[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

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 (7,677)

Search Parameters:
Keywords = phenolic extracts

Order results
Result details
Results per page
Select all
Export citation of selected articles as:
15 pages, 2290 KiB  
Article
Antimicrobial Activity of Leaf Aqueous Extract of Schinus polygamus (Cav.) Cabrera against Pathogenic Bacteria and Spoilage Yeasts
by Andrea Acuña-Fontecilla, Julio Bruna, María Angélica Ganga and Liliana Godoy
Plants 2024, 13(16), 2248; https://doi.org/10.3390/plants13162248 - 13 Aug 2024
Abstract
The antimicrobial activity of an aqueous extract of the leaves of Schinus polygamus (cav.) Cabrera against microorganisms of food importance was evaluated. First, the leaf aqueous extract of Schinus polygamus was characterized, quantifying hydroxycinnamic acids and phenolic compounds. Then, a battery of strains [...] Read more.
The antimicrobial activity of an aqueous extract of the leaves of Schinus polygamus (cav.) Cabrera against microorganisms of food importance was evaluated. First, the leaf aqueous extract of Schinus polygamus was characterized, quantifying hydroxycinnamic acids and phenolic compounds. Then, a battery of strains was tested, including Escherichia coli ATCC 25922, Salmonella Typhimurium ATCC 14028, and Listeria monocytogenes ATCC 13932. Also, we tested wine spoilage yeasts such as Brettanomyces bruxellensis LAMAP2480, B. bruxellensis LAMAP1359, B. bruxellensis CECT1451, and Pichia guilliermondii NPCC1051. Tests were conducted using the kinetic curve of growth and cell viability counts. The results indicate that with 10% v/v of concentrated extract, it is possible to observe growth inhibition of all microorganisms studied, with statistically significant differences during the whole measurement time (70 h for bacteria and 145 h for yeast). Full article
(This article belongs to the Topic Natural Compounds in Plants, 2nd Volume)
Show Figures

Figure 1

Figure 1
<p>HPLC chromatograms of volatile phenols (<b>A</b>) and hydroxycinnamic acids (<b>B</b>) present in <span class="html-italic">S. polygamus</span> aqueous extract.</p>
Full article ">Figure 2
<p>Structures of main identified compounds in <span class="html-italic">S. polygamus</span> leaf aqueous extract.</p>
Full article ">Figure 3
<p>Effect of aqueous extract of <span class="html-italic">S. polygamus</span> on the growth of bacteria. (<b>a</b>) <span class="html-italic">E. coli</span> ATCC25922; (<b>b</b>) <span class="html-italic">S. Typhimurium</span> ATCC14028; (<b>c</b>) <span class="html-italic">L. monocytogenes</span> ATCC13932. Control = medium of culture; AE = aqueous extract (10% water plus medium); AC = aqueous control (10% water plus medium); Amp = medium plus ampicillin.</p>
Full article ">Figure 4
<p>Effect of aqueous extract of <span class="html-italic">S. polygamus</span> on the growth of spoilage yeasts. (<b>a</b>) <span class="html-italic">B. bruxellensis</span> LAMAP2480; (<b>b</b>) <span class="html-italic">B. bruxellensis</span> LAMAP 1359; (<b>c</b>) <span class="html-italic">B. bruxellensis</span> CECT1451; (<b>d</b>) <span class="html-italic">P. guillermondii</span> NPCC1051. Control = medium of culture; AE = aqueous extract; AC = aqueous control (10% water); Hyg = medium plus hygromycin.</p>
Full article ">Figure 5
<p>Effect of aqueous extract of <span class="html-italic">S. polygamus</span> on cell viability of bacteria. (<b>a</b>) <span class="html-italic">E. coli</span> ATCC 25922; (<b>b</b>) <span class="html-italic">S. Typhimurium</span> ATCC14028; (<b>c</b>) <span class="html-italic">L. monocytogenes</span> ATCC13932. Control = medium of culture; AC = aqueous control (10% water plus medium); AE = aqueous extract (10% extract plus medium). Different letters above each column represent a significant difference (* <span class="html-italic">p</span> ≤ 0.05).</p>
Full article ">Figure 6
<p>Effect of aqueous extract of <span class="html-italic">S. polygamus</span> on cell viability of spoilage yeasts. (<b>a</b>) <span class="html-italic">B. bruxellensis</span> LAMAP2480; (<b>b</b>) <span class="html-italic">B. bruxellensis</span> LAMAP 1359; (<b>c</b>) <span class="html-italic">B. bruxellensis</span> CECT1451; (<b>d</b>) <span class="html-italic">P. guilliermondii</span> NPCC1051.Control = medium de culture; AC = aqueous control (10% water plus medium); AE = aqueous extract (10% extract plus medium). Different letter above each column represents a significant difference (* <span class="html-italic">p</span> ≤ 0.05).</p>
Full article ">
13 pages, 1159 KiB  
Article
Neuroprotective Effects of Phenolic Constituents from Drynariae Rhizoma
by Jin Sung Ahn, Chung Hyeon Lee, Xiang-Qian Liu, Kwang Woo Hwang, Mi Hyune Oh, So-Young Park and Wan Kyunn Whang
Pharmaceuticals 2024, 17(8), 1061; https://doi.org/10.3390/ph17081061 - 13 Aug 2024
Abstract
This study aimed to provide scientific data on the anti-Alzheimer’s disease (AD) effects of phenolic compounds from Drynariae Rhizoma (DR) extract using a multi-component approach. Screening of DR extracts, fractions, and the ten phenolic compounds isolated from DR against the key AD-related enzymes [...] Read more.
This study aimed to provide scientific data on the anti-Alzheimer’s disease (AD) effects of phenolic compounds from Drynariae Rhizoma (DR) extract using a multi-component approach. Screening of DR extracts, fractions, and the ten phenolic compounds isolated from DR against the key AD-related enzymes acetylcholinesterase (AChE), butyrylcholinesterase (BChE), β-site amyloid precursor protein cleaving enzyme 1 (BACE1), and monoamine oxidase-B (MAO-B) confirmed their significant inhibitory activities. The DR extract was confirmed to have BACE1-inhibitory activity, and the ethyl acetate and butanol fractions were found to inhibit all AD-related enzymes, including BACE1, AChE, BChE, and MAO-B. Among the isolated phenolic compounds, compounds (2) caffeic acid 4-O-β-D-glucopyranoside, (6) kaempferol 3-O-rhamnoside 7-O-glucoside, (7) kaempferol 3-o-b-d-glucopyranoside-7-o-a-L-arabinofuranoside, (8) neoeriocitrin, (9) naringin, and (10) hesperidin significantly suppressed AD-related enzymes. Notably, compounds 2 and 8 reduced soluble Amyloid Precursor Protein β (sAPPβ) and β-secretase expression by over 45% at a concentration of 1.0 μM. In the thioflavin T assay, compounds 6 and 7 decreased Aβ aggregation by approximately 40% and 80%, respectively, and degraded preformed Aβ aggregates. This study provides robust evidence regarding the potential of DR as a natural therapeutic agent for AD, highlighting specific compounds that may contribute to its efficacy. Full article
(This article belongs to the Section Natural Products)
Show Figures

Graphical abstract

Graphical abstract
Full article ">Figure 1
<p>Chemical structures of ten compounds isolated from <span class="html-italic">Drynaria fortunei</span>.</p>
Full article ">Figure 2
<p>The effects of six phenolic compounds (compounds <b>2</b>, <b>6</b>, <b>7</b>, <b>8</b>, <b>9</b>, and <b>10</b>) on sAPPβ and β-secretase production. (<b>A</b>) sAPPβ and β-secretase levels in APP-CHO cells treated with different concentrations (1.0 and 0.5 μM) of six compounds were determined by Western blot analysis. (<b>B</b>,<b>C</b>) Graphs show sAPPβ (<b>B</b>) and β-secretase (<b>C</b>) levels compared to DMSO-treated controls. Values are expressed as a percentage of DMSO-treated control. All data are presented as the mean ± standard deviation of three different experiments. * <span class="html-italic">p</span> &lt; 0.05: significant difference from the DMSO-treated control group.</p>
Full article ">Figure 3
<p>Inhibition of Aβ aggregation and degradation of preformed Aβ aggregates by six phenolic compounds (compounds <b>2</b>, <b>6</b>, <b>7</b>, <b>8</b>, <b>9</b>, and <b>10</b>). (<b>A</b>) Aβ was incubated with six phenolic compounds at concentrations of 50 μM and 10 μM. After 24 h, Aβ aggregation was assessed using the Th T assay. (<b>B</b>) Aβ pre-aggregated for 24 h was exposed to six phenolic compounds at concentrations of 1.0 and 0.5 μM. After another 24 h, Aβ disaggregation was evaluated using the Th T assay. All data are presented as the mean ± standard deviation of three independent experiments. * <span class="html-italic">p</span> &lt; 0.05: significant difference from the Aβ-only group.</p>
Full article ">
32 pages, 7447 KiB  
Article
Antioxidant Activity and UHPLC-MS/MS Characterization of Polyphenol and Nicotine Content in Nicotiana Glauca Leaf Extracts: A Comparative Study of Conventional and Deep Eutectic Solvent Extraction Methods
by Reem Issa, Faisal Al-Akayleh, Lilian Alnsour, Tabarak R. Al-Sammarraie, Khaled W. Omari and Shady H. Awwad
Plants 2024, 13(16), 2240; https://doi.org/10.3390/plants13162240 - 13 Aug 2024
Viewed by 45
Abstract
The leaves of Nicotiana glauca (N. glauca; Solanaceae) plant are a known, major human health concern. This study investigated the antioxidant activity and polyphenols composition of aerial parts of N. glauca collected from its wild habitat in Jordan, using Methanol-Conventional (MC) [...] Read more.
The leaves of Nicotiana glauca (N. glauca; Solanaceae) plant are a known, major human health concern. This study investigated the antioxidant activity and polyphenols composition of aerial parts of N. glauca collected from its wild habitat in Jordan, using Methanol-Conventional (MC) and deep eutectic solvents (DES) extraction methods in addition to nicotine content determination using UHPLC. Our results showed that the MC extract contains fewer total phenols and flavonoid content than the 90% DES extract, (0.1194 ± 0.009 and 0.311 ± 0.020 mg/mL equivalent to gallic acid) and (0.01084 ± 0.005 and 0.928 ± 0.09 mg/mL equivalent to rutin), respectively. Moreover, this study showed that the prepared MC extract contain 635.07 ppm nicotine, while the 90% DES extract contain 1194.91 ppm nicotine. Extracts prepared using the MC and the DES methods exhibited weak antioxidant activities; the highest was a 33% inhibition rate (equivalent to ascorbic acid), obtained by the 90% DES extract,. The performed UHPLC-MS/MS analysis in this study also revealed the presence of variations in the detected compounds between the two extraction methods. Furthermore, this study found that environmentally friendly DES extraction of N. glauca produced higher phenol and flavonoid content than the MC method; this highlights the superior efficiency and environmental benefits of sustainable chemistry methods for extracting valuable phytoconstituents. Full article
(This article belongs to the Section Phytochemistry)
Show Figures

Figure 1

Figure 1
<p>Total ion chromatograms for all compounds detected in <span class="html-italic">N. gluca</span> MC extract.</p>
Full article ">Figure 2
<p>The UHPLC-chromatograms show peaks and retention time of each compound detected in <span class="html-italic">N. glauca</span> prepared by MC extraction.</p>
Full article ">Figure 2 Cont.
<p>The UHPLC-chromatograms show peaks and retention time of each compound detected in <span class="html-italic">N. glauca</span> prepared by MC extraction.</p>
Full article ">Figure 3
<p>Total ion chromatograms for all compounds detected in <span class="html-italic">N. glauca</span> DES extract.</p>
Full article ">Figure 4
<p>The UHPLC-chromatograms show peaks and retention time of each compound detected in <span class="html-italic">N. glauca</span> DES extract.</p>
Full article ">Figure 4 Cont.
<p>The UHPLC-chromatograms show peaks and retention time of each compound detected in <span class="html-italic">N. glauca</span> DES extract.</p>
Full article ">Figure 5
<p>The UHPLC-chromatogram and mass spectra show peaks and retention time of nicotine detected in <span class="html-italic">N. glauca</span> extract.</p>
Full article ">Figure A1
Full article ">Figure A1 Cont.
Full article ">Figure A1 Cont.
Full article ">Figure A1 Cont.
Full article ">Figure A1 Cont.
Full article ">Figure A1 Cont.
Full article ">Figure A1 Cont.
Full article ">Figure A1 Cont.
Full article ">Figure A2
Full article ">Figure A2 Cont.
Full article ">Figure A2 Cont.
Full article ">Figure A2 Cont.
Full article ">Figure A2 Cont.
Full article ">Figure A2 Cont.
Full article ">Figure A2 Cont.
Full article ">
1827 KiB  
Proceeding Paper
Optimization of Extracted Phenolic Compounds from Oregano through Accelerated Solvent Extraction Using Response Surface Methodology
by Christina Panagiotidou, Elisavet Bouloumpasi, Maria Irakli and Paschalina Chatzopoulou
Eng. Proc. 2024, 67(1), 10; https://doi.org/10.3390/engproc2024067010 - 12 Aug 2024
Viewed by 12
Abstract
The current research focuses on the optimization of accelerated solvent extraction, a potential alternative to conventional solvent extraction, for the extraction of phenolics from Greek oregano. The response surface methodology based on central composite design was used to optimize methanol concentration (X1 [...] Read more.
The current research focuses on the optimization of accelerated solvent extraction, a potential alternative to conventional solvent extraction, for the extraction of phenolics from Greek oregano. The response surface methodology based on central composite design was used to optimize methanol concentration (X1, 40–80%), extraction time (X2, 3–9 min, 3 cycles), and extraction temperature (X3, 60–140 °C). Under the optimal extraction conditions (methanol concentration of 74%, extraction time of 9 min, extraction temperature of 140 °C), the experimental values for extraction yield (%), total phenolic (TPC) and flavonoid contents (TFC), and antioxidant capacity matched those predicted, therefore validating the model adequately. The oregano extracts were rich in phenolic compounds, with rosmarinic acid and salvianolic acid B being the most prevalent phenolic components. The results obtained revealed that ASE can be utilized for the extraction of bioactive compounds, and there are advantages to preserving phenolic content if optimization is applied. Full article
(This article belongs to the Proceedings of The 3rd International Electronic Conference on Processes)
Show Figures

Figure 1

Figure 1
<p>Response surface plots of ASE conditions for extraction yield (<b>a</b>), TPC (<b>b</b>), TFC (<b>c</b>), ABTS (<b>d</b>), and DPPH (<b>e</b>,<b>f</b>) contents of oregano extract, in the function of methanol concentration (% methanol), time of extraction (time), and extraction temperature (temper.). The values of the missing factor were kept at the center point.</p>
Full article ">
13 pages, 3469 KiB  
Article
Comparative Analysis of Phytochemical Composition and Antioxidant Properties of Smilax china Rhizome from Different Regions
by Chang-Dae Lee, Neil Patrick Uy, Yunji Lee, Dong-Ha Lee and Sanghyun Lee
Horticulturae 2024, 10(8), 850; https://doi.org/10.3390/horticulturae10080850 (registering DOI) - 12 Aug 2024
Viewed by 158
Abstract
This study aimed to investigate variations in the phytochemical compound contents and antioxidant potential of the ethanol rhizome extracts of Smilax china L., belonging to the Liliaceae family, from different parts of Korea, namely Uiwang (Mt. Gamnamugol), Gyeonggi Province (SC1); Geochang, Gyeongnam Province [...] Read more.
This study aimed to investigate variations in the phytochemical compound contents and antioxidant potential of the ethanol rhizome extracts of Smilax china L., belonging to the Liliaceae family, from different parts of Korea, namely Uiwang (Mt. Gamnamugol), Gyeonggi Province (SC1); Geochang, Gyeongnam Province (SC2); Yeongwol, Gangwon Province (SC3); and Chungju, Chungbuk Province (SC4). The phenolic and flavonoid contents, radical scavenging activity, and proximate composition of the ethanol extracts from the rhizome samples were determined. The total polyphenol content (TPC) of the extracts ranged between 13.6 and 67.5 mg tannic acid equivalent/g. TPC analysis showed that TPC was higher in SC2 than in SC3, SC4, or SC1. Among the rhizome samples, the SC3 rhizomes had the highest total flavonoid content (TFC) (5.2 mg quercetin equivalents/g). Additionally, SC2 showed the highest radical scavenging activity against DPPH and ABTS+ radicals. Chemical characterization using UPLC/UV revealed that the extracts contained compounds such as apiin, kaempferol-3-rutinoside, and chlorogenic acid. Specifically, in SC2, chlorogenic acid was the dominant compound, which supported the levels observed in the UPLC/UV and HPLC/ELSD investigations. Dioscin, another phytochemical, was detected in SC2, SC3, and SC4, indicating the diversity of compounds among the rhizome extracts. Variations in the phytochemical content and antioxidant activity were observed in the extracts from the different regions, underlining the role of geographical variation in the functional characteristics of S. china. The observed differences could have important implications for the medicinal use of S. china extracts in applications such as anti-inflammatory treatments, diabetes management, and potential anticancer therapies. This study underscores the critical need to consider geographical origin when sourcing and utilizing S. china for therapeutic purposes, as it may significantly impact its bioactive profile and efficacy. Full article
Show Figures

Graphical abstract

Graphical abstract
Full article ">Figure 1
<p>Geographic locations of <span class="html-italic">S. china</span> rhizome sample collection sites.</p>
Full article ">Figure 2
<p>Chemical structures of chlorogenic acid (<b>1</b>), apiin (<b>2</b>), afzelin (<b>3</b>), naringenin (<b>4</b>), and dioscin (<b>5</b>).</p>
Full article ">Figure 2 Cont.
<p>Chemical structures of chlorogenic acid (<b>1</b>), apiin (<b>2</b>), afzelin (<b>3</b>), naringenin (<b>4</b>), and dioscin (<b>5</b>).</p>
Full article ">Figure 3
<p>Results of the (<b>a</b>) TPC and (<b>b</b>) TFC analyses (<b>c</b>) DPPH, and (<b>d</b>) ABTS<sup>+</sup> assays. Each bar presents the mean ± SD. <sup>a–c</sup> indicates significant differences at <span class="html-italic">p</span> &lt; 0.0001. Ascorbic acid (AA) was used as the positive control.</p>
Full article ">Figure 4
<p>UPLC/UV chromatograms of four <span class="html-italic">S. china</span> rhizome samples: (<b>a</b>) SC1; (<b>b</b>) SC2; (<b>c</b>) SC3; (<b>d</b>) SC4. (<b>1</b>: chlorogenic acid, <b>2</b>: apiin, <b>3</b>: afzelin, <b>4</b>: naringenin).</p>
Full article ">Figure 4 Cont.
<p>UPLC/UV chromatograms of four <span class="html-italic">S. china</span> rhizome samples: (<b>a</b>) SC1; (<b>b</b>) SC2; (<b>c</b>) SC3; (<b>d</b>) SC4. (<b>1</b>: chlorogenic acid, <b>2</b>: apiin, <b>3</b>: afzelin, <b>4</b>: naringenin).</p>
Full article ">Figure 5
<p>Pearson’s correlation coefficient network (<span class="html-italic">r</span> ≥│1.00│) among the response variables in phytochemical concentrations and antioxidant activities measures in <span class="html-italic">S. china</span> rhizomes. The red and blue lines indicate the positive and negative correlation coefficients between variables, respectively.</p>
Full article ">Figure 6
<p>HPLC/ELSD chromatograms of four <span class="html-italic">S. china</span> rhizome samples: (<b>a</b>) SC1; (<b>b</b>) SC2; (<b>c</b>) SC3; (<b>d</b>) SC4. (<b>5</b>: dioscin).</p>
Full article ">
14 pages, 902 KiB  
Article
Solid-State Fermentation for Phenolic Compounds Recovery from Mexican Oregano (Lippia graveolens Kunth) Residual Leaves Applying a Lactic Acid Bacteria (Leuconostoc mesenteroides)
by Israel Bautista-Hernández, Ricardo Gómez-García, Cristóbal N. Aguilar, Guillermo C. G. Martínez-Ávila, Cristian Torres-León and Mónica L. Chávez-González
Agriculture 2024, 14(8), 1342; https://doi.org/10.3390/agriculture14081342 - 11 Aug 2024
Viewed by 527
Abstract
The Mexican oregano by-products are a source of bioactive molecules (polyphenols) that could be extracted using solid-state fermentation (SSF). This study fermented the by-products via SSF (120 h) with a lactic acid bacteria (LAB) Leuconostoc mesenteroides. Sequentially, a bioactive and chemical determination [...] Read more.
The Mexican oregano by-products are a source of bioactive molecules (polyphenols) that could be extracted using solid-state fermentation (SSF). This study fermented the by-products via SSF (120 h) with a lactic acid bacteria (LAB) Leuconostoc mesenteroides. Sequentially, a bioactive and chemical determination was made according to the phenolic content, antioxidant activity (DPPH/FRAP), bioactive properties (α-amylase inhibition and antimicrobial activity against Escherichia coli), and chemical composition (HPLC-MS). The results showed that the total phenolics and flavonoid content, as well as the antioxidant activity, increased (0.60, 2.55, and 3.01 times, respectively) during the SSF process compared with unfermented material. Also, the extracts showed antimicrobial activity against E. coli and α-amylase inhibition. These inhibitory results could be attributed to bioactive compounds identified via HPLC, such as gardenin B, trachelogenin, ferulic acid, and resveratrol 3-O-glucoside. Therefore, the application of L. mesenteroides under SSF on oregano by-products comprises an eco-friendly strategy for their valorization as raw materials for the recovery of phenolic compounds that could be natural alternatives against synthetic antioxidant and antimicrobial agents, promoting a more circular and sustainable supply system within the oregano industry. Full article
Show Figures

Figure 1

Figure 1
<p>General process diagram of bioactive activity and chemical evaluation of <span class="html-italic">Lippia graveolens</span> by-product valorization through SSF process.</p>
Full article ">Figure 2
<p>Polyphenolic compounds concentration in fermentative extracts obtained from SSF process using <span class="html-italic">L. mesenteroides</span>. (<b>A</b>) Total polyphenolic content (TPC) and (<b>B</b>) total flavonoid content (TFC). Different letters show significant differences (α = 0.05).</p>
Full article ">Figure 3
<p>Antioxidant activity of fermentative extracts via the SSF process using <span class="html-italic">L. mesenteroides</span>; (<b>A</b>) FRAP assay and (<b>B</b>) DPPH<sup>●</sup> assay. The different letters show significant differences (α = 0.05).</p>
Full article ">
15 pages, 16273 KiB  
Article
Xanthoxylin Attenuates Lipopolysaccharide-Induced Lung Injury through Modulation of Akt/HIF-1α/NF-κB and Nrf2 Pathways
by Fu-Chao Liu, Yuan-Han Yang, Chia-Chih Liao and Hung-Chen Lee
Int. J. Mol. Sci. 2024, 25(16), 8742; https://doi.org/10.3390/ijms25168742 (registering DOI) - 10 Aug 2024
Viewed by 287
Abstract
Xanthoxylin, a bioactive phenolic compound extracted from the traditional herbal medicine Penthorum Chinense Pursh, is renowned for its anti-inflammatory effects. While previous studies have highlighted the anti-inflammatory and antioxidant properties of Xanthoxylin, its precise mechanisms, particularly concerning immune response and organ protection, [...] Read more.
Xanthoxylin, a bioactive phenolic compound extracted from the traditional herbal medicine Penthorum Chinense Pursh, is renowned for its anti-inflammatory effects. While previous studies have highlighted the anti-inflammatory and antioxidant properties of Xanthoxylin, its precise mechanisms, particularly concerning immune response and organ protection, remain underexplored. This study aimed to elucidate the effects of Xanthoxylin on inflammation and associated signaling pathways in a mouse model of lipopolysaccharide (LPS)-induced acute lung injury (ALI). ALI was induced via intratracheal administration of LPS, followed by intraperitoneal injections of Xanthoxylin at doses of 1, 2.5, 5, and 10 mg/kg, administered 30 min post-LPS exposure. Lung tissues were harvested for analysis 6 h after LPS challenge. Xanthoxylin treatment significantly mitigated lung tissue damage, pathological alterations, immune cell infiltration, and the production of pro-inflammatory cytokines, including tumor necrosis factor-alpha (TNF-α) and interleukin-6 (IL-6). Additionally, Xanthoxylin modulated the expression of key proteins in the protein kinase B (Akt)/hypoxia-inducible factor 1-alpha (HIF-1α)/nuclear factor-kappa B (NF-κB) signaling pathway, as well as nuclear factor erythroid 2-related factor 2 (Nrf2) and oxidative markers such as superoxide dismutase (SOD) and malondialdehyde (MDA) in the context of LPS-induced injury. This study demonstrates that Xanthoxylin exerts protective and anti-inflammatory effects by down-regulating and inhibiting the Akt/HIF-1α/NF-κB pathways, suggesting its potential as a therapeutic target for the prevention and treatment of ALI or acute respiratory distress syndrome (ARDS). Full article
(This article belongs to the Special Issue New Insights in Natural Bioactive Compounds 3.0)
Show Figures

Figure 1

Figure 1
<p>The effect of Xanthoxylin on the viability and pro-inflammatory cytokine levels of RAW 264.7 cells in the presence and absence of LPS. (<b>A</b>) RAW 264.7 cells were treated with various concentrations of DMSO (0.01, 0.05, and 0.5 μL) or Xanthoxylin (0.1, 1, 5, 10, 20, and 50 μM) for 24 h. Results are expressed as a percentage relative to the control group and shown as mean ± SD (<span class="html-italic">n</span> = 6 per group). (<b>B</b>) RAW 264.7 cells were treated with Xanthoxylin (0, 5, and 10 μM) followed by LPS exposure for 48 h to assess cell viability. Results are expressed as a percentage relative to the control group and shown as mean ± SD (<span class="html-italic">n</span> = 12 per group). (<b>C</b>) Pro-inflammatory cytokines IL-1β, IL-6, and TNF-α levels in the supernatants of RAW 264.7 cells were measured after treatment with Xanthoxylin, followed by LPS exposure. ### <span class="html-italic">p</span> &lt; 0.005 vs. control group; * <span class="html-italic">p</span> &lt; 0.05, ** <span class="html-italic">p</span> &lt; 0.01, *** <span class="html-italic">p</span> &lt; 0.005 vs. LPS group.</p>
Full article ">Figure 2
<p>General lung appearance after LPS-induced injury and Xanthoxylin treatment. Mice received an intratracheal LPS challenge followed by intraperitoneal administration of Xanthoxylin (XT, 1, 2.5, 5, and 10 mg/kg) or saline. Lungs were collected 6 h post-LPS challenge for analysis.</p>
Full article ">Figure 3
<p>Histological examination of lung tissues stained with H&amp;E after LPS challenge and Xanthoxylin treatment. Mice received LPS intratracheally and were then treated with Xanthoxylin (XT, 1, 2.5, 5, and 10 mg/kg) or saline intraperitoneally. Lungs were harvested 6 h post-LPS challenge for H&amp;E staining. Representative images show ALI and histological changes (100× magnification, scar bar = 100 μm). Quantification of histologic lung injury was analyzed according to American Thoracic Society (ATS) scoring system (<span class="html-italic">n</span> = 6 per group). # <span class="html-italic">p</span> &lt; 0.05, ## <span class="html-italic">p</span> &lt; 0.01, ### <span class="html-italic">p</span> &lt; 0.005 vs. control group; * <span class="html-italic">p</span> &lt; 0.05, ** <span class="html-italic">p</span> &lt; 0.001 vs. LPS group.</p>
Full article ">Figure 4
<p>Neutrophil infiltration in lungs following LPS-induced injury and Xanthoxylin treatment. Mice were challenged with LPS intratracheally and treated with Xanthoxylin (XT, 1, 2.5, 5, and 10 mg/kg) or saline intraperitoneally. Lungs were collected 6 h post-LPS challenge and immunostained with Ly6G antibody (200× magnification, scar bar = 50 μm). Quantification of positive cells was analyzed under high power field (HPF). Data are mean ± SD (<span class="html-italic">n</span> = 6 per group). ### <span class="html-italic">p</span> &lt; 0.005 vs. control group; * <span class="html-italic">p</span> &lt; 0.05, *** <span class="html-italic">p</span> &lt; 0.005 vs. LPS group.</p>
Full article ">Figure 5
<p>Macrophage infiltration in lungs following LPS-induced injury and Xanthoxylin treatment. Mice received LPS intratracheally and were treated with Xanthoxylin (XT, 1, 2.5, 5, and 10 mg/kg) or saline intraperitoneally. Lungs were harvested 6 h post-LPS challenge and immunostained with Mac-2 antibody (200× magnification, scar bar = 50 μm). Quantification of positive cells was analyzed under HPF. Data are mean ± SD (<span class="html-italic">n</span> = 6 per group). ### <span class="html-italic">p</span> &lt; 0.005 vs. control group; *** <span class="html-italic">p</span> &lt; 0.005 vs. LPS group.</p>
Full article ">Figure 6
<p>Levels of (<b>A</b>) IL-6 and (<b>B</b>) TNF-α in lungs after LPS challenge and Xanthoxylin treatment. Mice were given intratracheal LPS challenge followed by Xanthoxylin (XT, 1, 2.5, 5, and 10 mg/kg) or saline intraperitoneally. Lungs were harvested 6 h post-LPS challenge for ELISA. Data are mean ± SD (<span class="html-italic">n</span> = 6 per group). # <span class="html-italic">p</span> &lt; 0.05, ## <span class="html-italic">p</span> &lt; 0.01, ### <span class="html-italic">p</span> &lt; 0.005 vs. control group; ** <span class="html-italic">p</span> &lt; 0.01, *** <span class="html-italic">p</span> &lt; 0.005 vs. LPS group.</p>
Full article ">Figure 7
<p>Levels of (<b>A</b>) MDA and (<b>B</b>) SOD in lungs after LPS challenge and Xanthoxylin treatment. Mice received intratracheal LPS challenge and were treated with Xanthoxylin (XT, 1, 2.5, 5, and 10 mg/kg) or saline intraperitoneally. Lungs were collected 6 h post-LPS challenge for oxidative stress assays. Data are mean ± SD (<span class="html-italic">n</span> = 6 per group). # <span class="html-italic">p</span> &lt; 0.05, ### <span class="html-italic">p</span> &lt; 0.005 vs. control group; * <span class="html-italic">p</span> &lt; 0.05, ** <span class="html-italic">p</span> &lt; 0.01, *** <span class="html-italic">p</span> &lt; 0.005 vs. LPS group.</p>
Full article ">Figure 8
<p>Effects of Xanthoxylin on expression of (<b>A</b>) Akt, (<b>B</b>) NF-κB, (<b>C</b>) HIF-1α, and (<b>D</b>) Nrf2 in lungs after LPS challenge. Mice were administered Xanthoxylin (XT, 2.5, 5, and 10 mg/kg) or saline intraperitoneally 30 min post-LPS challenge. Lungs were harvested 6 h later for Western blot analysis. Data are mean ± SD (<span class="html-italic">n</span> = 6 per group). ## <span class="html-italic">p</span> &lt; 0.01, ### <span class="html-italic">p</span> &lt; 0.005 vs. control group; * <span class="html-italic">p</span> &lt; 0.05, ** <span class="html-italic">p</span> &lt; 0.01, *** <span class="html-italic">p</span> &lt; 0.005 vs. LPS group.</p>
Full article ">Figure 9
<p>Nrf2 expression in lungs after LPS-induced injury and Xanthoxylin treatment. Mice received intratracheal LPS challenge followed by Xanthoxylin (XT, 1, 2.5, 5, and 10 mg/kg) or saline intraperitoneally. Lungs were collected 6 h post-LPS challenge and immunostained with Nrf2 antibody (400× magnification, scar bar = 25 μm). Quantification of positive cells was analyzed under HPF. Data are mean ± SD (<span class="html-italic">n</span> = 6 per group). ### <span class="html-italic">p</span> &lt; 0.005 vs. control group; *** <span class="html-italic">p</span> &lt; 0.005 vs. LPS group.</p>
Full article ">Figure 10
<p>A schematic representation of the involvement of Akt/HIF-1α/NF-κB and Nrf2 signaling pathways in the protective effects of Xanthoxylin against LPS-induced lung injury. Xanthoxylin modulates Akt expression, suppresses HIF-1α/NF-κB signaling, and activates Nrf2, thereby reducing cell damage and oxidative stress. It also inhibits TNF-α and IL-6 release from macrophages.</p>
Full article ">
19 pages, 777 KiB  
Review
Regulation of Intestinal Inflammation by Walnut-Derived Bioactive Compounds
by Kexin Dai, Neel Agarwal, Alexander Rodriguez-Palacios and Abigail Raffner Basson
Nutrients 2024, 16(16), 2643; https://doi.org/10.3390/nu16162643 - 10 Aug 2024
Viewed by 761
Abstract
Walnuts (Juglans regia L.) have shown promising effects in terms of ameliorating inflammatory bowel disease (IBD), attributed to their abundant bioactive compounds. This review comprehensively illustrates the key mechanisms underlying the therapeutic potential of walnuts in IBD management, including the modulation of [...] Read more.
Walnuts (Juglans regia L.) have shown promising effects in terms of ameliorating inflammatory bowel disease (IBD), attributed to their abundant bioactive compounds. This review comprehensively illustrates the key mechanisms underlying the therapeutic potential of walnuts in IBD management, including the modulation of intestinal mucosa permeability, the regulation of inflammatory pathways (such as NF-kB, COX/COX2, MAPCK/MAPK, and iNOS/NOS), relieving oxidative stress, and the modulation of gut microbiota. Furthermore, we highlight walnut-derived anti-inflammatory compounds, such as polyunsaturated fatty acids (PUFA; e.g., ω-3 PUFA), tocopherols, phytosterols, sphingolipids, phospholipids, phenolic compounds, flavonoids, and tannins. We also discuss unique anti-inflammatory compounds such as peptides and polysaccharides, including their extraction and preparation methods. Our review provides a theoretical foundation for dietary walnut supplementation in IBD management and provides guidance for academia and industry. In future, research should focus on the targeted isolation and purification of walnut-derived anti-inflammatory compounds or optimizing extraction methods to enhance their yields, thereby helping the food industry to develop dietary supplements or walnut-derived functional foods tailored for IBD patients. Full article
Show Figures

Figure 1

Figure 1
<p>The mechanism of walnuts regulating IBD. (1) An illustration of the intestinal mucosal barrier and the effect of walnuts on permeability. (2) A depiction of the antioxidant effects of walnuts on ROS. (3) A pathway map showing NF-κB, COX/COX-2 and MAPK signaling modulation by walnuts. (4) Diagram showing changes in gut microbiota composition due to walnut consumption.</p>
Full article ">
15 pages, 2283 KiB  
Article
Immunomodulatory Effects of Anadenanthera colubrina Bark Extract in Experimental Autoimmune Encephalomyelitis
by Karla A. Ramos, Igor G. M. Soares, Larissa M. A. Oliveira, Mariana A. Braga, Pietra P. C. Soares, Gracimerio J. Guarneire, Elaine C. Scherrer, Fernando S. Silva, Nerilson M. Lima, Felipe A. La Porta, Teresinha de Jesus A. S. Andrade, Gagan Preet, Sandra B. R. Castro, Caio César S. Alves and Alessandra P. Carli
Curr. Issues Mol. Biol. 2024, 46(8), 8726-8740; https://doi.org/10.3390/cimb46080515 (registering DOI) - 10 Aug 2024
Viewed by 388
Abstract
This study aimed to evaluate the efficacy of the ethanolic extract of Anadenanthera colubrina in modulating the immune response in the Experimental Autoimmune Encephalomyelitis (EAE) model. The ethanolic extract of the dried bark was analyzed by ESI (+) Orbitrap-MS to obtain a metabolite [...] Read more.
This study aimed to evaluate the efficacy of the ethanolic extract of Anadenanthera colubrina in modulating the immune response in the Experimental Autoimmune Encephalomyelitis (EAE) model. The ethanolic extract of the dried bark was analyzed by ESI (+) Orbitrap-MS to obtain a metabolite profile, demonstrating a wide variety of polyphenols, such as flavonoids and phenolic acids. Various parameters were evaluated, such as clinical signs, cytokines, cellular profile, and histopathology in the central nervous system (CNS). The ethanolic extract of A. colubrina demonstrated significant positive effects attenuating the clinical signs and pathological processes associated with EAE. The beneficial effects of the extract treatment were evidenced by reduced levels of pro-inflammatory cytokines, such as IL1β, IL-6, IL-12, TNF, IFN-γ, and a notable decrease in several cell profiles, including CD8+, CD4+, CD4+IFN-γ, CD4+IL-17+, CD11c+MHC-II+, CD11+CD80+, and CD11+CD86+ in the CNS. In addition, histological analysis revealed fewer inflammatory infiltrates and demyelination sites in the spinal cord of mice treated with the extract compared to the control model group. These results showed, for the first time, that the ethanolic extract of A. colubrina exerts a modulatory effect on inflammatory processes, improving clinical signs in EAE, in the acute phase of the disease, which could be further explored as a possible therapeutic alternative. Full article
Show Figures

Figure 1

Figure 1
<p>Clinical signs of EAE. Animals (n = 8/group) were monitored daily for clinical signs of EAE after immunization with 100 µg of MOG<sub>35–55</sub> peptide. Mice were treated with 200 mg/kg of the ethanolic extract of <span class="html-italic">A. colubrina</span> barks (EtAc) for six days. The dotted line indicates the start of treatment. Each dot represents the arithmetic mean ± SEM. * indicates <span class="html-italic">p</span> &lt; 0.05 compared to induced and PBS-treated animals (EAE), analyzed by two-way ANOVA with Dunnett’s correction. CN = negative control (not induced and treated with PBS).</p>
Full article ">Figure 2
<p>Histopathology of the spinal cord of mice. Histopathology of the spinal cord of mice immunized or not immunized with 100 µg of MOG<sub>35–55</sub> (n = 8/group). Figures are representative of the histological analysis of each experimental group: CN= non-immunized and PBS-treated group (<b>A</b>,<b>B</b>), EAE = immunized and PBS-treated group (<b>C</b>,<b>D</b>), EtAc = immunized and treated with 200 mg/kg ethanolic extract of <span class="html-italic">A. colubrina</span> barks for six days (<b>E</b>,<b>F</b>). The examined groups representative sections (5 µm) were stained with hematoxylin and eosin (H&amp;E) to analyze the cell infiltrate. Original magnification: 10× objective (<b>A</b>,<b>C</b>,<b>E</b>), 40× (<b>B</b>,<b>D</b>,<b>F</b>). Scale bars = 100 µm (10×) and 50 µm (40×). Arrows indicate cellular infiltrates.</p>
Full article ">Figure 3
<p>Demyelination of the spinal cord of mice. Histopathology of spinal cords of mice immunized or not immunized with 100 µg of MOG35–55 (n = 8/group). Figures are representative of the histological analysis of each experimental group: CN= non-immunized and PBS-treated group (<b>A</b>), EAE = immunized and PBS-treated group (<b>B</b>), EtAc = immunized and treated with 200 mg/kg ethanolic extract of <span class="html-italic">A. colubrina</span> barks for six days (<b>C</b>). Representative sections (8 µm) of the examined groups, stained with Luxol fast blue, for analysis of the demyelination. Original magnification: 10× objective. Scale bars = 100 µm. Delimited areas = areas of demyelination.</p>
Full article ">Figure 4
<p>Cellular profile. Mononuclear cell counts (<b>A</b>,<b>E</b>) and cellular profile determination (<b>B</b>–<b>D</b>,<b>F</b>–<b>H</b>) in the brains (<b>A</b>–<b>D</b>) and spinal cords (<b>E</b>–<b>H</b>) of mice immunized or not immunized with 100 µg of MOG<sub>35–55</sub> (n = 8/group). Mice were treated with 200 mg/kg of the ethanolic extract of <span class="html-italic">A. colubrina</span> barks (EtAc) for six days. Each bar represents the arithmetic mean ± SEM. * indicates <span class="html-italic">p</span> &lt; 0.05 compared to induced and PBS-treated animals (EAE). CN = negative control (not induced and treated with PBS).</p>
Full article ">Figure 5
<p>Absolute intensity of the most abundant phenolic acids (cinnamic acid, gallic acid, and <span class="html-italic">p</span>-coumaric acid) and flavonoids (apigenin, catechin, quercetin, and myricetin) annotated through ESI (+) Orbitrap-MS analysis of the ethanolic extract from <span class="html-italic">A. colubrina</span> bark.</p>
Full article ">Figure 6
<p>ESI (+) Orbitrap-MS-based metabolite profiling of <span class="html-italic">A. colubrina</span> showing the major classes identified in bark ethanolic extract.</p>
Full article ">
15 pages, 7966 KiB  
Article
Anticandidal Properties of Launaea sarmentosa among the Salt Marsh Plants Collected from Palk Bay and the Gulf of Mannar Coast, Southeastern India
by Smriti Das, Karuppannagounder Rajan Priyanka, Kolandhasamy Prabhu, Ramachandran Vinayagam, Rajendran Rajaram and Sang Gu Kang
Antibiotics 2024, 13(8), 748; https://doi.org/10.3390/antibiotics13080748 (registering DOI) - 9 Aug 2024
Viewed by 399
Abstract
Tidal wetlands, commonly known as salt marshes, are highly productive ecosystems in temperate regions worldwide. These environments constitute a unique flora composed primarily of salt-tolerant herbs, grasses, and shrubs. This study investigated the therapeutic properties of ten salt marsh plants collected mainly from [...] Read more.
Tidal wetlands, commonly known as salt marshes, are highly productive ecosystems in temperate regions worldwide. These environments constitute a unique flora composed primarily of salt-tolerant herbs, grasses, and shrubs. This study investigated the therapeutic properties of ten salt marsh plants collected mainly from Palk Bay and Mannar Gulf against Candida disease. This study examined the changes in natural plant products associated with their anti-Candida growth activity during two distinct seasonal changes—monsoon and summer. The potential of the salt marshes to inhibit the growth of five different Candida strains was assessed using four solvents. In phytochemical analysis, the extracts obtained from a Launaea sarmentosa exhibited the highest results compared to the other plant extracts. Fourier transform infrared spectroscopy revealed 12 peaks with alkane, aldehyde, amine, aromatic ester, phenol, secondary alcohol, and 1,2,3,4-tetrasubstituted. Gas-chromatography–mass spectrometry detected 30 compounds. Cyclotetracosane, lupeol, β-amyrin, and 12-oleanen-3-yl acetate showed the highest peak range. In particular, plant samples collected during the monsoon season were more effective in preventing Canda growth than the summer plant samples. In the monsoon season, the salt marsh plant extracted with ethyl acetate showed a high anti-Candida growth activity, while in the summer, the acetone extract exhibited a higher anti-Candida growth activity than the other solvents. The hexane extract of L. sarmentosa showed the highest inhibition zone against all Candidal strains. Furthermore, compounds, such as β-amyrin, lupeol, and oxirane, from the hexane extract of L. sarmentosa play a vital role in anti-Candida activity. This paper reports the potential of tidal marsh plant extracts for developing new antifungal agents for Candida infections. Full article
Show Figures

Figure 1

Figure 1
<p>Anticandial properties of salt marsh plants exhibited as the zone of inhibition against Candidal strains during the monsoon season ((<b>a</b>)—Acetone; (<b>b</b>)—Ethyl Acetate; (<b>c</b>)—Methonal and (<b>d</b>)—Hexane).</p>
Full article ">Figure 2
<p>Anticandidal properties of salt marsh plants exhibiting zone of inhibition against Candidal strains during the summer season ((<b>a</b>)—Acetone; (<b>b</b>)—Ethyl acetate; (<b>c</b>)—Methanol; and (<b>d</b>)—Hexane).</p>
Full article ">Figure 3
<p>Anticandidal activity of saltmarsh plant <span class="html-italic">Launaea sarmentosa</span> extract exhibits the highest inhibition (CA—<span class="html-italic">Candida albicans</span>; CR—<span class="html-italic">Candida kefyr</span>; CKr—<span class="html-italic">Candida krusei;</span> CT—<span class="html-italic">Candida tropicalis</span>; CP—<span class="html-italic">Candida parapsilosis</span>).</p>
Full article ">Figure 4
<p>FT−IR spectrum showing the peaks obtained from the hexane extract of salt marsh <span class="html-italic">Launaea sarmentosa</span>.</p>
Full article ">Figure 5
<p>GC-MS showing the peaks obtained from the hexane extract of saltmarsh <span class="html-italic">Launaea sarmentosa</span>.</p>
Full article ">Figure 6
<p>Sampling sites of salt marsh plants collected from Palk Bay and the Gulf of Mannar.</p>
Full article ">Figure 7
<p>Salt marsh plants collected from Palk Bay and the Gulf of Mannar ((<b>A</b>) <span class="html-italic">Ipomoea pes-caprae</span>, (<b>B</b>) <span class="html-italic">Suaeda maritima</span>, (<b>C</b>) <span class="html-italic">Sesuvium portulacastrum</span>, (<b>D</b>) <span class="html-italic">Heliotropium curassavicum</span>, (<b>E</b>) <span class="html-italic">Launaea sarmentosa</span>, (<b>F</b>) <span class="html-italic">Bulbostylis barbata</span>, (<b>G</b>) <span class="html-italic">Salicornia brachiata</span>, (<b>H</b>) <span class="html-italic">Spinifex littoreus</span>, (<b>I</b>) Fim<span class="html-italic">bristylis spathacea</span>, and (<b>J</b>) <span class="html-italic">Artiplex halimus</span>).</p>
Full article ">
17 pages, 9774 KiB  
Article
New Insights into Changes in DOM Fractions in a Crab Farming Park and Key Factors in the Removal Process Using Fluorescence Spectra with MW-2DCOS and SEM
by Ruijuan Zhou, Yan Hao, Benxin Yu, Junwen Hou, Kuotian Lu, Fang Yang and Qingqian Li
Water 2024, 16(16), 2249; https://doi.org/10.3390/w16162249 - 9 Aug 2024
Viewed by 352
Abstract
With the explosion of crab farming in China, the urgent need to treat crab wastewater can never be overemphasized. Hence, in this study, excitation–emission matrix (EEM) fluorescence spectroscopy with parallel factor analysis (PARAFAC), moving window two-dimensional correlation spectroscopy (MW-2DCOS) and structural equation modeling [...] Read more.
With the explosion of crab farming in China, the urgent need to treat crab wastewater can never be overemphasized. Hence, in this study, excitation–emission matrix (EEM) fluorescence spectroscopy with parallel factor analysis (PARAFAC), moving window two-dimensional correlation spectroscopy (MW-2DCOS) and structural equation modeling (SEM) were employed to identify changes in the dissolved organic matter (DOM) fractions in a crab farming park and reveal latent factors associated with removal processes. Seven components (C1–C7) were extracted from DOMs by EEM-PARAFAC as follows: C1: microbial byproduct-like substances, C2: visible-tryptophan-like substances, C3: fulvic-like substances, C4: phenolic-like substances, C5: ultraviolet tyrosine-like substances, C6: D-tryptophan-like substances and C7: L-tryptophan-like substances. Interestingly, C7 (39.20%), a representative component of DOM in the crab farming pond, was deeply degraded in the aeration pond by aerobic microbes, whereas C6 was absent in the crab pond. According to 2DCOS, the changing order of the components was C7 → C4 → C6 → C5 → C2 → C1 → C3, and the changing order of the functional groups was carboxylic → phenolic → aromatic. As assessed by MW-2DCOS, the Fmax of the components, especially components C2, C5 and C6 (and with the exception of C4 and C7) exponentially increased in the aeration pond, where an accumulative effect occurred. C2, C5 and C7 were removed by 24.26%, 39.42% and 98.25% in the crab farming system, and were deeply degraded in the paddy-field, purification pond and aeration pond, respectively. As assessed by SEM, the latent factors of organic matter removal were C1, C2, C4, C5, SUVA254, CODMn and DO. This study could be conducive to comprehensively characterizing the removal of components and functional groups of DOMs in crab farming parks. Full article
(This article belongs to the Special Issue Water Environment Pollution and Control, Volume III)
Show Figures

Figure 1

Figure 1
<p>Generalized diagram of the crab farming industry park and locations of sampling sites.</p>
Full article ">Figure 2
<p>Variations in water quality parameters in different treatment process sections of the crab farming park. (<b>a</b>) pH, (<b>b</b>) EC, (<b>c</b>) DO, (<b>d</b>) NTU, (<b>e</b>) TOC, (<b>f</b>) COD<sub>Cr</sub>, (<b>g</b>) COD<sub>Mn</sub>, (<b>h</b>) NH<sub>3</sub>-N, (<b>i</b>) TN, and (<b>j</b>) TP.</p>
Full article ">Figure 3
<p>EEM spectroscopies of DOMs from the crab farming wastewater at sampling site.</p>
Full article ">Figure 4
<p>UV-visible absorbing spectra at 200–700 nm (<b>a</b>) and 230–500 nm (<b>b</b>).</p>
Full article ">Figure 5
<p>PARAFAC components identified from EEM spectroscopies of DOMs in the crab farming industry park.</p>
Full article ">Figure 5 Cont.
<p>PARAFAC components identified from EEM spectroscopies of DOMs in the crab farming industry park.</p>
Full article ">Figure 6
<p>Fmax (<b>a</b>) and proportions (<b>b</b>) of DOM fractions in the crab farming industry park.</p>
Full article ">Figure 7
<p>Synchronous and asynchronous maps as described by 2DCOS of DOMs from the crab farming park between C1 and C2 (<b>a</b>,<b>b</b>), C1 and C3 (<b>c</b>,<b>d</b>), C2 and C5 (<b>e</b>,<b>f</b>), C5 and C6 (<b>g</b>,<b>h</b>), C4 and C6 (<b>i</b>,<b>j</b>), C4 and C7 (<b>k</b>,<b>l</b>).</p>
Full article ">Figure 8
<p>MW-2DCOS map of a given unit in the crab farming park of C1 (<b>a</b>), C2 (<b>b</b>), C3 (<b>c</b>), C4 (<b>d</b>), C5 (<b>e</b>), C6 (<b>f</b>), and C7 (<b>g</b>).</p>
Full article ">Figure 9
<p>Synchronous map (<b>a</b>) and asynchronous map (<b>b</b>) of 2DCOS using UV-visible absorbing spectra at 230–450 nm.</p>
Full article ">Figure 10
<p>Plots based on the RDA of the interactions between response variables and environmental explanatory variables (solid arrows with red fonts are the response variables and hollow arrows with black fonts are the environmental explanatory variables).</p>
Full article ">Figure 11
<p>SEM modeling for the relationship between fluorescent components (C1, C2, C4, C5), water quality parameters (CODMn, DO), and spectroscopic indices (SUVA254), and contributions to removal efficiencies of FDOMs and TOC (R-FDOM, R-TOC). Significance levels of standardized path coefficient are: *** <span class="html-italic">p</span> &lt; 0.001, ** <span class="html-italic">p</span> &lt; 0.01, * <span class="html-italic">p</span> &lt; 0.05.</p>
Full article ">
19 pages, 3795 KiB  
Article
It Is Not All about Alkaloids—Overlooked Secondary Constituents in Roots and Rhizomes of Gelsemium sempervirens (L.) J.St.-Hil
by Lilo K. Mailänder, Khadijeh Nosrati Gazafroudi, Peter Lorenz, Rolf Daniels, Florian C. Stintzing and Dietmar R. Kammerer
Plants 2024, 13(16), 2208; https://doi.org/10.3390/plants13162208 - 9 Aug 2024
Viewed by 354
Abstract
Gelsemium sempervirens (L.) J.St.-Hil. is an evergreen shrub occurring naturally in North and Middle America. So far, more than 120 alkaloids have been identified in this plant in addition to steroids, coumarins and iridoids, and its use in traditional medicine has been traced [...] Read more.
Gelsemium sempervirens (L.) J.St.-Hil. is an evergreen shrub occurring naturally in North and Middle America. So far, more than 120 alkaloids have been identified in this plant in addition to steroids, coumarins and iridoids, and its use in traditional medicine has been traced back to these compound classes. However, a comprehensive phytochemical investigation of the plant with a special focus on further compound classes has not yet been performed. Therefore, the present study aimed at an extensive HPLC-MSn characterization of secondary metabolites and, for the first time, reports the occurrence of various depsides and phenolic glycerides in G. sempervirens roots and rhizomes, consisting of benzoic and cinnamic acid derivatives as well as dicarboxylic acids. Furthermore, mono- and disaccharides were assigned by GC-MS. Applying the Folin–Ciocalteu assay, the phenolic content of extracts obtained with different solvents was estimated to range from 30 to 50% calculated as chlorogenic acid equivalents per g dry weight and was related to the DPPH radical scavenging activity of the respective extracts. Upon lactic acid fermentation of aqueous G. sempervirens extracts, degradation of phenolic esters was observed going along with the formation of low-molecular volatile metabolites. Full article
(This article belongs to the Section Phytochemistry)
Show Figures

Figure 1

Figure 1
<p>(<b>a</b>) Flowers of <span class="html-italic">G. sempervirens</span>. © Horst Arne Schneider. (<b>b</b>) Roots and rhizome of <span class="html-italic">G. sempervirens</span>. Photo: L. Mailänder.</p>
Full article ">Figure 2
<p>Total phenolic contents of dichloromethane (DCM), ethyl acetate (EtOAc), and <span class="html-italic">n</span>-butanol (<span class="html-italic">n</span>-BuOH) extracts of <span class="html-italic">G. sempervirens</span> roots and rhizomes. Results are expressed as µg gallic acid equivalents (GAE)/mg dry weight (dw) and µg chlorogenic acid equivalents (CAE)/mg dw, respectively; mean ± SD; <span class="html-italic">n</span> = 3.</p>
Full article ">Figure 3
<p>Percentage of DPPH scavenged by different solvent extracts from <span class="html-italic">G. sempervirens</span> roots and rhizomes in comparison to trolox as reference compound.</p>
Full article ">Figure 4
<p>GC-MS profiles of secondary constituents in (<b>a</b>) dichloromethane, (<b>b</b>) ethyl acetate, and (<b>c</b>) <span class="html-italic">n</span>-butanol extracts after silylation. Compound assignment is presented in <a href="#plants-13-02208-t001" class="html-table">Table 1</a>.</p>
Full article ">Figure 5
<p>HPLC-MS<sup>n</sup> base peak chromatograms of an (<b>a</b>) ethyl acetate and (<b>b</b>) <span class="html-italic">n</span>-butanol extract of dried <span class="html-italic">G. sempervirens</span> roots and rhizomes showing the occurrence of di- and tridepsides. The most abundant monomeric constituents are designated as follows: CA caffeic acid; CiA cinnamic acid; FA ferulic acid; GA gallic acid; gly glycerol; hex hexose; PA protocatechuic acid; QA quinic acid; SA syringic acid; sco scopoletin; SiA sinapinic acid; TA tartaric acid; VA veratric acid; x minor constituent (see <a href="#plants-13-02208-t002" class="html-table">Table 2</a>); dihydro derivatives are displayed in bordered boxes.</p>
Full article ">Figure 6
<p>UV spectral and mass spectrometric structure assignment exemplified for compounds <b>28</b> (<b>a</b>) and <b>41</b> (<b>b</b>).</p>
Full article ">Figure 7
<p>HPLC-MS<sup>n</sup> base peak chromatograms of fermentation samples on day 0 (light grey) and 30 (dark grey) were recorded in negative (ESI−; top) and positive (ESI+; bottom) ionization mode. Peak numbers correspond to <a href="#plants-13-02208-t002" class="html-table">Table 2</a>, LC-MS data of the alkaloids can be found in the <a href="#app1-plants-13-02208" class="html-app">Supporting Information (Table S4.1)</a>.</p>
Full article ">Figure 8
<p>GC-MS total ion current chromatogram of volatile compounds extracted with diethyl ether from six months old aqueous fermented <span class="html-italic">G. sempervirens</span> extracts.</p>
Full article ">Figure 9
<p>Dried <span class="html-italic">G. sempervirens</span> plant material used in this study.</p>
Full article ">
13 pages, 1789 KiB  
Article
Olive Mill Wastewater Extract: In Vitro Genotoxicity/Antigenotoxicity Assessment on HepaRG Cells
by Tommaso Rondini, Raffaella Branciari, Edoardo Franceschini, Mattia Acito, Cristina Fatigoni, Rossana Roila, David Ranucci, Milena Villarini, Roberta Galarini and Massimo Moretti
Int. J. Environ. Res. Public Health 2024, 21(8), 1050; https://doi.org/10.3390/ijerph21081050 - 9 Aug 2024
Viewed by 408
Abstract
Olive mill wastewater (OMWW), with its high level of phenolic compounds, simultaneously represents a serious environmental challenge and a great resource with potential nutraceutical activities. To increase the knowledge of OMWW’s biological effects, with an aim to developing a food supplement, we performed [...] Read more.
Olive mill wastewater (OMWW), with its high level of phenolic compounds, simultaneously represents a serious environmental challenge and a great resource with potential nutraceutical activities. To increase the knowledge of OMWW’s biological effects, with an aim to developing a food supplement, we performed a chemical characterisation of the extract using the Liquid Chromatography–Quadrupole Time-of-flight spectrometry (LC–QTOF) and an in vitro genotoxicity/antigenotoxicity assessment on HepaRG ™ cells. Chemical analysis revealed that the most abundant phenolic compound was hydroxytyrosol. Biological tests showed that the extract was not cytotoxic at the lowest tested concentrations (from 0.25 to 2.5 mg/mL), unlike the highest concentrations (from 5 to 20 mg/mL). Regarding genotoxic activity, when tested at non-cytotoxic concentrations, the extract did not display any effect. Additionally, the lowest tested OMWW concentrations showed antigenotoxic activity (J-shaped dose–response effect) against a known mutagenic substance, reducing the extent of DNA damage in the co-exposure treatment. The antigenotoxic effect was also obtained in the post-exposure procedure, although only at the extract concentrations of 0.015625 and 0.03125 mg/mL. This behaviour was not confirmed in the pre-exposure protocol. In conclusion, the present study established a maximum non-toxic OMWW extract dose for the HepaRG cell model, smoothing the path for future research. Full article
Show Figures

Figure 1

Figure 1
<p>Extract ion chromatograms of the three most abundant polyphenols: hydroxytyrosol, tyrosol, and vanillic acid. (<b>A</b>) Standard solution at 50 ng/mL; (<b>B</b>) extract (10,000-fold dilution).</p>
Full article ">Figure 2
<p>Effect of OMWW extract on cell viability in HepaRG cells after 4 h exposure. Cytotoxic effects were assessed with a AO/DAPI double staining test. Results are summarised as mean ± SEM of three independent experiments. Control = untreated cells. Results are expressed as normalised viability compared with the control (=100% of viability); * <span class="html-italic">p</span> &lt; 0.05 vs. control.</p>
Full article ">Figure 3
<p>Genotoxic effects of OMWW extract in HepaRG cells after 4 h of exposure. Each result is expressed as the mean ± SEM of three independent experiments. Control (untreated cells) = 2.89 ± 0.76%; Positive control (2 μM 4NQO).</p>
Full article ">Figure 4
<p>Antigenotoxic effects of OMWW extract on 4NQO-induced DNA damage in HepaRG cells (co-exposure protocol) and genotoxic inhibition ratio toward model mutagen 4NQO (2 μM). Each result is expressed as the mean ± SEM of three independent experiments. Control (untreated cells) = 3.45 ± 1.48%; positive control (2 μM 4NQO); * <span class="html-italic">p</span> &lt; 0.05 vs. 2 μM 4NQO.</p>
Full article ">Figure 5
<p>Antigenotoxic effects of OMWW extract on 4NQO-induced DNA damage in HepaRG cells: pre-exposure protocol. Each result is expressed as the mean ± SEM of three independent experiments. Control (untreated cells) = 3.91 ± 0.75%; positive control (2 µM 4NQO).</p>
Full article ">Figure 6
<p>Antigenotoxic effects of OMWW extract on 4NQO-induced DNA damage in HepaRG cells: post-exposure protocol. Each result is expressed as the mean ± standard error of the mean (SEM) of three independent experiments. Control (untreated cells) = 3.95 ± 0.75%; positive control (2 µM 4NQO); * <span class="html-italic">p</span> &lt; 0.05 vs. 2 µM 4NQO.</p>
Full article ">
26 pages, 6947 KiB  
Review
Citrus limon var. pompia Camarda var. nova: A Comprehensive Review of Its Botanical Characteristics, Traditional Uses, Phytochemical Profile, and Potential Health Benefits
by Anna Maria Posadino, Paola Maccioccu, Ali H. Eid, Roberta Giordo, Gianfranco Pintus and Grazia Fenu
Nutrients 2024, 16(16), 2619; https://doi.org/10.3390/nu16162619 - 8 Aug 2024
Viewed by 404
Abstract
Citrus limon var. pompia Camarda var. nova, commonly known as pompia, is a distinctive citrus ecotype native to Sardinia, notable for its unique botanical, phytochemical, and potential health benefits. It holds cultural significance as a traditional food product of Sardinia, recognized by [...] Read more.
Citrus limon var. pompia Camarda var. nova, commonly known as pompia, is a distinctive citrus ecotype native to Sardinia, notable for its unique botanical, phytochemical, and potential health benefits. It holds cultural significance as a traditional food product of Sardinia, recognized by the Italian Ministry of Agricultural Food and Forestry Policies. This comprehensive review examines pompia’s traditional uses, taxonomic classification, pomological characteristics, phytochemical profile, and potential health benefits. Pompia phytochemical analyses reveal a rich composition of flavonoids and terpenoids, with notable concentrations of limonene, myrcene, and various oxygenated monoterpenes. Pompia essential oils are primarily extracted from its peel and leaves. Peel essential oils exhibit a high concentration of the monoterpene limonene (82%) and significantly lower quantities of myrcene (1.8%), geranial (1.7%), geraniol (1.5%), and neral (1.4%). In its rind extract, flavanones such as naringin (23.77 µg/mg), neoeriocitrin (46.53 µg/mg), and neohesperidin (44.57 µg/mg) have been found, along with gallic acid (128.3 µg/mg) and quinic acid (219.67 µg/mg). The main compounds detected in the essential oils from pompia leaves are oxygenated monoterpenes (53.5%), with limonene (28.64%), α-terpineol (41.18%), geranial (24.44%), (E)-β-ocimene (10.5%), linalool (0.56%), and neryl acetate (13.56%) being particularly prominent. In pompia juice, the presence of phenolic compounds has been discovered, with a composition more similar to lemon juice than orange juice. The primary flavonoid identified in pompia juice is chrysoeriol-6,8-di-C-glucoside (stellarin-2) (109.2 mg/L), which has not been found in other citrus juices. The compound rhoifolin-4-glucoside (17.5 mg/L) is unique to pompia juice, whereas its aglycone, rhoifolin, is found in lemon juice. Other flavonoids identified in pompia juice include diosmetin 6,8-C-diglucoside (54.5 mg/L) and isorhamnetin 3-O-rutinoside (79.4 mg/L). These findings support the potential of pompia in developing nutraceuticals and natural health products, further confirmed by its compounds’ antioxidant, anti-inflammatory and antibacterial properties. Future research should focus on optimizing extraction methods, conducting clinical trials to evaluate efficacy and safety, and exploring sustainable cultivation practices. The potential applications of pompia extracts in food preservation, functional foods, and cosmetic formulations also warrant further investigation. Addressing these areas could significantly enhance pompia’s contribution to natural medicine, food science, and biotechnology. Full article
(This article belongs to the Section Phytochemicals and Human Health)
Show Figures

Figure 1

Figure 1
<p>Pompia plants in a citrus grove of the Baronia (<a href="https://www.biodiversitasardegna.it/laore/it/agrobiodiversita/repertorio-regionale/risorsa/Pompia/" target="_blank">https://www.biodiversitasardegna.it/laore/it/agrobiodiversita/repertorio-regionale/risorsa/Pompia/</a>) (accessed on 1 July 2024).</p>
Full article ">Figure 2
<p>Pompia fruit (<a href="https://www.biodiversitasardegna.it/laore/it/agrobiodiversita/repertorio-regionale/risorsa/Pompia" target="_blank">https://www.biodiversitasardegna.it/laore/it/agrobiodiversita/repertorio-regionale/risorsa/Pompia</a>) (accessed on 1 July 2024).</p>
Full article ">Figure 3
<p>An illustrated photo of opened pompia fruit (<a href="https://it.wikipedia.org/wiki/Pompia#/media/File:Pompia_frutto_aperto.jpg" target="_blank">https://it.wikipedia.org/wiki/Pompia#/media/File:Pompia_frutto_aperto.jpg</a>) (accessed on 1 July 2024).</p>
Full article ">Figure 4
<p>Special Sardinian dessert cake: “Sa Pompia Intrea” (<a href="https://it.wikipedia.org/wiki/Pompia#/media/File:Pompia_intrea.jpg" target="_blank">https://it.wikipedia.org/wiki/Pompia#/media/File:Pompia_intrea.jpg</a>) (accessed on 1 July 2024).</p>
Full article ">
16 pages, 2388 KiB  
Article
Anti-Growth and Anti-Metastatic Potential of Raw and Thermally Treated Eragrostis tef Extract in Human Cancer Cells
by Jina Seo, Hwa Jin Lee and Jihyeung Ju
Nutrients 2024, 16(16), 2612; https://doi.org/10.3390/nu16162612 - 8 Aug 2024
Viewed by 418
Abstract
Teff (Eragrostis tef), a gluten-free cereal crop cultivated originally in Northeast Africa, is increasingly utilized due to its nutritional and health benefits. The aim of the present study was to investigate the effects of ethanol extract obtained from raw and thermally [...] Read more.
Teff (Eragrostis tef), a gluten-free cereal crop cultivated originally in Northeast Africa, is increasingly utilized due to its nutritional and health benefits. The aim of the present study was to investigate the effects of ethanol extract obtained from raw and thermally treated teff, referred to as RTE and TTE, respectively, on uncontrolled growth and activated metastasis using human cancer cell lines. Both RTE and TTE contained flavones, such as orientin (luteolin 8-C-glucoside) and vitexin (apigenin 8-C-glucoside), and phenolic acids, such as protocatechuic acid and p-coumaric acid. TTE showed higher total phenol, protocatechuic acid, and p-coumaric acid contents, but lower orientin content compared to RTE. RTE and TTE significantly suppressed cell growth of H1299 human lung cancer cells, with TTE exhibiting more pronounced effects than RTE, while both extracts had only minimal effects on the growth of non-malignant human umbilical vein endothelial cells. The growth-inhibitory activities of RTE and TTE in H1299 cells were associated with apoptosis induction and cell cycle arrest at the G2/M phase. TTE produced an additional effect on inducing cell cycle arrest at the S phase in H1299 cells, potentially contributing to its stronger growth-inhibitory effects. Moreover, both RTE and TTE effectively inhibited key events in metastasis, such as invasion, migration, and adhesion, in H1299 cells under non-cytotoxic conditions, with TTE showing stronger effects. In HCT116 human colon cancer cells, a similar pattern of inhibition was demonstrated against the metastatic events, accompanied by reduced levels of matrix metalloproteinase-2 and -9. Our results indicate that teff extracts exhibit in vitro anti-growth and anti-metastatic activities, which are enhanced by thermal treatment of teff. Full article
Show Figures

Figure 1

Figure 1
<p>The HPLC chromatograms of RTE and TTE. 1: protocatechuic acid, 2: orientin, 3: <span class="html-italic">p</span>-coumaric acid, 4: vitexin.</p>
Full article ">Figure 2
<p>Effects of RTE and TTE on cell growth in H1299 human lung cancer cells and human umbilical vein endothelial cells. H1299 cells (<b>A</b>) and human umbilical vein endothelial cells (HUVEC) (<b>B</b>) were exposed to RTE and TTE at 0 (control, CTRL), 100, 250, and 500 μg/mL for 24 h, 48 h, and 72 h. The viability of cells was determined relative to the control. Within each treatment group (RTE and TTE), distinct letters (a–d) indicate significant differences among concentrations (<span class="html-italic">p</span> &lt; 0.05). Asterisks indicate significant differences between RTE and TTE treatment at a specific concentration (* <span class="html-italic">p</span> &lt; 0.05, ** <span class="html-italic">p</span> &lt; 0.01, *** <span class="html-italic">p</span> &lt; 0.001). NS: not significant.</p>
Full article ">Figure 3
<p>Effects of RTE and TTE on cell cycle distribution in H1299 human lung cancer cells. H1229 cells were exposed to 0 (control, CTRL) or 500 μg/mL of RTE and TTE for 48 h. The distribution of cells across sub-G1 (denoted as M4, apoptotic), G0/G1 (denoted as M1), S (denoted as M3), and G2/M (denoted as M2) phases was assessed using flow cytometry. Representative cell cycle distributions are shown in the right panel. Within each treatment group (RTE and TTE), distinct letters (a–c) indicate significant differences among concentrations (<span class="html-italic">p</span> &lt; 0.05).</p>
Full article ">Figure 4
<p>Effects of RTE and TTE on invasion in H1299 human lung cancer cells. H1299 were exposed to 0 (control, CTRL), 50, and 100 μg/mL of RTE and TTE for 16 h. The percentage of invasive cells was assessed using a transwell invasion assay, and representative images of stained invasive cells are shown in the right panel. Within each treatment group (RTE and TTE), distinct letters (a–c) indicate significant differences among concentrations (<span class="html-italic">p</span> &lt; 0.05). Asterisks indicate significant differences between RTE and TTE treatment at a specific concentration (*** <span class="html-italic">p</span> &lt; 0.001).</p>
Full article ">Figure 5
<p>Effects of RTE and TTE on migration in H1299 human lung cancer cells. H1229 cells were exposed to 0 (control, CTRL) or 100 μg/mL of RTE and TTE in scratch wound-healing assay, and the extent of cell migration was evaluated at 8 h by measuring the closure of wounds relative to the control. H1229 were exposed to 0 (control, CTRL), 50, and 100 μg/mL of RTE and TTE for 16 h in transwell migration assay, and the extent of cell migration was determined relative to the control. Within each treatment group (RTE and TTE), distinct letters (a–c) indicate significant differences among concentrations (<span class="html-italic">p</span> &lt; 0.05). Asterisks indicate significant differences between RTE and TTE treatment at a specific concentration (*** <span class="html-italic">p</span> &lt; 0.001).</p>
Full article ">Figure 6
<p>Effects of RTE and TTE on adhesion in H1299 human lung cancer cells. H1299 cells were exposed to 0 (control, CTRL) or 1000 μg/mL of RTE and TTE for 1 h, and the extent of cell adhesion to gelatin and fibronectin was determined relative to the control. Distinct letters (a–c) indicate significant differences among groups (<span class="html-italic">p</span> &lt; 0.05).</p>
Full article ">Figure 7
<p>Effects of RTE and TTE on invasion, migration, adhesion, and MMP levels in HCT116 human colon cancer cells. HCT116 cells were treated with RTE and TTE under the following conditions: 0, 50, and 100 μg/mL for 24 h in transwell invasion and migration assays; 0 or 1000 μg/mL for 1 h in adhesion assay using gelatin-coated plate; and 0 or 500 μg/mL for 24 h in ELISA for MMP-2 and -9. The percentage of invasive, migratory, adhesive cells, as well as the levels of MMP-2 and -9, were determined relative to the control. Within each treatment group (RTE and TTE), distinct letters (a–c) indicate significant differences among concentrations (<span class="html-italic">p</span> &lt; 0.05). Asterisks indicate significant differences between RTE and TTE treatment at a specific concentration (** <span class="html-italic">p</span> &lt; 0.01, *** <span class="html-italic">p</span> &lt; 0.001).</p>
Full article ">
Back to TopTop