[go: up one dir, main page]

 
 
Sign in to use this feature.

Years

Between: -

Subjects

remove_circle_outline
remove_circle_outline
remove_circle_outline
remove_circle_outline
remove_circle_outline
remove_circle_outline
remove_circle_outline
remove_circle_outline
remove_circle_outline

Journals

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

Article Types

Countries / Regions

remove_circle_outline
remove_circle_outline
remove_circle_outline
remove_circle_outline
remove_circle_outline
remove_circle_outline

Search Results (5,898)

Search Parameters:
Keywords = SOD

Order results
Result details
Results per page
Select all
Export citation of selected articles as:
13 pages, 3371 KiB  
Article
Safflower Yellow Injection Alleviates Myocardial Ischemia/Reperfusion Injury by Reducing Oxidative and Endoplasmic Reticulum Stress
by Wulin Liang, Mingqian Zhang, Jiahui Gao, Rikang Huang, Lu Cheng, Liyuan Zhang, Zhishan Huang, Zhanhong Jia and Shuofeng Zhang
Pharmaceuticals 2024, 17(8), 1058; https://doi.org/10.3390/ph17081058 (registering DOI) - 12 Aug 2024
Abstract
Safflower yellow is an extract of the famous Chinese medicine Carthamus tinctorious L, and safflower yellow injection (SYI) is widely used clinically to treat angina pectoris. However, there are few studies on the anti-myocardial ischemia/reperfusion (I/R) injury effect of SYI, and its mechanisms [...] Read more.
Safflower yellow is an extract of the famous Chinese medicine Carthamus tinctorious L, and safflower yellow injection (SYI) is widely used clinically to treat angina pectoris. However, there are few studies on the anti-myocardial ischemia/reperfusion (I/R) injury effect of SYI, and its mechanisms are unclear. In the present study, we aimed to investigate the protective effect of SYI on myocardial I/R injury and explore its underlying mechanisms. Male Sprague Dawley rats were randomly divided into a control group, sham group, model group, and SYI group (20 mg/kg, femoral vein injection 1 h before modeling). The left anterior descending coronary artery was ligated to establish a myocardial I/R model. H9c2 cells were exposed to oxygen–glucose deprivation/reoxygenation (OGD/R) after incubation with 80 μg/mL SYI for 24 h. In vivo, TsTC, HE, and TUNEL staining were performed to evaluate myocardial injury and apoptosis. A kit was used to detect superoxide dismutase (SOD) and malondialdehyde (MDA) to assess oxidative stress. In vitro, flow cytometry was used to detect the reactive oxygen species (ROS) content and apoptosis rate. Protein levels were determined via Western blotting. Pretreatment with SYI significantly reduced infarct size and pathological damage in rat hearts and suppressed cardiomyocyte apoptosis in vivo and in vitro. In addition, SYI inhibited oxidative stress by increasing SOD activity and decreasing MDA content and ROS production. Myocardial I/R and OGD/R activate endoplasmic reticulum (ER) stress, as evidenced by increased expression of activating transcription factor 6 (ATF6), glucose-regulated protein 78 (GRP78), cysteinyl aspartate-specific proteinase caspase-12, and C/EBP-homologous protein (CHOP), which were all inhibited by SYI. SYI ameliorated myocardial I/R injury by attenuating apoptosis, oxidative damage, and ER stress, which revealed new mechanistic insights into its application. Full article
(This article belongs to the Section Pharmacology)
Show Figures

Graphical abstract

Graphical abstract
Full article ">Figure 1
<p>SYI reduced myocardial I/R injury in rats. (<b>A</b>) Representative photographs of TTC-stained heart slices. (<b>B</b>) Quantitative analysis of the infarct area (<span class="html-italic">n</span> = 6). (<b>C</b>) LDH activity in serum (<span class="html-italic">n</span> = 6). (<b>D</b>) HE staining shows pathological changes in the myocardium (200× magnification; scale bar is 100 μm). <sup>###</sup> <span class="html-italic">p</span> &lt; 0.001 compared with the sham group. ** <span class="html-italic">p</span> &lt; 0.01 and *** <span class="html-italic">p</span> &lt; 0.001 compared with the model group.</p>
Full article ">Figure 2
<p>SYI attenuated I/R-induced myocardial apoptosis in rats. (<b>A</b>) Representative TUNEL staining (400× magnification; scale bar is 20 μm). (<b>B</b>) Quantitative analysis of apoptosis (<span class="html-italic">n</span> = 3). (<b>C</b>) The expression levels of Bax and Bcl−2 were analyzed via Western blotting. (<b>D</b>,<b>E</b>) The relative protein expression of Bax and Bcl−2 (<span class="html-italic">n</span> = 3). <sup>#</sup> <span class="html-italic">p</span> &lt; 0.05 and <sup>###</sup> <span class="html-italic">p</span> &lt; 0.001 compared with the sham group. * <span class="html-italic">p</span> &lt; 0.05 and ** <span class="html-italic">p</span> &lt; 0.01 compared with the model group.</p>
Full article ">Figure 3
<p>SYI attenuated I/R-induced oxidative and ER stress in rat hearts. (<b>A</b>,<b>B</b>) The SOD activity and MDA content in the myocardium (<span class="html-italic">n</span> = 6). (<b>C</b>) The expression levels of ER stress-related proteins (ATF6, GRP78, caspase-12, and CHOP) were analyzed via Western blotting. (<b>D</b>–<b>G</b>) The relative protein expression of ATF6, GRP78, caspase-12, and CHOP (<span class="html-italic">n</span> = 3). <sup>###</sup> <span class="html-italic">p</span> &lt; 0.001 compared with the sham group. * <span class="html-italic">p</span> &lt; 0.05 and ** <span class="html-italic">p</span> &lt; 0.01 compared with the model group.</p>
Full article ">Figure 4
<p>SYI reduced OGD/R-induced injury in H9c2 cells. (<b>A</b>) Cell viability of H9c2 cells incubated with different concentrations of SYI for 24 h. (<b>B</b>) Effects of SYI on cell viability after OGD/R (<span class="html-italic">n</span> = 6). (<b>C</b>) Effects of SYI on OGD/R-induced LDH leakage (<span class="html-italic">n</span> = 3). <sup>###</sup> <span class="html-italic">p</span> &lt; 0.001 compared with the sham group. * <span class="html-italic">p</span> &lt; 0.05 and *** <span class="html-italic">p</span> &lt; 0.01 compared with the model group.</p>
Full article ">Figure 5
<p>SYI reduced OGD/R-induced apoptosis in H9c2 cells. (<b>A</b>) Hoechst 33342 staining showing the anti-apoptotic potential of SYI. (<b>B</b>) Apoptosis in H9c2 cells was analyzed via flow cytometry. (<b>C</b>) Quantitative analysis of the apoptotic rate (<span class="html-italic">n</span> = 3). (<b>D</b>) Caspase-3 activity (<span class="html-italic">n</span> = 3). (<b>E</b>) The expression levels of Bax and Bcl-2 were analyzed via Western blotting. (<b>F</b>,<b>G</b>) The relative protein expression of Bax and Bcl-2 (<span class="html-italic">n</span> = 3). <sup>##</sup> <span class="html-italic">p</span> &lt; 0.01 and <sup>###</sup> <span class="html-italic">p</span> &lt; 0.001 compared with the control group. * <span class="html-italic">p</span> &lt; 0.05 and *** <span class="html-italic">p</span> &lt; 0.001 compared with the model group.</p>
Full article ">Figure 6
<p>SYI attenuated OGD/R-induced oxidative and ER stress in H9c2 cells. (<b>A</b>) The ROS expression level was observed. (<b>B</b>) Quantitative analysis of ROS fluorescence intensity. (<b>C</b>,<b>D</b>) SOD activity and MDA contents in cells (<span class="html-italic">n</span> = 3). (<b>E</b>) The level of ER stress was assessed. (<b>F</b>) Quantitative analysis of ER-Tracker Red staining (<span class="html-italic">n</span> = 3). (<b>G</b>) The expression levels of ER stress-related proteins (ATF6, GRP78, caspase-12, and CHOP) were analyzed via Western blotting. (<b>H</b>–<b>K</b>) The relative protein expression of ATF6, GRP78, caspase-12, and CHOP (<span class="html-italic">n</span> = 3). <sup>#</sup> <span class="html-italic">p</span> &lt; 0.05 and <sup>###</sup> <span class="html-italic">p</span> &lt; 0.001 compared with the sham group. * <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 compared with the model group.</p>
Full article ">
16 pages, 8211 KiB  
Article
Pyruvate Carboxylase Attenuates Myocardial Ischemia–Reperfusion Injury in Heart Transplantation via Wnt/β-Catenin-Mediated Glutamine Metabolism
by Zihao Wang, Hongwen Lan, Yixuan Wang, Qiang Zheng, Chenghao Li, Kan Wang, Tixiusi Xiong, Qingping Wu and Nianguo Dong
Biomedicines 2024, 12(8), 1826; https://doi.org/10.3390/biomedicines12081826 (registering DOI) - 12 Aug 2024
Abstract
The ischemia–reperfusion process of a donor heart during heart transplantation leads to severe mitochondrial dysfunction, which may be the main cause of donor heart dysfunction after heart transplantation. Pyruvate carboxylase (PC), an enzyme found in mitochondria, is said to play a role in [...] Read more.
The ischemia–reperfusion process of a donor heart during heart transplantation leads to severe mitochondrial dysfunction, which may be the main cause of donor heart dysfunction after heart transplantation. Pyruvate carboxylase (PC), an enzyme found in mitochondria, is said to play a role in the control of oxidative stress and the function of mitochondria. This research examined the function of PC and discovered the signaling pathways controlled by PC in myocardial IRI. We induced IRI using a murine heterotopic heart transplantation model in vivo and a hypoxia–reoxygenation cell model in vitro and evaluated inflammatory responses, oxidative stress levels, mitochondrial function, and cardiomyocyte apoptosis. In both in vivo and in vitro settings, we observed a significant decrease in PC expression during myocardial IRI. PC knockdown aggravated IRI by increasing MDA content, LDH activity, TUNEL-positive cells, serum cTnI level, Bax protein expression, and the level of inflammatory cytokines and decreasing SOD activity, GPX activity, and Bcl-2 protein expression. PC overexpression yielded the opposite findings. Additional research indicated that reducing PC levels could block the Wnt/β-catenin pathway and glutamine metabolism by hindering the movement of β-catenin to the nucleus and reducing the activity of complex I and complex II, as well as ATP levels, while elevating the ratios of NADP+/NADPH and GSSG/GSH. Overall, the findings indicated that PC therapy can shield the heart from IRI during heart transplantation by regulating glutamine metabolism through the Wnt/β-catenin pathway. Full article
(This article belongs to the Section Molecular and Translational Medicine)
Show Figures

Figure 1

Figure 1
<p>PC is significantly downregulated during myocardial IRI. (<b>A</b>) Enrichment analysis of top 10 GO terms for both the control and IRI groups. (<b>B</b>) Analysis of top 10 KEGG pathways in the control group and the IRI group. (<b>C</b>,<b>D</b>) Protein expression levels of PC in hearts subjected to IRI (<span class="html-italic">n</span> = 5). (<b>E</b>,<b>F</b>) Protein expression levels of PC in NRCMs subjected to H/R (<span class="html-italic">n</span> = 6). Data are shown as the means ± SD. *** <span class="html-italic">p</span> &lt; 0.001 for indicated comparisons.</p>
Full article ">Figure 2
<p>PC knockdown aggravates myocardial IRI in heart transplantation. (<b>A</b>,<b>B</b>) Protein levels of PC in cardiac tissue treated with AAV-sh-Con/AAV-sh-PC after IRI during heart transplantation (<span class="html-italic">n</span> = 6). The levels of (<b>C</b>) MDA content, (<b>D</b>) SOD activity, (<b>E</b>) GPX activity, and (<b>F</b>) LDH activity were detected in the hearts subjected to myocardial IRI (<span class="html-italic">n</span> = 5). (<b>G</b>,<b>H</b>) Heart inflammation was evaluated by H&amp;E staining at 20× magnification with a scale bar of 50 μm. (<b>I</b>–<b>L</b>) Serum levels of inflammatory cytokines such as IL-1β, IL-6, MCP-1, and TNF-α were measured in the heart after myocardial IRI (<span class="html-italic">n</span> = 5). (<b>M</b>) The levels of serum cTnI were evaluated in hearts following IRI (<span class="html-italic">n</span> = 5). (<b>N</b>,<b>O</b>) Fluorescent pictures from the TUNEL test were taken for every group. Apoptotic heart muscle cells were treated with TUNEL staining (red), while the nuclei of all heart muscle cells were stained with DAPI (blue). The apoptotic index was calculated by determining the percentage of cells undergoing apoptosis compared to the total number of cells. (<b>P</b>,<b>Q</b>) Fluorescent pictures showing the generation of reactive oxygen species in every category (magnified by 200 times; scale bar of 100 μm). ROS underwent DHE staining, appearing red, while DAPI was used to stain the nuclei of all cardiomyocytes, appearing blue. Data are shown as the means ± SD. * <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 for indicated comparisons.</p>
Full article ">Figure 3
<p>PC knockdown aggravates H/R injury in vitro. (<b>A</b>) Protein levels of PC expression in NRCMs transfected with NC siRNA or PC siRNA were measured after H/R (<span class="html-italic">n</span> = 6). (<b>B</b>) PC siRNA’s impact on NRCMs cell viability, (<b>C</b>) MDA content, (<b>D</b>) SOD activity, and (<b>E</b>) LDH activity was measured using kits (<span class="html-italic">n</span> = 5). (<b>F</b>) Western blotting was used to detect the protein levels of Bax and Bcl-2 (<span class="html-italic">n</span> = 3). Data are shown as the means ± SD. * <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 for indicated comparisons.</p>
Full article ">Figure 4
<p>PC knockdown inhibits Wnt/β-catenin pathway and glutamine metabolism during H/R. (<b>A</b>) The glutamine levels were measured in NRCMs transfected with either scramble siRNA or PC siRNA after exposure to H/R, with a sample size of 5. (<b>B</b>,<b>C</b>) Complex I and complex II activities were measured in all groups. (<b>D</b>–<b>F</b>) The levels of NADP+/NADPH, GSSG/GSH, and ATP content in NRCMs following H/R were measured using kits (<span class="html-italic">n</span> = 5). (<b>G</b>) Western blotting was used to detect the protein levels of β-catenin, c-Myc, and Cyclin D1 (<span class="html-italic">n</span> = 3). (<b>H</b>) The protein expressions of β-catenin in nuclei (<span class="html-italic">n</span> = 3). Data are shown as the means ± SD. * <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 for indicated comparisons. n.s. signifies no significant difference for indicated comparisons.</p>
Full article ">Figure 5
<p>PC overexpression activates Wnt/β-catenin pathway during H/R. (<b>A</b>,<b>C</b>) Protein levels of PC expression in NRCMs transfected with Ad-GFP or Ad-PC after exposure to H/R were measured (<span class="html-italic">n</span> = 6). (<b>B</b>,<b>D</b>–<b>F</b>) Western blotting was used to detect the protein levels of β-catenin, c-Myc, and Cyclin D1 (<span class="html-italic">n</span> = 3). (<b>G</b>–<b>J</b>) Ad-PC impact on NRCMs cell viability, MDA content, SOD activity, and LDH activity was measured using kits (<span class="html-italic">n</span> = 5). Data are shown as the means ± SD. * <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 for indicated comparisons.</p>
Full article ">Figure 6
<p>β-catenin upregulated GS expression by directly binding to its promoter. (<b>A</b>) Protein levels of GS and β-catenin were measured in NRCMs transfected with Ad-GFP or Ad-PC after H/R (<span class="html-italic">n</span> = 4). (<b>B</b>) Protein levels of GS and β-catenin were measured in NRCMs transfected with NC siRNA or β-catenin siRNA after H/R (<span class="html-italic">n</span> = 4). (<b>C</b>,<b>D</b>) GS mRNA expression (<span class="html-italic">n</span> = 9) in NRCMs. (<b>E</b>) A dual-luciferase assay was performed to assess β-catenin’s regulatory effect on the GS promoter. Data are shown as the means ± SD. ** <span class="html-italic">p</span> &lt; 0.01, *** <span class="html-italic">p</span> &lt; 0.001, and n.s. signifies no significant difference for indicated comparisons.</p>
Full article ">Figure 7
<p>Effect of MSAB administration on H/R injury after PC treatment. (<b>A</b>–<b>D</b>) Western blotting was used to detect the protein levels of β-catenin, Bax, and Bcl-2 (<span class="html-italic">n</span> = 3). (<b>E</b>,<b>F</b>) GS mRNA expression (<span class="html-italic">n</span> = 9) and glutamine level (<span class="html-italic">n</span> = 5) in NRCMs were evaluated with MSAB administration after PC treatment. (<b>G</b>–<b>I</b>) The levels of NADP+/NADPH, GSSG/GSH, and ATP content of NRCMs after H/R were measured by kits (<span class="html-italic">n</span> = 5). Data are shown as the means ± SD. * <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, and n.s. signifies no significant difference for indicated comparisons.</p>
Full article ">Figure 8
<p>PC overexpression alleviates myocardial IRI in heart transplantation. (<b>A</b>–<b>D</b>) The levels of MDA content, SOD activity, GPX activity, and LDH activity were detected in the hearts treated with AAV-GFP/AAV-PC after IRI during heart transplantation (<span class="html-italic">n</span> = 5). (<b>E</b>–<b>H</b>) Serum levels of inflammatory cytokines such as IL-1β, IL-6, MCP-1, and TNF-α were measured in the hearts after myocardial IRI (<span class="html-italic">n</span> = 5). (<b>I</b>) Inflammation in the cardiac tissue was evaluated using H&amp;E staining at a magnification of 20× with a scale bar of 50 μm. (<b>J</b>) Fluorescent pictures from the TUNEL test were taken for every group. Apoptotic heart muscle cells were treated with TUNEL staining (red), while the nuclei of all heart muscle cells were stained with DAPI (blue). The apoptotic index was calculated by determining the percentage of cells undergoing apoptosis compared to the total number of cells. (<b>K</b>) Serum cTnI levels were evaluated after IRI (<span class="html-italic">n</span> = 5). (<b>L</b>) Representative transmission electron microscopy images in each group. Data are shown as the means ± SD. ** <span class="html-italic">p</span> &lt; 0.01, *** <span class="html-italic">p</span> &lt; 0.001.</p>
Full article ">
12 pages, 2220 KiB  
Article
Lung EC-SOD Overexpression Prevents Hypoxia-Induced Platelet Activation and Lung Platelet Accumulation
by Daniel Colon Hidalgo, Mariah Jordan, Janelle N. Posey, Samuel D. Burciaga, Thi-Tina N. Nguyen, Christina Sul, Caitlin V. Lewis, Cassidy Delaney and Eva S. Nozik
Antioxidants 2024, 13(8), 975; https://doi.org/10.3390/antiox13080975 (registering DOI) - 10 Aug 2024
Viewed by 328
Abstract
Pulmonary hypertension (PH) is a progressive disease marked by pulmonary vascular remodeling and right ventricular failure. Inflammation and oxidative stress are critical in PH pathogenesis, with early pulmonary vascular inflammation preceding vascular remodeling. Extracellular superoxide dismutase (EC-SOD), a key vascular antioxidant enzyme, mitigates [...] Read more.
Pulmonary hypertension (PH) is a progressive disease marked by pulmonary vascular remodeling and right ventricular failure. Inflammation and oxidative stress are critical in PH pathogenesis, with early pulmonary vascular inflammation preceding vascular remodeling. Extracellular superoxide dismutase (EC-SOD), a key vascular antioxidant enzyme, mitigates oxidative stress and protects against inflammation and fibrosis in diverse lung and vascular disease models. This study utilizes a murine hypobaric hypoxia model to investigate the role of lung EC-SOD on hypoxia-induced platelet activation and platelet lung accumulation, a critical factor in PH-related inflammation. We found that lung EC-SOD overexpression blocked hypoxia-induced platelet activation and platelet accumulation in the lung. Though lung EC-SOD overexpression increased lung EC-SOD content, it did not impact plasma extracellular SOD activity. However, ex vivo, exogenous extracellular SOD treatment specifically blunted convulxin-induced platelet activation but did not blunt platelet activation with thrombin or ADP. Our data identify platelets as a novel target of EC-SOD in response to hypoxia, providing a foundation to advance the understanding of dysregulated redox signaling and platelet activation in PH and other chronic hypoxic lung diseases. Full article
(This article belongs to the Special Issue Role of Redox in Pulmonary Vascular Diseases)
Show Figures

Figure 1

Figure 1
<p>Platelet counts are similar across wildtype and EC-SOD overexpressing mice in normoxia and hypoxia, but platelet activation is attenuated by EC-SOD overexpression. (<b>A</b>) Using the Heska HT5 hematologic analyzer, total platelet counts were obtained from blood samples acquired from WT and lung EC-SOD overexpressing mice in both normoxia and hypoxia. N = 8–13. (<b>B</b>) P-selectin (Thermofisher) and (<b>C</b>) αIIBβ3 (Emfret) expression were assessed by flow cytometry in freshly isolated platelets from wildtype and lung EC-SOD overexpressing mice. N = 8, * <span class="html-italic">p</span> &lt; 0.05 by two-way ANOVA. (<b>D</b>) Platelet-leukocyte aggregation was measured by co-expression of CD41-BV421 (BioLegend) and CD45-APC (BioLegend) antibodies. N = 4–9, * <span class="html-italic">p</span> &lt; 0.05 by two-way ANOVA. (<b>E</b>) ELISA measuring PF4 (Abcam), a marker of platelet chemokine release, in platelet-poor plasma. N = 6–9, * <span class="html-italic">p</span> &lt; 0.05 by two-way ANOVA. WT = Wildtype, Tg+ = mice overexpressing lung EC-SOD, NMX = normoxia, HPX = hypoxia, PF4 = platelet factor 4.</p>
Full article ">Figure 2
<p>Hypoxia leads to increases in platelet infiltration and lung PF4 levels, which is prevented by EC-SOD overexpression. (<b>A</b>) Percentage of positive platelet marker CD41 (BioLegend) pixels in the lungs of wildtype mice and mice overexpressing lung EC-SOD in normoxia and hypoxia. N = 6–8, * <span class="html-italic">p</span> &lt; 0.05 by two-way ANOVA. (<b>B</b>) Lung PF4 ELISA showing the concentration of PF4 in the lungs of wildtype mice and mice overexpressing lung EC-SOD exposed to normoxia or hypoxia. N = 5–10, * <span class="html-italic">p</span> &lt; 0.05 by two-way ANOVA. (<b>C</b>) Representative histology (20x) images of CD41 (BioLegend) lung immunohistochemistry of the lungs of wildtype mice and mice overexpressing lung EC-SOD in normoxia and hypoxia. WT = Wildtype, Tg+ = mice overexpressing lung EC-SOD, NMX = normoxia, HPX = hypoxia, PF4 = platelet factor 4.</p>
Full article ">Figure 3
<p>Plasma SOD activity was similar between mice overexpressing lung EC-SOD (Tg) mice and wildtype mice. (<b>A</b>) Plasma SOD activity in normoxic and 3-day hypoxic WT and Tg mice using a colorimetric assay (Dojindo). N = 6. (<b>B</b>) Representative western blot of lung EC-SOD expression along with housekeeping protein, Vinculin in WT mice and mice with lung overexpression of EC-SOD. (<b>C</b>) Densitometry of lung EC-SOD expression. N = 6, * <span class="html-italic">p</span> &lt; 0.05 by two-way ANOVA. WT = Wildtype, Tg = mice overexpressing lung EC-SOD, NMX = normoxia, HPX = hypoxia.</p>
Full article ">Figure 4
<p>Ex vivo superoxide dismutase administration prevents convulxin-induced platelet activation. Freshly isolated platelets from wildtype mice were exposed to platelet agonists thrombin (0.1 U/mL, Chrono-log Corporation), adenosine diphosphate (ADP) (5 uM, Chrono-log Corporation) and Convulxin (350 ng/mL, Cayman Chemical), in the presence or absence of superoxide dismutase (SOD, 300 U/mL, Sigma Aldrich). Platelet activation was measured by expression (<b>A</b>) P-selectin and (<b>B</b>) activated αIIbβ3 via flow cytometry. N = 8, * <span class="html-italic">p</span> &lt; 0.05. SOD = superoxide dismutase, BL = baseline, THR = thrombin, ADP = adenosine diphosphate, CONV = Convulxin.</p>
Full article ">
15 pages, 16273 KiB  
Article
Xanthoxylin Attenuates Lipopolysaccharide-Induced Lung Injury through Modulation of Akt/HIF-1α/NF-κB and Nrf2 Pathways
by Fu-Chao Liu, Yuan-Han Yang, Chia-Chih Liao and Hung-Chen Lee
Int. J. Mol. Sci. 2024, 25(16), 8742; https://doi.org/10.3390/ijms25168742 (registering DOI) - 10 Aug 2024
Viewed by 263
Abstract
Xanthoxylin, a bioactive phenolic compound extracted from the traditional herbal medicine Penthorum Chinense Pursh, is renowned for its anti-inflammatory effects. While previous studies have highlighted the anti-inflammatory and antioxidant properties of Xanthoxylin, its precise mechanisms, particularly concerning immune response and organ protection, [...] Read more.
Xanthoxylin, a bioactive phenolic compound extracted from the traditional herbal medicine Penthorum Chinense Pursh, is renowned for its anti-inflammatory effects. While previous studies have highlighted the anti-inflammatory and antioxidant properties of Xanthoxylin, its precise mechanisms, particularly concerning immune response and organ protection, remain underexplored. This study aimed to elucidate the effects of Xanthoxylin on inflammation and associated signaling pathways in a mouse model of lipopolysaccharide (LPS)-induced acute lung injury (ALI). ALI was induced via intratracheal administration of LPS, followed by intraperitoneal injections of Xanthoxylin at doses of 1, 2.5, 5, and 10 mg/kg, administered 30 min post-LPS exposure. Lung tissues were harvested for analysis 6 h after LPS challenge. Xanthoxylin treatment significantly mitigated lung tissue damage, pathological alterations, immune cell infiltration, and the production of pro-inflammatory cytokines, including tumor necrosis factor-alpha (TNF-α) and interleukin-6 (IL-6). Additionally, Xanthoxylin modulated the expression of key proteins in the protein kinase B (Akt)/hypoxia-inducible factor 1-alpha (HIF-1α)/nuclear factor-kappa B (NF-κB) signaling pathway, as well as nuclear factor erythroid 2-related factor 2 (Nrf2) and oxidative markers such as superoxide dismutase (SOD) and malondialdehyde (MDA) in the context of LPS-induced injury. This study demonstrates that Xanthoxylin exerts protective and anti-inflammatory effects by down-regulating and inhibiting the Akt/HIF-1α/NF-κB pathways, suggesting its potential as a therapeutic target for the prevention and treatment of ALI or acute respiratory distress syndrome (ARDS). Full article
(This article belongs to the Special Issue New Insights in Natural Bioactive Compounds 3.0)
Show Figures

Figure 1

Figure 1
<p>The effect of Xanthoxylin on the viability and pro-inflammatory cytokine levels of RAW 264.7 cells in the presence and absence of LPS. (<b>A</b>) RAW 264.7 cells were treated with various concentrations of DMSO (0.01, 0.05, and 0.5 μL) or Xanthoxylin (0.1, 1, 5, 10, 20, and 50 μM) for 24 h. Results are expressed as a percentage relative to the control group and shown as mean ± SD (<span class="html-italic">n</span> = 6 per group). (<b>B</b>) RAW 264.7 cells were treated with Xanthoxylin (0, 5, and 10 μM) followed by LPS exposure for 48 h to assess cell viability. Results are expressed as a percentage relative to the control group and shown as mean ± SD (<span class="html-italic">n</span> = 12 per group). (<b>C</b>) Pro-inflammatory cytokines IL-1β, IL-6, and TNF-α levels in the supernatants of RAW 264.7 cells were measured after treatment with Xanthoxylin, followed by LPS exposure. ### <span class="html-italic">p</span> &lt; 0.005 vs. control group; * <span class="html-italic">p</span> &lt; 0.05, ** <span class="html-italic">p</span> &lt; 0.01, *** <span class="html-italic">p</span> &lt; 0.005 vs. LPS group.</p>
Full article ">Figure 2
<p>General lung appearance after LPS-induced injury and Xanthoxylin treatment. Mice received an intratracheal LPS challenge followed by intraperitoneal administration of Xanthoxylin (XT, 1, 2.5, 5, and 10 mg/kg) or saline. Lungs were collected 6 h post-LPS challenge for analysis.</p>
Full article ">Figure 3
<p>Histological examination of lung tissues stained with H&amp;E after LPS challenge and Xanthoxylin treatment. Mice received LPS intratracheally and were then treated with Xanthoxylin (XT, 1, 2.5, 5, and 10 mg/kg) or saline intraperitoneally. Lungs were harvested 6 h post-LPS challenge for H&amp;E staining. Representative images show ALI and histological changes (100× magnification, scar bar = 100 μm). Quantification of histologic lung injury was analyzed according to American Thoracic Society (ATS) scoring system (<span class="html-italic">n</span> = 6 per group). # <span class="html-italic">p</span> &lt; 0.05, ## <span class="html-italic">p</span> &lt; 0.01, ### <span class="html-italic">p</span> &lt; 0.005 vs. control group; * <span class="html-italic">p</span> &lt; 0.05, ** <span class="html-italic">p</span> &lt; 0.001 vs. LPS group.</p>
Full article ">Figure 4
<p>Neutrophil infiltration in lungs following LPS-induced injury and Xanthoxylin treatment. Mice were challenged with LPS intratracheally and treated with Xanthoxylin (XT, 1, 2.5, 5, and 10 mg/kg) or saline intraperitoneally. Lungs were collected 6 h post-LPS challenge and immunostained with Ly6G antibody (200× magnification, scar bar = 50 μm). Quantification of positive cells was analyzed under high power field (HPF). Data are mean ± SD (<span class="html-italic">n</span> = 6 per group). ### <span class="html-italic">p</span> &lt; 0.005 vs. control group; * <span class="html-italic">p</span> &lt; 0.05, *** <span class="html-italic">p</span> &lt; 0.005 vs. LPS group.</p>
Full article ">Figure 5
<p>Macrophage infiltration in lungs following LPS-induced injury and Xanthoxylin treatment. Mice received LPS intratracheally and were treated with Xanthoxylin (XT, 1, 2.5, 5, and 10 mg/kg) or saline intraperitoneally. Lungs were harvested 6 h post-LPS challenge and immunostained with Mac-2 antibody (200× magnification, scar bar = 50 μm). Quantification of positive cells was analyzed under HPF. Data are mean ± SD (<span class="html-italic">n</span> = 6 per group). ### <span class="html-italic">p</span> &lt; 0.005 vs. control group; *** <span class="html-italic">p</span> &lt; 0.005 vs. LPS group.</p>
Full article ">Figure 6
<p>Levels of (<b>A</b>) IL-6 and (<b>B</b>) TNF-α in lungs after LPS challenge and Xanthoxylin treatment. Mice were given intratracheal LPS challenge followed by Xanthoxylin (XT, 1, 2.5, 5, and 10 mg/kg) or saline intraperitoneally. Lungs were harvested 6 h post-LPS challenge for ELISA. Data are mean ± SD (<span class="html-italic">n</span> = 6 per group). # <span class="html-italic">p</span> &lt; 0.05, ## <span class="html-italic">p</span> &lt; 0.01, ### <span class="html-italic">p</span> &lt; 0.005 vs. control group; ** <span class="html-italic">p</span> &lt; 0.01, *** <span class="html-italic">p</span> &lt; 0.005 vs. LPS group.</p>
Full article ">Figure 7
<p>Levels of (<b>A</b>) MDA and (<b>B</b>) SOD in lungs after LPS challenge and Xanthoxylin treatment. Mice received intratracheal LPS challenge and were treated with Xanthoxylin (XT, 1, 2.5, 5, and 10 mg/kg) or saline intraperitoneally. Lungs were collected 6 h post-LPS challenge for oxidative stress assays. Data are mean ± SD (<span class="html-italic">n</span> = 6 per group). # <span class="html-italic">p</span> &lt; 0.05, ### <span class="html-italic">p</span> &lt; 0.005 vs. control group; * <span class="html-italic">p</span> &lt; 0.05, ** <span class="html-italic">p</span> &lt; 0.01, *** <span class="html-italic">p</span> &lt; 0.005 vs. LPS group.</p>
Full article ">Figure 8
<p>Effects of Xanthoxylin on expression of (<b>A</b>) Akt, (<b>B</b>) NF-κB, (<b>C</b>) HIF-1α, and (<b>D</b>) Nrf2 in lungs after LPS challenge. Mice were administered Xanthoxylin (XT, 2.5, 5, and 10 mg/kg) or saline intraperitoneally 30 min post-LPS challenge. Lungs were harvested 6 h later for Western blot analysis. Data are mean ± SD (<span class="html-italic">n</span> = 6 per group). ## <span class="html-italic">p</span> &lt; 0.01, ### <span class="html-italic">p</span> &lt; 0.005 vs. control group; * <span class="html-italic">p</span> &lt; 0.05, ** <span class="html-italic">p</span> &lt; 0.01, *** <span class="html-italic">p</span> &lt; 0.005 vs. LPS group.</p>
Full article ">Figure 9
<p>Nrf2 expression in lungs after LPS-induced injury and Xanthoxylin treatment. Mice received intratracheal LPS challenge followed by Xanthoxylin (XT, 1, 2.5, 5, and 10 mg/kg) or saline intraperitoneally. Lungs were collected 6 h post-LPS challenge and immunostained with Nrf2 antibody (400× magnification, scar bar = 25 μm). Quantification of positive cells was analyzed under HPF. Data are mean ± SD (<span class="html-italic">n</span> = 6 per group). ### <span class="html-italic">p</span> &lt; 0.005 vs. control group; *** <span class="html-italic">p</span> &lt; 0.005 vs. LPS group.</p>
Full article ">Figure 10
<p>A schematic representation of the involvement of Akt/HIF-1α/NF-κB and Nrf2 signaling pathways in the protective effects of Xanthoxylin against LPS-induced lung injury. Xanthoxylin modulates Akt expression, suppresses HIF-1α/NF-κB signaling, and activates Nrf2, thereby reducing cell damage and oxidative stress. It also inhibits TNF-α and IL-6 release from macrophages.</p>
Full article ">
23 pages, 6757 KiB  
Article
Function of the NAC1 Gene from Fraxinus mandshurica in Cold Resistance and Growth Promotion in Tobacco
by Yang Cao, Liming He, Shengdian Lu, Yuling Wang, Chenxi Zhang and Yaguang Zhan
Forests 2024, 15(8), 1405; https://doi.org/10.3390/f15081405 (registering DOI) - 10 Aug 2024
Viewed by 249
Abstract
To elucidate the function of the cold-resistance regulatory gene FmNAC1 from Fraxinus mandshurica Rupr., this study identified the role that overexpression of the FmNAC1 gene plays in tobacco growth and cold-stress regulation. The cloned FmNAC1 gene from F. mandshurica is 891 bp [...] Read more.
To elucidate the function of the cold-resistance regulatory gene FmNAC1 from Fraxinus mandshurica Rupr., this study identified the role that overexpression of the FmNAC1 gene plays in tobacco growth and cold-stress regulation. The cloned FmNAC1 gene from F. mandshurica is 891 bp in length and encodes 296 amino acids. Our subcellular localization analysis confirmed that FmNAC1 is primarily located in the nucleus and functions as a transcription factor. FmNAC1 is responsive to cold and NaCl stress, as well as to the induction of IAA, GA, and ABA hormone signals. To further elucidate its function in cold resistance, four transgenic tobacco lines expressing FmNAC1 (FmNAC1-OE) were generated through tissue culture after the Agrobacterium-mediated transformation of wild-type (WT) Nicotiana tabacum L.. These FmNAC1-OE plants exhibited accelerated growth after transplantation. When exposed to low-temperature conditions at −5 °C for 24 h, the rates of wilting and yellowing of the FmNAC1-OE plants were significantly lower than those of the WT tobacco plants. Additionally, the membrane integrity, osmotic regulation, and reactive oxygen species (ROS)-scavenging abilities of the FmNAC1-OE tobacco lines were better than those of the WT plants, indicating the potential of the FmNAC1 gene to improve plant cold resistance. The gene expression results further revealed that the FmNAC1 transcription factor exhibits regulatory interactions with growth-related genes such as IAA and AUX1; cold-resistance-related genes such as ICE, DREB, and CBF1; and genes involved in the clearance of reactive oxygen species (ROS), such as CAT and SOD. All of this evidence shows that the FmNAC1 transcription factor from F. mandshurica plays a key role in contributing to the enhancement of growth, cold resistance, and ROS clearance in transgenic tobacco plants. Full article
(This article belongs to the Section Genetics and Molecular Biology)
16 pages, 3974 KiB  
Article
Comparison of Festuca glauca ‘Uchte’ and Festuca amethystina ‘Walberla’ Varieties in a Simulated Extensive Roof Garden Environment
by Dóra Hamar-Farkas, Szilvia Kisvarga, Máté Ördögh, László Orlóci, Péter Honfi and Ildikó Kohut
Plants 2024, 13(16), 2216; https://doi.org/10.3390/plants13162216 (registering DOI) - 9 Aug 2024
Viewed by 175
Abstract
One of the most effective means of increasing urban green areas is the establishment of roof gardens. They have many positive properties and ecological functions, such as filling empty spaces with plants, protecting buildings, dust retention and air cleaning. In the case of [...] Read more.
One of the most effective means of increasing urban green areas is the establishment of roof gardens. They have many positive properties and ecological functions, such as filling empty spaces with plants, protecting buildings, dust retention and air cleaning. In the case of extensive constructions, mostly Sedum species are used, planted as carpet-like “grass” sods or by installing modular units as plugs; however, with the use of other plant genera, the efficiency of ecological services could be increased by expanding the diversity. Festuca taxa have good drought resistance, and these plants tolerate temperature alterations well. Their application would increase the biodiversity, quality and decorative value of roof gardens. Experiments were carried out on nursery benches imitating a roof garden, with the use of modular elements intended for Sedum species, which facilitate the establishment of green roofs. In our trial, varieties of two European native species, Festuca glauca Vill. ‘Uchte’ and F. amethystina L. ‘Walberla’, were investigated. In order to find and determine the differences between the cultivars and the effects of the media (leaf mold and rhyolite tuff), we drew inferences after morphological (height, circumference, root weight, fresh and dry weight) and physiological tests (peroxidase and proline enzyme activity). We concluded that F. glauca ‘Uchte’ is recommended for roof garden conditions, planted in modular elements. Although the specimens were smaller in the medium containing fewer organic components than in the version with larger amounts, they were less exposed to the effects of drought stress. This can be a key factor for survival in extreme roof gardens or even urban conditions for all plants. Full article
(This article belongs to the Section Horticultural Science and Ornamental Plants)
Show Figures

Figure 1

Figure 1
<p>The average heights of the <span class="html-italic">Festuca</span> plants with the measurements. The letters indicate the significance levels, separated by measurement. ANOVA: 08 November 2021. By medium: 75:25: 0.021; 50:50: 0.121; by plant: ‘Uchte’: 0.096; ‘Walberla’: 0.583; 11 April 2022. By medium: 75:25: 0.088; 50:50: 0.274; by plant: ‘Uchte’: 0.213; ‘Walberla’: 0.750; 18 May 2022. By medium: 75:25: 0.840; 50:50: 0.911; by plant: ‘Uchte’: 0.011; ‘Walberla’: 0.010; 15 October 2022. By medium: 75:25: 0.001; 50:50: 0.000; by plant: ‘Uchte’: 0.006; ‘Walberla’: 0.006.</p>
Full article ">Figure 2
<p>The average circumferences of the examined plants. The letters indicate the significance levels, separated by measurement. ANOVA: 08 November 2021. By medium: 75:25: 0.192; 50:50: 0.000; by plant: ‘Uchte’: 0.770; ‘Walberla’: 0.000; 11 April 2022. By medium: 75:25: 0.000; 50:50: 0.000; by plant: ‘Uchte’: 0.846; ‘Walberla’: 0.172; 18 May 2022. By medium: 75:25: 0.000; 50:50: 0.000; by plant: ‘Uchte’: 0.003; ‘Walberla’: 0.199; 15 October 2022. By medium: 75:25: 0.469; 50:50: 0.014; by plant: ‘Uchte’: 0.028; ‘Walberla’: 0.938.</p>
Full article ">Figure 3
<p>The moisture content of the plants (based on calculation with dry and fresh plant weight), separated by soil type.</p>
Full article ">Figure 4
<p>The peroxidase level per plant, separated by soil type. The letters indicate the significance levels. ANOVA. By medium: 75:25: 0.024; 50:50: 0.01; by plant: ‘Uchte’: 0.771; ‘Walberla’: 0.436.</p>
Full article ">Figure 5
<p>The proline level per plant, separated by soil type. The letters indicate the significance levels. ANOVA. By medium: 75:25: 0.375; 50:50: 0.01; by plant: ‘Uchte’: 0.898; ‘Walberla’: 0.041.</p>
Full article ">Figure 6
<p>Modular green roof element developed by Fito System Kft.</p>
Full article ">Figure 7
<p>The table modeling the roof garden, which was located 100 cm above the ground.</p>
Full article ">
20 pages, 3114 KiB  
Review
The Manganese–Bone Connection: Investigating the Role of Manganese in Bone Health
by Gulaim Taskozhina, Gulnara Batyrova, Gulmira Umarova, Zhamilya Issanguzhina and Nurgul Kereyeva
J. Clin. Med. 2024, 13(16), 4679; https://doi.org/10.3390/jcm13164679 - 9 Aug 2024
Viewed by 440
Abstract
The complex relationship between trace elements and skeletal health has received increasing attention in the scientific community. Among these minerals, manganese (Mn) has emerged as a key element affecting bone metabolism and integrity. This review examines the multifaceted role of Mn in bone [...] Read more.
The complex relationship between trace elements and skeletal health has received increasing attention in the scientific community. Among these minerals, manganese (Mn) has emerged as a key element affecting bone metabolism and integrity. This review examines the multifaceted role of Mn in bone health, including its effects on bone regeneration, mineralization, and overall skeletal strength. This review article is based on a synthesis of experimental models, epidemiologic studies, and clinical trials of the mechanisms of the effect of Mn on bone metabolism. Current research data show that Mn is actively involved in the processes of bone remodeling by modulating the activity of osteoblasts and osteoclasts, as well as the main cells that regulate bone formation and resorption. Mn ions have a profound effect on bone mineralization and density by intricately regulating signaling pathways and enzymatic reactions in these cells. Additionally, Mn superoxide dismutase (MnSOD), located in bone mitochondria, plays a crucial role in osteoclast differentiation and function, protecting osteoclasts from oxidative damage. Understanding the nuances of Mn’s interaction with bone is essential for optimizing bone strategies, potentially preventing and managing skeletal diseases. Key findings include the stimulation of osteoblast proliferation and differentiation, the inhibition of osteoclastogenesis, and the preservation of bone mass through the RANK/RANKL/OPG pathway. These results underscore the importance of Mn in maintaining bone health and highlight the need for further research into its therapeutic potential. Full article
Show Figures

Figure 1

Figure 1
<p>The role of manganese (Mn) in bone cellular and molecular functions. The trace element Mn, with its various biochemical and physiological effects, participates in the synthesis of bone matrix, the inhibition of the formation of osteoclast-like cells, antioxidant function with the enzyme Mn superoxide dismutase (MnSOD), and mRNA expression of RANKL receptors; it also contributes to cell adhesion with extracellular matrix proteins, regulating osteoid formation. It also protects cartilage and stimulates chondrocyte growth through ZIP14. This is important for its integrin-activating functions, which contribute to the adhesion, integrity, and proliferation of osteoblasts.</p>
Full article ">Figure 2
<p>Manganese (Mn) superoxide dismutase (MnSOD) in the bone resorption [<a href="#B52-jcm-13-04679" class="html-bibr">52</a>]. RANKL-induced differentiation of macrophages into osteoclasts and the role of MnSOD in managing oxidative stress during bone resorption are depicted. RANKL binds to the RANK receptors on these cells, promoting their maturation. During bone resorption, superoxide (O<sub>2</sub><sup>−</sup>) is produced as a byproduct, and the mitochondrial enzyme MnSOD catalyzes the conversion of O<sub>2</sub><sup>−</sup> into hydrogen peroxide (H<sub>2</sub>O<sub>2</sub>) and oxygen (O<sub>2</sub>), thereby reducing oxidative stress. H<sub>2</sub>O<sub>2</sub> is subsequently converted into water (H<sub>2</sub>O), providing cellular protection. This process underscores the critical role of MnSOD in maintaining the functionality and integrity of osteoclasts during bone resorption.</p>
Full article ">Figure 3
<p>Manganese (Mn) and bone remodeling [<a href="#B72-jcm-13-04679" class="html-bibr">72</a>]. The role of Mn in bone remodeling highlights its dual impact on osteoclasts and osteoblasts. Mn promotes osteoclast differentiation by enhancing the RANKL/RANK signaling pathway, where RANKL binds to RANK receptors on osteoclast progenitor cells, leading to their maturation into osteoclasts. Mature osteoclasts resorb bone, a process associated with oxidative stress, during which O<sub>2</sub><sup>−</sup> is converted into less harmful molecules by the mitochondrial enzyme Mn superoxide dismutase (MnSOD). Concurrently, Mn inhibits the PI3K/AKT and WNT/β-catenin signaling pathways in mesenchymal stem cells (MSCs), thereby reducing the differentiation and activity of osteoblasts. This dual mechanism underscores the essential role of Mn in maintaining bone mass and integrity, ensuring effective bone regeneration and homeostasis by balancing bone resorption and formation.</p>
Full article ">Figure 4
<p>Molecular mechanisms of manganese (Mn) metabolism. The molecular pathways involved in Mn metabolism highlight its absorption, transport, and accumulation in the body. Mn ions (Mn<sup>2+</sup>) are absorbed in the intestines through the divalent metal transporter 1 (DMT1). After absorption, Mn<sup>2+</sup> ions enter the bloodstream and are transported in a complex with proteins. The liver, considered the central organ in Mn metabolism, plays a crucial role in processing and regulating Mn levels. Mn is then distributed from the liver to various tissues throughout the body, with a significant accumulation in the bones. This high accumulation in bones underscores the essential role of Mn in skeletal health.</p>
Full article ">Figure 5
<p>Manganese (Mn) hemostasis in the bone [<a href="#B90-jcm-13-04679" class="html-bibr">90</a>]. The cellular mechanisms involved in maintaining Mn homeostasis in bones reveal the key physiological functions of Mn transporters and regulators, including ZIP8, ZNT10, and ZIP14. The process begins with the intake of Mn from food, where ZIP8 facilitates the intracellular accumulation of Mn<sup>2+</sup> ions. These Mn<sup>2+</sup> ions enter the bloodstream and are transported to various tissues, including bones and liver hepatocytes. The transport of Mn<sup>2+</sup> ions into bones and other tissues is facilitated by the ubiquitously expressed ZIP14. Mn<sup>2+</sup> ions reach the liver, where they undergo further processing and regulation. The ZIP10 transporter acts as an apical exporter, transporting Mn from the blood to the lumen of the small intestine for excretion in feces. These intricate regulatory mechanisms ensure the balance of Mn, which is crucial for maintaining bone health and overall metabolic homeostasis.</p>
Full article ">
19 pages, 4061 KiB  
Article
Dynamics of Physiological Properties and Endophytic Fungal Communities in the Xylem of Aquilaria sinensis (Lour.) with Different Induction Times
by Qingqing Zhang, Rongrong Li, Yang Lin, Weiwei Zhao, Qiang Lin, Lei Ouyang, Shengjiang Pang and Huahao Zeng
J. Fungi 2024, 10(8), 562; https://doi.org/10.3390/jof10080562 - 9 Aug 2024
Viewed by 412
Abstract
Xylem-associated fungus can secrete many secondary metabolites to help Aquilaria trees resist various stresses and play a crucial role in facilitating agarwood formation. However, the dynamics of endophytic fungi in Aquilaria sinensis xylem after artificial induction have not been fully elaborated. Endophytic fungi [...] Read more.
Xylem-associated fungus can secrete many secondary metabolites to help Aquilaria trees resist various stresses and play a crucial role in facilitating agarwood formation. However, the dynamics of endophytic fungi in Aquilaria sinensis xylem after artificial induction have not been fully elaborated. Endophytic fungi communities and xylem physio-biochemical properties were examined before and after induction with an inorganic salt solution, including four different times (pre-induction (0M), the third (3M), sixth (6M) and ninth (9M) month after induction treatment). The relationships between fungal diversity and physio-biochemical indices were evaluated. The results showed that superoxide dismutase (SOD) and peroxidase (POD) activities, malondialdehyde (MDA) and soluble sugar content first increased and then decreased with induction time, while starch was heavily consumed after induction treatment. Endophytic fungal diversity was significantly lower after induction treatment than before, but the species richness was promoted. Fungal β-diversity was also clustered into four groups according to different times. Core species shifted from rare to dominant taxa with induction time, and growing species interactions in the network indicate a gradual complication of fungal community structure. Endophytic fungi diversity and potential functions were closely related to physicochemical indices that had less effect on the relative abundance of the dominant species. These findings help assess the regulatory mechanisms of microorganisms that expedite agarwood formation after artificial induction. Full article
Show Figures

Figure 1

Figure 1
<p>Number of sequences (<b>A</b>), unique OTUs (<b>B</b>) and diversity indices (<b>C</b>) of endophytic fungi at different induction times. 0M, 3M, 6M and 9M represent pre-induction, the third month, the sixth month and the ninth month after artificial induction, respectively. Different letters indicate significant differences between induction time groups (<span class="html-italic">p</span> &lt; 0.05, Tukey’s test).</p>
Full article ">Figure 2
<p>The number (<b>A</b>,<b>C</b>) and relative abundance (<b>B</b>,<b>D</b>) of dominant fungi at order (<b>A</b>,<b>B</b>) and genus (<b>C</b>,<b>D</b>) levels under different induction times.</p>
Full article ">Figure 3
<p>NMDS ((<b>A</b>), based on Bray–Curtis) and ANOSIM ((<b>B</b>), based on weighted UniFrac distances) analyses of endophytic fungi of <span class="html-italic">A. sinensis</span> at different induction times.</p>
Full article ">Figure 4
<p>Changes in the relative abundance of dominant orders (<b>A</b>,<b>B</b>) and genera (<b>C</b>,<b>D</b>) in different induction times. The relative abundance of each indicator in the above figure was converted to log<sub>10</sub>(X + 1) standards. Different letters indicate significant differences between induction time groups (<span class="html-italic">p</span> &lt; 0.05, Tukey’s test).</p>
Full article ">Figure 5
<p>Co-occurrence network of endophytic fungi at the order level during different induction times. The size and color of each node depend on its abundance and phylum category. The red and green links indicate positive and negative correlations between nodes, respectively. 0M, 3M, 6M and 9M represent pre-induction, the third month, the sixth month and the ninth month after artificial induction, respectively.</p>
Full article ">Figure 6
<p>Function predicted of fungal community based on FUNGuild.</p>
Full article ">Figure 7
<p>The correlation heatmap between dominant species (<b>A</b>: order level) (<b>B</b>: genus level) and physio-biochemical properties in the xylem. SOD, POD and MDA represent superoxide dismutase, peroxidase and malondialdehyde, respectively. * <span class="html-italic">p</span> &lt; 0.05, ** <span class="html-italic">p</span> &lt; 0.01.</p>
Full article ">Figure 8
<p>Mantel test among diversity, community composition, function predicted of endophytic fungi and physio-biochemical properties in the xylem. SOD, POD and MDA represent superoxide dismutase, peroxidase and malondialdehyde, respectively. * <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.</p>
Full article ">
17 pages, 8251 KiB  
Article
Dynamic Changes in Endogenous Substances in Flowering Organs of Camellia drupifera during the Flowering Stage
by Zhen Liu, Jialu Tao, Chunhua Ma, Mengling Wen, Ruchun Xi and Xiaomei Deng
Forests 2024, 15(8), 1391; https://doi.org/10.3390/f15081391 - 9 Aug 2024
Viewed by 226
Abstract
Camellia drupifera is an important woody oil plant in South China, renowned for its seed oil that is rich in unsaturated fatty acids and possesses significant antioxidant, anti-cancer, and immune-enhancing properties. The low fruit-setting rate of C. drupifera is influenced by multiple factors, [...] Read more.
Camellia drupifera is an important woody oil plant in South China, renowned for its seed oil that is rich in unsaturated fatty acids and possesses significant antioxidant, anti-cancer, and immune-enhancing properties. The low fruit-setting rate of C. drupifera is influenced by multiple factors, including flowering stage climate, flowering habits, pollination biology, soil conditions, and self-incompatibility. Among these, large-scale pure forest plantations are the primary cause of the low fruit-setting rate. Although previous studies have explored the impact of self-incompatibility on fruit-setting in C. drupifera, research on the dynamic changes of endogenous substances during the flowering stage in pure forest environments remains limited. Research findings indicate that tannase activity is relatively high in the pistils of C. drupifera, creating a favorable environment for pollen tube growth. Plant hormones such as indole-3-acetic acid (IAA), cytokinin (CTK), gibberellin (GA), and ethylene (ETH) regulate the development and aging of floral organs through complex interactions. Specifically, high levels of IAA in the pistil promote pollen tube growth, while changes in ETH and ABA are closely related to the aging of floral organs. Under oxidative stress conditions, high levels of H2O2 in the pistil may contribute to self-incompatibility. The activity of superoxide dismutase (SOD) in the floral organs during the flowering stage is significantly higher compared to peroxidase (POD) and catalase (CAT), highlighting the critical role of SOD in regulating oxidative stress during this stage. This study provides new insights into the changes in endogenous substances in the floral organs of C. drupifera during the flowering stage. It offers theoretical references for understanding its sexual reproduction process and for the application of plant growth regulators to improve fruit setting. Full article
Show Figures

Figure 1

Figure 1
<p>Morphological Changes in Floral Organs of <span class="html-italic">C. drupifera</span> during the Flowering Stage. Note: Panels (<b>A</b>–<b>F</b>) illustrate the morphological changes at different stages: two days before flowering (K(−2)), on the day of flowering (K(0)), one day after flowering (K(1)), two days after flowering (K(2)), four days after flowering (K(4)), and six days after flowering (K(6)). (a) Petals (P); (b) Stamens (S); (c) Pistils (P); (d) Sepals (S); (e) Buds (B). The grid size varies among the images: the minimum square length is 1 cm, and this measurement remains consistent across all images.</p>
Full article ">Figure 2
<p>Heatmap of Endogenous Substance Abundance during Flowering Stage of <span class="html-italic">C. drupifera.</span> Note: (<b>A</b>); Pistils; (<b>B</b>): Stamens. The heatmap color gradient from red to blue represents the relative changes in endogenous substance content in floral organs during the flowering stage, with red indicating above-average levels and blue indicating below-average levels. The dendrogram on the left side of the heatmap shows the similarity in substance content changes, with closer branches representing more similar trends. The raw data were standardized using Z-score normalization. TNS: tannase, H<sub>2</sub>O<sub>2:</sub> Hydrogen peroxide, ROS: Reactive oxygen species, SP: Soluble proteins, SS: Soluble sugars, ABA: Abscisic acid, IAA: Indole-3-acetic acid, GA: Gibberellins, CTK: Cytokinin, ETH: Ethylene, BR: Brassinosteroids, POD: Peroxidase, SOD: Superoxide dismutase, CAT: Catalase.</p>
Full article ">Figure 3
<p>Dynamic Changes in Tannin Synthase Activity of Floral Organs during the Flowering Stage of <span class="html-italic">C. drupifera.</span> Note: different capital letters following data in the same row indicate significant differences (<span class="html-italic">p</span> &lt; 0.05) between different floral organs at the same sampling stage; different lowercase letters following data in the same column indicate significant differences (<span class="html-italic">p</span> &lt; 0.05) between different sampling stages of the same floral organ. Error bars represent the standard deviation (SD) of the mean. The red fill represents pistils, and the green dashed line represents stamens. The same notation and color coding apply to all relevant figures and tables in the text.</p>
Full article ">Figure 4
<p>Dynamic Changes in Hormone Content of Floral Organs during Flowering Stage of <span class="html-italic">C. drupifera.</span> Note: Panels (<b>A</b>–<b>F</b>) represent the content of IAA, CTK, GA, ETH, ABA, and BR, respectively.</p>
Full article ">Figure 5
<p>Dynamic Changes in Hormone Ratios of Floral Organs during Flowering Stage of <span class="html-italic">C. drupifera.</span> Note: Panels (<b>A</b>–<b>D</b>) represent the ratios of IAA/ABA, CTK/ABA, BR/ABA, and GA/ABA, respectively.</p>
Full article ">Figure 6
<p>Dynamic Changes in ROS, H<sub>2</sub>O<sub>2</sub> Content, and Antioxidant Enzyme Activity of Floral Organs during Flowering Stage of <span class="html-italic">C. drupifera.</span> Note: Panels (<b>A</b>–<b>E</b>) represent the ROS content, H<sub>2</sub>O<sub>2</sub> content, POD, SOD, and CAT activity, respectively.</p>
Full article ">Figure 7
<p>Dynamic Changes in Soluble Protein (SP) and Soluble Sugar (SS) Contents of Floral Organs during Flowering Stage of <span class="html-italic">C. drupifera.</span> Note: Panels (<b>A</b>,<b>B</b>) represent the Soluble Protein content and Soluble Sugar content, respectively.</p>
Full article ">Figure 8
<p>Correlation of Endogenous Substances in Floral Organs during Flowering Stage of <span class="html-italic">C. drupifera.</span> Note: (<b>A</b>): Correlation in pistils; (<b>B</b>): Correlation in stamens. Colors in the correlation analysis represent positive (red) and negative (blue) correlations, with darker colors indicating stronger correlations. “*” indicates significant correlation at the 0.05 level (<span class="html-italic">p</span> &lt; 0.05); “**” indicates highly significant correlation at the 0.01 level (<span class="html-italic">p</span> &lt; 0.01).</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 292
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 ">
18 pages, 4636 KiB  
Article
Exploring the Role of Bergamot Polyphenols in Alleviating Morphine-Induced Hyperalgesia and Tolerance through Modulation of Mitochondrial SIRT3
by Sara Ilari, Saverio Nucera, Lucia Carmela Passacatini, Federica Scarano, Roberta Macrì, Rosamaria Caminiti, Stefano Ruga, Maria Serra, Luigino Antonio Giancotti, Filomena Lauro, Concetta Dagostino, Valeria Mazza, Giovanna Ritorto, Francesca Oppedisano, Jessica Maiuolo, Ernesto Palma, Valentina Malafoglia, Carlo Tomino, Vincenzo Mollace and Carolina Muscoli
Nutrients 2024, 16(16), 2620; https://doi.org/10.3390/nu16162620 - 9 Aug 2024
Viewed by 341
Abstract
Morphine is an important pain reliever employed in pain management, its extended utilize is hindered by the onset of analgesic tolerance and oxidative stress. Long-term morphine administration causes elevated production of reactive oxygen species (ROS), disrupting mitochondrial function and inducing oxidation. Sirtuin 3 [...] Read more.
Morphine is an important pain reliever employed in pain management, its extended utilize is hindered by the onset of analgesic tolerance and oxidative stress. Long-term morphine administration causes elevated production of reactive oxygen species (ROS), disrupting mitochondrial function and inducing oxidation. Sirtuin 3 (SIRT3), a mitochondrial protein, is essential in modulating ROS levels by regulating mitochondrial antioxidant enzymes as manganese superoxide dismutase (MnSOD). Our investigation focused on the impact of SIRT3 on hyperalgesia and morphine tolerance in mice, as evaluating the antioxidant effect of the polyphenolic fraction of bergamot (BPF). Mice were administered morphine twice daily for four consecutive days (20 mg/kg). On the fifth day, mice received an acute dose of morphine (3 mg/kg), either alone or in conjunction with BPF or Mn (III)tetrakis (4-benzoic acid) porphyrin (MnTBAP). We evaluated levels of malondialdehyde (MDA), nitration, and the activity of SIRT3, MnSOD, glutamine synthetase (GS), and glutamate 1 transporter (GLT1) in the spinal cord. Our findings demonstrate that administering repeated doses of morphine led to the development of antinociceptive tolerance in mice, accompanied by increased superoxide production, nitration, and inactivation of mitochondrial SIRT3, MnSOD, GS, and GLT1. The combined administration of morphine with either BPF or MnTBAP prevented these effects. Full article
(This article belongs to the Special Issue Effects of Natural Bioactives on Pain and Neuroinflammation)
Show Figures

Figure 1

Figure 1
<p>Acute morphine injection. Administration of morphine (3 mg/kg), in mice, generated a considerable near-maximal antinociceptive response persisting for 60 min. Values are reported as mean ± SEM, based on 15 mice; * <span class="html-italic">p</span> &lt; 0.0001 vs. morphine 0 mg/kg.</p>
Full article ">Figure 2
<p>A significant loss to the antinociceptive effect of the acute injection of morphine was observed in animals that received repeated administration of morphine over 4 days. Concurrent administration of morphine with BPF (5–50 mg/kg) (<b>A</b>) or MnTBAP (5–30 mg/kg) (<b>B</b>) over a period of 4 days inhibited the development of tolerance in a dose-dependent manner. * <span class="html-italic">p</span> &lt; 0.001 compared to vehicle (Veh); † <span class="html-italic">p</span> &lt; 0.01; †† <span class="html-italic">p</span> &lt; 0.001 compared to vehicle + morphine.</p>
Full article ">Figure 3
<p>(<b>A</b>) Repeated administration of morphine for 4 days in mice caused an increased production of superoxide in the spinal cord compared to the control group (vehicle), as demonstrated by the oxidation of HE. Co-administration of morphine and BPF (25 mg/kg) or MnTBAP (10 mg/kg) was able to reduce the increase in ethidium and therefore in superoxide. Original magnification, ×10. Scale Bar 100 µm. Micrographs illustrate results from at least three distinct animals. (<b>B</b>) Persistent morphine treatment induced protein nitration in the spinal cord. Co-administration of morphine with BPF (25 mg/kg) and MnTBAP (10 mg/kg) inhibited nitrotyrosine formation. Original magnification, ×10. Scale Bar 100 µm. Micrographs illustrate results from at least three distinct animals in experiments conducted on separate days.</p>
Full article ">Figure 4
<p>Increased MDA levels in spinal cord represents the presence of oxidative stress during morphine tolerance in mice. Mice that received morphine for 4 days showed an amount of MDA level. Co-administration of morphine and BPF (25 mg/kg) or MnTBAP (10 mg/kg) resulted in a substantial decrease in MDA. The results are shown as the mean ± SEM for 6 mice. * <span class="html-italic">p</span> &lt; 0.0001 versus vehicle (Veh); † <span class="html-italic">p</span> &lt; 0.01 versus vehicle + morphine.</p>
Full article ">Figure 5
<p>(<b>A</b>,<b>B</b>) Nitration of GS and GLT1 proteins in spinal cord tissues as assessed by immunoprecipitation. Administering morphine for 4 days in combination with BPF (25 mg/kg) or MnTBAP (10 mg/kg) prevented the nitration of GS and GLT1. In these conditions, actin expression appeared statistically similar across the lanes. The reported data include densitometric analyses for all animals per group. GS, nitrated GS, GLT1 and nitrated GLT1 were first normalized to actin and then these values were used to obtain GS nitrated/GS and GLT1 nitrated/GLT1 ratio. The data are presented as the mean ± SEM for 6 mice; * <span class="html-italic">p</span> &lt; 0.001 versus Veh; † <span class="html-italic">p</span> &lt; 0.001 versus morphine.</p>
Full article ">Figure 6
<p>Chronic morphine administration in mice induced nitration on mitochondrial proteins as shown by WES methodology. Co-administration of BPF (25 mg/kg) or MnTBAP (10 mg/kg) attenuated mitochondrial proteins nitration. No statistically significant difference for the TOM20 value was identified in the lanes under these conditions. (<b>A</b>,<b>B</b>) Lanes and (<b>D</b>) electropherogram are representative of the results from six animals; (blue lanes: morphine groups; pink lanes: BPF group; green lanes: MnTBAP group; and light grey: vehicle group). (<b>C</b>) The reported data include densitometric analyses for all animals per group. The results are presented as the mean ± SEM for six mice. * <span class="html-italic">p</span> &lt; 0.05 versus Veh; † <span class="html-italic">p</span> &lt; 0.05 versus vehicle + morphine.</p>
Full article ">Figure 7
<p>(<b>A</b>) Nitration of MnSOD protein in spinal cord tissues as assessed by immunoprecipitation. Combined treatment with morphine and BPF (25 mg/kg) or MnTBAP (10 mg/kg), for four consecutive days, prevented MnSOD nitration. In these conditions, prohibitin expression appeared statistically similar across the lanes. Densitometric analyses for all animals in each group are reported. MnSOD and nitrated MnSOD were first normalized with prohibitin and then these values were used to obtain MnSOD nitrated/MnSOD ratio. (<b>B</b>) Nitration on MnSOD is linked to inactivation of its biological function, which is restored following the administration of BPF (25 mg/kg) or MnTBAP (10 mg/kg). The results are presented as the mean ± SEM for six mice; * <span class="html-italic">p</span> &lt; 0.001 compared to Veh; † <span class="html-italic">p</span> &lt; 0.001 compared to morphine.</p>
Full article ">Figure 8
<p>(<b>A</b>) Nitration of SIRT3 protein in spinal cord tissues is identified by immunoprecipitation. Combined treatment with morphine and BPF (25 mg/kg) or MnTBAP (10 mg/kg), for 4 consecutive days, blocked SIRT3 nitration. Prohibitin levels appeared statistically similar across the lanes. Densitometric analyses for all animals in each group are reported. MnSOD and nitrated MnSOD were initially normalized using prohibitin, and these values were then utilized to calculate the MnSOD nitrated/MnSOD ratio. (<b>B</b>) SIRT3 activation, expressed in arbitrary fluorescence units (AFU), is restored following the administration of BPF (25 mg/kg) or MnTBAP (10 mg/kg). The results are presented as the mean ± SEM for six mice; * <span class="html-italic">p</span> &lt; 0.001 compared to Veh; † <span class="html-italic">p</span> &lt; 0.001 compared to morphine.</p>
Full article ">Figure 9
<p>(<b>A</b>) SIRT3 inhibition induces acetylation on mitochondrial proteins during morphine tolerance in mice as shown by WES methodology. Co-administration of BPF (25 mg/kg) or MnTBAP (10 mg/kg) attenuated mitochondrial proteins acetylation. No statistically significant difference in TOM20 value was observed between the lanes under these conditions. (<b>A</b>,<b>B</b>) Lanes and (<b>D</b>) electropherogram are representative of results from six animals (green lanes: morphine groups; blue lanes: BPF group; dark grey lanes: MnTBAP group; and light grey lanes: vehicle group). (<b>C</b>) Densitometric analyses for all animals in each group are reported. Values are presented as the mean ± SEM for six mice. * <span class="html-italic">p</span> &lt; 0.05 vs Veh; † <span class="html-italic">p</span> &lt; 0.05 vs. vehicle + morphine.</p>
Full article ">
20 pages, 1219 KiB  
Review
Post-Translational Variants of Major Proteins in Amyotrophic Lateral Sclerosis Provide New Insights into the Pathophysiology of the Disease
by Léa Bedja-Iacona, Elodie Richard, Sylviane Marouillat, Céline Brulard, Tarek Alouane, Stéphane Beltran, Christian R. Andres, Hélène Blasco, Philippe Corcia, Charlotte Veyrat-Durebex and Patrick Vourc’h
Int. J. Mol. Sci. 2024, 25(16), 8664; https://doi.org/10.3390/ijms25168664 - 8 Aug 2024
Viewed by 398
Abstract
Post-translational modifications (PTMs) affecting proteins during or after their synthesis play a crucial role in their localization and function. The modification of these PTMs under pathophysiological conditions, i.e., their appearance, disappearance, or variation in quantity caused by a pathological environment or a mutation, [...] Read more.
Post-translational modifications (PTMs) affecting proteins during or after their synthesis play a crucial role in their localization and function. The modification of these PTMs under pathophysiological conditions, i.e., their appearance, disappearance, or variation in quantity caused by a pathological environment or a mutation, corresponds to post-translational variants (PTVs). These PTVs can be directly or indirectly involved in the pathophysiology of diseases. Here, we present the PTMs and PTVs of four major amyotrophic lateral sclerosis (ALS) proteins, SOD1, TDP-43, FUS, and TBK1. These modifications involve acetylation, phosphorylation, methylation, ubiquitination, SUMOylation, and enzymatic cleavage. We list the PTM positions known to be mutated in ALS patients and discuss the roles of PTVs in the pathophysiological processes of ALS. In-depth knowledge of the PTMs and PTVs of ALS proteins is needed to better understand their role in the disease. We believe it is also crucial for developing new therapies that may be more effective in ALS. Full article
(This article belongs to the Collection Feature Paper Collection in Biochemistry)
Show Figures

Figure 1

Figure 1
<p>Functional consequences of PTV (chemical modification or cleavage): WT protein or mutated protein (pathological variant) with a new PTV (black dots) and consequences on protein interaction, function, localization, and half-life.</p>
Full article ">Figure 2
<p>Schematic representations of SOD1, TDP-43, FUS, and TBK, with their post-translational modification sites.</p>
Full article ">
16 pages, 11940 KiB  
Article
Overexpression of Potato PYL16 Gene in Tobacco Enhances the Transgenic Plant Tolerance to Drought Stress
by Panfeng Yao, Chunli Zhang, Zhenzhen Bi, Yuhui Liu, Zhen Liu, Jia Wei, Xinglong Su, Jiangping Bai, Junmei Cui and Chao Sun
Int. J. Mol. Sci. 2024, 25(16), 8644; https://doi.org/10.3390/ijms25168644 - 8 Aug 2024
Viewed by 253
Abstract
PYR/PYL/RCAR proteins are abscisic acid (ABA) receptors that play a crucial role in plant responses to abiotic stresses. However, there have been no research reports on potato PYL so far. In this study, a potato PYL gene named StPYL16 was identified based on [...] Read more.
PYR/PYL/RCAR proteins are abscisic acid (ABA) receptors that play a crucial role in plant responses to abiotic stresses. However, there have been no research reports on potato PYL so far. In this study, a potato PYL gene named StPYL16 was identified based on transcriptome data under drought stress. Molecular characteristics analysis revealed that the StPYL16 protein possesses an extremely conserved PYL family domain. The tissue expression results indicated that the StPYL16 is predominantly expressed at high levels in the underground parts, particularly in tubers. Abiotic stress response showed that StPYL16 has a significant response to drought treatment. Further research on the promoter showed that drought stress could enhance the activation activity of the StPYL16 promoter on the reporter gene. Then, transient and stable expression of StPYL16 in tobacco enhanced the drought resistance of transgenic plants, resulting in improved plant height, stem thickness, and root development. In addition, compared with wild-type plants, StPYL16 transgenic tobacco exhibited lower malondialdehyde (MDA) content, higher proline accumulation, and stronger superoxide dismutase (SOD), peroxidase (POD), and catalase (CAT) activities. Meanwhile, StPYL16 also up-regulated the expression levels of stress-related genes (NtSOD, NtCAT, NtPOD, NtRD29A, NtLEA5, and NtP5CS) in transgenic plants under drought treatment. These findings indicated that the StPYL16 gene plays a positive regulatory role in potato responses to drought stress. Full article
(This article belongs to the Special Issue Physiology and Molecular Biology of Plant Stress Tolerance)
Show Figures

Figure 1

Figure 1
<p>Tissue-specific expression of <span class="html-italic">StPYL16</span> gene. Data from Spud DB Potato Genomics Resource website (<a href="http://spuddb.uga.edu/" target="_blank">http://spuddb.uga.edu/</a>, accessed on 23 May 2024).</p>
Full article ">Figure 2
<p>The relative expression level of the <span class="html-italic">StPYL16</span> gene under 100 µM ABA and 200 mM mannitol stress treatments. ‘Atl’ and ‘QS9’ represent drought-sensitive and drought-tolerant potato varieties, respectively. For the ABA treatment, 20-day-old potato seedlings were subjected to spraying with 100 M ABA, while other 20-day-old seedlings were treated with ethanol as a control. For the drought treatment, 20-day-old potato seedlings were transferred to a liquid MS medium containing 200 mM mannitol, with seedlings under normal liquid MS conditions serving as the control. Above-ground samples were collected for gene expression analysis at 0, 1, 3, 6, and 12 h post stress induction under both treatment conditions. Data represent the means ± SD of three replicates. * and ** indicate significant difference at <span class="html-italic">p</span> &lt; 0.05 and <span class="html-italic">p</span> &lt; 0.01 levels, respectively. ns indicates that the difference is not significant.</p>
Full article ">Figure 3
<p>Stress response analysis of <span class="html-italic">StPYL16</span> gene promoter. (<b>a</b>) Analysis of stress response elements in <span class="html-italic">StPYL16</span> gene promoter. (<b>b</b>) GUS histological staining analysis of transient transformation of tobacco with <span class="html-italic">StPYL16</span> gene promoter under 100 µM ABA and 200 mM mannitol treatment.</p>
Full article ">Figure 4
<p>Identification of drought resistance in tobacco after transient transformation of <span class="html-italic">StPYL16</span> gene. (<b>a</b>) GUS histochemical staining of <span class="html-italic">StPYL16</span> transgenic plants; (<b>b</b>) phenotypic collection of each genotype before and after drought stress; (<b>c</b>) determination of physiological indexes related to stress. *, **, ***, and **** indicate significant difference 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, and <span class="html-italic">p</span> &lt; 0.0001 levels, respectively. ns indicates that the difference is not significant.</p>
Full article ">Figure 5
<p>Positive identification of <span class="html-italic">StPYL16</span> transgenic tobacco. (<b>a</b>) The process of transforming <span class="html-italic">StPYL16</span> into tobacco; (<b>b</b>) PCR molecular identification of positive <span class="html-italic">StPYL16</span> transgenic lines; (<b>c</b>) GUS histochemical staining of <span class="html-italic">StPYL16</span> transgenic lines and control plants; (<b>d</b>) analysis of expression level of <span class="html-italic">StPYL16</span> in transgenic lines.</p>
Full article ">Figure 6
<p>Drought-resistance function of <span class="html-italic">StPYL16</span> after stable transformation of tobacco. Transgenic tobacco stem segments with consistent growth were transferred to MS medium as well as MS medium supplemented with 100 mM, 200 mM, and 300 mM mannitol for 30 days to stress, followed by the measurement of various phenotypic traits. Additionally, root traits of the plants were analyzed using a root scanner (LD-WinRHIZO). (<b>a</b>) Phenotypic identification of plants under drought stress; (<b>b</b>) statistics of phenotypic indicators of plants under drought stress; (<b>c</b>) root scanning diagram of each plant under different treatment conditions; (<b>d</b>) determination of root system related indexes of various plants under different treatment conditions. * and ** indicate significant differences at <span class="html-italic">p</span> &lt; 0.05 and <span class="html-italic">p</span> &lt; 0.01 levels, respectively. ns indicates that the difference is not significant.</p>
Full article ">Figure 7
<p>Determination of malondialdehyde (MDA) and proline (Pro) contents in different plants under different treatment conditions. Transgenic tobacco stem segments with consistent growth were transferred to MS medium as well as MS medium supplemented with 100 mM, 200 mM, and 300 mM mannitol. After 30 days of stress treatment, aboveground tissues were selected for the measurement of physiological indices. ** indicates significant difference at the <span class="html-italic">p</span> &lt; 0.01 level. ns indicates that the difference is not significant.</p>
Full article ">Figure 8
<p>Expression analysis of stress-related genes in transgenic plants and wild-type plants. Transgenic tobacco stem segments exhibiting consistent growth were transferred to MS medium and MS medium supplemented with 100 mM mannitol. Following a 30-day stress treatment period, aboveground tissues were collected for gene expression analysis. (<b>a</b>–<b>f</b>) represent <span class="html-italic">NtSOD</span>, <span class="html-italic">NtCAT</span>, <span class="html-italic">NtPOD</span>, <span class="html-italic">NtRD29A</span>, <span class="html-italic">NtLEA5</span>, and <span class="html-italic">NtP5CS</span> genes, respectively. ** represents significant differences between transgenic lines and WT at <span class="html-italic">p</span> &lt; 0.01. ns indicates that the difference is not significant.</p>
Full article ">
21 pages, 4578 KiB  
Article
Promoting Anthocyanin Biosynthesis in Purple Lettuce through Sucrose Supplementation under Nitrogen Limitation
by Chunhui Liu, Haiye Yu, Yucheng Liu, Lei Zhang, Dawei Li, Xiaoman Zhao, Junhe Zhang and Yuanyuan Sui
Horticulturae 2024, 10(8), 838; https://doi.org/10.3390/horticulturae10080838 - 8 Aug 2024
Viewed by 402
Abstract
Although nitrogen deficiency and sucrose are linked to anthocyanin synthesis, the potential role of sucrose in regulating anthocyanin biosynthesis under low nitrogen conditions (LN) in purple lettuce (Lactuca sativa L.) remains unclear. We found that adding exogenous sucrose enhanced anthocyanin biosynthesis but [...] Read more.
Although nitrogen deficiency and sucrose are linked to anthocyanin synthesis, the potential role of sucrose in regulating anthocyanin biosynthesis under low nitrogen conditions (LN) in purple lettuce (Lactuca sativa L.) remains unclear. We found that adding exogenous sucrose enhanced anthocyanin biosynthesis but significantly inhibited lettuce growth at high concentrations. Optimal results were obtained using 1 mmol/L sucrose in a low-nitrogen nutrient solution (LN + T1). Chlorophyll fluorescence imaging indicated that the addition of exogenous sucrose induced mild stress. Meanwhile, the activities of antioxidant enzymes (SOD, CAT, and POD) and antioxidant capacity were both enhanced. The mild stress activated the antioxidant system, thereby promoting the accumulation of anthocyanins induced by exogenous sucrose. LN + T1 (low nitrogen nutrient solution supplemented with 1 mmol/L sucrose) up-regulated enzyme genes in the biosynthetic pathway of anthocyanins, including phenylalanine ammonia-lyase (PAL), chalcone synthase (CHS), dihydroflavonol reductase (DFR), flavanone 3-hydroxylase (F3H), flavonoid 3′-hydroxylase (F3′H), flavone synthase II (FNSII), and anthocyanidin synthase (ANS). Additionally, various transcription factors such as AP2/ERF, MYB, bHLH, C2H2, NAC, C2C2, HB, MADS, bZIP, and WRKY were found to be up-regulated. This study elucidates the regulatory mechanism of anthocyanin metabolism in response to the addition of exogenous sucrose under low nitrogen conditions and provides a nutrient solution formula to enhance anthocyanin content in modern, high-quality agricultural cultivation. Full article
Show Figures

Figure 1

Figure 1
<p>The effect of exogenously applied sucrose combined with a low nitrogen nutrient solution on chlorophyll fluorescence imaging of the lettuce canopy. Purple lettuces were grown under various nutrient solution treatments: CK, CK + T1, CK + T2, CK + T3, LN, LN + T1, LN + T2, and LN + T3 (CK: normal nitrogen nutrient solution, LN: low nitrogen nutrient solution). The sucrose concentrations used were T1: 1 mmol/L, T2: 3 mmol/L, and T3: 5 mmol/L. LN + T1: low nitrogen nutrient solution supplemented with 1 mmol/L sucrose; LN + T2: low nitrogen nutrient solution supplemented with 3 mmol/L sucrose. LN + T3: low nitrogen nutrient solution supplemented with 5 mmol/L sucrose; CK + T1: normal nitrogen nutrient solution supplemented with 1 mmol/L sucrose; CK + T2: normal nitrogen nutrient solution supplemented with 3 mmol/L sucrose; CK + T3: normal nitrogen nutrient solution supplemented with 5 mmol/L sucrose.</p>
Full article ">Figure 2
<p>The effect of exogenously applied sucrose combined with a low nitrogen nutrient solution on the maximum quantum yield of photosystem II (Fv/Fm, (<b>a</b>)), actual photochemical efficiency of PSII under light adaptation (Fq′/Fm′, (<b>b</b>)), electron transport rate (ETR, (<b>c</b>)), and non-photochemical quenching coefficient (NPQ, (<b>d</b>)). Purple lettuces were grown under various nutrient solution treatments: CK, CK + T1, CK + T2, CK + T3, LN, LN + T1, LN + T2, and LN + T3 (CK: normal nitrogen nutrient solution, LN: low nitrogen nutrient solution). The sucrose concentrations used were T1: 1 mmol/L, T2: 3 mmol/L, and T3: 5 mmol/L. LN + T1: low nitrogen nutrient solution supplemented with 1 mmol/L sucrose; LN + T2: low nitrogen nutrient solution supplemented with 3 mmol/L sucrose. LN + T3: low nitrogen nutrient solution supplemented with 5 mmol/L sucrose; CK + T1: normal nitrogen nutrient solution supplemented with 1 mmol/L sucrose; CK + T2: normal nitrogen nutrient solution supplemented with 3 mmol/L sucrose; CK + T3: normal nitrogen nutrient solution supplemented with 5 mmol/L sucrose. Different lower-case letters above the bars indicated significant differences between treatments by Duncan’s multiple range test at a level of 0.05.</p>
Full article ">Figure 3
<p>The correlation analysis of samples (<b>a</b>), the volcanic map of differentially expressed genes (<b>b</b>), the clustering heat map of differentially expressed genes (<b>c</b>), C: normal nitrogen nutrient solution (CK), L: low nitrogen nutrient solution supplemented with 1 mmol/L sucrose (LN + T1).</p>
Full article ">Figure 4
<p>Functional analysis of Gene Ontology (GO) terms for the differentially expressed genes (DEGs) under low nitrogen nutrient solution supplemented with 1 mmol/L sucrose (LN + T1).</p>
Full article ">Figure 5
<p>GO enrichment analysis of the differentially expressed genes (DEGs) in purple lettuce treated with exogenously applied sucrose (1 mmol/L) in combination with a low nitrogen nutrient solution (LN + T1). GO classification of up-regulated DEGs in biological processes (<b>a</b>). GO classification of up-regulated DEGs in cellular processes (<b>b</b>).</p>
Full article ">Figure 6
<p>KEGG enrichment analysis of the differentially expressed genes (DEGs) under nitrogen deficiency nutrient solution supplemented with 1 mmol/L sucrose (LN + T1). KEGG classification of differentially expressed genes (<b>a</b>), where the vertical axis (left) represents secondary pathways, and the vertical axis (right) represents primary pathways. Enrichment map of differentially expressed genes in the top 20 gene pathways of KEGG (<b>b</b>).</p>
Full article ">Figure 7
<p>Anthocyanin biosynthesis pathway. The gene name in red font indicates that its expression level increased under nitrogen deficiency nutrient solution supplemented with 1 mmol/L sucrose (LN + T1), and blue indicates no significant change.</p>
Full article ">Figure 8
<p>Statistical map of differential expressions of transcription factors.</p>
Full article ">Figure 9
<p>The relative expression of the differentially expressed genes (DEGs) between exogenous sucrose coupled with low nitrogen nutrient solution (LN + T1) and CK (CK: normal nitrogen nutrient solution, LN + T1: low nitrogen nutrient solution supplemented with 1 mmol/L sucrose).</p>
Full article ">
14 pages, 3990 KiB  
Article
Berberine Attenuates Acetamiprid Exposure-Induced Mitochondrial Dysfunction and Apoptosis in Rats via Regulating the Antioxidant Defense System
by Annu Phogat, Jagjeet Singh, Reena Sheoran, Arun Hasanpuri, Aakash Chaudhary, Shakti Bhardwaj, Sandeep Antil, Vijay Kumar, Chandra Prakash and Vinay Malik
J. Xenobiot. 2024, 14(3), 1079-1092; https://doi.org/10.3390/jox14030061 - 7 Aug 2024
Viewed by 463
Abstract
Acetamiprid (ACMP) is a neonicotinoid insecticide that poses a significant threat to the environment and mankind. Oxidative stress and mitochondrial dysfunction are considered prime contributors to ACMP-induced toxic effects. Meanwhile, berberine (BBR) a natural plant alkaloid, is a topic of interest because of [...] Read more.
Acetamiprid (ACMP) is a neonicotinoid insecticide that poses a significant threat to the environment and mankind. Oxidative stress and mitochondrial dysfunction are considered prime contributors to ACMP-induced toxic effects. Meanwhile, berberine (BBR) a natural plant alkaloid, is a topic of interest because of its therapeutic and prophylactic actions. Therefore, this study evaluated the effects of BBR on ACMP-mediated alterations in mitochondrial functions and apoptosis in rat liver tissue. Male Wistar rats were divided into four groups: (I) control, (II) BBR-treated, (III) ACMP-exposed, and (IV) BBR+ACMP co-treated groups. The doses of BBR (150 mg/kg b.wt) and ACMP (1/10 of LD50, i.e., 21.7 mg/kg b.wt) were given intragastrically for 21 consecutive days. The results showed that the administration of ACMP diminished mitochondrial complex activity, downregulated complex I (ND1 and ND2) and complex IV (COX1 and COX4) subunit mRNA expression, depleted the antioxidant defense system, and induced apoptosis in rat liver. BBR pre-treatment significantly attenuated ACMP-induced mitochondrial dysfunction by maintaining mitochondrial complex activity and upregulating ND1, ND2, COX1, and COX4 mRNA expression. BBR reversed ACMP-mediated apoptosis by diminishing Bax and caspase-3 and increasing the Bcl-2 protein level. BBR also improved the mitochondrial antioxidant defense system by upregulating mRNA expression of PGC-1α, MnSOD, and UCP-2 in rat liver tissue. This study is the first to evaluate the protective potential of BBR against pesticide-induced mitochondrial dysfunction in liver tissue. In conclusion, BBR offers protection against ACMP-induced impairment in mitochondrial functions by maintaining the antioxidant level and modulating the apoptotic cascade. Full article
Show Figures

Figure 1

Figure 1
<p>Effects of BBR supplementation and ACMP exposure on (<b>a</b>) complex I, (<b>b</b>) complex II, and (<b>c</b>) complex IV activity of mitochondria isolated from liver tissue of different experimental groups. Results are expressed as mean ± SD of 5 rats. *** is significant at <span class="html-italic">p</span> &lt; 0.001 as compared with control rats. <sup>###</sup> is significant at <span class="html-italic">p</span> &lt; 0.001 as compared with ACMP-exposed rats.</p>
Full article ">Figure 2
<p>Effects of BBR supplementation and ACMP exposure on mRNA expression of mitochondrial complex subunits in liver tissue of different experimental groups. Mean relative mRNA level of (<b>b</b>) ND1, (<b>c</b>) ND2, (<b>d</b>) COX1, and (<b>e</b>) COX4 with respect to β-actin. Results are expressed as mean ± SD of 3 rats. ** is significant at <span class="html-italic">p</span> &lt; 0.01 and *** is significant at <span class="html-italic">p</span> &lt; 0.001 as compared with control rats. <sup>#</sup> is significant at <span class="html-italic">p</span> &lt; 0.05, <sup>##</sup> is significant at <span class="html-italic">p</span> &lt; 0.01, and NS is non-significant as compared with ACMP-exposed rats.</p>
Full article ">Figure 3
<p>Effects of BBR supplementation and ACMP exposure on mRNA expression of mitochondrial antioxidants in liver tissue of different experimental groups. Mean relative mRNA levels of (<b>b</b>) PGC-1α, (<b>c</b>) MnSOD, and (<b>d</b>) UCP-2 were analyzed using densitometric analysis with respect to β-actin. Results are expressed as mean ± SD of 3 rats. *** is significant at <span class="html-italic">p</span> &lt; 0.001 as compared with control rats. <sup>#</sup> is significant at <span class="html-italic">p</span> &lt; 0.05, <sup>##</sup> is significant at <span class="html-italic">p</span> &lt; 0.01, and <sup>###</sup> is significant at <span class="html-italic">p</span> &lt; 0.001 as compared with ACMP-exposed rats.</p>
Full article ">Figure 4
<p>(<b>a</b>) Western blotting depicting the effects of BBR supplementation and ACMP exposure on apoptotic marker proteins in liver tissue of different experimental groups. Relative expression of (<b>b</b>) Bax, (<b>c</b>) Bcl-2, and (<b>d</b>) caspase-3 with respect to β-actin. Results are expressed as mean ± SD of 3 rats. *** is significant at <span class="html-italic">p</span> &lt; 0.001 as compared with control rats. <sup>#</sup> is significant at <span class="html-italic">p</span> &lt; 0.05 as compared with ACMP-exposed rats.</p>
Full article ">Figure 4 Cont.
<p>(<b>a</b>) Western blotting depicting the effects of BBR supplementation and ACMP exposure on apoptotic marker proteins in liver tissue of different experimental groups. Relative expression of (<b>b</b>) Bax, (<b>c</b>) Bcl-2, and (<b>d</b>) caspase-3 with respect to β-actin. Results are expressed as mean ± SD of 3 rats. *** is significant at <span class="html-italic">p</span> &lt; 0.001 as compared with control rats. <sup>#</sup> is significant at <span class="html-italic">p</span> &lt; 0.05 as compared with ACMP-exposed rats.</p>
Full article ">Figure 5
<p>Photomicrographs representing the transmission electron microscopy of liver tissues of: (<b>a</b>) control groups; (<b>b</b>) BBR-treated group depicting nucleus (NU) normal distribution of mitochondria (Mt) and endoplasmic reticulum (ER); (<b>c</b>) ACMP-exposed group depicting chromatin condensation (red arrow), disruption of mitochondria (yellow arrow), loss of mitochondria (red star), and endoplasmic reticulum; and (<b>d</b>) BBR+ACMP co-treated group representing normal shape and distribution of mitochondria.</p>
Full article ">
Back to TopTop