[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

Article Types

Countries / Regions

remove_circle_outline
remove_circle_outline
remove_circle_outline
remove_circle_outline

Search Results (428)

Search Parameters:
Keywords = pyruvate dehydrogenase

Order results
Result details
Results per page
Select all
Export citation of selected articles as:
18 pages, 14538 KiB  
Article
The Protective Effect of Quercetin against the Cytotoxicity Induced by Fumonisin B1 in Sertoli Cells
by Jun Ma, Ruixue Huang, Huai Zhang, Dongju Liu, Xiaodong Dong, Yan Xiong, Xianrong Xiong, Daoliang Lan, Wei Fu, Honghong He, Jian Li and Shi Yin
Int. J. Mol. Sci. 2024, 25(16), 8764; https://doi.org/10.3390/ijms25168764 - 12 Aug 2024
Viewed by 221
Abstract
Fumonisin B1 (FB1), a mycotoxin produced by Fusarium species, is prevalent in crops and animal feed, posing significant health risks to livestock and humans. FB1 induces oxidative stress in Sertoli cells, destroys testicular structure, and affects spermatogenesis. However, methods to mitigate the reproductive [...] Read more.
Fumonisin B1 (FB1), a mycotoxin produced by Fusarium species, is prevalent in crops and animal feed, posing significant health risks to livestock and humans. FB1 induces oxidative stress in Sertoli cells, destroys testicular structure, and affects spermatogenesis. However, methods to mitigate the reproductive toxicity of FB1 in testes remain unknown. Quercetin, a natural flavonoid antioxidant, may offer protective benefits. This study investigated the protective effects and mechanisms of quercetin against FB1-induced reproductive toxicity in TM4 cells (a Sertoli cell line). The results indicated that 40 μM quercetin improved cell viability, reduced apoptosis, and preserved cell functions. Quercetin also decreased reactive oxygen species (ROS) levels in TM4 cells exposed to FB1, enhanced the expression of antioxidant genes, and improved mitochondrial membrane potential. Compared with FB1 alone, the combination of quercetin and FB1 increased ATP levels, as well as pyruvate and lactic acid, the key glycolysis products. Furthermore, this combination elevated the mRNA and protein expression of glycolysis-related genes, including glucose-6-phosphate isomerase 1 (Gpi1), hexokinase 2 (Hk2), aldolase (Aldoa), pyruvate kinase, muscle (Pkm), lactate dehydrogenase A (Ldha) and phosphofructokinase, liver, B-type (Pfkl). Quercetin also boosted the activity of PKM and LDHA, two crucial glycolytic enzymes. In summary, quercetin mitigates FB1-induced toxicity in TM4 cells by reducing ROS levels and enhancing glycolysis. This study offers new insights into preventing and treating FB1-induced toxic damage to the male reproductive system and highlights the potential application of quercetin. Full article
Show Figures

Figure 1

Figure 1
<p>Cytotoxicity detection of FB1 on the viability of the TM4 cell line. (<b>A</b>) The CCK-8 assay was performed to detect the proliferation rates of TM4 cells treated with different concentrations (0, 5, 10, 20, 40, and 80 μM) of FB1. * <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 indicated statistical significance at different levels. One-way ANOVA was employed (with Tukey’s multiple comparison test as the post hoc test). (<b>B</b>) Apoptotic cells were detected by TUNEL staining in the Control and 80 μM FB1 group. White arrow indicates the apoptotic cells. Bar: 100 μm. (<b>C</b>) Relative mRNA levels of several apoptosis and proliferation-related genes in Control and FB1 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; 0.001 indicated statistical significance at different levels; Student’s unpaired <span class="html-italic">t</span>-test was employed. Data presented in (<b>A</b>,<b>C</b>) are expressed as mean ± SEM from three independent experiments.</p>
Full article ">Figure 2
<p>Effect of quercetin on the viability of TM4 cell lines treated with fumonisin B1 (FB1). (<b>A</b>) The CCK-8 assay was performed to detect the proliferation rates of TM4 cells treated with different concentrations (0, 5, 10, 20, 40, 60, 80, and 100 μM) of quercetin. (<b>B</b>) Representative images of cell morphology in Control, 80 μM FB1 (FB1) and 80 μM FB1 + 40 μM quercetin (FB1 + QR) groups. Bar: 100 μm. (<b>C</b>) Cell viability in Control, FB1, and FB1 + QR groups. (<b>D</b>) Apoptotic cells detected by TUNEL staining in Control, FB1, and FB1 + QR groups. White arrows indicate the apoptotic cells. Bar: 100 μm. (<b>E</b>) Apoptosis ratio analysis in (<b>D</b>). (<b>F</b>) Relative mRNA levels of several proliferation and apoptosis-related genes in Control, FB1, and FB1 + QR groups. * <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 indicated statistical significance at different levels; one-way analysis of variance (ANOVA) was employed (with Tukey’s multiple comparisons test as the post hoc test). Data presented in (<b>A</b>,<b>C</b>,<b>E</b>,<b>F</b>) are expressed as mean ± standard error of the mean (SEM) from three independent experiments.</p>
Full article ">Figure 3
<p>The impact of quercetin treatment on mRNA expressions of Sertoli cell function-related genes in FB1-treated TM4 cells. (<b>A</b>) Relative mRNA levels of several Sertoli cell secretory factors responsible for spermatogenesis in Control, FB1, and FB1 + QR groups. (<b>B</b>) Relative mRNA levels of several genes encoding BTB-related proteins in Control, FB1, and FB1 + QR groups. * <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 indicated statistical significance at different levels; one-way ANOVA was employed (with Tukey’s multiple comparisons test as the post hoc test). Data presented are expressed as mean ± SEM from three independent experiments.</p>
Full article ">Figure 4
<p>Quercetin attenuates FB1-induced oxidative damage in the TM4 cell line. (<b>A</b>) Detection of ROS levels in Control, FB1, and FB1 + QR groups by treating cells with DCFH-DA diacetate. Bar: 100 μm. (<b>B</b>) Reactive oxygen species (ROS) levels in (<b>A</b>) were quantified by measuring the fluorescence intensity of DCFH-DA. (<b>C</b>) Detection of malondialdehyde (MDA) levels in Control, FB1, and FB1 + QR groups. (<b>D</b>) Relative mRNA levels of several antioxidant genes in Control, FB1, and FB1 + QR groups. (<b>E</b>) Protein expressions of antioxidant enzymes superoxide dismutase 1 (SOD1) and peroxiredoxin 1 (PRDX1) in Control, FB1, and FB1 + QR groups. BR1 and BR2 represents two independent biological replicates, the same as below. (<b>F</b>) Optical density analysis of protein expressions of SOD1 and PRDX1 in (<b>E</b>). (<b>G</b>) SOD activity detection in Control, FB1, and FB1 + QR groups. * <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 indicated statistical significance at different levels; one-way ANOVA was employed (with Tukey’s multiple comparisons test as the post hoc test). Data presented in (<b>B</b>–<b>E</b>,<b>G</b>) are expressed as mean ± SEM from three independent experiments.</p>
Full article ">Figure 5
<p>Quercetin restores mitochondrial membrane potential (MMP) in TM4 cells treated by FB1<b>.</b> (<b>A</b>) Mitochondrial membrane potential was detected by JC-1 fluorescent probe in Control, FB1, and FB1 + QR groups. JC1-monomers and JC1-aggregates produce green and red fluorescence, respectively. Bar: 50 μm. (<b>B</b>) Mitochondrial membrane potential (MMP) in (<b>A</b>) was quantified by measuring the ratio of fluorescence intensity of JC1-aggregates to JC1-monomers. *** <span class="html-italic">p</span> &lt; 0.001 indicated statistical significance at different levels; one-way ANOVA was employed (with Tukey’s multiple comparisons test as the post hoc test). Data presented in (<b>B</b>) are expressed as mean ± SEM from three independent experiments.</p>
Full article ">Figure 6
<p>Transcriptome analysis of TM4 cell line treated with FB1 and quercetin. (<b>A</b>) Venn Diagram of differentially expressed genes (DEGs) in Control versus (vs.) FB1, Control vs. FB1 + QR, and FB1 vs. FB1 + QR groups. (<b>B</b>–<b>D</b>) Volcano plot of all transcripts in Control vs. FB1 (<b>B</b>), Control vs. FB1 + QR (<b>C</b>), and FB1 vs. FB1 + QR groups (<b>D</b>). The Y-axis represents the negative logarithm value of the error detection rate. Blue, red, and green dots represent genes with no significant change, significantly up-regulated genes, and significantly down-regulated genes. (<b>E</b>) KEGG analysis of down-regulated gene enriched processes in FB1 vs. FB1 + QR groups. (<b>F</b>) KEGG analysis of up-regulated gene enriched processes in FB1 vs. FB1 + QR groups.</p>
Full article ">Figure 7
<p>Quercetin enhances glycolysis in FB1-treated TM4 Cells. (<b>A</b>–<b>D</b>) Intracellular contents of (<b>A</b>) ATP, (<b>B</b>) glucose, (<b>C</b>) pyruvate, (<b>D</b>) and lactic acid. * <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 indicated statistical significance at different levels; one-way ANOVA was employed (with Tukey’s multiple comparisons test as the post hoc test). Data presented are expressed as mean ± SEM from three independent experiments.</p>
Full article ">Figure 8
<p>Quercetin regulates the expression and activity of glycolysis-related enzymes. (<b>A</b>) Relative mRNA levels of several antioxidant glycolysis-related genes in Control, FB1, and FB1 + QR groups. (<b>B</b>) Relative protein expression of glycolysis-related proteins PKM2 and LDHA in Control, FB1, and FB1 + QR groups. (<b>C</b>) Optical density analysis of protein expressions of PKM2 and LDHA in (<b>B</b>). (<b>D</b>) Activities of PKM and LDHA in different groups. * <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 indicated statistical significance at different levels; one-way ANOVA was employed (with Tukey’s multiple comparisons test as the post hoc test). Data presented are expressed as mean ± SEM from three independent experiments.</p>
Full article ">
16 pages, 5630 KiB  
Article
Angiotensin II Alters Mitochondrial Membrane Potential and Lipid Metabolism in Rat Colonic Epithelial Cells
by Darby D. Toth, Christopher L. Souder, Sarah Patuel, Cole D. English, Isaac Konig, Emma Ivantsova, Wendi Malphurs, Jacqueline Watkins, Kaylie Anne Costa, John A. Bowden, Jasenka Zubcevic and Christopher J. Martyniuk
Biomolecules 2024, 14(8), 974; https://doi.org/10.3390/biom14080974 - 9 Aug 2024
Viewed by 341
Abstract
An over-active renin-angiotensin system (RAS) is characterized by elevated angiotensin II (Ang II). While Ang II can promote metabolic and mitochondrial dysfunction in tissues, little is known about its role in the gastrointestinal system (GI). Here, we treated rat primary colonic epithelial cells [...] Read more.
An over-active renin-angiotensin system (RAS) is characterized by elevated angiotensin II (Ang II). While Ang II can promote metabolic and mitochondrial dysfunction in tissues, little is known about its role in the gastrointestinal system (GI). Here, we treated rat primary colonic epithelial cells with Ang II (1–5000 nM) to better define their role in the GI. We hypothesized that Ang II would negatively affect mitochondrial bioenergetics as these organelles express Ang II receptors. Ang II increased cellular ATP production but reduced the mitochondrial membrane potential (MMP) of colonocytes. However, cells maintained mitochondrial oxidative phosphorylation and glycolysis with treatment, reflecting metabolic compensation with impaired MMP. To determine whether lipid dysregulation was evident, untargeted lipidomics were conducted. A total of 1949 lipids were detected in colonocytes spanning 55 distinct (sub)classes. Ang II (1 nM) altered the abundance of some sphingosines [So(d16:1)], ceramides [Cer-AP(t18:0/24:0)], and phosphatidylcholines [OxPC(16:0_20:5(2O)], while 100 nM Ang II altered some triglycerides and phosphatidylserines [PS(19:0_22:1). Ang II did not alter the relative expression of several enzymes in lipid metabolism; however, the expression of pyruvate dehydrogenase kinase 2 (PDK2) was increased, and PDK2 can be protective against dyslipidemia. This study is the first to investigate the role of Ang II in colonic epithelial cell metabolism. Full article
Show Figures

Figure 1

Figure 1
<p>Cytotoxicity of Ang II to colonocytes at 72 h. (<b>A</b>) Cytotoxicity, (<b>B</b>) Cell viability. The lysis control was used as a positive control for the assay (induces cell death of colonocytes). The columns represent the mean relative fluorescence ± standard deviation. Different letters denote significant differences from the media-only control (One-way ANOVA, Dunnett multiple comparison test, n = 4/experiment, significance determined at <span class="html-italic">p</span> &lt; 0.05).</p>
Full article ">Figure 2
<p>ATP levels after exposure to Ang II at 72 h. Carbonyl cyanide-4-phenylhydrazone (FCCP) was used as a positive control. Relative luminescence is graphed for each experimental group (horizontal bar represents mean relative luminescence ± standard deviation). Asterisks (****) denotes significant differences from the media-only control (One-way ANOVA followed by a Dunnett multiple comparison test, n = 4/experiment, significance determined at <span class="html-italic">p</span> &lt; 0.001).</p>
Full article ">Figure 3
<p>Mitochondrial membrane potential (MMP) after exposure to Ang II at 72 h. Carbonyl cyanide-4-phenylhydrazone (FCCP) was used as a positive control as it acts as an uncoupling agent for mitochondrial membranes. Relative fluorescence is based on the red/green signal intensity, and all data are normalized to the media-only control (mean relative fluorescence ± standard deviation). Asterisk denotes significant differences compared to the media-only control (one-way ANOVA followed by a Dunnett multiple comparison test, n = 4/experiment, significance determined at * <span class="html-italic">p</span> &lt; 0.05, ** <span class="html-italic">p</span> &lt; 0.01, *** <span class="html-italic">p</span> &lt; 0.001, **** <span class="html-italic">p</span> &lt; 0.0001).</p>
Full article ">Figure 4
<p>Normalized oxygen consumption rate and extracellular acidification rate for rat epithelial colonocytes after a 24 h exposure to Ang II. (<b>A</b>) Oxygen consumption rates over time (<b>B</b>) Acidification rates over time. Data are represented as mean ± standard deviation (one-way ANOVA followed by a Dunnett multiple comparison test, n = 4 replicates/groups, significance determined at <span class="html-italic">p</span> &lt; 0.05).</p>
Full article ">Figure 5
<p>Lipid abundance and categorical classification of lipids in rat epithelial colonocytes (all lipids detected in all treatments). Abbreviations: triglycerides (TG), plasmanyl-TG (plasmanyl-triglycerides), phosphatidylcholine (PC), phosphatidylethanolamines (PE), ceramide (Cer), diacylglycerol (DG), plasmanyl-PC (plasmanyl-phosphatidylcholine), plasmenyl-PE (plasmenyl- phosphatidylethanolamines), phosphatidylserines (PS), plasmenyl-PS (plasmenyl-phosphatidylethanolamines), oxidized phosphatidylcholines (OxPC), phosphatidylglycerol (PG), phosphoinositide (PI), oxidized phosphatidylethanolamines (PE), dimethyl-phosphatidylethanolamine (DMPE), hemibismonoacylglycerophosphate (HBMP), plasmenyl-PC (plasmenyl-phosphatidylcholine), polyethylene glycol (PEG), oxidized lysophosphatidylcholines (OxLPC), lysophosphatidylcholines (LPC), oxidized triglycerides (OxTG), cardiolipins (CL), monomethyl-phosphatidylethanolamine (MMPE), lysophosphatidylethanolamine (LPE), and glucosylceramide non-hydroxyfatty acid-sphingosine (HexCer-NS).</p>
Full article ">Figure 6
<p>(<b>A</b>-Top graph) Principal component analysis scores plot for rat colonocyte lipids with each point representing the lipids in a single sample, the ellipses representing the 95% confidence interval, and the colored groups representing the three different treatments (blue = control, red = low Ang II, and green = high Ang II). (<b>B</b>-bottom graph) Heatmap showing significant changes in the levels of lipids following exposure to Ang II. Data were subjected to ANOVA followed by Fisher’s least significant difference method (Fisher’s LSD), and significant changes were set at <span class="html-italic">p</span> &lt; 0.05.</p>
Full article ">Figure 7
<p>Volcano plots of (<b>A</b>) 1 nM of Ang II and (<b>B</b>) 100 nM of Ang II and the differentially abundant lipids (<span class="html-italic">p</span> &lt; 0.05) outlined in red and blue.</p>
Full article ">Figure 8
<p>Relative concentrations of lipid abundance by sample weight in rat epithelial colonocytes exposed to 1 nM of Ang II (<b>left panel</b>). The most abundant lipids measured include So(d16:1) and Cer-AP(t18:0/24:0). Abbreviations: sphingosine (So), alpha-hydroxy-fatty acid phytosphingosine ceramide (Cer-AP), oxidized lysophosphatidylcholines (OxLPC), and phosphatidylethanolamines (PE). Relative concentrations of lipid abundance by sample weight in rat epithelial colonocytes exposed to 100 nM of Ang II (<b>right panel</b>). The most abundant lipid measured was So(d16:1). Abbreviations: sphingosine (So), oxidized phosphatidylcholines (OxPC), phosphatidylethanolamines (PE), and phosphatidylserines (PS). The black dots represent the metabolite levels in all samples, and the yellow diamond represents the average value.</p>
Full article ">Figure 9
<p>Relative gene expression for rat colonocytes after exposure to Ang II. (<b>a</b>) <span class="html-italic">PDK1,</span> (<b>b</b>) <span class="html-italic">PDK2</span>, (<b>c</b>) <span class="html-italic">PDK4</span>. Data are represented as mean ± standard deviation. Asterisks (**) denote significant differences from the media-only control (data were evaluated using a Mann–Whitney U test, n = 4/experiment, significance determined at <span class="html-italic">p</span> &lt; 0.01).</p>
Full article ">
26 pages, 2104 KiB  
Article
Effects of Dietary Chitosan on Growth Performance, Serum Biochemical Indices, Antioxidant Capacity, and Immune Response of Juvenile Tilapia (Oreochromis niloticus) under Cadmium Stress
by Qin Zhang, Yi Xie, Jiaqiong Tang, Liuqing Meng, Enhao Huang, Dongsheng Liu, Tong Tong, Yongqiang Liu and Zhongbao Guo
Animals 2024, 14(15), 2259; https://doi.org/10.3390/ani14152259 - 3 Aug 2024
Viewed by 334
Abstract
The objective of this study was to examine the effects of varying levels of dietary chitosan supplementation on mitigating cadmium stress and its influence on growth performance, serum biochemical indices, antioxidant capacity, immune response, inflammatory response, and the expression of related genes in [...] Read more.
The objective of this study was to examine the effects of varying levels of dietary chitosan supplementation on mitigating cadmium stress and its influence on growth performance, serum biochemical indices, antioxidant capacity, immune response, inflammatory response, and the expression of related genes in juvenile Genetically Improved Farmed Tilapia (GIFT, Oreochromis niloticus). Five groups of juvenile tilapias (initial body weight 21.21 ± 0.24 g) were fed five diets with different levels (0%, 0.5%, 1.0%, 1.5%, and 2.0%) of chitosan supplementation for 60 days under cadmium stress (0.2 mg/L Cd2+). The findings indicated that, compared with the 0% chitosan group, dietary chitosan could significantly increase (p < 0.05) the final weight (Wf), weight gain rate (WGR), specific growth rate (SGR), daily growth index (DGI), and condition factor (CF), while the feed conversion ratio (FCR) expressed the opposite trend in juvenile GIFT. Dietary chitosan could significantly increase (p < 0.05) the activities (contents) of cholinesterase (CHE), albumin (ALB), lactate dehydrogenase (LDH), alkaline phosphatase (ALP), acid phosphatase (ACP), and lysozyme (LZM), while glutamic pyruvic transaminase (GPT), glutamic oxaloacetic transaminase (GOT), and complement 3 (C3) in the serum of juvenile GIFT expressed the opposite trend. Dietary chitosan could significantly increase (p < 0.05) the activities of superoxide dismutase (SOD) and catalase (CAT) and significantly decrease (p < 0.05) the activities (contents) of glutathione S-transferase (GST), glutathione peroxidase (GSH-Px), and malondialdehyde (MDA) in the serum of juvenile GIFT. Dietary chitosan could significantly increase (p < 0.05) the activities (contents) of CAT, GST, GSH-Px, and total antioxidant capacity (T-AOC) and significantly decrease (p < 0.05) the contents of MDA in the liver of juvenile GIFT. Dietary chitosan could significantly increase (p < 0.05) the activities (contents) of SOD, GSH-Px, T-AOC, Na+-K+-ATPase, and Ca2+-ATPase and significantly decrease (p < 0.05) the activities (contents) of CAT, GST, and MDA in the gills of juvenile GIFT. Dietary chitosan could significantly up-regulate (p < 0.05) the gene expression of cat, sod, gst, and gsh-px in the liver of juvenile GIFT. Dietary chitosan could significantly up-regulate (p < 0.05) the gene expression of interferon-γ (inf-γ) in the gills and spleen and significantly down-regulate (p < 0.05) the gene expression of inf-γ in the liver and head kidney of juvenile GIFT. Dietary chitosan could significantly down-regulate (p < 0.05) the gene expression of interleukin-6 (il-6), il-8, and tumor necrosis factor-α (tnf-α) in the liver, gills, head kidney, and spleen of juvenile GIFT. Dietary chitosan could significantly up-regulate (p < 0.05) the gene expression of il-10 in the liver, gills, head kidney, and spleen of juvenile GIFT. Dietary chitosan could significantly up-regulate (p < 0.05) the gene expression of transforming growth factor-β (tgf-β) in the liver and significantly down-regulate (p < 0.05) the gene expression of tgf-β in the head kidney and spleen of juvenile GIFT. In conclusion, dietary chitosan could mitigate the impact of cadmium stress on growth performance, serum biochemical indices, antioxidant capacity, immune response, inflammatory response, and related gene expression in juvenile GIFT. According to the analysis of second-order polynomial regression, it was found that the optimal dietary chitosan levels in juvenile GIFT was approximately 1.42% to 1.45%, based on its impact on Wf, WGR, SGR, and DGI. Full article
Show Figures

Figure 1

Figure 1
<p>Relationship between different dietary chitosan levels and the final weight (Wf), weight gain rate (WGR), specific growth rate (SGR), and daily growth index (DGI) of juvenile GIFT under cadmium stress based on second-order polynomial regression analysis.</p>
Full article ">Figure 2
<p>Effects of dietary chitosan on ATPase activity in the gills of juvenile GIFT under cadmium stress. All the above data are mean ± SE (<span class="html-italic">n</span> = 3). Different superscript letters in the figure indicate significant differences among the data (<span class="html-italic">p</span> &lt; 0.05).</p>
Full article ">Figure 3
<p>Effects of dietary chitosan on the relative expression of catalase (<span class="html-italic">cat</span>), superoxide dismutase (<span class="html-italic">sod</span>), glutathione S-transferase (<span class="html-italic">gst</span>), and glutathione peroxidase (<span class="html-italic">gsh-px</span>) in the liver of juvenile GIFT under cadmium stress. All the above data are mean ± SE (<span class="html-italic">n</span> = 3). Different superscript letters in the figure indicate significant differences among the data (<span class="html-italic">p</span> &lt; 0.05).</p>
Full article ">Figure 4
<p>Effects of dietary chitosan on the relative expression of interferon-γ (<span class="html-italic">inf-γ</span>) in the liver, gills, head kidney, and spleen of juvenile GIFT under cadmium stress. All the above data are mean ± SE (<span class="html-italic">n</span> = 3). Different superscript letters in the figure indicate significant differences among the data (<span class="html-italic">p</span> &lt; 0.05).</p>
Full article ">Figure 5
<p>Effects of dietary chitosan on the relative expression of interleukin-6 (<span class="html-italic">il-6</span>) in the liver, gills, head kidney, and spleen of juvenile GIFT under cadmium stress. All the above data are mean ± SE (<span class="html-italic">n</span> = 3). Different superscript letters in the figure indicate significant differences among the data (<span class="html-italic">p</span> &lt; 0.05).</p>
Full article ">Figure 6
<p>Effects of dietary chitosan on the relative expression of interleukin-8 (<span class="html-italic">il-8</span>) in the liver, gills, head kidney, and spleen of juvenile GIFT under cadmium stress. All the above data are mean ± SE (<span class="html-italic">n</span> = 3). Different superscript letters in the figure indicate significant differences among the data (<span class="html-italic">p</span> &lt; 0.05).</p>
Full article ">Figure 7
<p>Effects of dietary chitosan on the relative expression of interleukin-10 (<span class="html-italic">il-10</span>) in the liver, gills, head kidney, and spleen of juvenile GIFT under cadmium stress. All the above data are mean ± SE (<span class="html-italic">n</span> = 3). Different superscript letters in the figure indicate significant differences among the data (<span class="html-italic">p</span> &lt; 0.05).</p>
Full article ">Figure 8
<p>Effects of dietary chitosan on the relative expression of transforming growth factor-β (<span class="html-italic">tgf-β</span>) in the liver, gills, head kidney, and spleen of juvenile GIFT under cadmium stress. All the above data are mean ± SE (<span class="html-italic">n</span> = 3). Different superscript letters in the figure indicate significant differences among the data (<span class="html-italic">p</span> &lt; 0.05).</p>
Full article ">Figure 9
<p>Effects of dietary chitosan on the relative expression of tumor necrosis factor-α (<span class="html-italic">tnf-α</span>) in the liver, gills, head kidney, and spleen of juvenile GIFT under cadmium stress. All the above data are mean ± SE (<span class="html-italic">n</span> = 3). Different superscript letters in the figure indicate significant differences among the data (<span class="html-italic">p</span> &lt; 0.05).</p>
Full article ">
15 pages, 2731 KiB  
Article
Proteomic Analysis of the Characteristic Flavor Components in Bacillus subtilis BSNK-5-Fermented Soymilk
by Miao Hu, Jiao Wang, Yaxin Gao, Bei Fan, Fengzhong Wang and Shuying Li
Foods 2024, 13(15), 2399; https://doi.org/10.3390/foods13152399 - 29 Jul 2024
Viewed by 502
Abstract
Fermentation with Bacillus subtilis significantly enhances the physiological activity and bioavailability of soymilk, but the resulting characteristic flavor seriously affects its industrial promotion. The objective of this study was to identify key proteins associated with characteristic flavors in B. subtilis BSNK-5-fermented soymilk using [...] Read more.
Fermentation with Bacillus subtilis significantly enhances the physiological activity and bioavailability of soymilk, but the resulting characteristic flavor seriously affects its industrial promotion. The objective of this study was to identify key proteins associated with characteristic flavors in B. subtilis BSNK-5-fermented soymilk using tandem mass tag (TMT) proteomics. The results showed that a total of 765 differentially expressed proteins were identified. Seventy differentially expressed proteins related to characteristic flavor were screened through Gene Ontology (GO) and Kyoto Encyclopedia of Genes and Genomes (KEGG) enrichment analysis. After integrating metabolomics data, fifteen key proteases of characteristic flavor components in BSNK-5-fermented soymilk were further identified, and free ammonia was added. In addition, there were five main formation mechanisms, including the decomposition of urea to produce ammonia; the degradation of glutamate by glutamate dehydrogenase to produce ammonia; the degradation of threonine and non-enzymatic changes to form the derivative 2,5-dimethylpyrazine; the degradation of valine, leucine, and isoleucine to synthesize isovalerate and 2-methylbutyrate; and the metabolism of pyruvate and lactate to synthesize acetate. These results provide a theoretical foundation for the improvement of undesirable flavor in B. subtilis BSNK-5-fermented soy foods. Full article
Show Figures

Figure 1

Figure 1
<p>The concentration and SDS-PAGE of sample proteins. (<b>A</b>) The protein concentration; (<b>B</b>) SDS-PAGE of proteins with different fermentation times. The different letters represented the significant difference (<span class="html-italic">p</span> &lt; 0.05).</p>
Full article ">Figure 2
<p>(<b>A</b>) The distribution of protein molecular weight; (<b>B</b>) the distribution of peptide numbers; and (<b>C</b>) the extent of peptide sequence coverage.</p>
Full article ">Figure 3
<p>The volcano plots of differentially expressed proteins in different groups. (<b>A</b>) Comparison groups of 24 h and 48 h fermented soymilk (group 1); (<b>B</b>) 24 h and 84 h fermented soymilk (group 2); (<b>C</b>) 48 h and 84 h fermented soymilk (group 3).</p>
Full article ">Figure 4
<p>(<b>A</b>) Venn diagram and (<b>B</b>) hierarchical clustering analysis of differentially expressed proteins.</p>
Full article ">Figure 5
<p>GO enrichment analysis for different groups: (<b>A</b>) 1–2 comparison group: 24 h vs. 48 h; (<b>B</b>) 1–3 comparison group: 24 h vs. 84 h; (<b>C</b>) 2–3 comparison group: 48 h vs. 84 h.</p>
Full article ">Figure 6
<p>The KEGG bubble charts for different groups: (<b>A</b>) 1–2 comparison group: 24 h vs. 48 h; (<b>B</b>) 1–3 comparison group: 24 h vs. 84 h; (<b>C</b>) 2–3 comparison group: 48 h vs. 84 h.</p>
Full article ">
16 pages, 19441 KiB  
Article
Ocular Inflammation and Oxidative Stress as a Result of Chronic Intermittent Hypoxia: A Rat Model of Sleep Apnea
by Nina Donkor, Jennifer J. Gardner, Jessica L. Bradshaw, Rebecca L. Cunningham and Denise M. Inman
Antioxidants 2024, 13(7), 878; https://doi.org/10.3390/antiox13070878 - 22 Jul 2024
Viewed by 684
Abstract
Obstructive sleep apnea (OSA) is a sleep disorder characterized by intermittent complete or partial occlusion of the airway. Despite a recognized association between OSA and glaucoma, the nature of the underlying link remains unclear. In this study, we investigated whether mild OSA induces [...] Read more.
Obstructive sleep apnea (OSA) is a sleep disorder characterized by intermittent complete or partial occlusion of the airway. Despite a recognized association between OSA and glaucoma, the nature of the underlying link remains unclear. In this study, we investigated whether mild OSA induces morphological, inflammatory, and metabolic changes in the retina resembling those seen in glaucoma using a rat model of OSA known as chronic intermittent hypoxia (CIH). Rats were randomly assigned to either normoxic or CIH groups. The CIH group was exposed to periodic hypoxia during its sleep phase with oxygen reduction from 21% to 10% and reoxygenation in 6 min cycles over 8 h/day. The eyes were subsequently enucleated, and then the retinas were evaluated for retinal ganglion cell number, oxidative stress, inflammatory markers, metabolic changes, and hypoxic response modulation using immunohistochemistry, multiplex assays, and capillary electrophoresis. Statistically significant differences were observed between normoxic and CIH groups for oxidative stress and inflammation, with CIH resulting in increased HIF-1α protein levels, higher oxidative stress marker 8-OHdG, and increased TNF-α. Pyruvate dehydrogenase kinase-1 protein was significantly reduced with CIH. No significant differences were found in retinal ganglion cell number. Our findings suggest that CIH induces oxidative stress, inflammation, and upregulation of HIF-1α in the retina, akin to early-stage glaucoma. Full article
Show Figures

Figure 1

Figure 1
<p>Experimental design. (1) Rats were randomly assigned to either normoxia or CIH treatment groups. Seven days before the initiation of the CIH protocol, the rats’ home cages were placed into Oxycycler chambers to acclimatize the rats to the chambers under normoxic conditions (21% oxygen). (2) CIH was performed for the CIH group over 8 h starting at 2100 h during the sleep phase of the circadian cycle. The protocol consisted of oxygen reduction from 21% (room air) to 10% oxygen, then returned to 21% oxygen in 6 min cycles per hour (10 cycles/hour) over 8 h/day. For the remaining 16 h, animals were exposed to room air. Normoxic control rats remained in the Oxycycler chambers with room air (21% oxygen) for the duration. (3) Upon completion of the CIH protocol, rats were euthanized and the retinas were analyzed for markers of inflammation and oxidative stress using immunohistochemistry, capillary electrophoresis, thiobarbituric acid reactive substance (TBARS) assay, and Milliplex immunoassay. Schematic created using BioRender.</p>
Full article ">Figure 2
<p>(<b>A</b>) Representative immunohistochemistry images assessing expression of HIF-1α (green) in the normoxic (upper panels) and CIH group (lower panels). Sections were also immunolabeled for retinal ganglion cells (RBPMS, magenta) and cell nuclei (DAPI, blue). Scale bar = 50 μm. <span class="html-italic">n</span> (sample size) = 2 rats per group. (<b>B</b>) Capillary electrophoresis (CE) assessing the protein expression of HIF-1α in the CIH group compared to normoxia showed a significant increase in HIF-1α in the CIH group (* <span class="html-italic">p</span> = 0.0198). (<b>C</b>) SIRTUIN-1 (SIRT-1) expression was unchanged across groups (<span class="html-italic">p</span> = 0.5325), as measured by capillary electrophoresis; <span class="html-italic">n</span> = 4 rats per group. Error bars represent mean ± SEM. GCL = ganglion cell layer; INL = inner nuclear layer; ONL = outer nuclear layer. Electropherograms of HIF-1α and SIRTUIN-1 have been included as <a href="#app1-antioxidants-13-00878" class="html-app">Supplementary Materials (Supplementary Figures S1 and S2)</a>.</p>
Full article ">Figure 3
<p>CIH induces oxidative stress. (<b>A</b>) Representative images from immunohistochemistry assessing oxidative stress using an antibody directed against 8-OHdG (green) in the normoxic (upper panels) and CIH group (lower panels). Arrows (white) point to RGCs labeled with an antibody against RBPMS (magenta) in the GCL immunolabeled with 8-OHdG. Cell nuclei labeled with DAPI (blue). Scale bar = 50 μm, <span class="html-italic">n</span> = 2 rats per group. (<b>B</b>) Quantified fluorescence intensity of 8-OHdG showed increased nucleic acid-associated oxidative stress damage in the CIH compared to the normoxic group (* <span class="html-italic">p</span> = 0.0483). (<b>C</b>) A TBARS assay quantifying lipid peroxidation in the normoxic group compared to CIH showed no difference across groups (<span class="html-italic">p</span> = 0.8666); <span class="html-italic">n</span> = 6 per group. Error bars represent mean ± SEM. GCL = ganglion cell layer; INL = inner nuclear layer; ONL = outer nuclear layer.</p>
Full article ">Figure 4
<p>CIH induces retinal inflammation. (<b>A</b>) Representative immunohistochemistry images showing the expression of the cytokine TNF-α (green) in the normoxic (upper panels) and CIH groups (lower panels). Retinas were colabeled with RBPMS, specific for retinal ganglion cells (magenta), and also stained with DAPI for cell nuclei (blue). (<b>B</b>) Quantification of fluorescence intensity showed increased expression of the cytokine TNF-α in the CIH group compared to normoxia control (*** <span class="html-italic">p</span> = 0.0002; <span class="html-italic">n</span> = 2 rats per group. (<b>C</b>) ELISA showed a significant increase in TNF-α in the CIH group over normoxia (* <span class="html-italic">p</span> = 0.0379, <span class="html-italic">n</span> = 4 rats per group). (<b>D</b>) IHC images of IL-6 (green) in the normoxic (upper panels) and CIH group (lower panels). RBPMS (magenta) and DAPI (blue). (<b>E</b>) Quantifying IL-6 fluorescence intensity showed comparable expression of IL-6 in both normoxia and CIH groups. Error bars represent mean ± SEM. GCL = ganglion cell layer; INL = inner nuclear layer; ONL = outer nuclear layer. Scale bar = 50 μm, <span class="html-italic">n</span> = 2 rats per group.</p>
Full article ">Figure 5
<p>(<b>A</b>) Microglia immunolabeled using Iba-1 (green) in the normoxic (upper panels) and CIH group (lower panels). Retinal ganglion cells (magenta) and cell nuclei (DAPI, blue) are labeled for context. Arrows (white) point to Iba1-positive microglia somata. (<b>B</b>) Quantification of fluorescence intensity showed elevated levels of Iba-1 in the CIH group compared to the control (**** <span class="html-italic">p</span> = 0.0001). Error bars represent mean ± SEM. GCL = ganglion cell layer; INL = inner nuclear layer; ONL = outer nuclear layer. Scale bar = 50 μm, <span class="html-italic">n</span> = 2 rats per group.</p>
Full article ">Figure 6
<p>(<b>A</b>) Quantification of RGC somata in the normoxic and CIH groups (<span class="html-italic">p</span> = 0.3414) shows that this degree of hypoxia is not sufficient to lead to RGC apoptosis. RGCs were counted in retinal sections, with each point representing a separate section; numbers are expressed as cells per mm of GCL length. (<b>B</b>) Representative immunolabeling from normoxia and (<b>C</b>) CIH retina with RGCs immunolabeled with RBPMS (magenta) and DAPI for cell nuclei (blue). Scale bar = 20 μm.</p>
Full article ">Figure 7
<p>Effect of CIH on retinal metabolic enzymes and transporters as measured by capillary electrophoresis. (<b>A</b>) PDK-1 protein was significantly reduced in the CIH group retina compared to the control (* <span class="html-italic">p</span> = 0.03). (<b>B</b>) LDH-A protein levels were not affected by exposure to CIH (<span class="html-italic">p</span> = 0.9692). (<b>C</b>) The expression of GLUT-1 was significantly increased in CIH compared to the normoxia control group (* <span class="html-italic">p</span> = 0.0118). (<b>D</b>) GLUT-3 protein was not different across CIH and normoxia control groups (<span class="html-italic">p</span> = 0.2007). Error bars represent mean ± SEM; GCL = ganglion cell layer; INL = inner nuclear layer; ONL = outer nuclear layer; <span class="html-italic">n</span> = 4 rats per group. Electropherograms of LDH-A, PDK-1, GLUT1, and GLUT 3 have been included as <a href="#app1-antioxidants-13-00878" class="html-app">Supplementary Materials (Supplementary Figures S3–S6)</a>.</p>
Full article ">
14 pages, 10765 KiB  
Article
Causal Relationship between Mitochondrial Biological Function and Periodontitis: Evidence from a Mendelian Randomization Study
by Huan Zhou, Yan-Xin Qi, Ruo-Yan Cao, Xi-Xuan Zhang, Ang Li and Dan-Dan Pei
Int. J. Mol. Sci. 2024, 25(14), 7955; https://doi.org/10.3390/ijms25147955 - 21 Jul 2024
Viewed by 751
Abstract
A growing number of studies indicate that mitochondrial dysfunction serves as a pathological mechanism for periodontitis. Therefore, this two-sample Mendelian randomization (MR) study was carried out to explore the causal associations between mitochondrial biological function and periodontitis, because the specific nature of this [...] Read more.
A growing number of studies indicate that mitochondrial dysfunction serves as a pathological mechanism for periodontitis. Therefore, this two-sample Mendelian randomization (MR) study was carried out to explore the causal associations between mitochondrial biological function and periodontitis, because the specific nature of this causal relationship remains inconclusive in existing MR studies. Inverse variance weighting, Mendelian randomization-Egger, weighted mode, simple mode, and weighted median analyses were performed to assess the causal relationships between the exposure factors and periodontitis. The results of the present study revealed a causal association between periodontitis and medium-chain specific acyl-CoA dehydrogenase (MCAD), malonyl-CoA decarboxylase (MLYCD), glutaredoxin 2 (Grx2), oligoribonuclease (ORN), and pyruvate carboxylase (PC). Notably, MCAD and MLYCD are causally linked to periodontitis, and serve as protective factors. However, Grx2, ORN, and PC function as risk factors for periodontitis. Our study established a causal relationship between mitochondrial biological function and periodontitis, and such insights may provide a promising approach for treating periodontitis via mitochondrial regulation. Full article
(This article belongs to the Section Molecular Genetics and Genomics)
Show Figures

Figure 1

Figure 1
<p>Scatter plots for Mendelian randomization analyses of the causal effect of mitochondrial biological function on periodontitis. (<b>A</b>) Scatter plot of the causal relationship between medium−chain specific acyl−CoA dehydrogenase and periodontitis, evaluated by the IVW method. (<b>B</b>) Scatter plot of the causal relationship between malonyl−CoA decarboxylase and periodontitis, primarily evaluated via the IVW method. (<b>C</b>) Scatter plot of the causal relationship between glutaredoxin−2, mitochondrial and periodontitis, primarily evaluated using the IVW method. (<b>D</b>) Scatter plot of the causal relationship between oligoribonuclease and periodontitis using the IVW method. (<b>E</b>) Scatter plot of the causal relationship between pyruvate carboxylase and periodontitis, primarily evaluated with the IVW method. (<b>F</b>) Scatter plot of the causal relationship between mitochondrial DNA copy number and periodontitis, evaluated using the IVW method. IVW, inverse variance weighting.</p>
Full article ">Figure 2
<p>Forest map of the causal relationship between mitochondria and periodontitis (IVW method). (<b>A</b>) Forest map of the causal relationship between medium−chain specific acyl−CoA dehydrogenase and periodontitis. (<b>B</b>) Forest map of the causal relationship between malonyl−CoA decarboxylase and periodontitis. (<b>C</b>) Forest map of the causal relationship between glutaredoxin − 2 and periodontitis. (<b>D</b>) Forest map of the causal relationship between oligoribonuclease and periodontitis. (<b>E</b>) Forest map of the causal relationship between pyruvate carboxylase and periodontitis. (<b>F</b>) Forest map of the causal relationship between mitochondrial DNA copy number and periodontitis. IVW, inverse variance weighting.</p>
Full article ">Figure 3
<p>Funnel plot of the causal relationship between mitochondria and periodontitis (IVW method). (<b>A</b>) Funnel plot of the causal relationship between medium−chain specific acyl−CoA dehydrogenase and periodontitis. (<b>B</b>) Funnel plot of the causal relationship between malonyl-CoA decarboxylase and periodontitis. (<b>C</b>) Funnel plot of the causal relationship between glutaredoxin− 2 and periodontitis. (<b>D</b>) Funnel plot of the causal relationship between oligoribonuclease and periodontitis. (<b>E</b>) Funnel plot of the causal relationship between pyruvate carboxylase and periodontitis. (<b>F</b>) Funnel plot of the causal relationship between mitochondrial DNA copy number and periodontitis. IVW, inverse variance weighting.</p>
Full article ">Figure 4
<p>“Leave−one−out” forest map of the causal relationship between mitochondria and periodontitis (IVW method). (<b>A</b>) “Leave−one−out” forest map of the causal relationship between medium−chain specific acyl−CoA dehydrogenase and periodontitis. (<b>B</b>) “Leave−one−out” forest map of the causal relationship between malonyl−CoA decarboxylase and periodontitis. (<b>C</b>) “Leave−one−out” forest map of the causal relationship between glutaredoxin−2 and periodontitis. (<b>D</b>) “Leave−one−out” forest map of the causal relationship between oligoribonuclease and periodontitis. (<b>E</b>) “Leave−one−out” forest map of the causal relationship between pyruvate carboxylase and periodontitis. (<b>F</b>) “Leave−one−out” forest map of the causal relationship between mitochondrial DNA copy number and periodontitis. IVW, inverse variance weighting.</p>
Full article ">Figure 5
<p>Mendelian randomization design of the present study.</p>
Full article ">
22 pages, 7966 KiB  
Article
P38α MAPK Coordinates Mitochondrial Adaptation to Caloric Surplus in Skeletal Muscle
by Liron Waingerten-Kedem, Sharon Aviram, Achinoam Blau, Tony Hayek and Eyal Bengal
Int. J. Mol. Sci. 2024, 25(14), 7789; https://doi.org/10.3390/ijms25147789 - 16 Jul 2024
Viewed by 559
Abstract
Excessive calorie intake leads to mitochondrial overload and triggers metabolic inflexibility and insulin resistance. In this study, we examined how attenuated p38α activity affects glucose and fat metabolism in the skeletal muscles of mice on a high-fat diet (HFD). Mice exhibiting diminished p38α [...] Read more.
Excessive calorie intake leads to mitochondrial overload and triggers metabolic inflexibility and insulin resistance. In this study, we examined how attenuated p38α activity affects glucose and fat metabolism in the skeletal muscles of mice on a high-fat diet (HFD). Mice exhibiting diminished p38α activity (referred to as p38αAF) gained more weight and displayed elevated serum insulin levels, as well as a compromised response in the insulin tolerance test, compared to the control mice. Additionally, their skeletal muscle tissue manifested impaired insulin signaling, leading to resistance in insulin-mediated glucose uptake. Examination of muscle metabolites in p38αAF mice revealed lower levels of glycolytic intermediates and decreased levels of acyl-carnitine metabolites, suggesting reduced glycolysis and β-oxidation compared to the controls. Additionally, muscles of p38αAF mice exhibited severe abnormalities in their mitochondria. Analysis of myotubes derived from p38αAF mice revealed reduced mitochondrial respiratory capacity relative to the myotubes of the control mice. Furthermore, these myotubes showed decreased expression of Acetyl CoA Carboxylase 2 (ACC2), leading to increased fatty acid oxidation and diminished inhibitory phosphorylation of pyruvate dehydrogenase (PDH), which resulted in elevated mitochondrial pyruvate oxidation. The expected consequence of reduced mitochondrial respiratory function and uncontrolled nutrient oxidation observed in p38αAF myotubes mitochondrial overload and metabolic inflexibility. This scenario explains the increased likelihood of insulin resistance development in the muscles of p38αAF mice compared to the control mice on a high-fat diet. In summary, within skeletal muscles, p38α assumes a crucial role in orchestrating the mitochondrial adaptation to caloric surplus by promoting mitochondrial biogenesis and regulating the selective oxidation of nutrients, thereby preventing mitochondrial overload, metabolic inflexibility, and insulin resistance. Full article
Show Figures

Figure 1

Figure 1
<p>p38α<sup>AF</sup> mice presented worse metabolic parameters than control mice. (<b>A</b>) Six-week-old mice were fed with ND or an HFD for 10 weeks, and GC muscles were isolated (<span class="html-italic">n</span> = 5) from the control and p38α<sup>AF</sup> mice. Protein lysates from three of the mice per treatment were randomly analyzed by Western blotting with the designated antibodies. α Tubulin was used as the loading control. The quantification of relative p38α phosphorylation is presented in the histogram. (<b>B</b>) The mice underwent the diets described in (A), and the weight of each mouse was measured weekly (<span class="html-italic">n</span> = 5). The graphs represent the average percent change in the body weight of the two mouse groups (control and p38α<sup>AF</sup>), which were fed with ND or HFD. The weight was set to 100 on the first day of the diet. (<b>C</b>) The hematological parameters of control mice and p38α<sup>AF</sup> on an HFD. The glucose and cholesterol levels were measured in the serum of control and p38α<sup>AF</sup> mice after 10 weeks on an HFD (AML-central lab services). Insulin was measured (<span class="html-italic">n</span> = 3) using an ELISA kit (Millipore RAB0817). The significance probabilities between treatments were designated as numbers. (<b>D</b>) Insulin tolerance test (ITT): the graph displays the relative average glucose levels at 0, 30, 45, 60, 90, and 120 min following insulin injection (0.5 U/kg BW) in the blood of control and p38α<sup>AF</sup> mice after a 10-week HFD (<span class="html-italic">n</span> = 4 mice per group). The mice were deprived of chaw for 6 h before insulin was IP-injected. The glucose level before insulin injection was set to 100 percent, and all values were relative to 100. Data are presented as the mean ± SE. One-way ANOVA was followed by Tukey post-tests (<b>A</b>), two-way ANOVA was followed by Bonferroni post-tests (<b>B</b>,<b>D</b>) and a Student t-test (<b>C</b>). The <span class="html-italic">p</span> values for group difference are designated as follows: * <span class="html-italic">p</span> &lt; 0.05, ** <span class="html-italic">p</span> &lt; 0.01, and *** <span class="html-italic">p</span> &lt; 0.001.</p>
Full article ">Figure 2
<p>Block of the insulin-mediated 2 deoxy-glucose (2DG) uptake by the Tibialis muscle of p38α<sup>AF</sup> mice. (<b>A</b>) Experimental layout: saline or insulin (1 unit/kg) was IP-injected following a 3 h fasting of the mice previously fed with an HFD for 10 weeks. Ten min later, 5% 2DG was IP-injected (10 μL to 1 g weight). The mice were sacrificed one hour later, and the Tibialis muscles were frozen and used in the mass spectrometry (MS) analysis of metabolites, or to extract proteins for Western blotting analysis. (<b>B</b>) Peak area were analyzed by the MS values of 2- Deoxy –D Glucose (<span class="html-italic">n</span> = 4) that were normalized to mg tissue. (<b>C</b>) Protein extracts from the Tb muscles (<span class="html-italic">n</span> = 3) were analyzed by Western blotting with antibodies directed to phosphorylated Akt (Serine 473) and Pan Akt. Quantification of the relative phosphorylation (pAkt/Akt) is presented in the histogram. Data are presented as the mean ± SE. The Wilcoxon test and significance probabilities between treatments are designated as numbers in (<b>B</b>). One-way ANOVA was followed by Tukey post-tests. The <span class="html-italic">p</span> values for group difference are designated as follows: * <span class="html-italic">p</span> &lt; 0.05 (<b>C</b>).</p>
Full article ">Figure 3
<p>Reduced glycolytic metabolites and increased lactate-to-pyruvate ratio in the muscles of HFD-fed p38α<sup>AF</sup> mice. Extracted metabolites from the Tibialis muscles of 10-week HFD-fed mice that were IP-injected without or with insulin (<span class="html-italic">n</span> = 4). (<b>A</b>) The normalized peak areas (to mg tissue) that were analyzed by the MS of several glycolytic metabolites. (<b>B</b>) The normalized peak areas (to mg tissue) that were analyzed by the MS of pyruvate, lactate, and the ratio of lactate to pyruvate. (<b>C</b>) Analysis of the expression and the phosphorylation on serine 293 of the E1 subunit of pyruvate dehydrogenase (PDH) in the Tb muscles of control and p38α<sup>AF</sup> mice (<span class="html-italic">n</span> = 5) by Western blotting using antibodies to phospho-PDH (Ser293) and PDH. The quantification of relative phosphorylation (pPDH/PDH) is presented in the histogram. Data are presented as the mean ± SE. The Wilcoxon test and significance probabilities between treatments are designated as numbers in (<b>B</b>).</p>
Full article ">Figure 4
<p>Reduced β oxidation in the muscles of p38α<sup>AF</sup> mice relative to the muscles of control mice following a high-fat diet. Metabolites were extracted from the Tibialis muscles of 10-week HFD-fed mice that were IP-injected without or with insulin (<span class="html-italic">n</span> = 4). (<b>A</b>) The peak areas (normalized to mg tissue) of glycerol analyzed by MS are presented. (<b>B</b>) Analysis of the mRNA levels of FABP3 in the muscles of control and p38α<sup>AF</sup> mice by qPCR (<span class="html-italic">n</span> = 5). The β-actin housekeeping gene was used to normalize the mRNA levels. (<b>C</b>) Analysis of the mRNA levels of ACC2 in the muscles of control and p38α<sup>AF</sup> mice by qPCR (<span class="html-italic">n</span>= 4). The β-actin housekeeping gene was used to normalize mRNA levels. (<b>D</b>) Analysis of the expression and the phosphorylation on serine 212 of Acetyl CoA Carboxylase 2 (ACC2) in the muscles of control and p38α<sup>AF</sup> mice (<span class="html-italic">n</span> = 5) by Western blotting using antibodies to phospho-ACC2 (Ser212), ACC2, and αTubulin (which served as a loading control). The histograms present the relative expression of ACC2 (ACC2/Tubulin) and relative ACC2 phosphorylation on serine 212 (pACC2/ACC2). (<b>E</b>) The peak areas (normalized to mg tissue) of acyl-carnitines are presented. Values represent the means ± SEM. The Wilcoxon test and significance probabilities between treatments are designated as numbers (<b>A</b>,<b>E</b>). One-way ANOVA followed by Tukey post-tests (<b>B</b>,<b>C</b>). The <span class="html-italic">p</span> values for group difference are designated as follows: * <span class="html-italic">p</span> &lt; 0.05, ** <span class="html-italic">p</span> &lt; 0.01, and *** <span class="html-italic">p</span> &lt; 0.001.</p>
Full article ">Figure 5
<p>Severe mitochondrial defects in the muscles of p38α<sup>AF</sup> mice. (<b>A</b>) Transmission electron microscopy (TEM) analysis of the representative muscles from control and p38α<sup>AF</sup> mice fed with NDs and HFDs. The Tibialis muscles were isolated, and longitudinal sections were processed for TEM analysis (see <a href="#sec4dot10-ijms-25-07789" class="html-sec">Section 4.10</a>). Representative images are shown. Scale bar: 1 μm. Asterisks are adjacent to the mitochondria (<b>B</b>) Analysis of the mRNA levels of PGC1α in the muscles of control and p38α<sup>AF</sup> mice fed with NDs and HFDs by qPCR (<span class="html-italic">n</span> = 5). The β-actin housekeeping gene was used to normalize the mRNA levels. Data represent the means ± SEM. One-way ANOVA was followed by Tukey post-tests (B). The <span class="html-italic">p</span> values for group differences are designated as follows: * <span class="html-italic">p</span> &lt; 0.05.</p>
Full article ">Figure 6
<p>Biochemical and metabolic analysis of the myotubes derived from control and p38α<sup>AF</sup> mice. (<b>A</b>) p38 MAPK phosphorylation: Myotubes were grown for 24 h in the absence or presence of 0.4 mM of palmitate. Insulin (10 μg/mL) was added 30 min before the proteins were extracted and analyzed by Western blotting using the designated antibodies. (<b>B</b>) Insulin signaling pathway: The same protein samples as in (A) were analyzed by Western blotting using the designated antibodies. (<b>C</b>) Metabolism of the (U-<sup>13</sup>C<sub>6</sub>) glucose in myotubes: (U-<sup>13</sup>C<sub>6</sub>) glucose was introduced to the myotube media with or without 0.4 mM of palmitate for 24 h. The relative levels of glucose 6-phosphate (+6), fructose 6-phosphate (+6), and ribose phosphate (+5) isotopologues are presented. The peak area was normalized to protein concentration. (<b>D</b>) Medium acidification (ECAR) of myotubes in a “Seahorse” analysis: Myotubes were grown in glucose, or glucose and palmitate, for 24 h before analysis. (<b>E</b>) Metabolism of the (U-<sup>13</sup>C<sub>6</sub>) glucose in myotubes: The relative levels of the isotopologues of citrate are presented. The peak areas were normalized to protein concentration. (<b>F</b>) Mitochondrial enzymes: The same protein samples as in (A) were analyzed by Western blotting. The histograms present the relative expression and phosphorylation of PDH (Ser293), and the expression of citrate synthase. Data represent the means ± SEM. The Wilcoxon test and significance probabilities between treatments are designated as follows: * <span class="html-italic">p</span> &lt; 0.05, ** <span class="html-italic">p</span> &lt; 0.01, and *** <span class="html-italic">p</span> &lt; 0.001 (<b>C</b>,<b>E</b>). One-way ANOVA was followed by Tukey post-tests (<b>D</b>). The <span class="html-italic">p</span> values for group differences are designated as follows: * <span class="html-italic">p</span> &lt; 0.05 and *** <span class="html-italic">p</span> &lt; 0.001.</p>
Full article ">Figure 7
<p>Metabolism of palmitate in the myotubes derived from control and p38α<sup>AF</sup> mice. Myotubes were grown in a low-glucose DMEM supplemented with 0.4 mM of palmitate-<sup>13</sup>C<sub>16</sub> for 6 and 24 h. (<b>A</b>) The peak area (normalized to protein concentration) of palmitate (+16), the isotopologues of the TCA cycle, and the derived amino acids that originated from palmitate-<sup>13</sup>C<sub>16.</sub> FC: fold change in the palmitate derived (<sup>13</sup>C ≥ 2) metabolite abundance relative to a WT of 6 h or WT of 24 h. Dashed arrows indicate of missing stages in the TCA-cycle. (<b>B</b>) Myotubes were grown for 24 h in the absence or presence of 0.4 mM of palmitate. Insulin (10 μg/mL) was added 30 min before proteins were extracted and analyzed by Western blotting with the designated antibodies. The histograms present the relative expression of ACC2, the phosphorylation of ACC2 (Ser212), and the phosphorylation of AMPKα (Thr172). (<b>C</b>) The oxygen consumption rate (OCR) at the maximal respiration of myotubes that were grown on glucose, or glucose and palmitate, for 24 h. (<b>D</b>) Comparison of the mitochondrial membrane electrochemical potential in myotubes that were grown on glucose, or glucose and palmitate, for 24 h. JC-1 dye was used to monitor the mitochondrial membrane potential. FCCP disrupts the mitochondrial membrane potential. Data represent the means ± SEM. One-way ANOVA was followed by Tukey post-tests (<b>A</b>,<b>C</b>). The <span class="html-italic">p</span> values for group difference are designated as follows: * <span class="html-italic">p</span> &lt; 0.05, ** <span class="html-italic">p</span> &lt; 0.01, and *** <span class="html-italic">p</span> &lt; 0.001.</p>
Full article ">Figure 8
<p>A model for the role of p38α in insulin sensitivity. In the left panel, a high-fat diet activates p38α in wild-type mice, leading to an increased expression and activity of PGC1α and ACC2 in the skeletal muscles. PGC1α acts as a co-activator, increasing mitochondrial biogenesis and activity, while ACC2 regulates fatty acid transport into mitochondria. These activities of p38α help coordinate glucose and fat oxidation, preserving metabolic flexibility and preventing mitochondrial damage. Under these conditions, both energy balance and insulin sensitivity are preserved.</p>
Full article ">
13 pages, 2337 KiB  
Article
The Effects of Water Flow Speed on Swimming Capacity and Energy Metabolism in Adult Amur Grayling (Thymallus grubii)
by Cunhua Zhai, Yutao Li, Guanyu Zhu, Wenjie Peng, Qiuxu E, Ying Zhang and Bo Ma
Fishes 2024, 9(7), 272; https://doi.org/10.3390/fishes9070272 - 10 Jul 2024
Viewed by 549
Abstract
The present study aimed to explore whether water flow velocity could affect the swimming ability and overall energy metabolism of wild Amur grayling (Thymallus grubii). Swimming performance was assessed by measuring critical swimming speed (Ucrit), burst speed (Uburst [...] Read more.
The present study aimed to explore whether water flow velocity could affect the swimming ability and overall energy metabolism of wild Amur grayling (Thymallus grubii). Swimming performance was assessed by measuring critical swimming speed (Ucrit), burst speed (Uburst), and oxygen consumption rate (MO2) based on the stepped velocity test method. Our results showed that the absolute values of Ucrit and Uburst tended to increase with body length. In contrast, the relative values of Ucrit and Uburst tended to decrease and increase, respectively. MO2 in Amur grayling was elevated with increasing velocity, suggesting relatively high swimming efficiency. We also measured the biochemical indices related to energy metabolism. Lactate dehydrogenase, hexokinase, and pyruvate kinase activities significantly increased (p < 0.05). Hepatic glycogen, glucose, and muscle glycogen contents decreased with the increasing trend of velocity (p < 0.05), the lactic acid contents of the blood and muscles increased significantly with the increase in velocities (p < 0.05), and changes in creatine phosphate content showed no significant difference (p > 0.05). The results not only denote the relationship between body size and swimming speed but also show the effects of water flow velocity on energy metabolism in Amur grayling. The results provide basic data for the construction of fish passage. Full article
(This article belongs to the Section Biology and Ecology)
Show Figures

Figure 1

Figure 1
<p>Schematic diagram of the flume-type swimming respirometer: (A) fish swimming area, (B) rectifier, (C) DO probe, (D) temperature sensor, (E) propeller, (F) propeller motor, and (G) frequency changer.</p>
Full article ">Figure 2
<p>Relationships between body length and swimming speed for <span class="html-italic">T. grubii</span>: (<b>a</b>) for the linear regression of critical speed (U<sub>crit</sub>) and fish body length, the equation is y = −0.1448x<sup>3</sup> + 7.4003x<sup>2</sup> − 113.4x + 579.91 (<span class="html-italic">R</span><sup>2</sup> = 0.9586); (<b>b</b>) for the linear regression of burst speed (U<sub>burst</sub>) and fish body length, the equation is y = −0.0865x<sup>3</sup> + 4.8225x<sup>2</sup> − 82.031x + 545.38 (<span class="html-italic">R</span><sup>2</sup> = 0.9731); (<b>c</b>) for the linear regression of relative critical speed (U<sub>crit’</sub>) and fish body length, the equation is y = −0.0085x<sup>3</sup> + 0.4168x<sup>2</sup> − 6.308x + 33.259 (<span class="html-italic">R</span><sup>2</sup> = 0.8937); and (<b>d</b>) for the linear regression of relative burst speed (U<sub>burst’</sub>) and fish body length, the equation is y = −0.0056x<sup>3</sup> + 0.3132x<sup>2</sup> − 5.8463x + 43.188 (<span class="html-italic">R</span><sup>2</sup> = 0.4689).</p>
Full article ">Figure 3
<p>Relationships between oxygen consumption rate and water flow velocity (U, BL/s) for different body lengths: (<b>a</b>) the oxygen consumption rate of group 1 (body length: 16.42 ± 0.86 cm) at different water flow velocities (mean ± SD, <span class="html-italic">n</span> = 6); (<b>b</b>) the oxygen consumption rate of group 2 (body length: 20.50 ± 0.71 cm) at different water flow velocities (mean ± SD, <span class="html-italic">n</span> = 4); (<b>c</b>) the oxygen consumption rate of group 3 (body length: 23.70 ± 0.84 cm) at different water flow velocities (mean ± SD, <span class="html-italic">n</span> = 5); (<b>d</b>) effect of water flow velocity on MO<sub>2</sub>. The dot types represent the different body length categories. The equations in (<b>a</b>–<b>c</b>) are y = 0.04623 + 0.02452x<sup>0.39484</sup> (<span class="html-italic">R</span><sup>2</sup> = 0.94803), y = 0.03514 + 0.01284x<sup>0.80175</sup> (<span class="html-italic">R</span><sup>2</sup> = 0.91076), and y = 0.03214 + 0.00487x<sup>1.83557</sup> (<span class="html-italic">R</span><sup>2</sup> = 0.99464), respectively. Different lowercase letters above the line graphs represent significant differences (<span class="html-italic">p</span> &lt; 0.05) between the groups.</p>
Full article ">Figure 4
<p>Glycometabolism-related enzyme activities and physiological index content in the Amur grayling. LDH (<b>a</b>), PK (<b>b</b>), HK (<b>c</b>), and SDH (<b>d</b>) activities in the liver. Different lowercase letters indicated significant differences (<span class="html-italic">p</span> &lt; 0.05) among the different water flow velocities (U, BL/s). Data are presented as the mean ± SD (<span class="html-italic">n</span> = 3).</p>
Full article ">Figure 5
<p>Biochemical substance content in the Amur grayling at different water flow velocities (U, BL/s): (<b>a</b>–<b>f</b>) variations in the content of hepatic glycogen (<b>a</b>), glucose (<b>b</b>), muscle glycogen (<b>c</b>), creatine phosphate (CP, (<b>d</b>)), plasma lactic acid (<b>e</b>), and muscle lactic acid (<b>f</b>) of the Amur grayling at different flow velocities. The letters indicate a significant difference (<span class="html-italic">p</span> &lt; 0.05) between treatments. Data are presented as the mean ± SD (<span class="html-italic">n</span> = 3).</p>
Full article ">
20 pages, 6228 KiB  
Article
Effects of PEF on Cell and Transcriptomic of Escherichia coli
by Jinyan Kuang, Ying Lin, Li Wang, Zikang Yan, Jinmei Wei, Jin Du and Zongjun Li
Microorganisms 2024, 12(7), 1380; https://doi.org/10.3390/microorganisms12071380 - 7 Jul 2024
Viewed by 769
Abstract
Pulsed electric field (PEF) is an up-to-date non-thermal processing technology with a wide range of applications in the food industry. The inactivation effect of PEF on Escherichia coli was different under different conditions. The E. coli inactivated number was 1.13 ± 0.01 lg [...] Read more.
Pulsed electric field (PEF) is an up-to-date non-thermal processing technology with a wide range of applications in the food industry. The inactivation effect of PEF on Escherichia coli was different under different conditions. The E. coli inactivated number was 1.13 ± 0.01 lg CFU/mL when PEF was treated for 60 min and treated with 0.24 kV/cm. The treatment times were found to be positively correlated with the inactivation effect of PEF, and the number of E. coli was reduced by 3.09 ± 0.01 lg CFU/mL after 100 min of treatment. The inactivation assays showed that E. coli was inactivated at electrical intensity (0.24 kV/cm) within 100 min, providing an effective inactivating outcome for Gram-negative bacteria. The purpose of this work was to investigate the cellular level (morphological destruction, intracellular macromolecule damage, intracellular enzyme inactivation) as well as the molecular level via transcriptome analysis. Field Emission Scanning Electron Microscopy (TFESEM) and Transmission Electron Microscope (TEM) results demonstrated that cell permeability was disrupted after PEF treatment. Entocytes, including proteins and DNA, were markedly reduced after PEF treatment. In addition, the activities of Pyruvate Kinase (PK), Succinate Dehydrogenase (SDH), and Adenosine Triphosphatase (ATPase) were inhibited remarkably for PEF-treated samples. Transcriptome sequencing results showed that differentially expressed genes (DEGs) related to the biosynthesis of the cell membrane, DNA replication and repair, energy metabolism, and mobility were significantly affected. In conclusion, membrane damage, energy metabolism disruption, and other pathways are important mechanisms of PEF’s inhibitory effect on E. coli. Full article
(This article belongs to the Special Issue Microbial Safety and Biotechnology in Food Production and Processing)
Show Figures

Figure 1

Figure 1
<p>Inactivation of <span class="html-italic">E. coli</span> by PEF treatment. (<b>A</b>) Inactivation of <span class="html-italic">E. coli</span> in different media solutions by PEF. (<b>B</b>) Inactivation of <span class="html-italic">E. coli</span> in different NaCl mass fractions by PEF. (<b>C</b>) Inactivation of <span class="html-italic">E. coli</span> in different electric field intensities by PEF. (<b>D</b>) Inactivation of <span class="html-italic">E. coli</span> in different treatment times by PEF. (<b>E</b>) Inactivation of <span class="html-italic">E. coli</span> in different bacterial densities by PEF. (<b>F</b>) Inactivation of <span class="html-italic">E. coli</span> by PEF treatment. Different letters in the graph indicate significant differences (<span class="html-italic">p</span> &lt; 0.05). Data are shown as the mean ± SD (<span class="html-italic">n</span> = 3), with error bars representing standard errors. Different letters in the graphs indicate significant differences (<span class="html-italic">p</span> &lt; 0.05).</p>
Full article ">Figure 1 Cont.
<p>Inactivation of <span class="html-italic">E. coli</span> by PEF treatment. (<b>A</b>) Inactivation of <span class="html-italic">E. coli</span> in different media solutions by PEF. (<b>B</b>) Inactivation of <span class="html-italic">E. coli</span> in different NaCl mass fractions by PEF. (<b>C</b>) Inactivation of <span class="html-italic">E. coli</span> in different electric field intensities by PEF. (<b>D</b>) Inactivation of <span class="html-italic">E. coli</span> in different treatment times by PEF. (<b>E</b>) Inactivation of <span class="html-italic">E. coli</span> in different bacterial densities by PEF. (<b>F</b>) Inactivation of <span class="html-italic">E. coli</span> by PEF treatment. Different letters in the graph indicate significant differences (<span class="html-italic">p</span> &lt; 0.05). Data are shown as the mean ± SD (<span class="html-italic">n</span> = 3), with error bars representing standard errors. Different letters in the graphs indicate significant differences (<span class="html-italic">p</span> &lt; 0.05).</p>
Full article ">Figure 2
<p>Effect of PEF on the morphology of <span class="html-italic">E. coli</span> (FESEM). (<b>A</b>,<b>B</b>) Control. (<b>C</b>,<b>D</b>) PEF. The red arrows indicated the damaging bacteria.</p>
Full article ">Figure 3
<p>Effect of PEF on the morphology of <span class="html-italic">E. coli</span> (TEM). (<b>A</b>,<b>B</b>) Control. (<b>C</b>,<b>D</b>) PEF. Red arrows represent cell membrane, cell wall, and cytoplasmic changes in the cell.</p>
Full article ">Figure 4
<p>Effect of PEF on ATPase, SDH, and PK. (<b>A</b>) Effect of PEF on the cell membrane permeability of <span class="html-italic">E. coli</span>. (<b>B</b>) The effect of PEF on PK in <span class="html-italic">E. coli</span>. (<b>C</b>) The effect of PEF on SDH in <span class="html-italic">E. coli</span>. (<b>D</b>) The effect of PEF on Na<sup>+</sup> K<sup>+</sup> ATPase, Ca<sup>2+</sup> Mg<sup>2+</sup> ATPase in <span class="html-italic">E. coli</span>. Data are shown as the mean ± SD (<span class="html-italic">n</span> = 3), with error bars representing standard errors. Different letters in the graphs indicate significant differences (<span class="html-italic">p</span> &lt; 0.05).</p>
Full article ">Figure 5
<p>Gene expression analysis. (<b>A</b>) Correlation coefficient heat map. (<b>B</b>) Volcano plot of DEGs between PEF and control groups. The green and red colours of the horizontal and vertical coordinates in (<b>A</b>) indicate clustering.</p>
Full article ">Figure 6
<p>GO and KEGG enrichment analysis. (<b>A</b>) GO enrichment analysis (up-regulated). (<b>B</b>) GO enrichment analysis (down-regulated). (<b>C</b>) KEGG enrichment analysis (up-regulated). (<b>D</b>) KEGG enrichment analysis (down-regulated). (<b>E</b>) RT-qPCR verification of gene expression levels.</p>
Full article ">Figure 7
<p>Heat map of DEGs associated with GO and KEGG enrichment analysis in <span class="html-italic">E. coli</span>. (<b>A</b>) Cell structures. (<b>B</b>) DNA replication and repair. (<b>C</b>) ABC transporters. (<b>D</b>) Glycolysis and TCA cycle. (<b>E</b>) Mobility. Red indicates up-regulated genes, while blue indicates down-regulated genes.</p>
Full article ">
18 pages, 3283 KiB  
Article
CCL2 and Lactate from Chemotherapeutics-Treated Fibroblasts Drive Malignant Traits by Metabolic Rewiring in Low-Migrating Breast Cancer Cell Lines
by Maria Jesus Vera, Iván Ponce, Cristopher Almarza, Gonzalo Ramirez, Francisco Guajardo, Karen Dubois-Camacho, Nicolás Tobar, Félix A. Urra and Jorge Martinez
Antioxidants 2024, 13(7), 801; https://doi.org/10.3390/antiox13070801 - 1 Jul 2024
Viewed by 849
Abstract
While cytostatic chemotherapy targeting DNA is known to induce genotoxicity, leading to cell cycle arrest and cytokine secretion, the impact of these drugs on fibroblast–epithelial cancer cell communication and metabolism remains understudied. Our research focused on human breast fibroblast RMF-621 exposed to nonlethal [...] Read more.
While cytostatic chemotherapy targeting DNA is known to induce genotoxicity, leading to cell cycle arrest and cytokine secretion, the impact of these drugs on fibroblast–epithelial cancer cell communication and metabolism remains understudied. Our research focused on human breast fibroblast RMF-621 exposed to nonlethal concentrations of cisplatin and doxorubicin, revealing reduced proliferation, diminished basal and maximal mitochondrial respirations, heightened mitochondrial ROS and lactate production, and elevated MCT4 protein levels. Interestingly, RMF-621 cells enhanced glucose uptake, promoting lactate export. Breast cancer cells MCF-7 exposed to conditioned media (CM) from drug-treated stromal RMF-621 cells increased MCT1 protein levels, lactate-driven mitochondrial respiration, and a significantly high mitochondrial spare capacity for lactate. These changes occurred alongside altered mitochondrial respiration, mitochondrial membrane potential, and superoxide levels. Furthermore, CM with doxorubicin and cisplatin increased migratory capacity in MCF-7 cells, which was inhibited by MCT1 (BAY-8002), glutamate dehydrogenase (EGCG), mitochondrial pyruvate carrier (UK5099), and complex I (rotenone) inhibitors. A similar behavior was observed in T47-D and ZR-75-1 breast cancer cells. This suggests that CM induces metabolic rewiring involving elevated lactate uptake to sustain mitochondrial bioenergetics during migration. Treatment with the mitochondrial-targeting antioxidant mitoTEMPO in RMF-621 and the addition of an anti-CCL2 antibody in the CM prevented the promigratory MCF-7 phenotype. Similar effects were observed in THP1 monocyte cells, where CM increased monocyte recruitment. We propose that nonlethal concentrations of DNA-damaging drugs induce changes in the cellular environment favoring a promalignant state dependent on mitochondrial bioenergetics. Full article
(This article belongs to the Special Issue Oxidative Stress and Metabolite Signaling in the Heart and Cancer)
Show Figures

Figure 1

Figure 1
<p>DNA-damaging chemotherapeutics promote metabolic remodeling in fibroblast RMF-621 cells: (<b>A</b>) effect of BiBr (1.7 µM), Cisplatin (Cis, 3.0 µM), and Doxorubicin (Doxo, 18 nM) on sub-G1 population at 72 h of treatment, (<b>B</b>) relative <span class="html-italic">CCL2</span> gene expression measured by qPCR, (<b>C</b>) mitochondrial ROS levels (mtROS) measured using mitoSOX dye by flow cytometry, (<b>D</b>) oxygen consumption rate (OCR) profile, (<b>E</b>–<b>G</b>) changes in basal, ATP-driven, and maximal respirations of RMF-621 cells treated with BiBr, Cis, and Doxo for 48 h, (<b>H</b>) schematic diagram on lactate production, (<b>I</b>,<b>J</b>) changes in <span class="html-italic">Glut1</span> and <span class="html-italic">MCT4</span> gene expression in RMF-621 cells treated with BiBr (1.7 μM), Cisplatin (3 μM, Cis), and Doxorubicin (18 nM, Doxo) for 72 h, (<b>K</b>) lactate levels in the media of RMF-621 after treatment with DNA-damaging drugs. Data are shown as the mean ± SD of three or four independent experiments. * <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, vs. control (DMSO). n.s.: not significant.</p>
Full article ">Figure 2
<p>Conditioned media from DNA-damaging drugs increase motility in low-migrating epithelial breast cancer cells: (<b>A</b>–<b>C</b>) effect of CM on the migration of MCF-7, T47D, and ZR75-1 breast cancer cell lines. Data are shown as the mean ± SD of four independent experiments. ** <span class="html-italic">p</span> &lt; 0.01, *** <span class="html-italic">p</span> &lt; 0.001, vs. control (DMSO). n.s.: not significant.</p>
Full article ">Figure 3
<p>Conditioned media from DNA-damaging drugs increase metabolic plasticity dependent on lactate/MCT1 in MCF-7 cells: (<b>A</b>,<b>B</b>) effect of CM on MCT1 levels and glycolysis and glycolytic capacity in MCF-7 breast cancer cells at 72 h of exposition, (<b>C</b>–<b>E</b>) effect of CM on the profile of respiration, basal, and maximal respirations of MCF-7 exposed to Seahorse assay buffer containing 10 mM glucose and 4 mM glutamine, (<b>F</b>–<b>H</b>) effect of CM on the profile of respiration, basal, and maximal respirations of MCF-7 exposed to Seahorse assay buffer containing 10 mM lactate and 4 mM glutamine, (<b>I</b>) differences in the spare capacity dependent on glucose or lactate presence in MCF-7 cells exposed to CM by DNA-damaging drugs for 24 h. Data are shown as the mean ± SD, N = 4 independent experiments. * <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, vs. control (DMSO). n.s.: not significant.</p>
Full article ">Figure 4
<p>CCL2 and lactate from conditioned media by DNA-damaging drugs are essential for increased migration in MCF-7 cancer cells: (<b>A</b>) effect of increasing concentrations of CCL2 on MCT1 protein levels in MCF-7 cells at 72 h of exposition, (<b>B</b>) effect of lactate (20 mM) on migration of MCF-7 cells at 24 h of exposition, (<b>C</b>) effect of stromal CM prepared by DNA-damaging drugs (as explained above) and with a previous incubation (1 h) of mitoTEMPO (1 µM) on MCF-7 migration, (<b>D</b>,<b>E</b>) effect of CM produced by DNA-damaging drugs and then the addition of blocking antibody anti-CCL2 (10 µg/mL) or MCT1 inhibitor, BAY-8002 (100 nM) on MCF-7 migration. Data are shown as the mean ± SD, N = 4 independent experiments. * <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, vs. control (DMSO). n.s.: not significant.</p>
Full article ">Figure 5
<p>Mitochondrial pyruvate transport and glutaminolysis are required for maintaining the viability and migration of breast cancer cells exposed to CM from DNA-damaging drugs: (<b>A</b>) mitochondrial membrane potential (Δψm), (<b>B</b>) and mitochondrial superoxide production in MCF-7 cells exposed to CM from stromal cells for 24 h. FCCP (1 µM) and menadione (25 µM) were used as positive controls. (<b>C</b>) Vulnerability of MCF-7 cells exposed to CM from stromal cells for 48 h in the presence of a low concentration of menadione (0.5 µM), (<b>D</b>) diagram of mitochondrial utilization of energy substrates as pyruvate (which can be derived from glucose or lactate and incorporated into mitochondrion by mitochondrial pyruvate carrier, MPC, in IMM) and glutaminolysis, showing two inhibitors used in this study, (<b>E</b>) effect of MPC (5 µM UK5099) and glutamate dehydrogenase (25 µM EGCG) inhibitors on viability of MCF-7 cells exposed to CM from stromal cells treated with DNA-damaging drugs for 48 h, (<b>F</b>–<b>H</b>) effect of 25 µM EGCG and 5 µM UK5099 on the MCF-7, ZR75-1, and T47D cell migration stimulated by CM from stromal cells treated with doxorubicin (18 nM Doxo), (<b>I</b>) effect of rotenone (10 nM and 1 µM Rot, Complex I inhibitor) on MCF-7 migration. Data are shown as the mean ± SD, N = 4 independent experiments. * <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, vs. control (DMSO). n.s.: not significant.</p>
Full article ">Figure 6
<p>DNA-damaging drugs promote enhanced monocyte THP-1 recruitment: (<b>A</b>) diagram of experimental evaluation of conditioned media (CM) from fibroblasts RMF-621 treated with DNA-damaging drugs on migration, (<b>B</b>) effect of CM on THP-1 migration, (<b>C</b>) effect of CM produced by previous incubation (1 h) of mitoTEMPO (1 µM) and then DNA-damaging drugs on THP-1 migration, (<b>D</b>) effect of CM produced by DNA-damaging drugs and then the addition of blocking antibody anti-CCL2 (10 µg/mL) on THP-1 migration, (<b>E</b>) working model proposed in this work. Data are shown as the mean ± SD, N = 4–6 independent experiments. * <span class="html-italic">p</span> &lt; 0.05, *** <span class="html-italic">p</span> &lt; 0.001, vs. control (DMSO). n.s.: not significant.</p>
Full article ">
15 pages, 7025 KiB  
Article
Expression, Characterization, and Immobilization of a Novel D-Lactate Dehydrogenase from Salinispirillum sp. LH 10-3-1
by Jianguo Liu, Xuejiao Jiang, Yaru Zheng, Kaixuan Li, Ruixin Zhang, Jingping Xu, Zhe Wang, Yuxuan Zhang, Haoran Yin and Jing Li
Processes 2024, 12(7), 1349; https://doi.org/10.3390/pr12071349 - 28 Jun 2024
Viewed by 487
Abstract
Salinispirillum sp. LH 10-3-1 was newly isolated from the alkali lake water samples collected in Inner Mongolia. In this study, a gene coding for D-lactate dehydrogenase from the strain LH 10-3-1 (SaLDH) was cloned and characterized. The recombinant enzyme was a [...] Read more.
Salinispirillum sp. LH 10-3-1 was newly isolated from the alkali lake water samples collected in Inner Mongolia. In this study, a gene coding for D-lactate dehydrogenase from the strain LH 10-3-1 (SaLDH) was cloned and characterized. The recombinant enzyme was a tetramer with a native molecular mass of 146.2 kDa. The optimal conditions for SaLDH to reduce pyruvate and oxidize D-lactic acid were pH 8.0 and pH 5.0, at 25 °C. Cu2+ and Ca2+ slightly promoted the oxidation and reduction activities of SaLDH, respectively. To improve the stability of SaLDH, the enzyme was immobilized on Cu3(PO4)2-based inorganic hybrid nanoflowers. The results showed that the reduction activity of the hybrid nanoflowers disappeared, and the optimum temperature, specific activity, thermostability, and storage stability of the immobilized SaLDH were significantly improved. In addition, the biotransformation of D-lactic acid to pyruvate catalyzed by SaLDH and the hybrid nanoflowers was investigated. The maximum conversion of D-lactic acid catalyzed by the immobilized SaLDH was 25.7% higher than by free enzymes, and the immobilized SaLDH could maintain 84% of its initial activity after six cycles. Full article
Show Figures

Figure 1

Figure 1
<p>Alignment of <span class="html-italic">Sa</span>LDH from <span class="html-italic">Salinispirillum</span> sp. LH10-3-1 with LDH of <span class="html-italic">L. mesenteriodes</span> and <span class="html-italic">P. claussennii</span> based on the amino acid sequences. The LDH from <span class="html-italic">Pseudomonas aeruginosa</span> (PDB ID 6AJB) is used as top secondary structure. The nucleotide-binding signature domain GXGXXG is marked with blue square.</p>
Full article ">Figure 2
<p>Phylogenetic tree of <span class="html-italic">Sa</span>LDH, D-lactate dehydrogenases of the most related relatives and functional verified microorganisms based on their amino acid sequences. The phylogenetic tree was constructed using the neighbor-joining method with 1000 bootstrap replicates (using MEGA 11.0 software).</p>
Full article ">Figure 3
<p>(<b>a</b>) SDS-PAGE analysis of <span class="html-italic">Sa</span>LDH samples under reducing condition. M, protein marker; I, crude enzyme solution from induced cells; II, purified recombinant <span class="html-italic">Sa</span>LDH. (<b>b</b>) Molecular mass estimation of the native <span class="html-italic">Sa</span>LDH based on the gel filtration analysis using a Superdex 200 10/300 GL column. Four standard proteins were apoferritin (443 kDa), glucose oxidase (160 kDa), bovine serum albumin (66.4 kDa), and carbonic anhydrase (29 kDa).</p>
Full article ">Figure 4
<p>Enzymatic properties of recombinant <span class="html-italic">Sa</span>LDH. (<b>a</b>) Effect of pH on reduction and oxidation activity. (<b>b</b>) Effect of temperature on reduction and oxidation activity. Activity was measured at 15–60 °C for 15 min. (<b>c</b>) Thermostability of <span class="html-italic">Sa</span>LDH. (<b>d</b>) Effect of NaCl concentration on reduction and oxidation activity.</p>
Full article ">Figure 4 Cont.
<p>Enzymatic properties of recombinant <span class="html-italic">Sa</span>LDH. (<b>a</b>) Effect of pH on reduction and oxidation activity. (<b>b</b>) Effect of temperature on reduction and oxidation activity. Activity was measured at 15–60 °C for 15 min. (<b>c</b>) Thermostability of <span class="html-italic">Sa</span>LDH. (<b>d</b>) Effect of NaCl concentration on reduction and oxidation activity.</p>
Full article ">Figure 5
<p>SEM images (<b>a</b>,<b>b</b>) of <span class="html-italic">Sa</span>LDH/Cu<sub>3</sub>(PO<sub>4</sub>)<sub>2</sub> hybrid nanoflowers (Cu-<span class="html-italic">Sa</span>LDH HNF). FT-IR spectra (<b>c</b>) and XRD patterns (<b>d</b>) of free <span class="html-italic">Sa</span>LDH, Cu<sub>3</sub>(PO<sub>4</sub>)<sub>2</sub> particles, and Cu-<span class="html-italic">Sa</span>LDH HNF.</p>
Full article ">Figure 6
<p>Effects of pH (<b>a</b>) and temperature (<b>b</b>) on enzyme activity of the free <span class="html-italic">Sa</span>LDH and Cu-SaLDH HNF. Thermostability (<b>c</b>) and storage stability (<b>d</b>) of the free <span class="html-italic">Sa</span>LDH and Cu-<span class="html-italic">Sa</span>LDH HNF.</p>
Full article ">Figure 7
<p>(<b>a</b>) Time course of the biotransformation of D-lactic acid to pyruvate catalyzed by the free <span class="html-italic">Sa</span>LDH and Cu-<span class="html-italic">Sa</span>LDH HNF. Reaction conditions for using free <span class="html-italic">Sa</span>LDH: pH 5.0, 25 °C, enzyme content: 0.1 mg/mL; Reaction conditions for using Cu-<span class="html-italic">Sa</span>LDH HNF: pH 8.0, 50 °C, enzyme content: 0.1 mg/mL. (<b>b</b>) Reusability of Cu-<span class="html-italic">Sa</span>LDH HNF for catalyzing the biotransformation of D-lactic acid to pyruvate.</p>
Full article ">
19 pages, 4982 KiB  
Article
Skeletal Muscle Metabolism Is Dynamic during Porcine Postnatal Growth
by Linnea A. Rimmer, Erika R. Geisbrecht, Michael D. Chao, Travis G. O’Quinn, Jason C. Woodworth and Morgan D. Zumbaugh
Metabolites 2024, 14(7), 357; https://doi.org/10.3390/metabo14070357 - 26 Jun 2024
Viewed by 963
Abstract
Skeletal muscle metabolism has implications for swine feed efficiency (FE); however, it remains unclear if the metabolic profile of skeletal muscle changes during postnatal growth. To assess the metabolic changes, samples were collected from the longissimus dorsi (LD, glycolytic muscle), latissimus dorsi (LAT, [...] Read more.
Skeletal muscle metabolism has implications for swine feed efficiency (FE); however, it remains unclear if the metabolic profile of skeletal muscle changes during postnatal growth. To assess the metabolic changes, samples were collected from the longissimus dorsi (LD, glycolytic muscle), latissimus dorsi (LAT, mixed muscle), and masseter (MS, oxidative muscle) at 20, 53, 87, 120, and 180 days of age from barrows. Muscles were assessed to determine the abundance of several metabolic enzymes. Lactate dehydrogenase (LDHα) decreased in all muscles from 20 to 87 d (p < 0.01), which may be attributed to the muscles being more glycolytic at weaning from a milk-based diet. Pyruvate carboxylase (PC) increased in all muscles at 53 d compared to the other time points (p < 0.01), while pyruvate dehydrogenase α 1 (PDHα1) increased at 87 and 180 d in MS compared to LD (p < 0.05), indicating that potential changes occur in pyruvate entry into the tricarboxylic acid (TCA) cycle during growth. Isolated mitochondria from each muscle were incubated with 13C-labeled metabolites to assess isotopomer enrichment patterns of TCA intermediates. Citrate M + 2 and M + 4 derived from [13C3]-pyruvate increased at 87 d in LAT and MS mitochondria compared to LD mitochondria (p < 0.05). Regardless of the muscle, citrate M+3 increased at 87 d compared to 20, 53, and 120 d, while 180 d showed intermediate values (p < 0.01). These data support the notion that pyruvate metabolism is dynamic during growth. Our findings establish a metabolic fingerprint associated with postnatal muscle hypertrophy. Full article
(This article belongs to the Special Issue Unlocking the Mysteries of Muscle Metabolism in the Animal Sciences)
Show Figures

Figure 1

Figure 1
<p>Growth characteristics of pigs at each age point. (<b>A</b>) Body weight of pigs at 20, 53, 87, 120, and 180 d. (<b>B</b>) Muscle cross-sectional area (CSA) of LD, LAT, and LD muscles collected from pigs at each age point. Data are shown as means ± SE. Five barrows per age group (n = 5). Means lacking a common letter (a, b, c, d, e) differed within a time point (<span class="html-italic">p</span> &lt; 0.05). When the interaction was significant (muscle × age), means lacking a common letter (a, b) differ within a time point and (X, Y, Z) differ between ages.</p>
Full article ">Figure 2
<p>Relative abundance of enzymes involved in pyruvate metabolism. (<b>A</b>,<b>B</b>) Relative abundance of lactate dehydrogenase alpha (LDHα) in each (<b>A</b>) muscle and (<b>B</b>) age group. (<b>C</b>) Representative images of Western blots quantified in (<b>A</b>,<b>B</b>). (<b>D</b>) Relative abundance of pyruvate dehydrogenase alpha 1 (PDHα1). (<b>E</b>) Representative images of Western blots quantified in (<b>D</b>). (<b>F</b>–<b>H</b>) Relative abundance of pyruvate carboxylase (PC) in each (<b>F</b>) muscle and (<b>G</b>) age group. (<b>H</b>) Representative images of Western blots quantified in (<b>F</b>,<b>G</b>). Data are shown as means ± SE. Five barrows per age group (n = 5). Means lacking a common letter differed (a, b) within a time point (<span class="html-italic">p</span> &lt; 0.05).</p>
Full article ">Figure 3
<p>Relative abundance of enzymes involved in citrate metabolism. (<b>A</b>,<b>B</b>) Relative abundance of citrate synthase (CS) in each (<b>A</b>) muscle and (<b>B</b>) age group. (<b>C</b>) Representative images of Western blots quantified in (<b>A</b>,<b>B</b>). (<b>D</b>,<b>E</b>) Relative abundance of mitochondria citrate carrier (MCC) in each (<b>D</b>) muscle and (<b>E</b>) age group. (<b>F</b>) Representative images of Western blots quantified in (<b>D</b>,<b>E</b>). Data are shown as means ± SE. Five barrows per age group (n = 5). Means lacking a common letter (a, b, c) differed within a time point (<span class="html-italic">p</span> &lt; 0.05).</p>
Full article ">Figure 4
<p>Relative abundance of enzymes involved in amino acid metabolism. (<b>A</b>,<b>B</b>) Relative abundance of glutamate dehydrogenase (GDH) in each (<b>A</b>) muscle and (<b>B</b>) age group. (<b>C</b>) Representative images of Western blots quantified in (<b>A</b>,<b>B</b>). (<b>D</b>,<b>E</b>) Relative abundance of glutamic-oxaloacetic transaminase (GOT2) in each (<b>D</b>) muscle and (<b>E</b>) age group. (<b>F</b>) Representative images of Western blots quantified in (<b>D</b>,<b>E</b>). Data are shown as means ± SE. Five barrows per age group (n = 5). Means lacking a common letter (a, b) differed within a time point (<span class="html-italic">p</span> &lt; 0.05).</p>
Full article ">Figure 5
<p>Schematic of potential labeling patterns of TCA intermediates after isolated mitochondria were incubated with [<sup>13</sup>C<sub>3</sub>]-pyruvate. Black circles are <sup>13</sup>C derived from [<sup>13</sup>C<sub>3</sub>]-pyruvate that entered the TCA cycle through PDH. Grey circles indicate <sup>13</sup>C derived from [<sup>13</sup>C<sub>3</sub>]-pyruvate that entered the TCA cycle through PC. White circles represent unlabeled carbon atoms (<sup>12</sup>C).</p>
Full article ">Figure 6
<p>[<sup>13</sup>C<sub>3</sub>]-pyruvate derived isotopomer enrichment of oxaloacetate in isolated mitochondria. (<b>A</b>–<b>C</b>) Enrichment of (<b>A</b>) M + 2, (<b>B</b>) M + 3, and (<b>C</b>) M + 4 oxaloacetate in LD, LAT, and LD mitochondria isolated from pigs at each age point. Data are shown as means ± SE. Mitochondria were isolated from LD, LAT, and MS muscles of five barrows per age group (n = 5). Means lacking a common letter (a, b) differed within a time point (<span class="html-italic">p</span> &lt; 0.05). When the interaction was significant (muscle × age), means lacking a common letter (a, b) differ within a time point and (X, Y) differ between ages.</p>
Full article ">Figure 7
<p>[<sup>13</sup>C<sub>3</sub>]-pyruvate derived isotopomer enrichments of citrate in isolated mitochondria. (<b>A</b>–<b>C</b>) Enrichment of (<b>A</b>) M + 2, (<b>B</b>) M + 3, and (<b>C</b>) M + 4 citrate in LD, LAT, and LD mitochondria isolated from pigs at each age point. Data are shown as means ± SE. Mitochondria were isolated from LD, LAT, and MS muscles of five barrows per age group (n = 5). Means lacking a common letter (a, b, c) differed within a time point (<span class="html-italic">p</span> &lt; 0.05). When the interaction was significant (muscle × age), means lacking a common letter (a, b, c) differ within a time point and (X, Y, Z) differ between ages.</p>
Full article ">Figure 8
<p>[<sup>13</sup>C<sub>3</sub>]-pyruvate derived isotopomer enrichments of α-ketoglutarate in isolated mitochondria. (<b>A</b>–<b>C</b>) Enrichment of (<b>A</b>) M + 2, (<b>B</b>) M + 3, and (<b>C</b>) M + 4 α-ketoglutarate in LD, LAT, and LD mitochondria isolated from pigs at each age point. Data are shown as means ± SE. Mitochondria were isolated from LD, LAT, and MS muscles of five barrows per age group (n = 5). Means lacking a common letter (a, b) differed within a time point (<span class="html-italic">p</span> &lt; 0.05). When the interaction was significant (muscle × age), means lacking a common letter (a, b) differ within a time point and (X, Y, Z) differ between ages.</p>
Full article ">Figure 9
<p>(<b>A</b>) Schematic of potential labeling pattern of α-ketoglutarate from glutamate after isolated mitochondria were incubated with [<sup>13</sup>C<sub>5</sub>]-glutamate. Black circles are <sup>13</sup>C derived from [<sup>13</sup>C<sub>5</sub>]-glutamate that entered the TCA cycle as α-ketoglutarate. (<b>B</b>) [<sup>13</sup>C<sub>5</sub>]-glutamate-derived isotopomer enrichments of α-ketoglutarate in isolated mitochondria. Enrichment of (B) M + 5 α-ketoglutarate in LD, LAT, and LD mitochondria isolated from pigs at each age point. Data are shown as means ± SE. Mitochondria were isolated from LD, LAT, and MS muscles of five barrows per age group (n = 5). Means lacking a common letter (a, b, c) differed across the age and muscle groups (<span class="html-italic">p</span> &lt; 0.05).</p>
Full article ">
14 pages, 3984 KiB  
Article
Effects of Foliar Ca and Mg Nutrients on the Respiration of ‘Feizixiao’ Litchi Pulp and Identification of Differential Expression Genes Associated with Respiration
by Muhammad Sajjad, Hassam Tahir, Wuqiang Ma, Shi Shaopu, Muhammad Aamir Farooq, Muhammad Zeeshan Ul Haq, Shoukat Sajad and Kaibing Zhou
Agronomy 2024, 14(7), 1347; https://doi.org/10.3390/agronomy14071347 - 21 Jun 2024
Viewed by 593
Abstract
The ‘Feizixiao’ litchi cultivar, predominantly grown in Hainan Province, faces the issue of “sugar receding” during fruit ripening. The application of mixed foliar nutrients containing calcium and magnesium (Ca+Mg) during the fruit pericarp’s full coloring stage was investigated to overcome this issue. Experimental [...] Read more.
The ‘Feizixiao’ litchi cultivar, predominantly grown in Hainan Province, faces the issue of “sugar receding” during fruit ripening. The application of mixed foliar nutrients containing calcium and magnesium (Ca+Mg) during the fruit pericarp’s full coloring stage was investigated to overcome this issue. Experimental trials unveiled significant alterations in litchi pulp physiochemical properties, including the main nutrient and flavor quality, the total respiration rates of the main respiratory pathways, and the activities of some important enzymes associated with Embden–Meyerhof–Parnas (EMP), the tricarboxylic acid cycle (TCA) and the pentose phosphate pathway (PPP). The Ca+Mg treatment showed higher sugar levels than the control (CK) during ripening. Notably, the application of Ca+Mg in litchi pulp inhibited respiration rates through the EMP, TCA, and PPP pathways, resulting in a strong effect. RNA sequencing analysis revealed the impact of Ca+Mg treatment on respiratory pathways, revealing differentially expressed genes (DEGs) such as pyruvate PK1, PK2 (pyruvate kinase), and PDC (pyruvate dehydrogenase complex), validated through qRT-PCR with a significant correlation to RNA-seq results. In general, Ca+Mg treatment during litchi fruit ripening overcame “sugar receding” by inhibiting the expression of respiration key metabolic pathway genes. These findings provide insights for enhancing cultivation management strategies. Full article
(This article belongs to the Section Horticultural and Floricultural Crops)
Show Figures

Figure 1

Figure 1
<p>The effect of spraying calcium and magnesium fertilizer on litchi pulp as indicated by days after anthesis on the X-axis and respective sugar and acid contents on the Y-axis. (<b>a</b>) Total sugar, (<b>b</b>) total acid, and (<b>c</b>) sugar–acid ratio. Different letters represent the different kind of analysis in each figure and each portion.</p>
Full article ">Figure 2
<p>The effect of spraying calcium and magnesium fertilizer on litchi pulp as indicated by days after anthesis (DAA) on the x-axis and respective enzyme activities on the y-axis. (<b>a</b>) Total respiration rate: (<b>b</b>) EMP, (<b>c</b>) TCA, and (<b>d</b>) PPP. Different letters represent the different kind of analysis in each figure and each portion.</p>
Full article ">Figure 3
<p>The effect of spraying calcium and magnesium fertilizer on litchi pulp as indicated by days after anthesis (DAA) the X-axis and respective enzyme activities on the Y-axis. (<b>a</b>) SDH activity, (<b>b</b>) PDC activity, (<b>c</b>) GPI activity, (<b>d</b>) PK activity, (<b>e</b>) NAD-MDH activity, and (<b>f</b>) CCO activity. Different letters represent the different kind of analysis in each figure and each portion.</p>
Full article ">Figure 4
<p>Respiratory pathways map (EMP, TCA, and PPP). The ECs (enzyme numbers) marked by red boxes are related to up-regulated genes, and green boxes are related to down-regulated genes, whereas the ECs marked with blue boxes are related to both up-regulated and down-regulated genes.</p>
Full article ">Figure 5
<p>Differential gene expression heatmap of respiratory pathways indicating some important gene families with their respective ECs.</p>
Full article ">Figure 6
<p>Differentially expressed genes upon qRT-PCR validation related to respiratory pathways: (<b>a</b>) <span class="html-italic">PK1</span>, (<b>b</b>) <span class="html-italic">PK2</span>, and (<b>c</b>) <span class="html-italic">PDC</span> refer to real-time PCR trends from samples taken at 35d, 63d, and 70d of Ca+Mg treatment and CK. Different letters represent the different kind of analysis in each figure and each portion.</p>
Full article ">
12 pages, 3249 KiB  
Article
The Pathogenic Mechanism of Enterocytozoon hepatopenaei in Litopenaeus vannamei
by Rongrong Ma, Bo Zhu, Jinbo Xiong and Jiong Chen
Microorganisms 2024, 12(6), 1208; https://doi.org/10.3390/microorganisms12061208 - 15 Jun 2024
Viewed by 909
Abstract
Enterocytozoon hepatopenaei (EHP) is a parasite in shrimp farming. EHP mainly parasitizes the hepatopancreas of shrimp, causing slow growth, which severely restricts the economic income of shrimp farmers. To explore the pathogenic mechanism of EHP, the host subcellular construction, molecular biological characteristics, and [...] Read more.
Enterocytozoon hepatopenaei (EHP) is a parasite in shrimp farming. EHP mainly parasitizes the hepatopancreas of shrimp, causing slow growth, which severely restricts the economic income of shrimp farmers. To explore the pathogenic mechanism of EHP, the host subcellular construction, molecular biological characteristics, and mitochondrial condition of Litopenaeus vannamei were identified using transmission electron microscopy (TEM), real-time qPCR, an enzyme assay, and flow cytometry. The results showed that EHP spores, approximately 1 μm in size, were located on the cytoplasm of the hepatopancreas. The number of mitochondria increased significantly, and mitochondria morphology showed a condensed state in the high-concentration EHP-infected shrimp by TEM observation. In addition, there were some changes in mitochondrial potential, but apoptosis was not significantly different in the infected shrimp. The qPCR results showed that the gene expression levels of hexokinase and pyruvate kinase related to energy metabolism were both upregulated in the diseased L. vannamei. Enzymatic activity showed hexokinase and lactate dehydrogenase were significantly increased in the shrimp infected with EHP, indicating EHP infection can increase the glycolysis process and decrease the oxidative phosphorylation process of L. vannamei. Previous transcriptomic data analysis results also support this conclusion. Full article
(This article belongs to the Special Issue Current Insights into Host–Parasite Interactions)
Show Figures

Figure 1

Figure 1
<p>Shrimp with different concentrations of EHP infection. Note: A is the shrimp sample with a high concentration of EHP infection; B is the shrimp sample with a low concentration of EHP infection.</p>
Full article ">Figure 2
<p>Flow chart of glucose metabolism. Note: The text marked in red indicates the enzyme activity detected in this study.</p>
Full article ">Figure 3
<p>Histopathological analysis of hepatopancreatic tissues at different concentrations of EHP infection by transmission electron microscopy. Notes: (<b>A</b>,<b>B</b>) represent low-concentration EHP infection groups; (<b>C</b>,<b>D</b>) represent high-concentration EHP infection groups; NC means cell nucleus; red arrows represent EHP; white arrows represent mitochondria.</p>
Full article ">Figure 4
<p>HK and PK gene expression in the hepatopancreas of <span class="html-italic">L. vannamei</span>. Notes: HK represents the hexokinase gene; PK represents the pyruvate kinase gene; *** means <span class="html-italic">p</span> ≤ 0.001.</p>
Full article ">Figure 5
<p>Enzyme activity analysis in the infected shrimp in the first week. Notes: The control group represents the healthy group (orange ) and the experiment group represents the artificial infection group (blue); ** means <span class="html-italic">p</span> ≤ 0.01, *** means <span class="html-italic">p</span> ≤ 0.001; HK is hexokinase, PK is pyruvate kinase, SDH is succinate dehydrogenase, CCO is cytochrome c oxidase, and LDH is lactate dehydrogenase.</p>
Full article ">Figure 6
<p>Enzyme activity analysis in the infected shrimp in the second week. Notes: The relevant annotations are consistent with those in <a href="#microorganisms-12-01208-f005" class="html-fig">Figure 5</a>; ** means <span class="html-italic">p</span> ≤ 0.01, *** means <span class="html-italic">p</span> ≤ 0.001.</p>
Full article ">Figure 7
<p>DEGs involved in oxidative phosphorylation pathway in group C vs. group A. Notes: Red indicates upregulated DEGs; dark green represents downregulated DEGs.</p>
Full article ">Figure 8
<p>The analysis of cell apoptosis using flow cytometry.</p>
Full article ">
13 pages, 1638 KiB  
Article
Antioxidant and Metabolic Response to Acute Acidification Stress of Juvenile Yellowfin Tuna (Thunnus albacares)
by Xiaoyan Wang, Rui Yang, Zhengyi Fu, Lei Zhao and Zhenhua Ma
J. Mar. Sci. Eng. 2024, 12(6), 970; https://doi.org/10.3390/jmse12060970 - 8 Jun 2024
Viewed by 653
Abstract
This study aimed to explore the impact of acute acidification on the antioxidant, metabolic performance, and liver histology of juvenile yellowfin tuna. The experiment subjected juvenile yellowfin tuna to a pH gradient environment of 8.1, 7.6, 7.1, and 6.6 for 48 h. The [...] Read more.
This study aimed to explore the impact of acute acidification on the antioxidant, metabolic performance, and liver histology of juvenile yellowfin tuna. The experiment subjected juvenile yellowfin tuna to a pH gradient environment of 8.1, 7.6, 7.1, and 6.6 for 48 h. The findings indicate that a seawater pH of 7.1 significantly impacts the antioxidant and metabolic systems of the juvenile yellowfin tuna in comparison to the control group. At pH 7.1, there were observed increases in glutathione reductase (GR), total antioxidant capacity (T-AOC), lactate dehydrogenase (LDH), hexokinase (HK), pyruvate kinase (PK), sodium-potassium ATPase (Na+K+-ATP), and calcium-magnesium ATPase (Ca2+Mg2+-ATP). Conversely, low-density lipoprotein cholesterol (LDL-C), high-density lipoprotein cholesterol (HDL-C), and triglycerides (TGs) were not significantly different across the treatment groups. However, an increase in transaminases at pH 7.1 suggested potential liver damage, which was further supported by observed structural liver tissue degeneration and hepatocyte vacuolation. In conclusion, under conditions of acute acidification stress, there is a decrease in antioxidant capacity and a suppression of metabolic levels in juvenile yellowfin tuna, leading to oxidative damage. This study lays the foundation for an in-depth understanding of the response mechanisms of juvenile yellowfin tuna in response to seawater acidification as well as healthy tuna farming in the broader context of seawater acidification. Full article
(This article belongs to the Special Issue New Techniques and Equipment in Large Offshore Aquaculture Platform)
Show Figures

Figure 1

Figure 1
<p>Effect of acute acidification stress on antioxidant parameters in juvenile yellowfin tuna liver. The bar graph represents the mean ± SD of the measurements taken at 48 h. (<b>a</b>) Glutathione reductase (GR) activity (ANOVA, GR: F = 10.968, <span class="html-italic">p</span> &lt; 0.05), (<b>b</b>) total antioxidant (T-AOC) capacity (ANOVA, T-AOC: F = 2.396, <span class="html-italic">p</span> &gt; 0.05), and (<b>c</b>) lipid peroxidation (LPO) (ANOVA, LPO: F = 10.714, <span class="html-italic">p</span> &lt; 0.05). Different letters on the columns indicate significant differences between groups (<span class="html-italic">p</span> &lt; 0.05), and the same letters indicate non-significant differences between groups (<span class="html-italic">p</span> &gt; 0.05).</p>
Full article ">Figure 2
<p>Effect of acute acidification stress on metabolic enzymes in juvenile yellowfin tuna liver. The bar graph represents the mean ± SD of the measurements taken at 48 h. (<b>a</b>) Lactate dehydrogenase (LDH) activity (ANOVA, LDH: F = 24.458, <span class="html-italic">p</span> &lt; 0.05), (<b>b</b>) hexokinase (HK) activity (ANOVA, HK: F = 12.447, <span class="html-italic">p</span> &lt; 0.05), (<b>c</b>) pyruvate kinase (PK) activity (ANOVA, PK: F = 28.622, <span class="html-italic">p</span> &lt; 0.05), (<b>d</b>) sodium-potassium ATPase (Na<sup>+</sup>K<sup>+</sup>-ATP) activity (ANOVA, Na<sup>+</sup>K<sup>+</sup>-ATP: F = 25.508, <span class="html-italic">p</span> &lt; 0.05), and (<b>e</b>) calcium-magnesium ATPase (Ca<sup>2+</sup>Mg<sup>2+</sup>-ATP) activity (ANOVA, Ca<sup>2</sup><sup>+</sup>Mg<sup>2</sup><sup>+</sup>-ATP: F = 347.865, <span class="html-italic">p</span> &lt; 0.05). Different letters on the columns indicate significant differences between groups (<span class="html-italic">p</span> &lt; 0.05), and the same letters indicate non-significant differences between groups (<span class="html-italic">p</span> &gt; 0.05).</p>
Full article ">Figure 3
<p>Effect of acidification stress on serum indices of juvenile yellowfin tuna. The bar graph represents the mean ± SD of the measurements taken at 48 h. (<b>a</b>) Glucose (GLU) content (ANOVA, GLU: F = 148.637, <span class="html-italic">p</span> &lt; 0.05), (<b>b</b>) low-density lipoprotein cholesterol (LDL-C) content (ANOVA, LDH-C: F = 3.385, <span class="html-italic">p</span> &gt; 0.05), (<b>c</b>) high-density lipoprotein cholesterol (HDL-C) content (ANOVA, HDH-C: F = 1.148, <span class="html-italic">p</span> &gt; 0.05), (<b>d</b>) triglycerides (TGs) content (ANOVA, TG: F = 2.688, <span class="html-italic">p</span> &gt; 0.05), (<b>e</b>) total cholesterol (TCH) content (ANOVA, TCH: F = 1.209, <span class="html-italic">p</span> &gt; 0.05), (<b>f</b>) glutamic oxaloacetic transaminase (GOT) activity (ANOVA, GOT: F = 57.086, <span class="html-italic">p</span> &lt; 0.05), (<b>g</b>) glutamine pyruvate transaminase (GPT) activity (ANOVA, GPT: F = 4.197, <span class="html-italic">p</span> &lt; 0.05), and (<b>h</b>) alkaline phosphatase (AKP) activity (ANOVA, AKP: F = 18.039, <span class="html-italic">p</span> &lt; 0.05). Different letters on the columns indicate significant differences between groups (<span class="html-italic">p</span> &lt; 0.05), and the same letters indicate non-significant differences between groups (<span class="html-italic">p</span> &gt; 0.05).</p>
Full article ">Figure 4
<p>Effect of acute acidification stress of yellowfin tuna on liver histology (400×). (<b>a</b>) The pH 8.1 treatment group (control), (<b>b</b>) the pH 7.6 treatment group, (<b>c</b>) the pH 7.1 treatment group, and (<b>d</b>) the pH 6.6 treatment group. Red arrows indicate hepatocyte nuclei; red circles indicate blurred hepatocyte structures; yellow arrows indicate sinusoidal gaps; yellow circles indicate loss of nuclei; green arrows indicate rounded vacuoles; green circles indicate abnormal cellular morphology; and blue circles indicate nuclear excursions.</p>
Full article ">
Back to TopTop