[go: up one dir, main page]

 
 
Sign in to use this feature.

Years

Between: -

Subjects

remove_circle_outline
remove_circle_outline
remove_circle_outline
remove_circle_outline
remove_circle_outline
remove_circle_outline
remove_circle_outline
remove_circle_outline
remove_circle_outline

Journals

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

Search Results (19,661)

Search Parameters:
Keywords = fatty acids

Order results
Result details
Results per page
Select all
Export citation of selected articles as:
13 pages, 11706 KiB  
Article
Chemical Profile and Potential Applications of Sclerocarya birrea (A.Rich.) Hochst. subsp. caffra (Sond.) Kokwaro Kernel Oils: Analysis of Volatile Compounds and Fatty Acids
by Callistus Bvenura and Learnmore Kambizi
Molecules 2024, 29(16), 3815; https://doi.org/10.3390/molecules29163815 (registering DOI) - 11 Aug 2024
Abstract
Sclerocarya birrea kernel volatile compounds and fatty acid methyl esters (FAMEs) from the Bubi district in Matabeleland North province of Zimbabwe were characterised by GC–MS. The volatile compounds of the oil include 65 different compounds from 24 distinct classes, dominated by 13 alcohols [...] Read more.
Sclerocarya birrea kernel volatile compounds and fatty acid methyl esters (FAMEs) from the Bubi district in Matabeleland North province of Zimbabwe were characterised by GC–MS. The volatile compounds of the oil include 65 different compounds from 24 distinct classes, dominated by 13 alcohols and 14 aldehydes (42%). Other classes include carboxylic acids, phenols, sesquiterpenes, lactones, pyridines, saturated fatty acids, ketones, and various hydrocarbons. The kernel oils revealed essential fatty acids such as polyunsaturated (α-linolenic and linoleic acids) and monounsaturated fatty acids (palmitic, palmitoleic, and oleic acids). Notably, oleic acid is the predominant fatty acid at 521.61 mg/g, constituting approximately 73% of the total fatty acids. Linoleic acid makes up 8%, and saturated fatty acids make up about 7%, including significant amounts of stearic (42.45 mg/g) and arachidic (3.46 mg/g) acids. These results validate the use of marula oils in food, pharmaceutical, and health industries, as well as in the multibillion USD cosmetics industry. Therefore, the potential applications of S. berria kernel oils are extensive, necessitating further research and exploration to fully unlock their capabilities. Full article
(This article belongs to the Special Issue Functional Evaluation of Bioactive Compounds from Natural Sources)
12 pages, 2904 KiB  
Article
Sodium Acetate Enhances Neutrophil Extracellular Trap Formation via Histone Acetylation Pathway in Neutrophil-like HL-60 Cells
by Hiroyuki Yasuda, Yutaka Takishita, Akihiro Morita, Tomonari Tsutsumi, Naoya Nakagawa and Eisuke F. Sato
Int. J. Mol. Sci. 2024, 25(16), 8757; https://doi.org/10.3390/ijms25168757 (registering DOI) - 11 Aug 2024
Abstract
Neutrophil extracellular trap formation has been identified as a new cell death mediator, termed NETosis, which is distinct from apoptosis and necrosis. NETs capture foreign substances, such as bacteria, by releasing DNA into the extracellular environment, and have been associated with inflammatory diseases [...] Read more.
Neutrophil extracellular trap formation has been identified as a new cell death mediator, termed NETosis, which is distinct from apoptosis and necrosis. NETs capture foreign substances, such as bacteria, by releasing DNA into the extracellular environment, and have been associated with inflammatory diseases and altered immune responses. Short-chain fatty acids, such as acetate, are produced by the gut microbiota and reportedly enhance innate immune responses; however, the underlying molecular mechanisms remain unclear. Here, we investigated the effects of sodium acetate, which has the highest SCFA concentration in the blood and gastrointestinal tract, on NETosis by focusing on the mechanisms associated with histone acetylation in neutrophil-like HL-60 cells. Sodium acetate enhanced NETosis, as shown by fluorescence staining with SYTOX green, and the effect was directly proportional to the treatment duration (16–24 h). Moreover, the addition of sodium acetate significantly enhanced the acetylation of Ace-H3, H3K9ace, and H3K14ace. Sodium acetate-induced histone acetylation rapidly decreased upon stimulation with the calcium ionophore A23187, whereas histone citrullination markedly increased. These results demonstrate that sodium acetate induces NETosis via histone acetylation in neutrophil-like HL-60 cells, providing new insights into the therapeutic effects based on the innate immunity-enhancing effect of dietary fiber. Full article
Show Figures

Figure 1

Figure 1
<p>Sodium acetate increases A23187-induced neutrophil extracellular trap formation (NETosis) (NOX-independent NETosis). (<b>a</b>) After treatment with or without 10 mM sodium acetate during differentiation, NETosis levels in nHL-60 cells treated with (+) or without (−) 10 μM A23187 for 4 h were analyzed through SYTOX green assay (<span class="html-italic">n</span> = 5). Control was not treated with sodium acetate. 16 hr and 24 hr: incubation time of sodium acetate. After sodium acetate incubation, cells were stimulated by A23187. (<b>b</b>) After treatment with or without 10 mM sodium acetate during differentiation, NETosis images were obtained using confocal microscopy of SYTOX green (DNA)-stained nHL-60 cells treated with (+) or without (−) 10 μM A23187 for 2 h.</p>
Full article ">Figure 2
<p>Sodium acetate treatment did not increase ROS production in nHL-60 cells after treatment with or without 10 mM sodium acetate during differentiation. Control was not treated with sodium acetate. 16 hr and 24 hr: incubation time of sodium acetate. After sodium acetate incubation, cells were stimulated by A23187. Quantitative chemiluminescence analysis of ROS production in nHL-60 cell culture incubated with or without 10 μM A23187 for 90 min was performed. Data are shown as mean ± SD (<span class="html-italic">n</span> = 3).</p>
Full article ">Figure 3
<p>Effect of ACSS2 inhibitor on release of extracellular DNA. After treating nHL-60 cells with or without 10 μ M ACSS2 inhibitor during sodium acetate treatment, extracellular DNA was isolated and measured by performing extracellular DNA quantification assay. +: Addition of reagents, −: No treatment. Data are shown as mean ± SD (<span class="html-italic">n</span> = 3).</p>
Full article ">Figure 4
<p>Expression of ACSS2, MCT-1, and MCT-4 in nHL-60 cells. After treatment with or without sodium acetate (1 or 10 mM) for 24 h, ACSS2, MCT-1, and MCT-4 expression levels in nHL-60 cells were analyzed using Western blotting. −: No treatment. Data are shown as mean ± SD (<span class="html-italic">n</span> = 3).</p>
Full article ">Figure 5
<p>Sodium acetate enhanced histone acetylation. The nHL-60 cells were incubated with or without sodium acetate for 1 or 24 h. −: The control was not treated with sodium acetate. Following histone extraction, histone acetylation levels were analyzed using Western blotting. Loading of the histones was monitored using Coomassie staining (denoted as CBB). The data are shown as the mean ± SD (<span class="html-italic">n</span> = 3).</p>
Full article ">Figure 6
<p>Histone H3 acetylation in nHL-60 cells. The nHL-60 cells were incubated with or without sodium acetate for 24 h. Following histone extraction, histone acetylation (denoted as H3K9ace, H3K14ace, H3K18ace, and H3K27ace) levels were analyzed by Western blotting using specific antibodies. −: The control was not treated with sodium acetate. +: Addition of 10 mM CH<sub>3</sub>COONa. Loading of the histones was monitored using Coomassie staining (denoted as CBB). The data are shown as the mean ± SD (<span class="html-italic">n</span> = 3).</p>
Full article ">Figure 7
<p>Histone acetylation levels before and after A23187 stimulation. The nHL-60 cells were incubated with or without 10 mM sodium acetate for 24 h. Then, the cells were stimulated with 10 μM A23187. Following histone extraction, histone acetylation levels were analyzed by Western blotting. The control was not treated with sodium acetate. +: Addition of 10 mM CH<sub>3</sub>COONa, −: No treatment. Loading of the histones was monitored using Coomassie staining (denoted as CBB). The data are shown as the mean ± SD (<span class="html-italic">n</span> = 3).</p>
Full article ">Figure 8
<p>Sodium acetate enhanced histone citrullination but not PAD4 expression. (<b>a</b>) nHL-60 cells (treated with or without 10 mM sodium acetate) were treated with or without 10 μM A23187 for 3 h. Following histone extraction, citrullinated histone H3 (denoted as citH3) protein levels were analyzed via Western blotting. Loading of histones was monitored using Coomassie staining (denoted as CBB). +: Addition of reagents, −: No treatment. (<b>b</b>) nHL-60 cells were incubated with or without 10 mM sodium acetate. After cell extraction, PAD4 expression was determined through Western blotting. Ct: Control was not treated with sodium acetate. Data are shown as mean ± SD (<span class="html-italic">n</span> = 3).</p>
Full article ">
24 pages, 711 KiB  
Review
Effects of Physical Exercise on the Microbiota in Irritable Bowel Syndrome
by Chunpeng Li, Jianmin Li, Qiaorui Zhou, Can Wang, Jiahui Hu and Chang Liu
Nutrients 2024, 16(16), 2657; https://doi.org/10.3390/nu16162657 (registering DOI) - 11 Aug 2024
Abstract
Irritable bowel syndrome (IBS) is a prevalent functional gastrointestinal disorder characterized by abdominal pain, bloating, diarrhea, and constipation. Recent studies have underscored the significant role of the gut microbiota in the pathogenesis of IBS. Physical exercise, as a non-pharmacological intervention, has been proposed [...] Read more.
Irritable bowel syndrome (IBS) is a prevalent functional gastrointestinal disorder characterized by abdominal pain, bloating, diarrhea, and constipation. Recent studies have underscored the significant role of the gut microbiota in the pathogenesis of IBS. Physical exercise, as a non-pharmacological intervention, has been proposed to alleviate IBS symptoms by modulating the gut microbiota. Aerobic exercise, such as running, swimming, and cycling, has been shown to enhance the diversity and abundance of beneficial gut bacteria, including Lactobacillus and Bifidobacterium. These bacteria produce short-chain fatty acids that possess anti-inflammatory properties and support gut barrier integrity. Studies involving IBS patients participating in structured aerobic exercise programs have reported significant improvements in their gut microbiota’s composition and diversity, alongside an alleviation of symptoms like abdominal pain and bloating. Additionally, exercise positively influences mental health by reducing stress and improving mood, which can further relieve IBS symptoms via the gut–brain axis. Long-term exercise interventions provide sustained benefits, maintaining the gut microbiota’s diversity and stability, supporting immune functions, and reducing systemic inflammation. However, exercise programs must be tailored to individual needs to avoid exacerbating IBS symptoms. Personalized exercise plans starting with low-to-moderate intensity and gradually increasing in intensity can maximize the benefits and minimize risks. This review examines the impact of various types and intensities of physical exercise on the gut microbiota in IBS patients, highlighting the need for further studies to explore optimal exercise protocols. Future research should include larger sample sizes, longer follow-up periods, and examine the synergistic effects of exercise and other lifestyle modifications. Integrating physical exercise into comprehensive IBS management plans can enhance symptom control and improve patients’ quality of life. Full article
(This article belongs to the Section Nutritional Immunology)
Show Figures

Figure 1

Figure 1
<p>Effects of physical exercise on the gut microbiota in irritable bowel syndrome and on symptoms. IBS—irritable bowel syndrome; IS—immune system; PC—pro-inflammatory cytokines.</p>
Full article ">
21 pages, 6311 KiB  
Article
Investigation of Antioxidant Activity of Protein Hydrolysates from New Zealand Commercial Low-Grade Fish Roes
by Shuxian Li, Alan Carne and Alaa El-Din Ahmed Bekhit
Mar. Drugs 2024, 22(8), 364; https://doi.org/10.3390/md22080364 (registering DOI) - 11 Aug 2024
Abstract
The objective of this study was to investigate the nutrient composition of low-grade New Zealand commercial fish (Gemfish and Hoki) roe and to investigate the effects of delipidation and freeze-drying processes on roe hydrolysis and antioxidant activities of their protein hydrolysates. Enzymatic hydrolysis [...] Read more.
The objective of this study was to investigate the nutrient composition of low-grade New Zealand commercial fish (Gemfish and Hoki) roe and to investigate the effects of delipidation and freeze-drying processes on roe hydrolysis and antioxidant activities of their protein hydrolysates. Enzymatic hydrolysis of the Hoki and Gemfish roe homogenates was carried out using three commercial proteases: Alcalase, bacterial protease HT, and fungal protease FP-II. The protein and lipid contents of Gemfish and Hoki roes were 23.8% and 7.6%; and 17.9% and 10.1%, respectively. The lipid fraction consisted mainly of monounsaturated fatty acid (MUFA) in both Gemfish roe (41.5%) and Hoki roe (40.2%), and docosahexaenoic (DHA) was the dominant polyunsaturated fatty acid (PUFA) in Gemfish roe (21.4%) and Hoki roe (18.6%). Phosphatidylcholine was the main phospholipid in Gemfish roe (34.6%) and Hoki roe (28.7%). Alcalase achieved the most extensive hydrolysis, and its hydrolysate displayed the highest 2,2-dipheny1-1-picrylhydrazyl (DPPH)˙ and 2,2′-azino-bis(3-ethylbenzothiazoline-6-sulfonic acid) (ABTS) radical scavenging activities and ferric reducing antioxidant power (FRAP). The combination of defatting and freeze-drying treatments reduced DPPH˙ scavenging activity (by 38%), ABTS˙ scavenging activity (by 40%) and ferric (Fe3+) reducing power by18% (p < 0.05). These findings indicate that pre-processing treatments of delipidation and freeze-drying could negatively impact the effectiveness of enzymatic hydrolysis in extracting valuable compounds from low grade roe. Full article
(This article belongs to the Special Issue The Bioactive Potential of Marine-Derived Peptides and Proteins)
Show Figures

Figure 1

Figure 1
<p>Time course hydrolysis of frozen-thawed Hoki and Gemfish roes with lipid present. Homogenates of fresh Hoki and Gemfish roe were prepared by adding 7.3 g and 5.5 g of roe to 200 mL potassium phosphate buffer (pH 7) to achieve 6.5 mg/mL protein concentration, and aliquots of 20 mL were subject to hydrolysis with three different concentrations (2%, 6%, and 10%) of either microbial protease Alcalase (<span class="html-italic">v</span>/<span class="html-italic">w</span>), bacterial protease HT (<span class="html-italic">w</span>/<span class="html-italic">w</span>), or fungal protease FP-II (<span class="html-italic">w</span>/<span class="html-italic">w</span>), respectively, at 45 °C. Alcalase is commercially available as a solution, and HT and FP-II are commercially available as powders. The degree of hydrolysis was determined based on the L-serine equivalent. The data were obtained from three independent hydrolyses for each roe and protease combination. Analysis of variance (ANOVA) was carried out. Letters a–c indicate significant differences among different samples prepared with varying enzyme concentrations at the same time point. (<span class="html-italic">p</span> &lt; 0.05).</p>
Full article ">Figure 2
<p>Time course hydrolysis of fresh Hoki and Gemfish roes after removing the lipid fraction. Before hydrolysis, the Hoki roe and Gemfish roe homogenates were delipidated by ethanol and hexane using the ETHEX lipid extraction method. For hydrolysis, homogenates of delipidated frozen-thawed Hoki and Gemfish roes were prepared by adding 7.3 g and 5.5 g of thawed delipidation roe to 200 mL of potassium phosphate buffer (pH 7) to achieve 6.5 mg/mL protein concentration and aliquots of 20 mL were subject to hydrolysis at three different concentrations (2%, 6%, and 10%) of either Alcalase (<span class="html-italic">v/w</span>), bacterial protease HT (<span class="html-italic">w/w</span>), or fungal protease FP-II(<span class="html-italic">w/w</span>) at 45 °C. The degree of hydrolysis of the samples was determined based on the L-serine equivalent method. The data were obtained from three independent hydrolyses for each roe and protease combination. Analysis of variance (ANOVA) was carried out. Letters a–c indicate significant differences among different samples prepared with varying enzyme concentrations at the same time point (<span class="html-italic">p</span> &lt; 0.05).</p>
Full article ">Figure 3
<p>Time course hydrolysis of freeze-dried Hoki and Gemfish roe without lipid extraction. Before hydrolysis, the Hoki roe and Gemfish roe homogenates were freeze-dried. For hydrolysis, homogenates of freeze-dried Hoki and Gemfish roe were prepared by adding 2.76 g and 1.96 g of freeze-dried roe powder to 200 mL of potassium phosphate buffer (pH 7) to achieve the protein concentration 6.5 mg/mL, and aliquots of 20 mL were subjected to hydrolysis with three different concentrations (2%, 6%, and 10%) of either Alcalase (v/w), bacterial protease HT (<span class="html-italic">w/w</span>), or fungal protease FP-II (<span class="html-italic">w/w</span>) at 45 °C. The data were obtained from three independent hydrolyses for each roe and protease combination. Analysis of variance (ANOVA) was carried out. Letters a–c indicate significant differences among different samples prepared with varying enzyme concentrations at the same time point (<span class="html-italic">p</span> &lt; 0.05).</p>
Full article ">Figure 4
<p>Time course hydrolysis of freeze-dried Hoki and Gemfish roe without lipid present. Before hydrolysis, Hoki roe and Gemfish roe homogenate delipidated and freeze-dried by the ETHEX method and the freeze drier. For hydrolysis, homogenates of freeze-dried Hoki and Gemfish roe were prepared by adding 2.07 g and 2.12 g of delipidated freeze-dried roe powder to 200 mL of potassium phosphate buffer (pH 7), and aliquots of 20 mL were subject to hydrolysis with three different concentrations (2%, 6%, and 10%) of either plant-based protease Alcalase (<span class="html-italic">v/w</span>), bacterial protease HT (<span class="html-italic">w/w</span>), or fungal protease FP-II (<span class="html-italic">w/w</span>), respectively, at 45 °C. Analysis of variance (ANOVA) was carried out. Letters a–c indicate significant differences among different samples prepared with varying enzyme concentrations at the same time point (<span class="html-italic">p</span> &lt; 0.05).</p>
Full article ">Figure 5
<p>Time course hydrolysis of freeze-dried Hoki and Gemfish roe without lipid present. 1D SDS-PAGE of protein hydrolysis profiles of three proteases (Alcalase, HT, and FP-II) with the concentration [2%, 6%, and 10% (<span class="html-italic">w</span>/<span class="html-italic">v</span> or <span class="html-italic">v</span>/<span class="html-italic">v</span>)] control before 24 h incubation (<b>a</b>), control after 24 h incubation (<b>b</b>), and the sample control of Hoki and Gemfish roe homogenate with 4 treatments before hydrolysis (<b>c</b>). F = fresh, dL = delipidation, FD = freeze-dried, and FD-dL = freeze-dried with delipidation. A darker blue band indicates a higher concentration of protein.</p>
Full article ">Figure 6
<p>1D SDS-PAGE of protein hydrolysis profiles of Hoki roe (<b>a</b>,<b>c</b>,<b>e</b>) and Gemfish roe (<b>b</b>,<b>d</b>,<b>f</b>) homogenate. Alcalase (<b>a</b>,<b>b</b>), HT (<b>c</b>,<b>d</b>), and FP-II (<b>e</b>,<b>f</b>) all with the same amount of protease concentration (10% <span class="html-italic">v</span>/<span class="html-italic">w</span> or <span class="html-italic">w</span>/<span class="html-italic">w</span>) and effect of 4 treatments (F, dL, FD, and dL-FD) on protein hydrolysis of Hoki and Gemfish roe homogenate. F = fresh, dL = delipidation, FD = freeze-dried, and FD-dL: freeze-dried with delipidation. A darker blue band indicates a higher concentration of protein.</p>
Full article ">Figure 7
<p>DPPH radical scavenging activity (<b>a</b>), ABTS radical scavenging activity (<b>b</b>), and Ferric (Fe<sup>3+</sup>) reducing power (<b>c</b>). Different enzyme treatments and fish roe homogenates (<b>I</b>), different enzyme treatments and fish roe homogenates (<b>II</b>), and different treatments and enzyme treatments (<b>III</b>). (a–f) indicating a significant difference in antioxidant activity among different treatments (<span class="html-italic">p</span> &lt; 0.05). F = fresh, dL = delipidation, FD = freeze-dried, and FD-dL: freeze-dried with delipidation.</p>
Full article ">
25 pages, 1102 KiB  
Article
Co-Digestion and Mono-Digestion of Sewage Sludge and Steam-Pretreated Winter Wheat Straw in Continuous Stirred-Tank Reactors—Nutrient Composition and Process Performance
by Emma Kreuger, Virginia Tosi, Maja Lindblad and Åsa Davidsson
Fermentation 2024, 10(8), 414; https://doi.org/10.3390/fermentation10080414 (registering DOI) - 10 Aug 2024
Viewed by 334
Abstract
Wheat straw (WS) constitutes a considerable biomass resource and can be used to produce the energy carrier methane through anaerobic digestion. Due to the low contents of several nutrient elements and water in harvested WS, the use of sewage sludge (SS), consisting of [...] Read more.
Wheat straw (WS) constitutes a considerable biomass resource and can be used to produce the energy carrier methane through anaerobic digestion. Due to the low contents of several nutrient elements and water in harvested WS, the use of sewage sludge (SS), consisting of primary sludge and waste-activated sludge, as a nutrient source in co-digestion with steam-pretreated wheat straw (PWS) was investigated theoretically and practically. WS was steam-pretreated, with acetic acid as the catalyst, at 190 °C for 10 min, ending with a rapid reduction in pressure. Process stability and specific methane production were studied for the mono-digestion and co-digestion of PWS and SS in continuous stirred-tank reactors for 208 days. The HRT was 22 days and the OLR 2.1 gVS L−1 d−1. In co-digestion, the OLR was increased to 2.8 gVS L−1 d−1 for one week. Nutrient elements were added to PWS mono-digestion at two different concentration levels. Co-digestion was stable, with a total concentration of short-chain fatty acids (SCFAs) at a safe level below 0.35 g L−1 at both OLRs. The higher OLR during co-digestion would require an increase in reactor volume of 14%, compared to the mono-digestion of SS, but would increase the annual production of methane by 26%. The specific methane production levels for PWS mono-digestion, SS mono-digestion, and co-digestion were 170, 320, and 260 mL g−1VS, respectively. Co-digestion did not result in a synergistic increase in the methane yield. SCFAs accumulated in the mono-digestion of PWS when using lower levels of nutrient supplements, and the concentrations fluctuated at higher nutrient levels. The main conclusion is that PWS and SS can be co-digested with long-term process stability, without the addition of chemicals other than water and acetic acid. The specific methane production for mono-digestion of PWS was relatively low. The effect of using higher concentrations of micronutrients in PWS mono-digestion should be evaluated in future studies. Full article
(This article belongs to the Special Issue Biofuels Production and Processing Technology: 3rd Edition)
15 pages, 1286 KiB  
Article
Cultivated Winter-Type Lunaria annua L. Seed: Deciphering the Glucosinolate Profile Integrating HPLC, LC-MS and GC-MS Analyses, and Determination of Fatty Acid Composition
by Gina Rosalinda De Nicola, Sabine Montaut, Kayla Leclair, Joëlle Garrioux, Xavier Guillot and Patrick Rollin
Molecules 2024, 29(16), 3803; https://doi.org/10.3390/molecules29163803 (registering DOI) - 10 Aug 2024
Viewed by 341
Abstract
Lunaria annua L. (Brassicaceae) is an ornamental plant newly identified in Europe as a promising industrial oilseed crop for its valuable very-long-chain monounsaturated fatty acids (MUFAs), especially erucic acid (EA) and nervonic acid (NA). L. annua seeds were obtained from annual winter-type plants [...] Read more.
Lunaria annua L. (Brassicaceae) is an ornamental plant newly identified in Europe as a promising industrial oilseed crop for its valuable very-long-chain monounsaturated fatty acids (MUFAs), especially erucic acid (EA) and nervonic acid (NA). L. annua seeds were obtained from annual winter-type plants selected and cultivated in Northern France. Using a systematic multiple-method approach, we set out to determine the profile and content of glucosinolates (GSLs), which are the relevant chemical tag of Brassicaceae. Intact GSLs were analyzed through a well-established LC-MS method. Identification and quantification were performed by HPLC-PDA of desulfo-GSLs (dGLs) according to the official EU ISO method. Moreover, GSL structures were confirmed by GC-MS analysis of the related isothiocyanates (ITCs). Seven GSLs were identified, directly or indirectly, as follows: 1-methylethyl GSL, (1S)-1-methylpropyl GSL, (Rs)-5-(methylsulfinyl)pentyl GSL, (Rs)-6-(methylsulfinyl)hexyl GSL, (2S)-2-hydroxy-4-pentenyl GSL, 2-phenylethyl GSL, and 1-methoxyindol-3-ylmethyl GSL. In other respects, the FA composition of the seed oil was determined. Results revealed cultivated L. annua seed to be a source of NA-rich oil, and presscake as a valuable coproduct. This presscake is indeed rich in GSLs (4.3% w/w), precursors of promising bioactive molecules for agricultural and nutraceutical applications. Full article
Show Figures

Figure 1

Figure 1
<p>Structures of GSLs (R<sup>2</sup> = SO<sub>3</sub><sup>−</sup>) (<b>1</b>–<b>5</b>) and dGSLs (R<sup>2</sup> = H) (<b>d1</b>–<b>d7</b>) identified and quantified in cultivated <span class="html-italic">Lunaria annua</span> seed and presscake by LC-MS and HPLC-PDA analysis. Quantification of GSLs was performed by HPLC-PDA analysis of the corresponding dGSLs. Legend: 1-methylethyl GSL (glucoputranjivin) (<b>1</b>), (1<span class="html-italic">S</span>)-1-methylpropyl GSL (glucocochlearin) (<b>2</b>), (<span class="html-italic">Rs</span>)-5-(methylsulfinyl)pentyl GSL (glucoalyssin) (<b>3</b>), (<span class="html-italic">Rs</span>)-6-(methylsulfinyl)hexyl GSL (glucohesperin) (<b>4</b>), (2<span class="html-italic">S</span>)-2-hydroxy-4-pentenyl GSL (gluconapoleiferin) (<b>5</b>), 2-phenylethyl GSL (gluconasturtiin) (<b>6</b>), 1-methoxyindol-3-ylmethyl GSL (neoglucobrassicin) (<b>7</b>), d-glucoputranjivin (<b>d1</b>), d-glucocochlearin (<b>d2</b>), d-glucoalyssin (<b>d3</b>), d-glucohesperin (<b>d4</b>), d-gluconapoleiferin (<b>d5</b>), 2-phenylethyl dGSL (d-gluconasturtiin) (<b>d6</b>), and 1-methoxyindol-3-ylmethyl dGSL (d-neoglucobrassicin) (<b>d7</b>).</p>
Full article ">Figure 2
<p>HLPC chromatogram of the ethanolic seed extract of <span class="html-italic">Lunaria annua</span> L. Detection at 220 nm. Legend: 1-methylethyl GSL (glucoputranjivin) (<b>1</b>), (1<span class="html-italic">S</span>)-1-methylpropyl GSL (glucocochlearin) (hypothesis) (<b>2</b>), (<span class="html-italic">Rs</span>)-5-(methylsulfinyl)pentyl GSL (glucoalyssin) (<b>3</b>), (<span class="html-italic">Rs</span>)-6-(methylsulfinyl)hexyl GSL (glucohesperin) (<b>4</b>), (2<span class="html-italic">S</span>)-2-hydroxy-4-pentenyl GSL (gluconapoleiferin) (hypothesis) (<b>5</b>).</p>
Full article ">Figure 3
<p>HLPC-PDA chromatogram of dGSLs obtained from the ethanolic seed extract of <span class="html-italic">Lunaria annua</span> L (Laboratory A). Detection at 229 nm. Legend: 1-methylethyl dGSL (d-glucoputranjivin) (<b>d1</b>), (1<span class="html-italic">S</span>)-1-methylpropyl dGSL (d-glucocochlearin) (<b>d2</b>), (<span class="html-italic">Rs</span>)-5-(methylsulfinyl)pentyl dGSL (d-glucoalyssin) (<b>d3</b>), (<span class="html-italic">Rs</span>)-6-(methylsulfinyl)hexyl dGSL (d-glucohesperin) (<b>d4</b>).</p>
Full article ">Figure 4
<p>Structures of ITCs issued from enzymatic hydrolysis of GSLs and identified by GC-MS analysis. Legend: 1-methylethyl ITC (putranjivin) (<b>i1</b>), (1<span class="html-italic">S</span>)-1-methylpropyl ITC (<b>i2</b>), (<span class="html-italic">Rs</span>)-5-(methylsulfinyl)pentyl ITC (<span class="html-italic">R</span>-alyssin) (<b>i3</b>), (<span class="html-italic">Rs</span>)-6-(methylsulfinyl)hexyl ITC (<span class="html-italic">R</span>-hesperin) (<b>i4</b>).</p>
Full article ">Figure 5
<p>Proposed mechanistic scheme for the thermal decomposition of (<span class="html-italic">Rs</span>)-5-(methylsulfinyl)pentyl ITC (<span class="html-italic">R</span>-alyssin) (<b>i3</b>, n = 5) and (<span class="html-italic">Rs</span>)-6-(methylsulfinyl)hexyl ITC (<span class="html-italic">R</span>-hesperin) (<b>i4</b>, n = 6) to the corresponding 4-pentenyl ITC (brassicanapin, n = 5) and 5-hexenyl ITC (n = 6) detected by GC-MS in the experiment with the injector port set at 250 °C.</p>
Full article ">
22 pages, 825 KiB  
Article
The Impact of Freeze-Dried Tenebrio molitor Larvae on the Quality, Safety Parameters, and Sensory Acceptability of Wheat Bread
by Agnė Jankauskienė, Aistė Kabašinskienė, Dominykas Aleknavičius, Sandra Kiseliovienė, Sigita Kerzienė, Vytautė Starkutė, Elena Bartkienė, Monika Zimkaitė, Ignė Juknienė and Paulina Zavistanavičiūtė
Insects 2024, 15(8), 603; https://doi.org/10.3390/insects15080603 (registering DOI) - 10 Aug 2024
Viewed by 195
Abstract
The research context involves analyzing the potential benefits derived from integrating insect protein into everyday food items. Utilizing methods consistent with established food science protocols, wheat bread was prepared with variations of 0%, 5%, 10%, and 15% Tenebrio molitor larvae powder, derived from [...] Read more.
The research context involves analyzing the potential benefits derived from integrating insect protein into everyday food items. Utilizing methods consistent with established food science protocols, wheat bread was prepared with variations of 0%, 5%, 10%, and 15% Tenebrio molitor larvae powder, derived from larvae cultivated on brewery spent grain. A substrate selected for its superior nutritional content and a substrate with agar–agar gels were used. The tests included basic bread tests; sugar, acrylamide, amino, and fatty acid (FA) tests; and sensory acceptability. The results have shown that the acrylamide levels in bread with larvae remained below harmful thresholds, suggesting that using T. molitor can be a safe alternative protein source. The incorporation of powdered T. molitor larvae (p-TMLs) into bread was observed to increase certain sugar levels, such as glucose, particularly at higher larval concentrations. The addition of T. molitor significantly raised the protein and fat levels in bread. The inclusion of larvae enriched the bread with essential amino acids, enhancing the nutritional value of the bread significantly. The FA profile of the bread was altered by the inclusion of p-TMLs, increasing the levels of monounsaturated FAs. Despite the nutritional benefits, higher concentrations of larvae decreased the sensory acceptability of the bread. This suggests that there is a balance to be found between enhancing the nutritional content and maintaining consumer appeal. These findings highlight the potential for using p-TMLs as a sustainable, nutritious ingredient in bread making, although the sensory qualities at higher concentrations might limit consumer acceptance. 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 (registering DOI) - 10 Aug 2024
Viewed by 399
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 ">
17 pages, 18568 KiB  
Article
A Mixture of Soybean Oil and Lard Alleviates Postpartum Cognitive Impairment via Regulating the Brain Fatty Acid Composition and SCFA/ERK(1/2)/CREB/BDNF Pathway
by Runjia Shi, Xiaoying Tian, Andong Ji, Tianyu Zhang, Huina Xu, Zhongshi Qi, Liying Zhou, Chunhui Zhao and Duo Li
Nutrients 2024, 16(16), 2641; https://doi.org/10.3390/nu16162641 (registering DOI) - 10 Aug 2024
Viewed by 251
Abstract
Lard is highly appreciated for its flavor. However, it has not been elucidated how to consume lard while at the same time eliminating its adverse effects on postpartum cognitive function. Female mice were divided into three groups (n = 10): soybean oil [...] Read more.
Lard is highly appreciated for its flavor. However, it has not been elucidated how to consume lard while at the same time eliminating its adverse effects on postpartum cognitive function. Female mice were divided into three groups (n = 10): soybean oil (SO), lard oil (LO), and a mixture of soybean oil and lard at a ratio of 1:1 (LS). No significant difference was observed between the SO and LS groups in behavioral testing of the maternal mice, but the LO group was significantly worse compared with these two groups. Moreover, the SO and LS supplementation increased docosahexaenoic acid (DHA) and total n-3 polyunsaturated fatty acid (PUFA) levels in the brain and short-chain fatty acid (SCFA)-producing bacteria in feces, thereby mitigating neuroinflammation and lowering the p-ERK(1/2)/ERK(1/2), p-CREB/CREB, and BDNF levels in the brain compared to the LO group. Collectively, the LS group inhibited postpartum cognitive impairment by regulating the brain fatty acid composition, neuroinflammation, gut microbiota, and the SCFA/ERK(1/2)/CREB/BDNF signaling pathway compared to lard. Full article
(This article belongs to the Section Lipids)
Show Figures

Graphical abstract

Graphical abstract
Full article ">Figure 1
<p>The schematic depicting the present study design.</p>
Full article ">Figure 2
<p>Effect of a mixture of soybean oil and lard on postpartum cognitive function. (<b>A</b>–<b>C</b>) The detection results and (<b>D</b>–<b>F</b>) the representative traveled path of maternal mice in the open-field test. (<b>G</b>–<b>I</b>) The detection results and (<b>J</b>–<b>L</b>) the representative traveled path of maternal mice in the Y-maze test. (<b>M</b>–<b>O</b>) The detection results and (<b>P</b>–<b>R</b>) the representative traveled path of maternal mice in the Morris water maze test. Data represent mean ± SD (<span class="html-italic">n</span> = 7). * <span class="html-italic">p</span> &lt; 0.05 and ** <span class="html-italic">p</span> &lt; 0.01 represent the significant difference.</p>
Full article ">Figure 3
<p>Effect of a mixture of soybean oil and lard on brain fatty acid profile and its correlation with behavioral testing outcome indicators. (<b>A</b>,<b>B</b>) Brain fatty acid proportions (%). (<b>C</b>–<b>H</b>) Pearson’ correlations between DHA and <span class="html-italic">n</span>-3 PUFA levels in the maternal mice brain and behavioral testing outcome indicators. Data represent median (interquartile range) (<span class="html-italic">n</span> = 6). * <span class="html-italic">p</span> &lt; 0.05 and ** <span class="html-italic">p</span> &lt; 0.01 represent the significant difference.</p>
Full article ">Figure 4
<p>Effect of a mixture of soybean oil and lard on brain histopathology. (<b>A</b>) The H&amp;E and (<b>B</b>) Nissl staining diagrams in the brain (scale bar = 20 μm). Red arrows: nuclei pyknosis.</p>
Full article ">Figure 5
<p>Effect of a mixture of soybean oil and lard on the activation of neuroglial cells. Brain GFAP expression was analyzed via (<b>A</b>) immunofluorescence and (<b>B</b>) Western blotting (<span class="html-italic">n</span> = 4). IBA1 expression in the brain was analyzed via (<b>C</b>) immunofluorescence and (<b>D</b>) Western blotting (<span class="html-italic">n</span> = 4). Scale bar = 20 μm. Data represent median (interquartile range). * <span class="html-italic">p</span> &lt; 0.05 and ** <span class="html-italic">p</span> &lt; 0.01 represent the significant difference.</p>
Full article ">Figure 6
<p>Effect of a mixture of soybean oil and lard on brain neuroinflammation. The levels of (<b>A</b>–<b>D</b>) inflammatory cytokines (pg/g, <span class="html-italic">n</span> = 3) and (<b>E</b>–<b>H</b>) NLRP3 inflammasome complex-related proteins (<span class="html-italic">n</span> = 4). Data represent mean ± SD or median (interquartile range). * <span class="html-italic">p</span> &lt; 0.05 and ** <span class="html-italic">p</span> &lt; 0.01 represent the significant difference.</p>
Full article ">Figure 7
<p>Effect of a mixture of soybean oil and lard on the gut microbiota composition. (<b>A</b>) Chao1 index; (<b>B</b>) ACE index; (<b>C</b>) Shannon index; (<b>D</b>) Simpson index of each group. (<b>E</b>) Principal coordinates analysis (PCoA) of weighted unifrac. All phyla (<b>F</b>) and genera (<b>G</b>) of gut microbiota. Data represent median (interquartile range) (<span class="html-italic">n</span> = 8). * <span class="html-italic">p</span> &lt; 0.05 represents the significant difference.</p>
Full article ">Figure 8
<p>Effect of a mixture of soybean oil and lard on the gut microbiota composition. The (<b>A</b>) cladogram (LDA &gt; 3) and (<b>B</b>) LDA score of the taxa obtained from LEfSe analysis. (<b>C</b>) The Pearson correlation analysis between behavioral testing outcome indicators and the biomarkers in microbiota from each group. * <span class="html-italic">p</span> &lt; 0.05.</p>
Full article ">Figure 9
<p>Effect of a mixture of soybean oil and lard on the SCFA and ERK(1/2)/CREB/BDNF pathway-related protein levels. (<b>A</b>) The levels of SCFAs in the feces of maternal mice (µg/g, <span class="html-italic">n</span> = 3). (<b>B</b>–<b>F</b>) The relative protein levels of p-ERK(1/2)/ERK(1/2), p-CREB/CREB, BDNF, and PSD-95 in brain (<span class="html-italic">n</span> = 4). (<b>G</b>) The ultrastructure of synapses on the transmission electron micrograph in the hippocampus (Scale bars = 2 μm and 1 μm). Red arrows: postsynaptic density. Data represent mean ± SD or median (interquartile range). * <span class="html-italic">p</span> &lt; 0.05 and ** <span class="html-italic">p</span> &lt; 0.01 represent the significant difference.</p>
Full article ">
18 pages, 5596 KiB  
Article
Analysis of BnGPAT9 Gene Expression Patterns in Brassica napus and Its Impact on Seed Oil Content
by Man Xing, Bo Hong, Mengjie Lv, Xueyi Lan, Danhui Zhang, Chunlei Shu, Shucheng Qi, Zechuan Peng, Chunyun Guan, Xinghua Xiong and Luyao Huang
Agriculture 2024, 14(8), 1334; https://doi.org/10.3390/agriculture14081334 (registering DOI) - 10 Aug 2024
Viewed by 279
Abstract
Glycerol-3-phosphate acyltransferase (GPAT) genes encode enzymes involved in the biosynthesis of plant oils. Rapeseed has four BnGPAT9 genes, but the expression patterns and functions of these four homologous copies in rapeseed for seed oil accumulation are not well understood. In this [...] Read more.
Glycerol-3-phosphate acyltransferase (GPAT) genes encode enzymes involved in the biosynthesis of plant oils. Rapeseed has four BnGPAT9 genes, but the expression patterns and functions of these four homologous copies in rapeseed for seed oil accumulation are not well understood. In this study, we cloned the four BnGPAT9 genes and their promoters from Brassica napus and found significant differences in the expression of BnGPAT9 genes among different rapeseed varieties. We confirmed that BnGPAT9-A01/C01 are highly conserved in rapeseed, with high expression levels in various tissues, especially during the late stages of silique development and seed maturation. All four BnGPAT9 genes (BnGPAT9-A01/C01/A10/C09) can promote seed oil accumulation, but BnGPAT9-A01/C01 have a greater effect. Overexpression in Arabidopsis and rapeseed increased seed oil content and altered fatty acid composition, significantly increasing linolenic acid content. Transcriptome analysis revealed that BnGPAT9 genes promote the upregulation of genes related to oil synthesis, particularly those in the Plant–pathogen interaction, alpha-Linolenic acid metabolism, MAPK signaling pathway—plant, and Glutathione metabolism pathways. In summary, these results indicate that the four BnGPAT9 genes in rapeseed have different expression patterns and roles in regulating seed oil accumulation, with BnGPAT9-A01/C01 contributing the most to promoting oil accumulation. Full article
(This article belongs to the Special Issue Molecular Breeding and Genetic Improvement of Oilseed Crops)
Show Figures

Figure 1

Figure 1
<p>Characterization of the BnGPAT9 protein sequences. (<b>A</b>) The analysis of conserved motifs in BnGPAT9 and AtGPAT9 proteins; (<b>B</b>) the multiple sequence alignment of plant BnGPAT9 proteins.</p>
Full article ">Figure 2
<p>Cis elements detected in the promoter of the <span class="html-italic">BnGPAT9</span> genes. (<b>A</b>) Promoter element distribution, where different colors correspond to different elements in the figure below; (<b>B</b>) the heat map shows the number of promoter elements, and the gray square indicates that the elements could not be detected.</p>
Full article ">Figure 3
<p>Tissue expression patterns of <span class="html-italic">BnGPAT9</span> genes. (<b>A</b>) The GUS staining results of <span class="html-italic">Arabidopsis thaliana</span>. (<b>A1</b>,<b>B1</b>,<b>C1</b>,<b>D1</b>,<b>E1</b>,<b>F1</b>): GUS staining in stems; (<b>A2</b>,<b>B2</b>,<b>C2</b>,<b>D2</b>,<b>E2</b>,<b>F2</b>): GUS staining in leaves; (<b>A3</b>,<b>B3</b>,<b>C3</b>,<b>D3</b>,<b>E3</b>,<b>F3</b>): GUS staining in inflorescences; (<b>A4</b>,<b>B4</b>,<b>C4</b>,<b>D4</b>,<b>E4</b>,<b>F4</b>): GUS staining in flowers; (<b>A5</b>,<b>B5</b>,<b>C5</b>,<b>D5</b>,<b>E5</b>,<b>F5</b>): GUS staining in siliques. Bar = 1 mm. (<b>B</b>) Expression profiles of <span class="html-italic">BnGPAT9</span> genes in ZS11. Data sourced from BnTIR (<a href="https://yanglab.hzau.edu.cn/BnTIR" target="_blank">https://yanglab.hzau.edu.cn/BnTIR</a> (accessed on 15 October 2019)).</p>
Full article ">Figure 4
<p>Analysis of tissue-specific expression patterns of <span class="html-italic">BnGPAT9</span> genes and seed oil accumulation. (<b>A</b>) qRT-PCR detection of <span class="html-italic">BnGPAT9</span> expression in roots, stems, leaves, flowers, and seeds at 1–7 weeks of development, as well as siliques of XY15. (<b>B</b>–<b>D</b>) qRT-PCR detection of <span class="html-italic">BnGPAT9</span> expression in seeds and siliques at 1–7 weeks of development in transgenic lines overexpressing the <span class="html-italic">BnGPAT9-C01</span> gene. (<b>E</b>) Oil content in the seeds of XY15 and the three transgenic lines.</p>
Full article ">Figure 5
<p>Analysis of the thousand-seed weight and oil content in <span class="html-italic">BnGPAT9</span> transgenic Arabidopsis seeds. (<b>A</b>) Thousand-seed weight of seeds from four <span class="html-italic">BnGPAT9</span> transgenic lines. (<b>B</b>) Oil content of seeds from four <span class="html-italic">BnGPAT9</span> transgenic lines. “*” indicates highly significant differences, <span class="html-italic">p</span> &lt; 0.05. “**” indicates highly significant differences, <span class="html-italic">p</span> &lt; 0.01.</p>
Full article ">Figure 6
<p>An analysis of differential gene expression in siliques of <span class="html-italic">BnGPAT9</span> transgenic Arabidopsis. (<b>A</b>) A statistical analysis of the number of DEGs. The red bars represent upregulated genes, while the blue bars represent downregulated genes. (<b>B</b>) A Venn diagram analysis of upregulated DEGs. This diagram illustrates the common and unique upregulated genes across the different <span class="html-italic">BnGPAT9</span> transgenic lines. (<b>C</b>) A Venn diagram analysis of downregulated DEGs. This diagram shows the common and unique downregulated genes among the various <span class="html-italic">BnGPAT9</span> transgenic lines. (<b>D</b>) A clustering heatmap of DEGs. Each column represents a sample, and each row represents a gene. The color in the heatmap indicates the normalized expression level of the gene in each sample, with red representing higher expression levels and blue representing lower expression levels.</p>
Full article ">Figure 7
<p>KEGG enrichment analysis of DEGs in four <span class="html-italic">BnGPAT9</span> transgenic <span class="html-italic">Arabidopsis thaliana.</span> (<b>A</b>) KEGG enrichment analysis of DEGs between <span class="html-italic">BnGPAT9-A01</span> transgenic <span class="html-italic">Arabidopsis thaliana</span> and wild type. (<b>B</b>) KEGG enrichment analysis of DEGs between <span class="html-italic">BnGPAT9-A10</span> transgenic <span class="html-italic">Arabidopsis thaliana</span> and wild type. (<b>C</b>) KEGG enrichment analysis of DEGs between <span class="html-italic">BnGPAT9-C01</span> transgenic <span class="html-italic">Arabidopsis thaliana</span> and wild type. (<b>D</b>) KEGG enrichment analysis of DEGs between <span class="html-italic">BnGPAT9-C09</span> transgenic <span class="html-italic">Arabidopsis thaliana</span> and wild type. (<b>E</b>) Heatmap of expression of DEGs enriched in alpha-Linolenic acid metabolism pathway. “**” indicates highly significant differences, <span class="html-italic">p</span> &lt; 0.01.</p>
Full article ">
21 pages, 2255 KiB  
Article
Kalanchoe tomentosa: Phytochemical Profiling, and Evaluation of Its Biological Activities In Vitro, In Vivo, and In Silico
by Jorge L. Mejía-Méndez, Gildardo Sánchez-Ante, Yulianna Minutti-Calva, Karen Schürenkämper-Carrillo, Diego E. Navarro-López, Ricardo E. Buendía-Corona, Ma. del Carmen Ángeles González-Chávez, Angélica Lizeth Sánchez-López, J. Daniel Lozada-Ramírez, Eugenio Sánchez-Arreola and Edgar R. López-Mena
Pharmaceuticals 2024, 17(8), 1051; https://doi.org/10.3390/ph17081051 (registering DOI) - 9 Aug 2024
Viewed by 366
Abstract
In this work, the leaves of K. tomentosa were macerated with hexane, chloroform, and methanol, respectively. The phytochemical profiles of hexane and chloroform extracts were unveiled using GC/MS, whereas the chemical composition of the methanol extract was analyzed using UPLC/MS/MS. The antibacterial activity [...] Read more.
In this work, the leaves of K. tomentosa were macerated with hexane, chloroform, and methanol, respectively. The phytochemical profiles of hexane and chloroform extracts were unveiled using GC/MS, whereas the chemical composition of the methanol extract was analyzed using UPLC/MS/MS. The antibacterial activity of extracts was determined against gram-positive and gram-negative strains through the minimal inhibitory concentration assay, and in silico studies were implemented to analyze the interaction of phytoconstituents with bacterial peptides. The antioxidant property of extracts was assessed by evaluating their capacity to scavenge DPPH, ABTS, and H2O2 radicals. The toxicity of the extracts was recorded against Artemia salina nauplii and Caenorhabditis elegans nematodes. Results demonstrate that the hexane and chloroform extracts contain phytosterols, triterpenes, and fatty acids, whereas the methanol extract possesses glycosidic derivatives of quercetin and kaempferol together with sesquiterpene lactones. The antibacterial performance of extracts against the cultured strains was appraised as weak due to their MIC90 values (>500 μg/mL). As antioxidants, treatment with extracts executed high and moderate antioxidant activities within the range of 50–300 μg/mL. Extracts did not decrease the viability of A. salina, but they exerted a high toxic effect against C. elegans during exposure to treatment. Through in silico modeling, it was recorded that the flavonoids contained in the methanol extract can hamper the interaction of the NAM/NAG peptide, which is of great interest since it determines the formation of the peptide wall of gram-positive bacteria. This study reports for the first time the biological activities and phytochemical content of extracts from K. tomentosa and proposes a possible antibacterial mechanism of glycosidic derivatives of flavonoids against gram-positive bacteria. Full article
19 pages, 1341 KiB  
Review
The Impacts of Dietary Intervention on Brain Metabolism and Neurological Disorders: A Narrative Review
by Priya Rathor and Ratnasekhar Ch
Dietetics 2024, 3(3), 289-307; https://doi.org/10.3390/dietetics3030023 (registering DOI) - 9 Aug 2024
Viewed by 145
Abstract
Neurological disorders are increasing globally due to their complex nature, influenced by genetics and environmental factors. Effective treatments remain limited, and early diagnosis is challenging. Recent evidence indicates that metabolic activities play a crucial role in the onset of neural defects. Molecular changes [...] Read more.
Neurological disorders are increasing globally due to their complex nature, influenced by genetics and environmental factors. Effective treatments remain limited, and early diagnosis is challenging. Recent evidence indicates that metabolic activities play a crucial role in the onset of neural defects. Molecular changes offer new diagnostic markers and dietary targets for disease management. Diets such as MIND, DASH, omega-3 fatty acids, and polyphenols show promise in protecting brain metabolism through their anti-inflammatory properties. Personalized dietary interventions could mitigate neurodegenerative diseases. This review highlights the effects of various dietary interventions, including calorie restriction, fasting, and ketogenic diets, on neurological disorders. Additionally, it emphasizes the nutritional impacts on immunomodulation and the underlying mechanisms, including the influence of gut microbiota on brain function. Dietary interventions could serve as adjunctive therapies in disease management. Full article
Show Figures

Figure 1

Figure 1
<p>Effects of different diets on neurological disease.</p>
Full article ">Figure 2
<p>Examples of macronutrients and micronutrients.</p>
Full article ">Figure 3
<p>Different dietary intervention controls the brain pathways.</p>
Full article ">
18 pages, 8359 KiB  
Article
Membrane Damage and Metabolic Disruption as the Mechanisms of Linalool against Pseudomonas fragi: An Amino Acid Metabolomics Study
by Jiaxin Cai, Haiming Chen, Runqiu Wang, Qiuping Zhong, Weijun Chen, Ming Zhang, Rongrong He and Wenxue Chen
Foods 2024, 13(16), 2501; https://doi.org/10.3390/foods13162501 - 9 Aug 2024
Viewed by 430
Abstract
Pseudomonas fragi (P. fragi) is usually detected in low-temperature meat products, and seriously threatens food safety and human health. Therefore, the study investigated the antibacterial mechanism of linalool against P. fragi from membrane damage and metabolic disruption. Results from field-emission transmission [...] Read more.
Pseudomonas fragi (P. fragi) is usually detected in low-temperature meat products, and seriously threatens food safety and human health. Therefore, the study investigated the antibacterial mechanism of linalool against P. fragi from membrane damage and metabolic disruption. Results from field-emission transmission electron microscopy (FETEM) and atomic force microscopy (AFM) showed that linalool damage membrane integrity increases surface shrinkage and roughness. According to Fourier transform infrared (FTIR) spectra results, the components in the membrane underwent significant changes, including nucleic acid leakage, carbohydrate production, protein denaturation and modification, and fatty acid content reduction. The data obtained from amino acid metabolomics indicated that linalool caused excessive synthesis and metabolism of specific amino acids, particularly tryptophan metabolism and arginine biosynthesis. The reduced activities of glucose 6-phosphate dehydrogenase (G6PDH), malate dehydrogenase (MDH), and phosphofructokinase (PFK) suggested that linalool impair the respiratory chain and energy metabolism. Meanwhile, genes encoding the above enzymes were differentially expressed, with pfkB overexpression and zwf and mqo downregulation. Furthermore, molecular docking revealed that linalool can interact with the amino acid residues of G6DPH, MDH and PFK through hydrogen bonds. Therefore, it is hypothesized that the mechanism of linalool against P. fragi may involve cell membrane damage (structure and morphology), disturbance of energy metabolism (TCA cycle, EMP and HMP pathway) and amino acid metabolism (cysteine, glutamic acid and citrulline). These findings contribute to the development of linalool as a promising antibacterial agent in response to the food security challenge. Full article
Show Figures

Figure 1

Figure 1
<p>AFM images and FETEM micrographs of <span class="html-italic">P. fragi</span> treated with or without linalool for 4 h. MIC: cells treated with 1.5 mL/L of linalool. Ethanol: cells treated with ethanol. Control: cells treated with water.</p>
Full article ">Figure 2
<p>Calcein-AM/PI dual-stained confocal laser scanning microscopy (<b>A</b>). Representative FTIR spectra (4000–400 cm<sup>−1</sup>) of <span class="html-italic">P. fragi</span> (<b>B</b>). MIC: cells treated with 1.5 mL/L of linalool. Ethanol: cells treated with ethanol. Control: cells treated with water.</p>
Full article ">Figure 3
<p>SDS-PAGE analysis of intracellular soluble proteins of <span class="html-italic">P. fragi</span> (<b>A</b>). DNA release results based on gel electrophoresis in <span class="html-italic">P. fragi</span> (<b>B</b>). MIC: cells treated with 1.5 mL/L of linalool. Ethanol: cells treated with ethanol. Control: cells treated with water.</p>
Full article ">Figure 4
<p>Inhibitory effect of linalool on the enzymatic activity: G6DPH activity (<b>A</b>), MDH activity (<b>B</b>), PFK activity (<b>C</b>). MIC: cells treated with 1.5 mL/L of linalool. Ethanol: cells treated with ethanol. Control: cells treated with water.</p>
Full article ">Figure 5
<p>Score plots of PCA (<b>A</b>) and OPLS-DA (<b>B</b>). Volcano plot (<b>C</b>). Upregulated, downregulated and non-significant differential metabolites are represented by red, blue, grey dots, respectively. Heat map (<b>D</b>).</p>
Full article ">Figure 6
<p>Bubble plot (<b>A</b>); each bubble represents a metabolic pathway. KEGG pathway classification (<b>B</b>). Enrichment analysis of the amino acids present as a tree map (<b>C</b>).</p>
Full article ">Figure 7
<p>Pathway analysis of enzymes and amino acids related to amino acid metabolism, EMP (green arrows), HMP (orange arrows) and TCA cycle (purple arrows) in <span class="html-italic">P. fragi</span> with linalool treatment; red color represents upregulation of amino acids and blue color represents downregulation of amino acids.</p>
Full article ">Figure 8
<p>Validation of selected DEGs using real-time PCR. *** <span class="html-italic">p</span> &lt; 0.001.</p>
Full article ">Figure 9
<p>Molecular docking perspective of linalool with the binding sites of G6PDH (<b>A</b>), MDH (<b>B</b>) and PFK (<b>C</b>). Ligand: green color, macromolecule: pink color, receptor: orange color.</p>
Full article ">
14 pages, 1897 KiB  
Article
The Association between IL-1β and IL-18 Levels, Gut Barrier Disruption, and Monocyte Activation during Chronic Simian Immunodeficiency Virus Infection and Long-Term Suppressive Antiretroviral Therapy
by Siva Thirugnanam, Chenxiao Wang, Chen Zheng, Brooke F. Grasperge, Prasun K. Datta, Jay Rappaport, Xuebin Qin and Namita Rout
Int. J. Mol. Sci. 2024, 25(16), 8702; https://doi.org/10.3390/ijms25168702 - 9 Aug 2024
Viewed by 217
Abstract
HIV-induced persistent immune activation is a key mediator of inflammatory comorbidities such as cardiovascular disease (CVD) and neurocognitive disorders. While a preponderance of data indicate that gut barrier disruption and microbial translocation are drivers of chronic immune activation, the molecular mechanisms of this [...] Read more.
HIV-induced persistent immune activation is a key mediator of inflammatory comorbidities such as cardiovascular disease (CVD) and neurocognitive disorders. While a preponderance of data indicate that gut barrier disruption and microbial translocation are drivers of chronic immune activation, the molecular mechanisms of this persistent inflammatory state remain poorly understood. Here, utilizing the nonhuman primate model of Human Immunodeficiency Virus (HIV) infection with suppressive antiretroviral therapy (ART), we investigated activation of inflammasome pathways and their association with intestinal epithelial barrier disruption (IEBD). Longitudinal blood samples obtained from rhesus macaques with chronic SIV infection and long-term suppressive ART were evaluated for IEBD biomarkers, inflammasome activation (IL-1β and IL-18), inflammatory cytokines, and triglyceride (TG) levels. Activated monocyte subpopulations and glycolytic potential were investigated in peripheral blood mononuclear cells (PBMCs). During the chronic phase of treated SIV infection, elevated levels of plasma IL-1β and IL-18 were observed following the hallmark increase in IEBD biomarkers, intestinal fatty acid-binding protein (IFABP) and LPS-binding protein (LBP). Further, significant correlations of plasma IFABP levels with IL-1β and IL-18 were observed between 10 and 12 months of ART. Higher levels of sCD14, IL-6, and GM-CSF, among other inflammatory mediators, were also observed only during the long-term SIV + ART phase along with a trend of increase in the frequencies of activated CD14+CD16+ intermediate monocyte subpopulations. Lastly, we found elevated levels of blood TG and higher glycolytic capacity in PBMCs of chronic SIV-infected macaques with long-term ART. The increase in circulating IL-18 and IL-1β following IEBD and their significant positive correlation with IFABP suggest a connection between gut barrier disruption and inflammasome activation during chronic SIV infection, despite viral suppression with ART. Additionally, the increase in markers of monocyte activation, along with elevated TG and enhanced glycolytic pathway activity, indicates metabolic remodeling that could fuel metabolic syndrome. Further research is needed to understand the mechanisms by which gut dysfunction and inflammasome activation contribute to HIV-associated metabolic complications, enabling targeted interventions in people with HIV. Full article
Show Figures

Figure 1

Figure 1
<p>Study design and viral loads. (<b>A</b>) Six rhesus macaques (RM) were inoculated with SIV and treated with combination ART consisting of daily subcutaneous injection of Tenofovir Disoproxil Fumarate (TDF), Emtricitabine (FTC), and Dolutegravir (DTG). (<b>B</b>) SIV RNA quantification in plasma over 14 months of infection in the study. The dashed line represents the beginning of ART. Each symbol represents an animal and the red line is the average viral load for all animals.</p>
Full article ">Figure 2
<p>Plasma levels of inflammasome/caspase-1 pathway activation and IEBD biomarkers. Longitudinal plasma levels of inflammasome activated cytokines IL-1β (<b>A</b>) and IL-18 (<b>B</b>) measured by Luminex assay in SIV-infected rhesus macaques. Plasma levels of IFABP (<b>C</b>), a marker of enterocyte loss and generalized damage to the intestinal epithelium and LBP (<b>D</b>), a marker of host response to LPS via microbial translocation measured at serial time-points pre-SIV baseline and post-SIV + ART. Two technical replicates were used for each time-point. One-way ANOVA with Tukey’s multiple comparisons test was used to determine significant differences between baseline and different time points post-SIV infection and ART. Asterisks indicate significant differences between time points (* <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). Associations of IFABP with IL-1β (<b>E</b>) and IL-18 (<b>F</b>) showing significant positive correlations during 10–12 months of ART. Non-significant correlations between LBP and IL-1β (<b>G</b>) and IL-18 (<b>H</b>) in the bottom panel. Statistical correlations were investigated by the Spearman correlation coefficient and 95% confidence limits. Statistical significance (<span class="html-italic">p</span> values) and Spearman’s coefficient of rank correlation (r) are shown for significant correlations.</p>
Full article ">Figure 3
<p>Circulating levels of caspase-1 and associations with IEBD biomarkers. (<b>A</b>) Longitudinal serum levels of caspase-1 (Casp-1) measured by ELISA. Associations of caspase-1 with IFABP (<b>B</b>) and sCD14 (<b>C</b>) showing significant positive correlation for IFABP during 10–12 months of ART. Statistical significance (<span class="html-italic">p</span> value) and Spearman’s coefficient of rank correlation (r) are shown for significant correlation.</p>
Full article ">Figure 4
<p>Circulating levels of inflammatory cytokines and frequencies of monocyte subpopulations in PBMC. Plasma levels of sCD14 (<b>A</b>), IL-6 (<b>B</b>), TNF-<span class="html-italic">α</span> (<b>C</b>), and MCP-1 (<b>D</b>) at pre-SIV, 1 month post-SIV and 12 months post-SIV infection. Asterisks indicate significant differences between time-points (* <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). Frequencies of circulating monocyte subsets comprising CD14<sup>+</sup>CD16<sup>+</sup> intermediate monocytes (<b>E</b>), CD14<sup>+</sup>CD16<sup>−</sup> classical monocytes (<b>F</b>), and CD14<sup>lo/neg</sup>CD16<sup>++</sup> nonclassical monocytes (<b>G</b>) in PBMC at the 0-, 1-, and 12-month post-SIV time-points. (<b>H</b>) Association between plasma IFABP levels and frequencies of intermediate monocyte at 12 months of ART. Statistical significance (<span class="html-italic">p</span> value) and Spearman’s coefficient of rank correlation (r) are shown.</p>
Full article ">Figure 5
<p>Circulating levels of triglycerides and PBMC glycolytic capacity during chronic SIV-infection under long-term virus-suppressive ART. (<b>A</b>) Triglyceride (TG) levels evaluated by blood chemistry on serial blood draws from 2 to 12 months of ART in SIV-infected macaques (* <span class="html-italic">p</span> &lt; 0.05, ** <span class="html-italic">p</span> &lt; 0.01; ns, not significant). (<b>B</b>,<b>C</b>) Representative extracellular acidification rate (ECAR) measurements in one SIV-naïve control rhesus macaque (RM1) and one SIV + RM at 12 months of ART. (<b>D</b>) Increased glycolytic capacity in Ox-LDL-treated and -untreated PBMCs from SIV-infected ART-treated RMs (blue bars) in comparison to Ox-LDL-treated and -untreated PBMCs from SIV-naïve RMs (red bars). The extracellular acidification rate (ECAR) was measured in 0.2 × 10<sup>5</sup> PBMCs from both groups that were left untreated or treated with Ox-LDL (10 μg/mL) for 24 h. Each dot represents the mean of 5 replicates from 2 animals. Asterisk (*) represents <span class="html-italic">p</span> &lt; 0.05 significant differences determined using Student’s <span class="html-italic">t</span>-test.</p>
Full article ">
15 pages, 2334 KiB  
Article
A Comparative Study on the Muscle and Gut Microbiota of Opsariichthys bidens from Rice Field and Pond Culture Breeding Modes
by Fan Zhou, Weichao Bu, Hongjie Fan, Shuirong Guo, Ming Qi, Gaohua Yao, Yijiang Bei, Yuanfei Huang, Shicheng Zhu, Xueyan Ding and Xingwei Xiang
Metabolites 2024, 14(8), 443; https://doi.org/10.3390/metabo14080443 - 9 Aug 2024
Viewed by 327
Abstract
To investigate difference in the quality of the different parts (back, tail muscles, and fish skin) of Opsariichthys bidens from pond and rice field cultures, a comparative study was conducted in terms of nutritional composition, volatile flavor profiles and gut microbiota. In detail, [...] Read more.
To investigate difference in the quality of the different parts (back, tail muscles, and fish skin) of Opsariichthys bidens from pond and rice field cultures, a comparative study was conducted in terms of nutritional composition, volatile flavor profiles and gut microbiota. In detail, the texture, free amino acids, fatty acids were further assessed. The results suggested that the moisture content, crude protein and crude fat content in the skin of O. bidens are higher than those in the back and tail muscles, regardless of breeding modes. The fish cultured in the rice field had a higher protein content than those from the pond culture, while the fat content of the rice field-cultured fish was significantly low compared to the fish from the pond culture, especially in the back and tail parts. A total of 43 volatile components were detected by Gas Chromatography–Mass Spectrometry (GC-MS), with a maximum of 18 types of aldehydes and the highest concentration being nonanal. Compared to pond cultures, the fish from the rice field cultures showed more abundant flavor composition and odor-active compounds. The total content of DHA (Docosahexaenoic Acid) and EPA (Eicosapentaenoic Acid) in the rice field-cultured fish was higher than that of the pond group, while no significant disparity in amino acid composition was observed (p > 0.05). Comparative and clustering analyses of gut microbiota revealed notable discrepancies in the gut microbiota of O. bidens from two aquaculture systems. However, an inherent correlation between the gut microbiome and meat quality would be further emphasized in further studies. This study can offer a theoretical reference for the development of high-quality aquatic products by selecting the appropriate aquaculture models. Full article
(This article belongs to the Special Issue Metabolism and Nutrition in Fish)
Show Figures

Figure 1

Figure 1
<p>(<b>A</b>) Chord chart of the proportion of volatile components in <span class="html-italic">Opsariichthys bidens</span> under different cultivation modes; (<b>B</b>) quantity of volatile flavor compounds in <span class="html-italic">Opsariichthys bidens</span> under different cultivation modes. PB: back muscle of pond-cultured fish; PT: tail of pond-cultured fish; PS: skin of pond-cultured fish; RFB: back muscle of rice-bred fish; RFT: tail of rice-bred fish; RFS: skin of rice-bred fish.</p>
Full article ">Figure 2
<p>(<b>A</b>) Venn diagram of intestinal microflora of <span class="html-italic">Opsariichthys bidens</span> in two farming modes; (<b>B</b>) Community barbolt analysis of intestinal microflora in two farming modes of <span class="html-italic">Opsariichthys bidens.</span> RF: rice field aquaculture; PD: pond aquaculture.</p>
Full article ">Figure 3
<p>(<b>A</b>) Principal component analysis of intestinal microflora in two farming methods; PC1 (Principal Coordinate 1) and PC2 (Principal Coordinate 2) are used to describe the two main dimensions of differences and similarities between samples. (<b>B</b>) According to the hierarchical clustering tree of the two breeding modes, the samples are clustered according to the similarity, and the branch length between the samples is negatively correlated with the similarity. RF: rice field aquaculture; PD: pond aquaculture.</p>
Full article ">Figure 4
<p>Two-level diagram of bacterial distribution in 12 samples. The heatmap represents the relative abundance of each bacterial genus (variables clustered on the vertical axis) in each sample (horizontal clustering). The values of bacterial genera are expressed as color intensity, as shown in the legend on the right side of the figure. RF: rice field aquaculture; PD: pond aquaculture.</p>
Full article ">
Back to TopTop