactions with COX-2, comparable to the reference ligand S58. Overall, the study underscores the pharmacological potential of A. arpilobus extract, particularly chrysin, as a source of bioactive compounds with antioxidant and anti-inflammatory properties. Further research is warranted to elucidate the therapeutic mechanisms and clinical implications of these natural compounds. Full article
(This article belongs to the Special Issue Antioxidant and Protective Effects of Plant Extracts)
Show Figures

Figure 1

Figure 1
<p>2D interaction modes of S58 and the best docked phytocompounds with COX-2 (PDB:1CX).</p>
Full article ">Figure 2
<p>Pharmacophore modeling of S58 and the best docked phytocompounds. The red contour represents acceptor/donor H-bond (Acc/Don), the green contour represents hydrophobic region, the orange contour represents aromatic/hydrophobic regions, and purple contour represents Pi-charge electrostatic forming regions.</p>
Full article ">Figure 3
<p>(<b>A</b>) RMSD of the complexes 1CX2-S58 and 1CX2-Chrysin observed during 100 ns MD simulation. (<b>B</b>) RMSF of the complexes 1CX2-S58 and 1CX2-Chrysin during 100 ns MD simulation. (<b>C</b>) The radius of gyration of the complexes 1CX2-S58 and 1CX2-Chrysin during 100 ns MD simulation. (<b>D</b>) SASA of the complexes 1CX2-S58 and 1CX2-Chrysin observed during 100 ns MD simulation. (<b>E</b>) Number of H-bonds of the two complexes 1CX2-S58 and 1CX2-Chrysin during 100 ns MD simulation.</p>
Full article ">Figure 3 Cont.
<p>(<b>A</b>) RMSD of the complexes 1CX2-S58 and 1CX2-Chrysin observed during 100 ns MD simulation. (<b>B</b>) RMSF of the complexes 1CX2-S58 and 1CX2-Chrysin during 100 ns MD simulation. (<b>C</b>) The radius of gyration of the complexes 1CX2-S58 and 1CX2-Chrysin during 100 ns MD simulation. (<b>D</b>) SASA of the complexes 1CX2-S58 and 1CX2-Chrysin observed during 100 ns MD simulation. (<b>E</b>) Number of H-bonds of the two complexes 1CX2-S58 and 1CX2-Chrysin during 100 ns MD simulation.</p>
Full article ">
11 pages, 3121 KiB  
Article
A Comparison Study on the Metabolites in PC-3, RWPE-1, and Chrysin-Treated PC-3 Cells
by Jae-Hyeon Lee, Jung-Eun Kim, Eun-Ok Lee and Hyo-Jeong Lee
Appl. Sci. 2024, 14(10), 4255; https://doi.org/10.3390/app14104255 - 17 May 2024
Viewed by 732
Abstract
Prostate cancer is frequently diagnosed and the leading cause of death in men worldwide. Prostate-specific antigen (PSA) blood tests and biopsies are the primary methods for diagnosing prostate cancer; however, their accuracy is less than 50%. Therefore, there is a need to develop [...] Read more.
Prostate cancer is frequently diagnosed and the leading cause of death in men worldwide. Prostate-specific antigen (PSA) blood tests and biopsies are the primary methods for diagnosing prostate cancer; however, their accuracy is less than 50%. Therefore, there is a need to develop diagnostic tests that minimize patient discomfort during examination and adequate biomarkers that are more accurate, sensitive, and specific for the detection of prostate cancer. This study investigated the application of metabolomics to identify biomarkers in prostate cancer biofluids. In addition, changes in prostate cancer metabolite levels induced by chrysin, a natural anticancer compound, were evaluated and compared with those in non-treated prostate cancer cells. Gas chromatography-mass spectrometry (GC-MS)-based metabolomic profiling was performed to investigate the differences in metabolic alterations among prostate cancer, normal prostate, and chrysin-treated prostate cancer cells. Pairwise comparisons of the extracellular fluid metabolomes were performed using principal component analysis (PCA), partial least squares–discriminant analysis (PLS-DA), and Student’s t-test. The results revealed significantly different patterns among the metabolite groups, including alcohols, amino acids, carboxylic acids, organic acids, sugars, and urea. The RWPE-1- and chrysin-treated PC-3 (PC-3 Chr) cell groups showed similar tendencies for 23 metabolites, while the groups showed significant differences from the PC-3 group. Most amino acids showed higher concentrations in PC-3 cells than in the normal cell line RWPE-1 cells and PC-3 Chr cells. Our results revealed that GC-MS might be an effective diagnostic tool to detect prostate cancer and contribute to finding new tumor markers for prostate cancer as the basis for new ideas. Full article
(This article belongs to the Special Issue Advances in Biological Activities of Natural Products)
Show Figures

Figure 1

Figure 1
<p>PCA plot for each group. The sample numbers used in the PCA plot for group 1: RWPE-1 cells, 2; PC-3 cells, 3; chrysin-treated PC-3 cells, 2; respectively. After drawing the peak information matrix extracted from the total ion graph, the PCA plot was generated by analyzing the principal component with Simca-P 17.0. PCA, principal component analysis.</p>
Full article ">Figure 2
<p>Bar graphs of the 18 most significant metabolites in the analysis of variance results comparing the two groups (i.e., RWPE-1 and PC-3 cells). Data represent the mean ± SD. * <span class="html-italic">p</span> &lt; 0.05, ** <span class="html-italic">p</span> &lt; 0.01, and *** <span class="html-italic">p</span> &lt; 0.001 versus RWPE-1 cell group.</p>
Full article ">Figure 3
<p>Bar graphs of the 15 most significant metabolites in the analysis of variance results comparing the two groups (PC-3 and chrysin-treated PC-3 cells (PC-3 Chr)). Data represent the mean ± SD. * <span class="html-italic">p</span> &lt; 0.05, ** <span class="html-italic">p</span> &lt; 0.01, and *** <span class="html-italic">p</span> &lt; 0.001 versus the PC-3 cell group.</p>
Full article ">Figure 4
<p>Bar graphs of the 23 most significant metabolites in the analysis of variance results comparing the three groups (RWPE-1, PC-3, and chrysin-treated PC-3 cells (PC-3 Chr)). Data represent the mean ± SD. a~c means in a row by different superscripts are significantly different by LSD (least significant difference) at <span class="html-italic">p</span> &lt; 0.05.</p>
Full article ">
24 pages, 5150 KiB  
Article
Phytochemical, In Vitro, In Vivo, and In Silico Research on the Extract of Ajuga chamaepitys (L.) Schreb.
by Elis Ionus, Verginica Schröder, Carmen Lidia Chiţescu, Laura Adriana Bucur, Carmen Elena Lupu, Denisa-Elena Dumitrescu, Liliana Popescu, Dragoș Paul Mihai, Octavian Tudorel Olaru, George Mihai Nițulescu, Rica Boscencu and Cerasela Elena Gîrd
Plants 2024, 13(9), 1192; https://doi.org/10.3390/plants13091192 - 25 Apr 2024
Cited by 2 | Viewed by 1214
Abstract
The present study focuses on the chemical characterization of a dry extract obtained from the species Ajuga chamaepitys (L.) Schreb, evaluating its antioxidant properties, toxicity, and in silico profile. Quantitative analysis of the dry extract revealed a notable amount of phytochemical compounds: 59.932 [...] Read more.
The present study focuses on the chemical characterization of a dry extract obtained from the species Ajuga chamaepitys (L.) Schreb, evaluating its antioxidant properties, toxicity, and in silico profile. Quantitative analysis of the dry extract revealed a notable amount of phytochemical compounds: 59.932 ± 21.167 mg rutin equivalents (mg REs)/g dry weight, 45.864 ± 4.434 mg chlorogenic acid equivalents (mg ChAEs)/g dry weight and, respectively, 83.307 ± 3.989 mg tannic acid equivalents (TAEs)/g dry weight. By UHPLC-HRMS/MS, the following were quantified as major compounds: caffeic acid (3253.8 μg/g extract) and kaempherol (3041.5 μg/g extract); more than 11 types of polyphenolic compounds were quantified (genistin 730.2 μg/g extract, naringenin 395 μg/g extract, apigenin 325.7 μg/g extract, galangin 283.3 μg/g extract, ferulic acid 254.3 μg/g extract, p-coumaric acid 198.2 μg/g extract, rutin 110.6 μg/g extract, chrysin 90.22 μg/g extract, syringic acid 84.2 μg/g extract, pinocembrin 32.7 μg/g extract, ellagic acid 18.2 μg/g extract). The antioxidant activity was in accordance with the amount of phytochemical compounds: IC50DPPH = 483.6 ± 41.4 µg/mL, IC50ABTS•+ = 127.4 ± 20.2 µg/mL, and EC50FRAP = 491.6 ± 2 µg/mL. On the larvae of Artemia sp., it was found that the extract has a low cytotoxic action. In silico studies have highlighted the possibility of inhibiting the activity of protein kinases CDK5 and GSK-3b for apigenin, galangin, and kaempferol, with possible utility for treating neurodegenerative pathologies and neuropathic pain. Further studies are warranted to confirm the predicted molecular mechanisms of action and to further investigate the therapeutic potential in animal models of neurological disorders. Full article
Show Figures

Figure 1

Figure 1
<p>ACH located in Dobrogea Gorges, Romania.</p>
Full article ">Figure 2
<p>Results of the spectrophotometric determinations of the dry extract of ACHE. The data are represented as mean ± SD and analyzed using ANOVA with a Tukey post hoc test. * <span class="html-italic">p</span> &lt; 0.05, ns: <span class="html-italic">p</span> &gt; 0.05 no statistical differences.</p>
Full article ">Figure 3
<p>UHPLC-HRMS/MS chromatogram for ACHE in which the following were identified (top to bottom): apigenin (<span class="html-italic">m</span>/<span class="html-italic">z</span>: 269.04502, Rt: 24.11), kaempferol (<span class="html-italic">m</span>/<span class="html-italic">z</span>: 285.04049, Rt: 23.2), naringenin (<span class="html-italic">m</span>/<span class="html-italic">z</span>: 271.06122, Rt: 22.71), chrysin (<span class="html-italic">m</span>/<span class="html-italic">z</span>: 253.05066, Rt: 25.72), chlorogenic/neochlorogenic acid (<span class="html-italic">m</span>/<span class="html-italic">z</span>: 353.08783, Rt: 10.64/13.86), caffeic acid (<span class="html-italic">m</span>/<span class="html-italic">z</span>: 179.03501, Rt: 14.48).</p>
Full article ">Figure 4
<p>UHPLC-HRMS/MS chromatogram for ACHE in which the following were identified (top to bottom): biochanin A (<span class="html-italic">m</span>/<span class="html-italic">z</span>: 283.06122, Rt: 26.21), pratensein (<span class="html-italic">m</span>/<span class="html-italic">z</span>: 299.05614, Rt: 24.28), apigenin-7-O-glucosylglucoside (<span class="html-italic">m</span>/<span class="html-italic">z</span>: 593.1512176, Rt: 17.90), luteolin (<span class="html-italic">m</span>/<span class="html-italic">z</span>: 285.04049, Rt: 23.20), genistin (<span class="html-italic">m</span>/<span class="html-italic">z</span>: 431.09837, Rt: 19.64), vitexin (apigenin-8-C-glucoside)/isovitexin (<span class="html-italic">m</span>/<span class="html-italic">z</span>: 431.09839, Rt: 20.19/21.37), cynaroside (luteolin-7-O-glucoside) (<span class="html-italic">m</span>/<span class="html-italic">z</span>: 447.093284, Rt: 20.31).</p>
Full article ">Figure 5
<p>Antioxidant activity of ACHE. The data are represented as mean ± SD and analyzed using ANOVA with a Tukey post hoc test. *** <span class="html-italic">p</span> &lt; 0.001, ns: <span class="html-italic">p</span> &gt; 0.05 no statistical differences.</p>
Full article ">Figure 6
<p>The lethality curves obtained after 48 h of exposure of <span class="html-italic">Daphnia pulex</span> to ACHE; error bars represent the SE of two replicates.</p>
Full article ">Figure 7
<p><span class="html-italic">Daphnia magna</span> embryo test: (<b>a</b>)—embryos at 0 h; (<b>b</b>)—intermediary larval stage treated with 500 µg/mL ACHE at 24 h; (<b>c</b>)—intermediary larval stage treated with 1000 µg/mL ACHE at 24 h; (<b>d</b>)—normal larvae treated with 500 µg/mL ACHE at 48 h; (<b>e</b>)—undeveloped egg treated with 1000 µg/mL ACHE at 48 h; (<b>f</b>)—larvae treated with 1000 µg/mL ACHE at 48 h; (<b>g</b>)—untreated control at 24 h; (<b>h</b>)—untreated control at 48 h.</p>
Full article ">Figure 8
<p>(<b>a</b>) Superposition of predicted pose (green) of CDK5 inhibitor on experimental conformation (pink); (<b>b</b>) Superposition of predicted pose (green) of GSK-3β inhibitor on experimental conformation (pink); (<b>c</b>) Docked poses of all screened phytochemicals into CDK5 ATP-binding site; (<b>d</b>) Docked poses of all screened phytochemicals into GSK-3β ATP-binding site.</p>
Full article ">Figure 9
<p>(<b>a</b>) Predicted conformation of apigenin in complex with CDK5; (<b>b</b>) 2D diagram illustrating predicted interactions between apigenin and CDK5 ATP-binding site; (<b>c</b>) Predicted conformation of galangin in complex with CDK5; (<b>d</b>) 2D diagram illustrating predicted interactions between galangin and CDK5 ATP-binding site.</p>
Full article ">Figure 10
<p>(<b>a</b>) Predicted conformation of apigenin in complex with GSK-3β; (<b>b</b>) 2D diagram illustrating predicted interactions between apigenin and GSK-3β ATP-binding site; (<b>c</b>) Predicted conformation of kaempferol in complex with GSK-3β; (<b>d</b>) 2D diagram illustrating predicted interactions between kaempferol and GSK-3β ATP-binding site.</p>
Full article ">
19 pages, 6916 KiB  
Article
Chrysin Inhibits TAMs-Mediated Autophagy Activation via CDK1/ULK1 Pathway and Reverses TAMs-Mediated Growth-Promoting Effects in Non-Small Cell Lung Cancer
by Xinglinzi Tang, Xiaoru Luo, Xiao Wang, Yi Zhang, Jiajia Xie, Xuan Niu, Xiaopeng Lu, Xi Deng, Zheng Xu and Fanwei Wu
Pharmaceuticals 2024, 17(4), 515; https://doi.org/10.3390/ph17040515 - 17 Apr 2024
Viewed by 1250
Abstract
The natural flavonoid compound chrysin has promising anti-tumor effects. In this study, we aimed to investigate the mechanism by which chrysin inhibits the growth of non-small cell lung cancer (NSCLC). Through in vitro cell culture and animal models, we explored the impact of [...] Read more.
The natural flavonoid compound chrysin has promising anti-tumor effects. In this study, we aimed to investigate the mechanism by which chrysin inhibits the growth of non-small cell lung cancer (NSCLC). Through in vitro cell culture and animal models, we explored the impact of chrysin on the growth of NSCLC cells and the pro-cancer effects of tumor-associated macrophages (TAMs) and their mechanisms. We observed that M2-TAMs significantly promoted the growth and migration of NSCLC cells, while also markedly activating the autophagy level of these cells. Chrysin displayed a significant inhibitory effect on the growth of NSCLC cells, and it could also suppress the pro-cancer effects of M2-TAMs and inhibit their mediated autophagy. Furthermore, combining network pharmacology, we found that chrysin inhibited TAMs-mediated autophagy activation in NSCLC cells through the regulation of the CDK1/ULK1 signaling pathway, rather than the classical mTOR/ULK1 signaling pathway. Our study reveals a novel mechanism by which chrysin inhibits TAMs-mediated autophagy activation in NSCLC cells through the regulation of the CDK1/ULK1 pathway, thereby suppressing NSCLC growth. This discovery not only provides new therapeutic strategies for NSCLC but also opens up new avenues for further research on chrysin. Full article
Show Figures

Graphical abstract

Graphical abstract
Full article ">Figure 1
<p>The impact of M2-TAMs on the proliferation and migration abilities of NSCLC cells was assessed. (<b>A</b>) Macrophage polarization was detected using WB; (<b>B</b>) macrophage polarization was evaluated through flow cytometry; (<b>C</b>) cell counting was used to examine the influence of M2-TAMs on the proliferation ability of A549 and H157 NSCLC cells; (<b>D</b>) clone formation experiment was conducted to assess the effect of M2-TAMs on the clonogenic potential of A549 and H157 NSCLC cells; (<b>E</b>) scratch and transwell assays were performed to investigate the influence of M2-TAMs on the migratory ability of A549 cells; (<b>F</b>) scratch and transwell assays were carried out to determine the impact of M2-TAMs on the migration ability of H157 cells. Scale bars in the images for both scratch and transwell assays represent 100 μm. All values are presented as mean ± SD (<span class="html-italic">n</span> = 3, * <span class="html-italic">p</span> &lt; 0.05, ** <span class="html-italic">p</span> &lt; 0.01 vs. Ctrl group).</p>
Full article ">Figure 2
<p>Chrysin inhibited the growth of NSCLC cells and reversed the pro-tumor effect of TAMs. (<b>A</b>,<b>B</b>) CCK-8 assay demonstrated that chrysin (0–20 µM) suppressed the proliferation of A549 and H157 cells; (<b>C</b>) CCK-8 assay indicated that chrysin did not exhibit a significant inhibitory effect on the normal human lung epithelial cell BEAS-2B; (<b>D</b>) cell counting and (<b>E</b>) colony formation assay showed that chrysin (8 µM) attenuated the TAMs-induced proliferation of both NSCLC cell lines; (<b>F</b>) scratch and transwell assays revealed that chrysin (8 µM) reversed the TAMs-induced migration-promoting effect on A549 cells; (<b>G</b>) scratch and transwell assays demonstrated that chrysin (8 µM) also reversed the TAMs-induced migration-promoting effect on H157 cells. Scale bars in the images for both scratch and transwell assays represent 100 μm. All values are presented as mean ± SD (<span class="html-italic">n</span> = 3, * <span class="html-italic">p</span> &lt; 0.05, ** <span class="html-italic">p</span> &lt; 0.01 vs. TAMs group).</p>
Full article ">Figure 3
<p>Chrysin inhibited the polarization of macrophages towards the M2 phenotype and inhibited autophagy activation mediated by TAMs. (<b>A</b>) Flow cytometry results demonstrated that chrysin inhibited macrophage polarization into M2-TAMs; (<b>B</b>) LC3-mRFP-GFP lentiviral vectors transfection was used to investigate the effect of chrysin (8 µM) and 3-MA (10 mM) on autophagic flux in A549 cells treated with TAMs-conditioned media; (<b>C</b>) Western blotting assessed the expression of autophagy-related proteins, LC3 and p62, in A549 cells treated with TAMs-conditioned media and treated with chrysin (8 µM) or 3-MA (10 mM); (<b>D</b>) Western blotting examined the expression of early autophagy proteins, mTOR, p-mTOR, ULK1, and p-ULK1, in A549 cells treated with TAM-conditioned media and treated with chrysin (8 µM) or 3-MA (10 mM). Scale bars in the immunofluorescence images represent 10 μm. All values are presented as mean ± SD (<span class="html-italic">n</span> = 3, * <span class="html-italic">p</span> &lt; 0.05, ** <span class="html-italic">p</span> &lt; 0.01 vs. TAMs group, NS indicates nonsignificant).</p>
Full article ">Figure 4
<p>Chrysin reversed the pro-cancer effects of TAMs by inhibiting autophagy. (<b>A</b>) The colony formation assay reflected the effects of chrysin (8 µM), 3-MA (10 mM), and aloperine (100 µM) on the clonogenicity of A549 and H157 cells; (<b>B</b>) representative images of scratch assays conducted on A549 and H157 cells after the indicated treatments; (<b>C</b>) representative images of scratch and transwell assays conducted on A549 and H157 cells after the respective treatments. Scale bars in the images for both scratch and transwell assays represent 100 μm. All values are presented as mean ± SD (<span class="html-italic">n</span> = 3, * <span class="html-italic">p</span> &lt; 0.05, ** <span class="html-italic">p</span> &lt; 0.01 vs. TAMs+ chrysin group, NS indicates nonsignificant).</p>
Full article ">Figure 5
<p>KEGG pathway enrichment and GO analysis for “Chrysin”, “Non-small Cell Lung Cancer”, “Tumor Associated Macrophages”, and “Autophagy”. (<b>A</b>) Venn diagram illustrating the overlap among the targets of “Chrysin”, “Non-small Cell Lung Cancer”, “Tumor Associated Macrophages”, and “Autophagy”; (<b>B</b>) bubble chart showing the KEGG terms for enrichment analysis; (<b>C</b>) bubble chart depicting the GO biological processes analysis.</p>
Full article ">Figure 6
<p>Molecular docking. (<b>A</b>) The image shows chrysin docked into the active site of CCNB1; (<b>B</b>) the image depicts chrysin docked in the active site of CCNB2; (<b>C</b>) the image displays chrysin docked within the active site of CDK6; (<b>D</b>) the image illustrates chrysin docked at the active site of CDK1.</p>
Full article ">Figure 7
<p>Chrysin inhibited autophagy through the CDK1/ULK1 pathway. (<b>A</b>) The expression of CDK1, CDK6, CCNB1, and CCNB2 in A549 cells was detected via Western blot after the specified treatments. (<b>B</b>) The expression of CDK1, LC3, p62, mTOR, p-mTOR, ULK1, and p-ULK1 in A549 cells was examined by Western blot after treatment with TAMs and exposure to chrysin (8 μM) and/or TC11 (5 μM). (<b>C</b>) Histogram of the grayscale values for CDK1 in (<b>B</b>). (<b>D</b>) Histogram of the grayscale values for p-mTOR/mTOR and p-ULK/ULK1 in (<b>B</b>). (<b>E</b>) Histogram of the grayscale values for LC3 II/LC3 I and p62 in (<b>B</b>). All values are presented as mean ± SD (<span class="html-italic">n</span> = 3, * <span class="html-italic">p</span> &lt; 0.05, ** <span class="html-italic">p</span> &lt; 0.01 vs. TAMs + chrysin group, NS indicates nonsignificant).</p>
Full article ">Figure 8
<p>Chrysin reversed the pro-tumor effect of TAMs in vivo. (<b>A</b>) Photographic images of mouse tumors; (<b>B</b>) tumor weight of mice after treatment with TAMs and chrysin (30 mg/kg); (<b>C</b>) tumor volume change in mice over 21 days; (<b>D</b>) body weight change in mice over 21 days; (<b>E</b>) H&amp;E staining and IHC analysis investigating the effects of TAMs and chrysin on the proliferation of lung cancer cells, as well as the expression of LC3, P62, and CDK1 proteins. Scale bars in the images for both H&amp;E staining and IHC assays represent 100 μm. All values are presented as mean ± SD (* <span class="html-italic">p</span> &lt; 0.05, ** <span class="html-italic">p</span> &lt; 0.01 vs. TAMs group).</p>
Full article ">
14 pages, 6545 KiB  
Article
Exploring the Chemical Profile, In Vitro Antioxidant and Anti-Inflammatory Activities of Santolina rosmarinifolia Extracts
by Janos Schmidt, Kata Juhasz and Agnes Bona
Molecules 2024, 29(7), 1515; https://doi.org/10.3390/molecules29071515 - 28 Mar 2024
Viewed by 845
Abstract
In this study, the phytochemical composition, in vitro antioxidant, and anti-inflammatory effects of the aqueous and 60% ethanolic (EtOH) extracts of Santolina rosmarinifolia leaf, flower, and root were examined. The antioxidant activity of S. rosmarinifolia extracts was determined by 2,2’-azino-bis(3-ethylbenzothiazoline-6-sulfonic acid) (ABTS) and [...] Read more.
In this study, the phytochemical composition, in vitro antioxidant, and anti-inflammatory effects of the aqueous and 60% ethanolic (EtOH) extracts of Santolina rosmarinifolia leaf, flower, and root were examined. The antioxidant activity of S. rosmarinifolia extracts was determined by 2,2’-azino-bis(3-ethylbenzothiazoline-6-sulfonic acid) (ABTS) and 2,2-diphenyl-1-picrylhydrazyl (DPPH) radical scavenging assays. The total phenolic content (TPC) of the extracts was measured by the Folin–Ciocalteu assay. The anti-inflammatory effect of the extracts was monitored by the Griess assay. The chemical composition of S. rosmarinifolia extracts was analysed using the LC-MS technique. According to our findings, 60% EtOH leaf extracts showed the highest Trolox equivalent antioxidant capacity (TEAC) values in both ABTS (8.39 ± 0.43 µM) and DPPH (6.71 ± 0.03 µM) antioxidant activity assays. The TPC values of the samples were in good correspondence with the antioxidant activity measurements and showed the highest gallic acid equivalent value (130.17 ± 0.01 µg/mL) in 60% EtOH leaf extracts. In addition, the 60% EtOH extracts of the leaves were revealed to possess the highest anti-inflammatory effect. The LC-MS analysis of S. rosmarinifolia extracts proved the presence of ascorbic acid, catalpol, chrysin, epigallocatechin, geraniol, isoquercitrin, and theanine, among others, for the first time. However, additional studies are needed to investigate the direct relationship between the chemical composition and physiological effects of the herb. The 60% EtOH extracts of S. rosmarinifolia leaves are potential new sources of natural antioxidants and anti-inflammatory molecules in the production of novel nutraceutical products. Full article
Show Figures

Graphical abstract

Graphical abstract
Full article ">Figure 1
<p>The ABTS radical scavenging activity of <span class="html-italic">S. rosmarinifolia</span> 60% EtOH and aqueous extracts expressed as Trolox equivalent antioxidant capacity (TEAC) in micromolar concentration (µM). The error bars show the standard deviation calculated from a triplicate of samples. Significance levels were determined using two-way ANOVA with Tukey’s test and are indicated as **** = <span class="html-italic">p</span> &lt; 0.0001.</p>
Full article ">Figure 2
<p>The DPPH radical scavenging activity of <span class="html-italic">S. rosmarinifolia</span> 60% EtOH and aqueous extracts expressed as Trolox equivalent antioxidant capacity (TEAC) in micromolar concentration (µM). The error bars show the standard deviation calculated from a triplicate of samples. Significance levels were determined using two-way ANOVA with Tukey’s test and are indicated as **** = <span class="html-italic">p</span> &lt; 0.0001.</p>
Full article ">Figure 3
<p>The total phenolic content (TPC) of <span class="html-italic">S. rosmarinifolia</span> 60% EtOH and aqueous extracts expressed as gallic acid equivalent (GAE) in µg/mL. The error bars show the standard deviation calculated from a triplicate of samples. Significance levels were determined using two-way ANOVA with Tukey’s test and are indicated as **** = <span class="html-italic">p</span> &lt; 0.0001.</p>
Full article ">Figure 4
<p>Anti-inflammatory activity of aqueous and 60% EtOH <span class="html-italic">S. rosmarinifolia</span> extracts: (<b>A</b>) Effect of <span class="html-italic">S. rosmarinifolia</span> aqueous extracts on NO production in LPS-treated RAW 264.7 cells. Murine macrophage cells were pretreated with the indicated dilutions of leaf, flower, or root stock extracts of <span class="html-italic">S. rosmarinifolia</span> for 6 h, and then cells were challenged by 1 µg/mL LPS for 24 h. NO production was measured using the Griess reagent. (<b>B</b>) Effect of <span class="html-italic">S. rosmarinifolia</span> 60% EtOH extracts on NO production in LPS-treated RAW 264.7 cells. Murine macrophage cells were pretreated with the indicated dilutions of leaf, flower, or root stock extracts of <span class="html-italic">S. rosmarinifolia</span> for 6 h, and then cells were challenged by 1 µg/mL LPS for 24 h. NO production was measured using the Griess reagent. The error bars show the standard deviation calculated from five parallel samples. Significance levels were determined using two-way ANOVA with Tukey’s test and are indicated as ** = <span class="html-italic">p</span> &lt; 0.01, *** = <span class="html-italic">p</span> &lt; 0.001, **** = <span class="html-italic">p</span> &lt; 0.0001, <sup>##</sup> = <span class="html-italic">p</span> &lt; 0.01, <sup>###</sup> = <span class="html-italic">p</span> &lt; 0.001, <sup>####</sup> = <span class="html-italic">p</span> &lt; 0.0001.</p>
Full article ">Figure 5
<p>Identification of α-cyperone in <span class="html-italic">S. rosmarinifolia</span> using MS-DIAL ESI(+)-MS/MS database from authentic standards. The MS<sup>2</sup> fragment spectrum of the detected component (measurement) was matched (match score: 1.5827) with the MS<sup>2</sup> fragment spectrum of the molecule from the database (reference).</p>
Full article ">Figure 6
<p>The identified components and their relative abundance in <span class="html-italic">S. rosmarinifolia</span> leaf, flower, and root extracts. The colours represent the ion intensity of individual components; low or ND (white) indicate molecules with relative intensity values &lt; 1 × 10<sup>3</sup>, medium (light green) includes molecules within the relative intensity range of 1 × 10<sup>3</sup> and 1 × 10<sup>4</sup>, and those above 1 × 10<sup>4</sup> relative intensity values are marked as high (dark green).</p>
Full article ">
23 pages, 636 KiB  
Review
Flavonoids as CYP3A4 Inhibitors In Vitro
by Martin Kondža, Ivica Brizić and Stela Jokić
Biomedicines 2024, 12(3), 644; https://doi.org/10.3390/biomedicines12030644 - 13 Mar 2024
Viewed by 1794
Abstract
Flavonoids, a diverse group of polyphenolic compounds found abundantly in fruits, vegetables, and beverages like tea and wine, offer a plethora of health benefits. However, they have a potential interaction with drug metabolism, particularly through the inhibition of the cytochrome P450 3A4 enzyme, [...] Read more.
Flavonoids, a diverse group of polyphenolic compounds found abundantly in fruits, vegetables, and beverages like tea and wine, offer a plethora of health benefits. However, they have a potential interaction with drug metabolism, particularly through the inhibition of the cytochrome P450 3A4 enzyme, the most versatile and abundant enzyme in the liver. CYP3A4 is responsible for metabolizing approximately 50% of clinically prescribed drugs across diverse therapeutic classes, so these interactions have raised concerns about potential adverse effects. This review delves into the scientific evidence surrounding flavonoid-mediated CYP3A4 inhibition, exploring the inhibitory potential of investigated flavonoids and future implications. Kusehnol I, chrysin, leachianone A, and sophoraflavone G showed the largest inhibitory potentials and lowest IC50 values. While the clinical significance of flavonoid-mediated CYP3A4 inhibition in dietary contexts is generally considered low due to moderate intake and complex interactions, it poses a potential concern for individuals consuming high doses of flavonoid supplements or concurrently taking medications metabolized by CYP3A4. This can lead to increased drug exposure, potentially triggering adverse reactions or reduced efficacy. Full article
(This article belongs to the Special Issue Phytochemicals: Current Status and Future Prospects)
Show Figures

Figure 1

Figure 1
<p>Hem structure formula.</p>
Full article ">Figure 2
<p>Flavonoid basic structure.</p>
Full article ">Figure 3
<p>Basic structures of flavonoid subgroups.</p>
Full article ">
16 pages, 3771 KiB  
Article
CM1, a Chrysin Derivative, Protects from Endotoxin-Induced Lethal Shock by Regulating the Excessive Activation of Inflammatory Responses
by Jae-Hyung Lee, Young-Bok Ko, Yong-Min Choi, Jinju Kim, Hwan-Doo Cho, Hyeonil Choi, Ha-Yeon Song, Jeong-Moo Han, Guang-Ho Cha, Young-Ha Lee, Jin-Man Kim, Woo-Sik Kim, Eui-Baek Byun and Jae-Min Yuk
Nutrients 2024, 16(5), 641; https://doi.org/10.3390/nu16050641 - 25 Feb 2024
Cited by 1 | Viewed by 1361
Abstract
Sepsis, a leading cause of death worldwide, is a harmful inflammatory condition that is primarily caused by an endotoxin released by Gram-negative bacteria. Effective targeted therapeutic strategies for sepsis are lacking. In this study, using an in vitro and in vivo mouse model, [...] Read more.
Sepsis, a leading cause of death worldwide, is a harmful inflammatory condition that is primarily caused by an endotoxin released by Gram-negative bacteria. Effective targeted therapeutic strategies for sepsis are lacking. In this study, using an in vitro and in vivo mouse model, we demonstrated that CM1, a derivative of the natural polyphenol chrysin, exerts an anti-inflammatory effect by inducing the expression of the ubiquitin-editing protein TNFAIP3 and the NAD-dependent deacetylase sirtuin 1 (SIRT1). Interestingly, CM1 attenuated the Toll-like receptor 4 (TLR4)-induced production of inflammatory cytokines by inhibiting the extracellular-signal-regulated kinase (ERK)/MAPK and nuclear factor kappa B (NF-κB) signalling pathways. In addition, CM1 induced the expression of TNFAIP3 and SIRT1 on TLR4-stimulated primary macrophages; however, the anti-inflammatory effect of CM1 was abolished by the siRNA-mediated silencing of TNFAPI3 or by the genetic or pharmacologic inhibition of SIRT1. Importantly, intravenous administration of CM1 resulted in decreased susceptibility to endotoxin-induced sepsis, thereby attenuating the production of pro-inflammatory cytokines and neutrophil infiltration into the lung compared to control mice. Collectively, these findings demonstrate that CM1 has therapeutic potential for diverse inflammatory diseases, including sepsis. Full article
(This article belongs to the Special Issue The Role of Nutrition and Diet in Immune Regulation)
Show Figures

Figure 1

Figure 1
<p>Chemical structures and cytotoxicity in chrysin and its derivatives. (<b>A</b>–<b>C</b>) BMDMs were incubated with different concentration of chrysin, CM1, or CM2 for 18 h. Cell viability measured by a Cell Counting Kit 8 (CCK-8) assay. Data are representative of three independent experiments and are presented as means ± standard deviation (SD). *** <span class="html-italic">p</span> &lt; 0.001, compared with control cells (two-tailed Student’s <span class="html-italic">t</span>-test). SC, solvent control (0.01% dimethylsulfoxide).</p>
Full article ">Figure 2
<p>CM1 inhibits lipopolysaccharide (LPS)-induced inflammatory responses in bone marrow-derived macrophages (BMDMs). (<b>A</b>–<b>C</b>) BMDMs were incubated with LPS (100 ng/mL) for the indicated times. (<b>A</b>,<b>B</b>) mRNA levels of <span class="html-italic">Tnfα</span> and <span class="html-italic">Il6</span> analysed by reverse-transcription polymerase chain reaction (RT-PCR) and real-time PCR (qPCR). (<b>C</b>) Protein levels of TNF-α and IL-6 in culture medium measured by enzyme-linked immunosorbent assay (ELISA). (<b>D</b>,<b>E</b>) BMDMs were stimulated with LPS and co-treated with chrysin (0.5 or 2.5 µg/mL), CM1 (0.5 or 2.5 µg/mL), or CM2 (5 or 10 µg/mL) for 18 h. (<b>D</b>) mRNA levels of <span class="html-italic">Tnfα</span> and <span class="html-italic">Il6</span> evaluated by RT-PCR (top) and qPCR (bottom). (<b>E</b>) Protein levels of TNF-α and IL-6 in culture medium investigated by ELISA. Data are representative of three independent experiments and are presented as means ± SD. *** <span class="html-italic">p</span> &lt; 0.001, compared with control cells (two-tailed Student’s <span class="html-italic">t</span>-test). U, untreated cells; SC, solvent control (0.01% DMSO).</p>
Full article ">Figure 3
<p>CM1 inhibits LPS-induced extracellular-signal-regulated kinase (ERK) phosphorylation and nuclear factor kappa B (NF-κB) activation. (<b>A</b>,<b>D</b>) BMDMs were incubated with LPS (100 ng/mL) for the indicated times. Mitogen-activated protein kinase (MAPK) and NF-κB activation were determined by immunoblotting. (<b>B</b>,<b>C</b>,<b>E</b>) BMDMs were stimulated with LPS only, LPS and chrysin (0.1, 0.5, or 2.5 µg/mL), or LPS and CM1 (0.1, 0.5, or 2.5 µg/mL) for 30 min. (<b>B</b>,<b>E</b>) Protein levels determined by immunoblotting. (<b>C</b>) Densitometric analysis of p-ERK, p-JNK, and p-p38 expression with normalisation to β–tubulin. (<b>F</b>) Cells were fixed and stained for NF-κB p65 (green); nuclei were stained with 4′,6-diamidino-2-phenylindol (DAPI) (blue). Nuclear translocation of NF-κB p65 analysed by confocal microscopy. Scale bar: 20 µm. Data are representative of three independent experiments and are presented as means ± SD. *** <span class="html-italic">p</span> &lt; 0.001, compared with control cells (two-tailed Student’s <span class="html-italic">t</span>-test). U, untreated cells; SC, solvent control (0.01% DMSO).</p>
Full article ">Figure 4
<p>CM1 attenuates LPS-induced inflammatory responses through TNFAIP3 upregulation. (<b>A</b>,<b>B</b>) BMDMs were stimulated with LPS only (100 ng/mL) or LPS with CM1 (1 µg/mL) for the indicated times. (<b>A</b>) Protein levels of tumour necrosis factor alpha-induced protein 3 (TNFAIP3) determined by immunoblotting. (<b>B</b>) Densitometric analysis of TNFAIP3 expression with normalisation to β–tubulin. (<b>C</b>,<b>D</b>) HeLa cells were transfected with control siRNA or siTNFAIP3. (<b>C</b>) Transfected cells were stimulated with LPS only or LPS and CM1 for 18 h. mRNA levels of <span class="html-italic">Tnfα</span> and <span class="html-italic">IL6</span> were determined by qPCR. RT-PCR was performed to assess transfection efficiency (inset). (<b>D</b>) Transfected cells were co-stimulated with LPS and CM1 for 30 min. Protein levels of p-ERK, p-IKKαβ, total IκBα, and TNFAIP3 were evaluated by immunoblotting. Data are representative of three independent experiments and are presented as means ± SD. *** <span class="html-italic">p</span> &lt; 0.001, compared with control cells (two-tailed Student’s <span class="html-italic">t</span>-test). U, untreated cells; SC, solvent control (0.01% DMSO); siNS, non-specific siRNA; siTNFAIP3, specific siRNA for TNFAIP3.</p>
Full article ">Figure 5
<p>CM1 reduces LPS-induced inflammatory responses by enhancing SIRT1 activity. (<b>A</b>) BMDMs were stimulated with LPS only (100 ng/mL) or LPS and CM1 (1 µg/mL) for the indicated times. NF-κB p65 acetylation and SIRT1 expression were determined by immunoblotting. (<b>B</b>–<b>E</b>) BMDMs were pre-treated with increasing concentrations of sirtinol (5, 15, or 30 µM; B and D) or EX-527 (5, 15, or 30 µM; (<b>C</b>,<b>E</b>)) for 2 h, followed by exposure to LPS only or LPS and CM1 for 30 min (<b>B</b>,<b>C</b>) or 18 h (<b>D</b>,<b>E</b>). (<b>B</b>,<b>C</b>) Immunoblotting was performed to evaluate the acetylation of NF-κB p65. (<b>D</b>,<b>E</b>) mRNA levels of <span class="html-italic">Tnfα</span> and <span class="html-italic">Il6</span> determined by qPCR. (<b>F</b>,<b>G</b>) BMDMs from m<span class="html-italic">Sirt1<sup>+/+</sup></span> and m<span class="html-italic">Sirt1<sup>−/−</sup></span> mice stimulated with LPS only or LPS and CM1 (0.1, 0.5, or 2.5 µg/mL) for 30 min (for (<b>F</b>)) or 18 h (for (<b>G</b>)). (<b>F</b>) NF-κB p65 acetylation was determined by immunoblotting. (<b>G</b>) mRNA levels of <span class="html-italic">Tnfα</span> and <span class="html-italic">Il6</span> measured by qPCR. Data are representative of three independent experiments and are presented as means ± SD. ** <span class="html-italic">p</span> &lt; 0.01, *** <span class="html-italic">p</span> &lt; 0.001, compared with control cells (for (<b>D</b>,<b>E</b>)) or cells isolated from <span class="html-italic">Sirt1<sup>+</sup></span><sup>/<span class="html-italic">+</span></sup> mouse (two-tailed Student’s <span class="html-italic">t</span>-test). U, untreated cells; SC, solvent control (0.01% DMSO).</p>
Full article ">Figure 6
<p>CM1 contributes to the protection of mice against lethal shock. (<b>A</b>–<b>E</b>) Mice (<span class="html-italic">n</span> = 10 per group) were intravenously injected with either vehicle control or CM1 (10 mg/kg) once daily for 3 days before endotoxin stimulation (intraperitoneal injection, 30 mg/kg). (<b>A</b>) Survival rates of each group were monitored for 168 h. (<b>B</b>–<b>E</b>) Mice were sacrificed at 24 h post-LPS injection (<span class="html-italic">n</span> = 5 per group). (<b>B</b>) Serum samples were collected from vehicle control-treated or CM1-treated mice. Levels of TNFα and IL-6 were determined using ELISA. (<b>C</b>) The expression of <span class="html-italic">Tnfα</span> and <span class="html-italic">Il6</span> in lung (left) and spleen (right) was analysed using real-time qPCR. (<b>D</b>) Immunohistochemical analysis of the lung tissue was performed to determine neutrophil infiltration. Scale bar: 50 µm. (<b>E</b>) The number of infiltrating neutrophils was counted from 8 random fields. The experiments were conducted in triplicate to ensure reproducibility, with the results expressed as the means ± SD. Statistical significance of mean differences was determined using either a log-rank test (<b>A</b>) or a two-tailed Student’s <span class="html-italic">t</span>-test (<b>B</b>,<b>C</b>,<b>E</b>). * <span class="html-italic">p</span> &lt; 0.05, *** <span class="html-italic">p</span> &lt; 0.001, compared with control mice stimulated with LPS. U, untreated; SC, solvent control.</p>
Full article ">Figure 7
<p>Inhibitory effect of CM1 on TLR4-induced inflammation and its molecular mechanisms. Upon the binding of LPS to the TLR4/MD2/CD14 complex, the downstream signalling cascade is activated, leading to the ubiquitination of TRAF6 and subsequently activation of TAK1 and the IKK complex, which phosphorylates IκB, leading to the release of NF-κB p65 into the nucleus. This process results in the transcription of pro-inflammatory cytokines such as TNF-α and IL-6. CM1 directly upregulates the expression of TNFAIP3/A20, which inhibits the NF-κB p65 pathway and enhances the activity of SIRT1. Increased SIRT1 activity, in turn, deacetylates p65, suppressing its ability to transcribe pro-inflammatory genes. The dual action of CM1 suggests its therapeutic potential for LPS-induced inflammation.</p>
Full article ">
14 pages, 2120 KiB  
Article
Surface Modification Strategies for Chrysin-Loaded Iron Oxide Nanoparticles to Boost Their Anti-Tumor Efficacy in Human Colon Carcinoma Cells
by Aynura Karimova, Sabina Hajizada, Habiba Shirinova, Sevinj Nuriyeva, Lala Gahramanli, Mohammed M. Yusuf, Stefano Bellucci, Christoph Reissfelder and Vugar Yagublu
J. Funct. Biomater. 2024, 15(2), 43; https://doi.org/10.3390/jfb15020043 - 13 Feb 2024
Cited by 2 | Viewed by 2816
Abstract
Enhancing nanoparticles’ anti-cancer capabilities as drug carriers requires the careful adjustment of formulation parameters, including loading efficiency, drug/carrier ratio, and synthesis method. Small adjustments to these parameters can significantly influence the drug-loading efficiency of nanoparticles. Our study explored how chitosan and polyethylene glycol [...] Read more.
Enhancing nanoparticles’ anti-cancer capabilities as drug carriers requires the careful adjustment of formulation parameters, including loading efficiency, drug/carrier ratio, and synthesis method. Small adjustments to these parameters can significantly influence the drug-loading efficiency of nanoparticles. Our study explored how chitosan and polyethylene glycol (PEG) coatings affect the structural properties, drug-loading efficiency, and anti-cancer efficacy of Fe3O4 nanoparticles (NPs). The loading efficiency of the NPs was determined using FTIR spectrometry and XRD. The quantity of chrysin incorporated into the coated NPs was examined using UV–Vis spectrometry. The effect of the NPs on cell viability and apoptosis was determined by employing the HCT 116 human colon carcinoma cell line. We showed that a two-fold increase in drug concentration did not impact the loading efficiency of Fe3O4 NPs coated with PEG. However, there was a 33 Å difference in the crystallite sizes obtained from chitosan-coated Fe3O4 NPs and drug concentrations of 1:0.5 and 1:2, resulting in decreased system stability. In conclusion, PEG coating exhibited a higher loading efficiency of Fe3O4 NPs compared to chitosan, resulting in enhanced anti-tumor effects. Furthermore, variations in the loaded amount of chrysin did not impact the crystallinity of PEG-coated NPs, emphasizing the stability and regularity of the system. Full article
(This article belongs to the Special Issue Nanoparticles and Nanocompounds for Cancer Therapy)
Show Figures

Figure 1

Figure 1
<p>X-ray diffractogram of Fe<sub>3</sub>O<sub>4</sub> NPs coated with PEG and chitosan. The crystal structure of the magnetite (Fe<sub>3</sub>O<sub>4</sub>) NPs synthesized via chemical co-precipitation and coated with PEG and chitosan was analyzed. Characteristic X-ray lines of magnetite NPs were observed in both samples. Therefore, the d-spacing values of the significant peaks match well (ICDD DB card number 01-073-9877).</p>
Full article ">Figure 2
<p>FTIR spectrum of pure chrysin. FTIR spectrum for pure chrysin allows the evaluation of its molecular state.</p>
Full article ">Figure 3
<p>FTIR spectra of the samples obtained after attaching chrysin on Fe<sub>3</sub>O<sub>4</sub>@PEG and Fe<sub>3</sub>O<sub>4</sub>@chitosan in different ratios: (<b>A</b>) Fe<sub>3</sub>O<sub>4</sub>@PEG250 with a 1:0.5 PEG-chrysin ratio; (<b>B</b>) Fe<sub>3</sub>O<sub>4</sub>@PEG500 with a 1:1 PEG-chrysin ratio; (<b>C</b>) Fe<sub>3</sub>O<sub>4</sub>@PEG1000 with a 1:2 PEG-chrysin ratio; (<b>D</b>) Fe<sub>3</sub>O<sub>4</sub>@Chitosan250 with a 1:0.5 chitosan-chrysin ratio; (<b>E</b>) Fe<sub>3</sub>O<sub>4</sub>@Chitosan500 with a 1:1 chitosan-chrysin ratio; (<b>F</b>) Fe<sub>3</sub>O<sub>4</sub>@Chitosan1000 with a 1:2 chitosan-chrysin ratio. The characteristic peaks in the spectra indicate that PEG and chitosan-coated magnetite NPs were successfully synthesized, and the chrysin drug was loaded into them.</p>
Full article ">Figure 4
<p>X-ray diffractograms of the obtained samples based on Fe<sub>3</sub>O<sub>4</sub> nanoparticles coated with different polymers and loaded with various amounts of chrysin and pure chrysin. Fe<sub>3</sub>O<sub>4</sub>@PEG250, Fe<sub>3</sub>O<sub>4</sub>@PEG500 and Fe<sub>3</sub>O<sub>4</sub>@PEG1000 accordingly correspond to the Fe<sub>3</sub>O<sub>4</sub>@PEG and chrysin ratios of 1:0.5, 1:1, and 1:2. Fe<sub>3</sub>O<sub>4</sub>@Chtiosan250, Fe<sub>3</sub>O<sub>4</sub>@Chitosan500, and Fe<sub>3</sub>O<sub>4</sub>@Chitosan1000 accordingly correspond to Fe<sub>3</sub>O<sub>4</sub>@Chitosan and chrysin ratios of 1:0.5, 1:1, and 1:2.</p>
Full article ">Figure 5
<p>Cell viability and apoptosis rate under treatment with Fe<sub>3</sub>O<sub>4</sub> NPs coated with PEG and chitosan. (<b>A</b>) Cell viability assessment using the MTT test with a 5 µg/mL treatment of Fe<sub>3</sub>O<sub>4</sub>@PEG500 with a 1:1 PEG-chrysin ratio and with 5µg/mL of Fe<sub>3</sub>O<sub>4</sub>@Chitosan500 with a 1:1 chitosan-chrysin ratio (P<sub>NP_PEG_Chry</sub> vs. <sub>NP_Chit_Chry</sub> = 0.0033). NC = negative control, where no treatment was administered. PC = positive control, where a treatment with 5 μg/mL of pure chrysin was applied. (<b>B</b>) Apoptosis rates following 48 h of treatment with Fe<sub>3</sub>O<sub>4</sub>@PEG500 and Fe<sub>3</sub>O<sub>4</sub>@Chitosan500 NPs, comparing the effects of coating with PEG versus chitosan. NC = negative control, representing the condition where no treatment was administered, serving as a control group. PC = positive control, representing the positive control group where a treatment of 5 μg/mL of chrysin was applied, serving as a reference for assessing the impact of the treatment with pure chrysin. (<b>C</b>) Representative images from the FACS experiments conducted after 48 h of treatment with Fe<sub>3</sub>O<sub>4</sub>@PEG500, inducing apoptosis in 61.7% of cells. (<b>D</b>) Representative images from the FACS experiments conducted after 48 h of treatment with Fe<sub>3</sub>O<sub>4</sub>@Chitosan500, inducing apoptosis in 36.8% of cells. Viable cells stain negative for both PI and Annexin V (Q4); early apoptotic cells stain positive for Annexin V and negative for PI (Q3); necrotic cells stain positive for PI only (Q1); and late apoptotic cells stain positive for both Annexin-V and PI (Q2). In order to make the text shorter we decided not include, because the image is self-descriptive.</p>
Full article ">
2 pages, 137 KiB  
Abstract
The Anti-Inflammatory Action of Artichoke, Fenugreek and Caigua (AFC) Original Blend in an Inflammatory Bowel Disease In Vitro Model
by Paola Palestini, Elena Lonati, Paolo Corbetta, Stefania Pagliari, Emanuela Cazzaniga, Luca Campone and Alessandra Bulbarelli
Proceedings 2023, 91(1), 291; https://doi.org/10.3390/proceedings2023091291 - 7 Feb 2024
Viewed by 679
Abstract
Background and objectives: The incidence of chronic inflammatory pathologies has incrementally increased in recent years, as in the case of inflammatory bowel disease (IBD), which is characterized by intestinal epithelial barrier disruption, increased inflammatory mediator production and excessive tissue injury. Changes in eating [...] Read more.
Background and objectives: The incidence of chronic inflammatory pathologies has incrementally increased in recent years, as in the case of inflammatory bowel disease (IBD), which is characterized by intestinal epithelial barrier disruption, increased inflammatory mediator production and excessive tissue injury. Changes in eating habits might have played a key role in this scenario. Therefore, the interest in specific diet development and in functional food formulation has been growing. Phytoextracts from several origins, from plants to waste, enriched in bioactive molecules, alone or combined, might be a resource for the obtainment of an efficient synergistic beneficial. Thus, the aim of this study consists of evaluating the protective effects of artichoke, fenugreek and caigua (AFC) phytoextract original blend. Methods: In order to mimic the intestinal barrier’s inflammatory environment, Caco-2 cells were cultured and polarized on a transwell system and then exposed to a pro-inflammatory cytokine cocktail (TNFα and IL-1β). Before being exposed to an inflammatory stimulus, cells were pre-treated with an AFC digested blend, according to the INFOGEST in vitro static digestion protocol. After digestion, the content of active substances within the blended extract (ACFB) was revealed by UHPLC–ESI–HRMS analysis. The AFC digested extract’s protective effect was evaluated by measuring the transepithelial resistance (TEER) as a marker of barrier integrity and analysing the nuclear factor kappa B (NF-κB) pathway. Results: The TEER values improved in cells which were pre-treated with the AFC blend, relative to inflamed cells, suggesting a regulation in tight junction protein expression and/or localization. The transcription factor p65NF-κB is activated by phosphorylation under cytokine exposure, with a 160% increase in its target COX-2. Moreover, a 40-fold increase in IL-8 release was observed. Interestingly, in cells pre-treated with the AFC blend, the activated p65NF-κB was halved, compared to inflamed cells only. Furthermore, a consequent reduction by about 50% for COX-2 and by 30% for IL-8 was observed. Discussion: Taken together, these results highlight the anti-inflammatory potential of the AFC blend, probably due to the presence of flavonoids such as luteolin, apigenin and chrysin. This experimental evidence suggests that an AFC blend could be a good ingredient for food functionalization if further used in nutritional strategies. Full article
(This article belongs to the Proceedings of The 14th European Nutrition Conference FENS 2023)
25 pages, 1146 KiB  
Article
Phytochemical Profiles and Antimicrobial Activity of Selected Populus spp. Bud Extracts
by Piotr Okińczyc, Jarosław Widelski, Kinga Nowak, Sylwia Radwan, Maciej Włodarczyk, Piotr Marek Kuś, Katarzyna Susniak and Izabela Korona-Głowniak
Molecules 2024, 29(2), 437; https://doi.org/10.3390/molecules29020437 - 16 Jan 2024
Cited by 3 | Viewed by 1449
Abstract
Buds of poplar trees (Populus species) are often covered with sticky, usually polyphenol-rich, exudates. Moreover, accessible data showed that some Populus bud extracts may be excellent antibacterial agents, especially against Gram-positive bacteria. Due to the fragmentary nature of the data found, we [...] Read more.
Buds of poplar trees (Populus species) are often covered with sticky, usually polyphenol-rich, exudates. Moreover, accessible data showed that some Populus bud extracts may be excellent antibacterial agents, especially against Gram-positive bacteria. Due to the fragmentary nature of the data found, we conducted a systematic screening study. The antimicrobial activity of two extract types (semi-polar—ethanolic and polar—ethanolic-water (50/50; V/V)) from 27 bud samples of different poplar taxons were compared. Antimicrobial assays were performed against Gram-positive (five strains) and Gram-negative (six strains) bacteria as well as fungi (three strains) and covered the determination of minimal inhibitory, bactericidal, and fungicidal concentrations. The composition of extracts was later investigated by ultra-high-performance liquid chromatography coupled with ultraviolet detection (UHPLC-DAD) and with electrospray-quadrupole-time-of-flight tandem mass spectrometry (UHPLC-ESI-qTOF-MS). As a result, most of the extracts exhibited good (MIC ≤ 62.5 µg/mL) or moderate (62.5 < MIC ≤ 500 µg/mL) activity against Gram-positives and Helicobacter pylori, as well as fungi. The most active were ethanolic extracts from P. trichocarpa, P. trichocarpa clone ‘Robusta’, and P. tacamahaca × P. trichocarpa. The strongest activity was observed for P. tacamahaca × P. trichocarpa. Antibacterial activity was supposedly connected with the abundant presence of flavonoids (pinobanksin, pinobanksin 3-acetate, chrysin, pinocembrin, galangin, isosakuranetin dihydrochalcone, pinocembrin dihydrochalcone, and 2′,6′-dihydroxy-4′-methoxydihydrochalcone), hydroxycinnamic acids monoesters (p-methoxycinnamic acid cinnamyl ester, caffeic acid phenethylate and different isomers of prenyl esters), and some minor components (balsacones). Full article
(This article belongs to the Special Issue Antibacterial Agents from Natural Source, 2nd Edition)
Show Figures

Graphical abstract

Graphical abstract
Full article ">Figure 1
<p>LC-MS chromatograms of <span class="html-italic">Populus</span> buds EtOH extracts representing five chemical groups (negative mode, base peak chromatograms). Figure legend: <span class="underline">Lanes</span>: orange—<b>P.N.3</b>—<span class="html-italic">P. nigra</span>, sample 3 (flavonoid type); blue—<b>P.N.1</b>—<span class="html-italic">P</span>. <span class="html-italic">nigra</span>, sample 1 (hydroxycinnamic monoesters type); green—<b>P.BA</b>—<span class="html-italic">P</span>. <span class="html-italic">balsamifera</span> (hydroxycinnamic monoesters + flavonoid type); black—<b>P.TA</b> × <b>P.TRI.2</b>—<span class="html-italic">P</span>. <span class="html-italic">tacamahaca</span> × <span class="html-italic">P</span>.<span class="html-italic">trichocarpa</span>, sample 2 (mixed type); red—<b>P.LAS</b>—<span class="html-italic">P</span>. <span class="html-italic">lasiocarpa</span> (hydrocinnamic acids glycerides type). <span class="underline">Component abbreviations:</span> <b>A-CA-<span class="html-italic">p</span>-COG</b>—2-Acetyl-1-caffeoyl-3-<span class="html-italic">p</span>-coumaroylglycerol; <b>A-d-CAG</b>—2-Acetyl-1,3-di-caffeoylglycerol; <b>A-d-<span class="html-italic">p</span>-COG</b>—2-Acetyl-1,3-di-<span class="html-italic">p</span>-coumaroylglycerol; <b>API</b>—apigenin; <b>BA-A-F.1</b>—Balsacone A/B/E/F isomer I; <b>BA-A-F.2</b>—Balsacone A/B/E/F isomer II; <b>BA-C/B</b>—Balsacone C or D; <b>CA.A</b>—Caffeic acid; <b>CA-2M-2BE</b>—Caffeic acid 2-methyl-2-butenyl ester; <b>CA-3M-2BE</b>—Caffeic acid 3-methyl-2-butenyl ester; <b>CA-3M-3BE</b>—Caffeic acid 3-methyl-3-butenyl ester; <b>CA-B/I.1</b>—Caffeic acid butyl or isobutyl ester isomer I; <b>CA-B/I.2</b>—Caffeic acid butyl or isobutyl ester isomer II; <b>CABE</b>—Caffeic acid benzyl ester; <b>CAL</b>—2′,6′-Dihydroxy-4′,4-dimethoxydihydrochalcone (calomelanone); <b>CA-<span class="html-italic">p</span>COG</b>—Caffeoyl-<span class="html-italic">p</span>-coumaroylglycerol; <b>CAPE</b>—Caffeic acid benzyl ester; <b>CHR</b>—Chrysin; <b>d-CAG</b>—di-Caffeoylglycerol; <b>DHMC</b>—2′,6′-Dihydroxy-4′-methoxydihydrochalcone; <b>DHMPPh</b>—2′,6′-Dihydroxy-4′-methoxypentanophenone; <b>DISOC</b>—Isosakuranetin dihydrochalcone; <b>d-<span class="html-italic">p</span>-COG</b>—1,3-di-<span class="html-italic">p</span>-Coumaroylglycerol; <b>GAL</b>—Galangin; <b>ISA</b>—Isosakuranetin; <b>KAE</b>—Kaempferol; <b>LUT-5-ME</b>—Luteolin 5-methyl ether; <b>M-CA.CE</b>—Metoxycinnamic acid cinnamyl ester; <b>M-CHR</b>—Methoxychrysin; <b>P.C</b>—Pinocembrin chalcone; <b><span class="html-italic">p</span>-CO.A</b>—<span class="html-italic">p</span>-Coumaric acid; <b><span class="html-italic">p</span>-CO-3M-3BE</b>—<span class="html-italic">p</span>-Coumaric acid 3-methyl-3-butenyl ester; <b><span class="html-italic">p</span>-CO-BE</b>—<span class="html-italic">p</span>-Coumaric acid benzyl ester; <b><span class="html-italic">p</span>-CO-CE</b>—<span class="html-italic">p</span>-Coumaric acid cinnamyl ester; <b><span class="html-italic">p</span>-CO-MB.I</b>—<span class="html-italic">p</span>-Coumaric acid 3-methyl-2-butenyl or 2-methyl-2-butenyl ester isomer I; <b><span class="html-italic">p</span>-CO-MB.II</b>—<span class="html-italic">p</span>-Coumaric acid 3-methyl-2-butenyl or 2-methyl-2-butenyl isomer II; <b><span class="html-italic">p</span>-CO-PE</b>—<span class="html-italic">p</span>-Coumaric acid phenethyl ester; <b>PIN</b>—Pinobanksin; <b>PIN-3-A</b>—Pinobanksin 3-acetate; <b>PIN-3-P</b>—Pinobanksin 3-propanoate; <b>PIN-3-B/I</b>—Pinobanksin 3-butanoate or -isobutanoate; <b>PIN-3-P/I.1</b>—Pinobanksin 3-pentanoate or -isopentanoate isomer I; <b>PIN-3-P/I.2</b>—Pinobanksin 3-pentanoate or -isopentanoate isomer II; <b>PIN-5-ME</b>—Pinobanksin 5-methyl ether; <b>PM.DC</b>—Pinocembrin dihydrochalcone; <b>PNM</b>—Pinocembrin; <b>PS.CH</b>—Pinostrobin chalcone.</p>
Full article ">Figure 2
<p>Antimicrobial agents of <span class="html-italic">Populus</span> buds.</p>
Full article ">
Back to TopTop