[go: up one dir, main page]

 
 
Sign in to use this feature.

Years

Between: -

Subjects

remove_circle_outline
remove_circle_outline
remove_circle_outline
remove_circle_outline
remove_circle_outline
remove_circle_outline
remove_circle_outline
remove_circle_outline
remove_circle_outline

Journals

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

Article Types

Countries / Regions

remove_circle_outline
remove_circle_outline
remove_circle_outline
remove_circle_outline
remove_circle_outline
remove_circle_outline
remove_circle_outline
remove_circle_outline
remove_circle_outline
remove_circle_outline

Search Results (16,920)

Search Parameters:
Keywords = ROS

Order results
Result details
Results per page
Select all
Export citation of selected articles as:
21 pages, 6102 KiB  
Article
Mechanisms of Action of Sea Cucumber Triterpene Glycosides Cucumarioside A0-1 and Djakonovioside A Against Human Triple-Negative Breast Cancer
by Ekaterina S. Menchinskaya, Ekaterina A. Chingizova, Evgeny A. Pislyagin, Ekaterina A. Yurchenko, Anna A. Klimovich, Elena. A. Zelepuga, Dmitry L. Aminin, Sergey A. Avilov and Alexandra S. Silchenko
Mar. Drugs 2024, 22(10), 474; https://doi.org/10.3390/md22100474 (registering DOI) - 17 Oct 2024
Abstract
Breast cancer is the most prevalent form of cancer in women worldwide. Triple-negative breast cancer is the most unfavorable for patients, but it is also the most sensitive to chemotherapy. Triterpene glycosides from sea cucumbers possess a high therapeutic potential as anticancer agents. [...] Read more.
Breast cancer is the most prevalent form of cancer in women worldwide. Triple-negative breast cancer is the most unfavorable for patients, but it is also the most sensitive to chemotherapy. Triterpene glycosides from sea cucumbers possess a high therapeutic potential as anticancer agents. This study aimed to identify the pathways triggered and regulated in MDA-MB-231 cells (triple-negative breast cancer cell line) by the glycosides cucumarioside A0-1 (Cuc A0-1) and djakonovioside A (Dj A), isolated from the sea cucumber Cucumaria djakonovi. Using flow cytometry, fluorescence microscopy, immunoblotting, and ELISA, the effects of micromolar concentrations of the compounds on cell cycle arrest, induction of apoptosis, the level of reactive oxygen species (ROS), mitochondrial membrane potential (Δψm), and expression of anti- and pro-apoptotic proteins were investigated. The glycosides caused cell cycle arrest, stimulated an increase in ROS production, and decreased Δψm in MDA-MB-231 cells. The depolarization of the mitochondrial membrane caused by cucumarioside A0-1 and djakonovioside A led to an increase in the levels of APAF-1 and cytochrome C. This, in turn, resulted in the activation of caspase-9 and caspase-3 and an increase in the level of their cleaved forms. Glycosides also affected the expression of Bax and Bcl-2 proteins, which are associated with mitochondria-mediated apoptosis in MDA-MB-231 cells. These results indicate that cucumarioside A0-1 and djakonovioside A activate the intrinsic apoptotic pathway in triple-negative breast cancer cells. Additionally, it was found that treatment with Cuc A0-1 resulted in in vivo inhibition of tumor growth and metastasis of murine solid Ehrlich adenocarcinoma. Full article
(This article belongs to the Collection Marine Compounds and Cancer)
Show Figures

Graphical abstract

Graphical abstract
Full article ">Figure 1
<p>Chemical structures of triterpene glycosides: cucumarioside A<sub>0</sub>-1 (<b>a</b>) and djakonovioside A (<b>b</b>) isolated from the sea cucumber <span class="html-italic">Cucumaria djakonovi</span>.</p>
Full article ">Figure 2
<p>Distribution of MDA-MB-231 cells according to the phases of the cell cycle after treatment with various concentrations of Cuc A<sub>0</sub>-1 and Dj A for 24 h.</p>
Full article ">Figure 3
<p>Visualization of cyclin B and A contents and cyclin-dependent kinases in MDA-MB-231 cells treated with triterpene glycosides Cuc A<sub>0</sub>-1 and Dj A at different concentrations. Representative Western blot membranes showing the effect of glycosides on cyclin and CDK protein expression levels (<b>a</b>). Processed data on cyclin B content in MDA-MB-231 cells treated with Cuc A<sub>0</sub>-1 (<b>b</b>). Processed data on cyclin A content in MDA-MB-231 cells treated with Dj A (<b>c</b>). Processed data on CDK-1 content in MDA-MB-231 cells treated with Cuc A<sub>0</sub>-1 (<b>d</b>). Processed data on CDK-2 content in MDA-MB-231 cells treated with Dj A (<b>e</b>). All data were normalized to the β-actin levels. Data are presented as means ± SEM. * <span class="html-italic">p</span> value &lt; 0.05 was considered significant.</p>
Full article ">Figure 4
<p>Analysis of apoptosis induced by triterpene glycosides in MDA-MB-231 cells after 24 h of incubation. Flow cytometry assay for Annexin V-FITC/PI staining (<b>a</b>). Quantitative calculation of the data obtained via flow cytometry: Cuc A<sub>0</sub>-1 (0.5 and 1 μM)—(<b>b</b>) and Dj A (1 and 2 μM)—(<b>c</b>). Data are presented as means ± SEM. * <span class="html-italic">p</span> value &lt; 0.05 was considered significant. Apoptosis assay using Hoechst 33342 in a fluorescent microscopy analysis (<b>d</b>). Hoechst 33342 staining showed an increase in chromatin condensation and DNA fragmentation in apoptotic cells treated with Cuc A<sub>0</sub>-1 (1 μM) and Dj A (2 μM) compared with untreated control cells. Arrows indicate nuclei with condensed chromatin.</p>
Full article ">Figure 5
<p>Quantitative evaluation of ROS levels in MDA-MB-231 cells after incubation with Cuc A<sub>0</sub>-1 (<b>a</b>) and Dj A (<b>b</b>) for different times (6, 12, and 24 h) using the fluorescent dye H<sub>2</sub>DCF-DA. Glycosides Cuc A<sub>0</sub>-1 (<b>c</b>) and Dj A (<b>d</b>), at various concentrations, reduced the mitochondrial membrane potential (Δψm), as measured using the fluorescent dye TMRE. Data are presented as means ± SEM. * <span class="html-italic">p</span> value &lt; 0.05 was considered significant. Staining of MDA-MB-231 cells with the fluorescent dye JC-1 showed a change in the mitochondrial membrane potential (<b>e</b>).</p>
Full article ">Figure 6
<p>Western blot analysis of cytoplasmic proteins: apoptosis promoter Bax (<b>a</b>,<b>b</b>) and apoptosis inhibitor Bcl-2 (<b>a</b>,<b>c</b>) with β-actin as a protein loading control under the treatment of MDA-MB-231 cells with different concentrations of Cuc A<sub>0</sub>-1 and Dj A. Cytoplasmic protein levels were normalized to the control group (untreated cells). * <span class="html-italic">p</span> &lt; 0.05 compared with untreated MDA-MB-231 cells.</p>
Full article ">Figure 7
<p>Quantitative assessment of the contents of cytochrome C (<b>a</b>,<b>b</b>) and APAF-1 (<b>c</b>,<b>d</b>) in MDA-MB-231 cells after treatment with different concentrations of glycosides Cuc A<sub>0</sub>-1 (<b>a</b>,<b>c</b>) and Dj A (<b>b</b>,<b>d</b>) at different times (6, 12, and 24 h) using ELISA kits. * <span class="html-italic">p</span> &lt; 0.05 compared with untreated MDA-MB-231 cells.</p>
Full article ">Figure 8
<p>Caspase-3/7 activity in the control cells and cells treated with triterpene glycosides for 12 and 24 h was measured using the Muse™ Caspase-3/7 Kit and flow cytometry in MDA-MB-231 cells.</p>
Full article ">Figure 9
<p>Western blot analysis of apoptotic markers (<b>a</b>) and quantitative analysis of the levels of cleaved caspase-9 (<b>b</b>), cleaved caspase-3 (<b>c</b>), and cleaved PARP-1 (<b>d</b>) in MDA-MB-231 cells treated with different concentrations of Cuc A<sub>0</sub>-1 and Dj A. β-actin was used as a protein loading control (<b>a</b>). The levels of apoptotic markers were normalized to those of the control group (untreated cells). * <span class="html-italic">p</span> &lt; 0.05 compared to untreated MDA-MB-231 cells.</p>
Full article ">Figure 10
<p>Influence of Cuc A<sub>0</sub>-1 on the area (<b>a</b>) and integrated density (<b>b</b>) of the fluorescence zone detected by in vivo fluorescence imager, Fluor I. Effect of Cuc A<sub>0</sub>-1 (0.4 µg/mL) on tumor volume (<b>c</b>) and tumor growth index (<b>d</b>). The data are presented as a mean ± SEM (n = 7). Asterisks indicate the significance of the differences at <span class="html-italic">p</span> ≤ 0.05 * and <span class="html-italic">p</span> ≤ 0.01 ** according to one-factor analysis of variance (ANOVA) with Tukey’s correction.</p>
Full article ">Figure 11
<p>The visualization of tumor cells labeled with PKH800 NIR fluorescent dye using the fluorescence imager system, “Fluor I IN VIVO”, in untreated mice (<b>a</b>), mice treated with Cuc A<sub>0</sub>-1 in group II (<b>b</b>), mice treated with Cuc A<sub>0</sub>-1 in group III (<b>c</b>), and mice treated with doxorubicin in group IV (<b>d</b>). On day 12, the tumor area was visualized in live mice; afterward, the mice were euthanized, the skin was opened, and tumor cells were visualized again. Arrows indicate tumor metastasis in the abdominal cavity.</p>
Full article ">Figure 12
<p>Three-dimensional plot of cytotoxic activity (pIC<sub>50</sub>) dependence on the principal component values (PCA1–PCA3) calculated for 25 conformational forms of 20 glycosides tested against MDA-MB-231 cells. The glycosides demonstrating cytotoxic activity with IC<sub>50</sub> ≤ 10 μM were outlined as active and are marked in red, while inactive glycosides are marked in violet.</p>
Full article ">Figure 13
<p>The PLS QSAR model correlation plot reflecting the relationship of predicted and experimental cytotoxicity of the glycosides against MDA-MB-231 cells. The cytotoxic action was expressed as pIC<sub>50</sub>.</p>
Full article ">
15 pages, 1789 KiB  
Article
A Comparison-Based Framework for Argument Quality Assessment
by Jianzhu Bao, Bojun Jin, Yang Sun, Yice Zhang, Yuhang He and Ruifeng Xu
Electronics 2024, 13(20), 4088; https://doi.org/10.3390/electronics13204088 (registering DOI) - 17 Oct 2024
Abstract
Assessing the quality of arguments is both valuable and challenging. Humans often find that making pairwise comparisons between a target argument and several reference arguments facilitates a more precise judgment of the target argument’s quality. Inspired by this, we propose a comparison-based framework [...] Read more.
Assessing the quality of arguments is both valuable and challenging. Humans often find that making pairwise comparisons between a target argument and several reference arguments facilitates a more precise judgment of the target argument’s quality. Inspired by this, we propose a comparison-based framework for argument quality assessments (CompAQA), which scores the quality of an argument through multiple pairwise comparisons. Additionally, we introduce an argument order-based data augmentation strategy to enhance CompAQA’s relative quality comparison ability. By introducing multiple reference arguments for pairwise comparisons, CompAQA improves the objectivity and precision of argument quality assessments. Another advantage of CompAQA is its ability to integrate both pairwise argument quality classification and argument quality ranking tasks into a unified framework, distinguishing it from existing methods. We conduct extensive experiments using various pre-trained encoder-only models. Our experiments involve two argument quality ranking datasets (IBM-ArgQ-5.3kArgs and IBM-Rank-30k) and one pairwise argument quality classification dataset (IBM-ArgQ-9.1kPairs). Overall, CompAQA significantly outperforms several strong baselines. Specifically, when using the RoBERTa model as a backbone, CompAQA outperforms the previous best method on the IBM-Rank-30k dataset, improving Pearson correlation by 0.0203 and Spearman correlation by 0.0148. On the IBM-ArgQ-5.3kArgs dataset, it shows improvements of 0.0069 in Pearson correlation and 0.0208 in Spearman correlation. Furthermore, CompAQA demonstrates a 4.71% increase in accuracy over the baseline method on the IBM-ArgQ-9.1kPairs dataset. We also show that CompAQA can be effectively applied to fine-tune larger decoder-only pre-trained models, such as Llama. Full article
(This article belongs to the Special Issue New Advances in Affective Computing)
Show Figures

Figure 1

Figure 1
<p>The architecture of CompAQA. For the sake of simplicity, we omit the topic corresponding to each argument. Here, we assume that the quality of the “Target Argument <math display="inline"><semantics> <msup> <mi>a</mi> <mi>t</mi> </msup> </semantics></math>” is superior to that of the “Reference Argument <math display="inline"><semantics> <msubsup> <mi>a</mi> <mi>i</mi> <mi>r</mi> </msubsup> </semantics></math>”.</p>
Full article ">Figure 2
<p>Prompt for fine-tuning decoder-only language models.</p>
Full article ">
17 pages, 1606 KiB  
Article
Dopaminergic- and Serotonergic-Dependent Behaviors Are Altered by Lanthanide Series Metals in Caenorhabditis elegans
by Anthony Radzimirski, Michael Croft, Nicholas Ireland, Lydia Miller, Jennifer Newell-Caito and Samuel Caito
Toxics 2024, 12(10), 754; https://doi.org/10.3390/toxics12100754 - 17 Oct 2024
Abstract
The lanthanide series elements are transition metals used as critical components of electronics, as well as rechargeable batteries, fertilizers, antimicrobials, contrast agents for medical imaging, and diesel fuel additives. With the surge in their utilization, lanthanide metals are being found more in our [...] Read more.
The lanthanide series elements are transition metals used as critical components of electronics, as well as rechargeable batteries, fertilizers, antimicrobials, contrast agents for medical imaging, and diesel fuel additives. With the surge in their utilization, lanthanide metals are being found more in our environment. However, little is known about the health effects associated with lanthanide exposure. Epidemiological studies as well as studies performed in rodents exposed to lanthanum (La) suggest neurological damage, learning and memory impairment, and disruption of neurotransmitter signaling, particularly in serotonin and dopamine pathways. Unfortunately, little is known about the neurological effects of heavier lanthanides. As dysfunctions of serotonergic and dopaminergic signaling are implicated in multiple neurological conditions, including Parkinson’s disease, depression, generalized anxiety disorder, and post-traumatic stress disorder, it is of utmost importance to determine the effects of La and other lanthanides on these neurotransmitter systems. We therefore hypothesized that early-life exposure of light [La (III) or cerium (Ce (III))] or heavy [erbium (Er (III)) or ytterbium (Yb (III))] lanthanides in Caenorhabditis elegans could cause dysregulation of serotonergic and dopaminergic signaling upon adulthood. Serotonergic signaling was assessed by measuring pharyngeal pump rate, crawl-to-swim transition, as well as egg-laying behaviors. Dopaminergic signaling was assessed by measuring locomotor rate and egg-laying and swim-to-crawl transition behaviors. Treatment with La (III), Ce (III), Er (III), or Yb (III) caused deficits in serotonergic or dopaminergic signaling in all assays, suggesting both the heavy and light lanthanides disrupt these neurotransmitter systems. Concomitant with dysregulation of neurotransmission, all four lanthanides increased reactive oxygen species (ROS) generation and decreased glutathione and ATP levels. This suggests increased oxidative stress, which is a known modifier of neurotransmission. Altogether, our data suggest that both heavy and light lanthanide series elements disrupt serotonergic and dopaminergic signaling and may affect the development or pharmacological management of related neurological conditions. Full article
(This article belongs to the Special Issue Heavy Metal Induced Neurotoxicity)
Show Figures

Figure 1

Figure 1
<p>Lanthanide-induced toxicity in <span class="html-italic">C. elegans</span>. L1-stage N2 worms were treated with increasing concentrations of (<b>top-left</b>) LaCl<sub>3</sub>, (<b>top-right</b>) CeCl<sub>2</sub>, (<b>bottom-left</b>) ErCl<sub>3</sub>, or (<b>bottom-right</b>) YbCl<sub>3</sub> for 30 min, washed, and transferred to agar plate supplemented with OP50 <span class="html-italic">Escherichia coli</span>. Survival was manually counted 48 h post-exposure and fit to a non-linear sigmoidal dose response. LD50 values were calculated and reported. Data points represent the mean + SEM from 5 independent experiments.</p>
Full article ">Figure 2
<p>Lanthanides decrease DAergic function in <span class="html-italic">C. elegans</span>. L1-stage N2 worms were treated with lanthanides for 30 min, washed, and maintained on NGM agar plates seeded with OP50 <span class="html-italic">E.</span> coli. A total of 72 h post-exposure, worms were analyzed for the (<b>A</b>) BSR or (<b>B</b>) swim-to-crawl transition. (<b>C</b>) Alternatively, 30 h after reaching the L4 stage, eggs were quantified on agar or unlaid inside the worm. Data are presented as average (<b>A</b>) change in body bends on vs. off <span class="html-italic">E. coli</span>, (<b>B</b>) time to crawl, or (<b>C</b>) % eggs laid ± SEM of five independent experiments. * <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 as compared to untreated control. Horizontal bars represent comparisons between concentrations of lanthanide-treated worms. # <span class="html-italic">p</span> &lt; 0.05 and ### <span class="html-italic">p</span> &lt; 0.001 as compared between lower and higher concentrations of lanthanides.</p>
Full article ">Figure 3
<p>Lanthanides decrease serotonergic function in <span class="html-italic">C. elegans</span>. L1-stage (<b>A</b>) PE254 or (<b>B</b>) N2 worms were treated with lanthanides for 30 min, washed, and maintained on NGM agar plates seeded with OP50 <span class="html-italic">E. coli</span>. A total of 72 h post-exposure, worms were analyzed for (<b>A</b>) pharynx pump rate or (<b>B</b>) crawl-to-swim transition. Data are presented as average (<b>A</b>) rate of luminescence for 1 h and (<b>B</b>) time to swim ± SEM of five independent experiments. * <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 as compared to untreated control. Horizontal bars represent comparisons between concentrations of lanthanide-treated worms. ## <span class="html-italic">p</span> &lt; 0.01 as compared between lower and higher concentrations of lanthanides.</p>
Full article ">Figure 4
<p>Lanthanides decrease dopamine and serotonin in <span class="html-italic">C. elegans</span>. (<b>A</b>) DA or (<b>B</b>) 5-HT was quantified immediately following exposure to La (III), Ce (III), Er (III), or Yb (III). Data are expressed as means ± SEM of five independent experiments. ** <span class="html-italic">p</span> &lt; 0.01, and *** <span class="html-italic">p</span> &lt; 0.001 as compared to untreated control. Horizontal bars represent comparisons between concentrations of lanthanide-treated worms. ## <span class="html-italic">p</span> &lt; 0.01 and ### <span class="html-italic">p</span> &lt; 0.001 as compared between lower and higher concentrations of lanthanides.</p>
Full article ">Figure 5
<p>Lanthanides induce oxidative stress in <span class="html-italic">C. elegans</span>. Worms were treated for 30 min with La (III), Ce (III), Er (III), or Yb (III). Measures of oxidative stress were either assessed immediately (<b>A</b>) or 24 h following exposure (<b>B</b>–<b>E</b>). (<b>A</b>) ROS levels were assessed by DCFDA fluorescence. Data are expressed as mean fluorescence ± SEM for five independent experiments (<b>B</b>) AGE protein adducts were measured and normalized to protein content. Data are expressed as means ± SEM from 5 independent experiments. (<b>C</b>) Total GSH levels were quantified and normalized to protein content. Data are expressed as mean ± SEM from five independent experiments. (<b>D</b>) Quantification of GFP fluorescence of VP596 transgenic worms. Data are expressed as mean fluorescence ± SEM from 5 independent experiments. (<b>E</b>) ATP levels were quantified and normalized to protein content. Data are expressed as mean ± SEM from five independent experiments. * <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 as compared with untreated control. Horizontal bars represent comparisons between concentrations of lanthanide-treated worms. # <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 as compared between lower and higher concentrations of lanthanides.</p>
Full article ">
23 pages, 17221 KiB  
Article
Aged Gut Microbiome Induces Metabolic Impairment and Hallmarks of Vascular and Intestinal Aging in Young Mice
by Chak-Kwong Cheng, Lianwei Ye, Yuanyuan Zuo, Yaling Wang, Li Wang, Fuyong Li, Sheng Chen and Yu Huang
Antioxidants 2024, 13(10), 1250; https://doi.org/10.3390/antiox13101250 - 17 Oct 2024
Viewed by 112
Abstract
Aging, an independent risk factor for cardiometabolic diseases, refers to a progressive deterioration in physiological function, characterized by 12 established hallmarks. Vascular aging is driven by endothelial dysfunction, telomere dysfunction, oxidative stress, and vascular inflammation. This study investigated whether aged gut microbiome promotes [...] Read more.
Aging, an independent risk factor for cardiometabolic diseases, refers to a progressive deterioration in physiological function, characterized by 12 established hallmarks. Vascular aging is driven by endothelial dysfunction, telomere dysfunction, oxidative stress, and vascular inflammation. This study investigated whether aged gut microbiome promotes vascular aging and metabolic impairment. Fecal microbiome transfer (FMT) was conducted from aged (>75 weeks old) to young C57BL/6 mice (8 weeks old) for 6 weeks. Wire myography was used to evaluate endothelial function in aortas and mesenteric arteries. ROS levels were measured by dihydroethidium (DHE) staining and lucigenin-enhanced chemiluminescence. Vascular and intestinal telomere function, in terms of relative telomere length, telomerase reverse transcriptase expression and telomerase activity, were measured. Systemic inflammation, endotoxemia and intestinal integrity of mice were assessed. Gut microbiome profiles were studied by 16S rRNA sequencing. Some middle-aged mice (40–42 weeks old) were subjected to chronic metformin treatment and exercise training for 4 weeks to evaluate their anti-aging benefits. Six-week FMT impaired glucose homeostasis and caused vascular dysfunction in aortas and mesenteric arteries in young mice. FMT triggered vascular inflammation and oxidative stress, along with declined telomerase activity and shorter telomere length in aortas. Additionally, FMT impaired intestinal integrity, and triggered AMPK inactivation and telomere dysfunction in intestines, potentially attributed to the altered gut microbial profiles. Metformin treatment and moderate exercise improved integrity, AMPK activation and telomere function in mouse intestines. Our data highlight aged microbiome as a mechanism that accelerates intestinal and vascular aging, suggesting the gut-vascular connection as a potential intervention target against cardiovascular aging and complications. Full article
Show Figures

Figure 1

Figure 1
<p>Effects of aged-to-young FMT on body parameters. (<b>A</b>) Schematic overview on FMT protocol from aged and young donor mice to young recipient mice. (<b>B</b>) Body weights of aged donor mice (Aged), young-transplanted (Young (Control)) and aged-transplanted young mice (Young (FMT)) after 6-week FMT protocol. (<b>C</b>) Body weight changes and (<b>D</b>) percentage changes in body weights of mice in (<b>B</b>) during the 6-week FMT. (<b>E</b>) Weights of indicated organs of mice in (<b>B</b>) postmortem after the 6-week FMT. (<b>F</b>) Weights of inguinal subcutaneous adipose tissue (ingSAT), perigonadal visceral adipose tissue (pgVAT) and brown adipose tissue (BAT) of mice in (<b>B</b>). (<b>G</b>) Gross appearance of adipose tissues of mice in (<b>B</b>). (<b>H</b>) Glucose tolerance test (GTT) on mice in (<b>B</b>) at week 6 of FMT, and (<b>I</b>) corresponding area under curve (AUC) analysis of glucose over time. (<b>J</b>) Insulin tolerance test (ITT) of mice in (<b>B</b>) at week 6 of FMT, and (<b>K</b>) corresponding AUC analysis of glucose over time. <span class="html-italic">N</span> = 10 per group. Data are mean ± SD. * <span class="html-italic">p</span> &lt; 0.05; Brown-Forsythe and Welch ANOVA and Dunnett T3 test.</p>
Full article ">Figure 2
<p>Effects of aged-to-young FMT on endothelial function. Representative traces for endothelium-dependent relaxations (EDRs) in (<b>A</b>) aortas and (<b>B</b>) mesenteric arteries of Aged, young-transplanted (Young (Control)) and aged-transplanted mice (Young (FMT)). Summary statistics of wire myography on EDRs in (<b>C</b>) aortas and (<b>D</b>) mesenteric arteries from different mouse groups. (<b>E</b>) Dihydroethidium (DHE) staining on en face endothelium of different mouse groups, and (<b>F</b>) corresponding quantification of DHE fluorescence. (<b>G</b>) Lucigenin-enhanced chemiluminescence on aortic ROS levels of different mouse groups. (<b>H</b>) Nitrite levels in aortas of different mouse groups. <span class="html-italic">N</span> = 8 per group. (<b>I</b>) Representative Western blots, and (<b>J</b>,<b>K</b>) quantification of Western blotting on expression of AMPK, p-AMPK at Thr172, eNOS and p-eNOS at Ser1177 in aortas of different mouse groups. <span class="html-italic">N</span> = 6 per group. Data are mean ± SD. * <span class="html-italic">p</span> &lt; 0.05; Brown-Forsythe and Welch ANOVA and Dunnett T3 test.</p>
Full article ">Figure 3
<p>Effects of aged-to-young FMT on vascular and systemic inflammation, and vascular telomere function. (<b>A</b>) RT-PCR on mRNA levels of pro-inflammatory genes in aortas of Aged, young-transplanted (Young (Control)) and aged-transplanted mice (Young (FMT)). (<b>B</b>) ELISA on circulating inflammatory markers of different mouse groups. (<b>C</b>) ELISA on circulating GLP-1 levels of different mouse groups. (<b>D</b>) Tert mRNA level in aortas of different mouse groups. (<b>E</b>) Telomerase activities in aortas of different mouse groups. (<b>F</b>) Relative telomere length in aortas of different mouse groups. <span class="html-italic">N</span> = 8 per group. Data are mean ± SD. * <span class="html-italic">p</span> &lt; 0.05; Brown-Forsythe and Welch ANOVA and Dunnett T3 test.</p>
Full article ">Figure 4
<p>Effects of aged-to-young FMT on intestinal inflammation, telomere function and barrier function. (<b>A</b>) RT-PCR on mRNA levels of pro-inflammatory genes in intestines of Aged, young-transplanted (Young (Control)) and aged-transplanted mice (Young (FMT)). (<b>B</b>) Lucigenin-enhanced chemiluminescence on intestinal ROS levels of different mouse groups. (<b>C</b>) Tert mRNA level in intestines of different mouse groups. (<b>D</b>) Telomerase activities in intestines of different mouse groups. (<b>E</b>) Relative telomere length in intestines of different mouse groups. Endotoxin levels in (<b>F</b>) feces and (<b>G</b>) sera of different mouse groups. ELISA on serum levels of (<b>H</b>) LBP and (<b>I</b>) I-FABP of different mouse groups. (<b>J</b>) Proglucagon mRNA level in intestines of different mouse groups. <span class="html-italic">N</span> = 8 per group. (<b>K</b>) Representative Western blots and quantification of Western blotting on expression of AMPK and p-AMPK at Thr172 in intestines of different mouse groups. <span class="html-italic">N</span> = 6 per group. Data are mean ± SD. * <span class="html-italic">p</span> &lt; 0.05; Brown-Forsythe and Welch ANOVA and Dunnett T3 test.</p>
Full article ">Figure 5
<p>Effects of aged-to-young FMT on gut microbial profiles in young host mice. (<b>A</b>) Principal component analysis (PCA) plot revealing distinct clusters for fecal microbiome samples obtained from young (depicted in blue) and aged (in brown) mice before antibiotic treatment and FMT, highlighting the species contributing to this clustering. <span class="html-italic">N</span> = 8 per group. (<b>B</b>) PCA plot showing the clustering of fecal microbiome samples from young-transplanted (Young (Control); depicted in blue) and aged-transplanted mice (Young (FMT); in red). <span class="html-italic">N</span> = 8 per group. (<b>C</b>) Non-Metric Multi-Dimensional Scaling (NMDS) plot displaying the clustering of microbiome across various mouse groups. <span class="html-italic">N</span> = 6–8 per group. (<b>D</b>) Differential abundance analysis on the mean difference in centered log ratio for enriched species in young, aged, Young (Control) and Young (FMT) mice. <span class="html-italic">N</span> = 8 per group.</p>
Full article ">Figure 6
<p>Effects of chronic metformin treatment and moderate exercise training on intestinal homeostasis. (<b>A</b>) Schematic diagram on chronic metformin treatment and moderate exercise training with the presence and absence of compound C (CC) treatment in middle-aged C57BL/6 mice. Representative Western blots and quantification of Western blotting on expression of AMPK and p-AMPK at Thr172 in intestines of (<b>B</b>) metformin-treated mice, and (<b>C</b>) exercise-trained mice. Lucigenin-enhanced chemiluminescence on intestinal ROS levels of (<b>D</b>) metformin-treated mice, and (<b>E</b>) exercise-trained mice. RT-PCR on mRNA levels of pro-inflammatory genes in intestines of (<b>F</b>) metformin-treated mice, and (<b>G</b>) exercise-trained mice. (<b>H</b>) Tert mRNA level in intestines of different mouse groups. (<b>I</b>) Telomerase activities in intestines of different mouse groups. (<b>J</b>) Relative telomere length in intestines of different mouse groups. ELISA on serum levels of (<b>K</b>) LBP and (<b>L</b>) I-FABP of different mouse groups. <span class="html-italic">N</span> = 6 per group. Data are mean ± SD. * <span class="html-italic">p</span> &lt; 0.05; Brown-Forsythe and Welch ANOVA and Dunnett T3 test.</p>
Full article ">Figure 7
<p>Schematic overview of the study. Aged microbiome induces metabolic impairments and vascular dysfunction in young mice. Aged microbiome causes telomere dysfunction, oxidative stress, and inflammation in intestines and vasculature of young mice. Metformin and moderate exercise potentially retard hallmarks of intestinal aging through AMPK activation. The study highlights the network among multiple aging hallmarks, including dysbiosis, deregulated nutrient sensing, chronic inflammation and telomere attrition, in terms of gut-vascular connection.</p>
Full article ">
18 pages, 2875 KiB  
Article
Formulation and Evaluation of Radiance Serum Containing Astaxanthin–Zeaxanthin Nanoemulsions as an Anti-Wrinkle Agent: Stability, Ex Vivo, and In Vivo Assessments
by Lusi Nurdianti, Fajar Setiawan, Ira Maya, Taofik Rusdiana, Cahya Khairani Kusumawulan, Dolih Gozali and Dewi Peti Virgianti
Cosmetics 2024, 11(5), 182; https://doi.org/10.3390/cosmetics11050182 - 17 Oct 2024
Viewed by 109
Abstract
Reactive oxygen species (ROS), commonly known as free radicals, induced by UV radiation can compromise the dermal structure, leading to a loss of skin elasticity and subsequent wrinkle formation. A promising strategy to prevent and mitigate skin aging involves the use of topical [...] Read more.
Reactive oxygen species (ROS), commonly known as free radicals, induced by UV radiation can compromise the dermal structure, leading to a loss of skin elasticity and subsequent wrinkle formation. A promising strategy to prevent and mitigate skin aging involves the use of topical formulations with potent antioxidant properties. Secondary metabolites such as astaxanthin and zeaxanthin are known for their robust antioxidant activities, which surpass those of tocopherol, offering significant benefits for skin health and protection against UV-induced damage. These properties suggest their potential application in anti-aging products. This study aims to evaluate the stability, ex vivo penetration, and in vivo efficacy of a radiance serum containing an astaxanthin–zeaxanthin nanoemulsion (AZ-NE) designed as an anti-wrinkle agent for topical application. The research was conducted in four stages: production of the astaxanthin–zeaxanthin nanoemulsion (AZ-NE), formulation of the AZ-NE radiance serum, stability, and efficacy testing. In this study, the formulated radiance serum demonstrated stability over three months under specified storage conditions. Ex vivo penetration studies indicated efficient diffusion of the active ingredients, with astaxanthin showing a penetration rate of 25.95%/cm2 and zeaxanthin at 20.80%/cm2 after 120 min. In vivo irritation tests conducted on human subjects revealed no adverse effects. Moreover, the serum exhibited substantial anti-wrinkle efficacy, with 15 female participants experiencing a wrinkle reduction of 80% to 93% over a 28-day period. Full article
(This article belongs to the Special Issue Bioactive Compounds From Natural Resources Against Skin Aging)
Show Figures

Figure 1

Figure 1
<p>Astaxanthin–zeaxanthin nanoemulsion.</p>
Full article ">Figure 2
<p>AZ-NE radiance serum.</p>
Full article ">Figure 3
<p>Results of the accelerated stability study on viscosity parameters.</p>
Full article ">Figure 4
<p>Results of the accelerated stability study on pH parameters.</p>
Full article ">Figure 5
<p>The amount of blank control that penetrated over a duration of 120 min is presented as a percentage per cm<sup>2</sup>.</p>
Full article ">Figure 6
<p>The amount of free astaxanthin in the radiance serum that penetrated over a duration of 120 min is presented as a percentage per cm<sup>2</sup>.</p>
Full article ">Figure 7
<p>The amount of free zeaxanthin in the radiance serum that penetrated over a duration of 120 min is presented as a percentage per cm<sup>2</sup>.</p>
Full article ">Figure 8
<p>The amount of astaxanthin nanoemulsions (A-NE) in the AZ-NE radiance serum that penetrated over a duration of 120 min is presented as a percentage per cm<sup>2</sup>.</p>
Full article ">Figure 9
<p>The amount of zeaxanthin nanoemulsions (Z-NE) in the AZ-NE radiance serum that penetrated over a duration of 120 min is presented as a percentage per cm<sup>2</sup>.</p>
Full article ">
20 pages, 15952 KiB  
Article
Immp2l Deficiency Induced Granulosa Cell Senescence Through STAT1/ATF4 Mediated UPRmt and STAT1/(ATF4)/HIF1α/BNIP3 Mediated Mitophagy: Prevented by Enocyanin
by Xiaoya Qu, Pengge Pan, Sinan Cao, Yan Ma, Jinyi Yang, Hui Gao, Xiuying Pei and Yanzhou Yang
Int. J. Mol. Sci. 2024, 25(20), 11122; https://doi.org/10.3390/ijms252011122 - 16 Oct 2024
Viewed by 243
Abstract
Dysfunctional mitochondria producing excessive ROS are the main factors that cause ovarian aging. Immp2l deficiency causes mitochondrial dysfunction and excessive ROS production, leading to ovarian aging, which is attributed to granulosa cell senescence. The pathway controlling mitochondrial proteostasis and mitochondrial homeostasis of the [...] Read more.
Dysfunctional mitochondria producing excessive ROS are the main factors that cause ovarian aging. Immp2l deficiency causes mitochondrial dysfunction and excessive ROS production, leading to ovarian aging, which is attributed to granulosa cell senescence. The pathway controlling mitochondrial proteostasis and mitochondrial homeostasis of the UPRmt and mitophagy are closely related with the ROS and cell senescence. Our results suggest that Immp2l knockout led to granulosa cell senescence, and enocyanin treatment alleviated Immp2l deficiency-induced granulosa cell senescence, which was accompanied by improvements in mitochondrial function and reduced ROS levels. Interestingly, redox-related protein modifications, including S-glutathionylation and S-nitrosylation, were markedly increased in Immp2l-knockout granulosa cells, and were markedly reduced by enocyanin treatment. Furthermore, STAT1 was significantly increased in Immp2l-knockout granulosa cells and reduced by enocyanin treatment. The co-IP results suggest that the expression of STAT1 was controlled by S-glutathionylation and S-nitrosylation, but not phosphorylation. The UPRmt was impaired in Immp2l-deficient granulosa cells, and unfolded and misfolded proteins aggregated in mitochondria. Then, the HIF1α/BNIP3-mediated mitophagy pathway was activated, but mitophagy was impaired due to the reduced fusion of mitophagosomes and lysosomes. The excessive aggregation of mitochondria increased ROS production, leading to senescence. Hence, Enocyanin treatment alleviated granulosa cell senescence through STAT1/ATF4-mediated UPRmt and STAT1/(ATF4)/HIF1α/BNIP3-mediated mitophagy. Full article
(This article belongs to the Section Biochemistry)
Show Figures

Figure 1

Figure 1
<p>Enocyanin delays <span class="html-italic">Immp2l</span> knockdown-induced primary granulosa cell senescence. (<b>A</b>) Western blot analysis of cell senescence marker molecules in <span class="html-italic">Immp2l</span>-knockdown granulosa cells treated with enocyanin. (<b>B</b>) Staining of senescence-associated β-galactosidase in <span class="html-italic">Immp2l</span>-knockdown granulosa cells treated with enocyanin, at least 50 different cells in different fields were analyzed, and the dots were blindly counted by three different individuals. (<b>C</b>) Western blot analysis of cell STAT1 and UPR<sup>mt</sup> molecules in <span class="html-italic">Immp2l</span> knockdown granulosa cells treated with enocyanin. (<b>D</b>) Detection of the colocalization of UPR<sup>mt</sup> molecules ATF4, ATF5, CHOP and Mito-tracker. (<b>E</b>) Detection of protein aggresomes with the PROTEOSTAT Aggresome Detection Kit in <span class="html-italic">Immp2l</span>-knockdown granulosa cells treated with enocyanin. (<b>F</b>) Western blot analysis of mitophagy molecules and lysosome membrane protein Lamp2 in <span class="html-italic">Immp2l</span>-knockdown granulosa cells treated with enocyanin. (<b>G</b>) Detection of autophagic flux with the mCherry-GFP-LC3 lentivirus in <span class="html-italic">Immp2l</span>-knockdown granulosa cells treated with enocyanin; at least 50 different cells in different fields were analyzed, and the dots were blindly counted by three different individuals. (<b>H</b>) Detection of mitROS and mitochondrial membrane potential in <span class="html-italic">Immp2l</span>-knockdown granulosa cells treated with enocyanin. (<b>I</b>) Western blot analysis of S-glutathionylation and S-nitrosylation in <span class="html-italic">Immp2l</span>-knockdown granulosa cells treated with enocyanin. (* <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), (Scale bar: 10 μm).</p>
Full article ">Figure 1 Cont.
<p>Enocyanin delays <span class="html-italic">Immp2l</span> knockdown-induced primary granulosa cell senescence. (<b>A</b>) Western blot analysis of cell senescence marker molecules in <span class="html-italic">Immp2l</span>-knockdown granulosa cells treated with enocyanin. (<b>B</b>) Staining of senescence-associated β-galactosidase in <span class="html-italic">Immp2l</span>-knockdown granulosa cells treated with enocyanin, at least 50 different cells in different fields were analyzed, and the dots were blindly counted by three different individuals. (<b>C</b>) Western blot analysis of cell STAT1 and UPR<sup>mt</sup> molecules in <span class="html-italic">Immp2l</span> knockdown granulosa cells treated with enocyanin. (<b>D</b>) Detection of the colocalization of UPR<sup>mt</sup> molecules ATF4, ATF5, CHOP and Mito-tracker. (<b>E</b>) Detection of protein aggresomes with the PROTEOSTAT Aggresome Detection Kit in <span class="html-italic">Immp2l</span>-knockdown granulosa cells treated with enocyanin. (<b>F</b>) Western blot analysis of mitophagy molecules and lysosome membrane protein Lamp2 in <span class="html-italic">Immp2l</span>-knockdown granulosa cells treated with enocyanin. (<b>G</b>) Detection of autophagic flux with the mCherry-GFP-LC3 lentivirus in <span class="html-italic">Immp2l</span>-knockdown granulosa cells treated with enocyanin; at least 50 different cells in different fields were analyzed, and the dots were blindly counted by three different individuals. (<b>H</b>) Detection of mitROS and mitochondrial membrane potential in <span class="html-italic">Immp2l</span>-knockdown granulosa cells treated with enocyanin. (<b>I</b>) Western blot analysis of S-glutathionylation and S-nitrosylation in <span class="html-italic">Immp2l</span>-knockdown granulosa cells treated with enocyanin. (* <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), (Scale bar: 10 μm).</p>
Full article ">Figure 2
<p>Enocyanin delays <span class="html-italic">Immp2l</span> deficiency-induced granulosa cell senescence. (<b>A</b>) Western blot analysis of cell senescence marker molecules in <span class="html-italic">Immp2l</span>-deficient granulosa cells treated with enocyanin. (<b>B</b>) Staining of senescence-associated β-galactosidase in <span class="html-italic">Immp2l</span>-deficient granulosa cells treated with enocyanin; at least 50 different cells in different fields were analyzed, and the dots were blindly counted by three different individuals. (<b>C</b>) Detection of mitROS and mitochondrial membrane potential in <span class="html-italic">Immp2l</span>-deficient granulosa cells treated with enocyanin. (<b>D</b>) Western blot analysis of S-glutathionylation and S-nitrosylation in Immp2l-deficient granulosa cells treated with enocyanin. (* <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), (Scale bar: 10 μm).</p>
Full article ">Figure 3
<p>The translational modification of GST and SNO but not phosphorylation regulates STAT1. (<b>A</b>) Western blot analysis of STAT1 in <span class="html-italic">Immp2l</span>-deficient granulosa cells treated with enocyanin. (<b>B</b>) Colocalization of STAT1 and Mito-tracker in <span class="html-italic">Immp2l</span>-deficient granulosa cells treated with enocyanin. (<b>C</b>) Western blot analysis of phosphorylated STAT1 in <span class="html-italic">Immp2l</span>-deficient granulosa cells. (<b>D</b>) Co-IP confirmation of the interaction of STAT1 and S-glutathionylation and S-nitrosylation; at least 50 different cells in different fields were analyzed, and the dots were blindly counted by three different individuals. (* <span class="html-italic">p</span> &lt; 0.05), (Scale bar: 10 μm).</p>
Full article ">Figure 4
<p>UPR<sup>mt</sup> pathway proteins were altered in <span class="html-italic">Immp2l</span>-deficient granulosa cells and regulated by enocyanin. (<b>A</b>) Western blot analysis of UPR<sup>mt</sup> marker molecules in <span class="html-italic">Immp2l</span>-deficient granulosa cells treated with enocyanin. (<b>B</b>) Colocalization of the UPR<sup>mt</sup> molecules ATF4, ATF5, CHOP and Mito-tracker in <span class="html-italic">Immp2l</span>-deficient granulosa cells treated with enocyanin. (<b>C</b>) Detection of protein aggresomes with the PROTEOSTAT Aggresome Detection Kit in <span class="html-italic">Immp2l</span>-deficient granulosa cells treated with enocyanin. (* <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) (scale bar: 10 μm).</p>
Full article ">Figure 5
<p>Mitophagy pathway proteins are altered in <span class="html-italic">Immp2l</span>-deficient granulosa cells and regulated by enocyanin. (<b>A</b>) Western blot analysis of mitophagy marker molecules in <span class="html-italic">Immp2l</span>-deficient granulosa cells treated with enocyanin. (<b>B</b>) Western blot analysis of the lysosome marker molecules Lamp1, Lamp2 and Galectin3 in <span class="html-italic">Immp2l</span>-deficient granulosa cells treated with enocyanin. (<b>C</b>) Detection of autophagic flux with the mCherry-GFP-LC3 lentivirus in <span class="html-italic">Immp2l</span>-deficient granulosa cells treated with enocyanin; at least 50 different cells in different fields were analyzed, and the dots were blindly counted by three different individuals. (* <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) (scale bar: 10 μm).</p>
Full article ">Figure 6
<p>The UPR<sup>mt</sup> and senescence proteins regulated by the STAT1 inhibitor fludarabine (FLU). (<b>A</b>) Western blot analysis of cell senescence marker molecules in <span class="html-italic">Immp2l</span>-deficient granulosa cells treated with the STAT1 inhibitor fludarabine (FLU). (<b>B</b>) Staining of senescence-associated β-galactosidase in <span class="html-italic">Immp2l</span>-deficient granulosa cells treated with the STAT1 inhibitor fludarabine (FLU); at least 50 different cells in different fields were analyzed, and the dots were blindly counted by three different individuals. (<b>C</b>) Detection of mitROS and mitochondrial membrane potential in <span class="html-italic">Immp2l</span>-deficient granulosa cells treated with the STAT1 inhibitor fludarabine (FLU); at least 50 different cells in different fields were analyzed, and the dots were blindly counted by three different individuals. (<b>D</b>) Western blot analysis of UPR<sup>mt</sup> marker molecules in <span class="html-italic">Immp2l</span>-deficient granulosa cells treated with the STAT1 inhibitor fludarabine (FLU). (<b>E</b>) Colocalization of the UPR<sup>mt</sup> marker molecules ATF4, ATF5, CHOP and Mito-tracker. (<b>F</b>) Detection of protein aggresome with the PROTEOSTAT Aggresome Detection Kit; at least 50 different cells in different fields were analyzed, and the dots were blindly counted by three different individuals. (<b>G</b>) Co-IP detection and analysis of the interaction of UPR<sup>mt</sup> ATF4 with STAT1, ATF4 and HIF1α, HIF1α and BNIP3, STAT1 and HIF1α. (* <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), (scale bar: 10 μm).</p>
Full article ">Figure 7
<p>Mitophagy and lysosomal proteins regulated by the STAT1 inhibitor fludarabine (FLU). (<b>A</b>) Western blot analysis of mitophagy and lysosome molecule in <span class="html-italic">Immp2l</span>-deficient granulosa cells treated by STAT1 inhibitor fludarabine (FLU). (<b>B</b>) Detection of autophagic flux with the mCherry-GFP-LC3 lentivirus in <span class="html-italic">Immp2l</span>-deficient granulosa cells treated by STAT1 inhibitor fludarabine (FLU); at least 50 different cells in different fields were analyzed, and the dots were blindly counted by three different individuals. (<b>C</b>) Western blot analysis of UPR<sup>mt</sup> and mitophagy marker molecules in <span class="html-italic">Immp2l</span>-deficient granulosa cells treated by HIF1α inhibitor 2-Methoxyestradiol (2-MeOE2). (<b>D</b>) Detection of the colocalization of UPR<sup>mt</sup> molecules ATF4, ATF5, CHOP and Mito-tracker. (<b>E</b>) Detection of protein aggresome with PROTEOSTAT Aggresome Detection Kit in <span class="html-italic">Immp2l</span>-deficient granulosa cells treated by HIF1α inhibitor 2-Methoxyestradiol (2-MeOE2); at least 50 different cells in different fields were analyzed, and the dots were blindly counted by three different individuals. (* <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), (Scale bar: 10 μm).</p>
Full article ">Figure 8
<p>UPR<sup>mt</sup> regulated by mitophagy inhibitor cyclosporin A (CsA) in <span class="html-italic">Immp2l</span>-deficient granulosa cells. (<b>A</b>) Western blot analysis of UPR<sup>mt</sup> marker molecules in <span class="html-italic">Immp2l</span>-deficient granulosa cells treated by mitophagy inhibitor cyclosporin A (CsA). (<b>B</b>) Detection of the colocalization of UPR<sup>mt</sup> molecules ATF4, ATF5, CHOP and Mito-tracker. (<b>C</b>) Detection of protein aggresome with PROTEOSTAT Aggresome Detection Kit in <span class="html-italic">Immp2l</span>-deficient granulosa cells treated by mitophagy inhibitor cyclosporin A (CsA); at least 50 different cells in different fields were analyzed, and the dots were blindly counted by three different individuals. (* <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), (scale bar: 10 μm).</p>
Full article ">Figure 9
<p>The pathway by which enocyanin alleviates <span class="html-italic">Immp2l</span> deficiency-induced granulosa cell senescence. Under normal physiological conditions, mitochondrial proteostasis and mitochondrial homeostasis in granulosa cells are maintained by the UPR<sup>mt</sup> and mitophagy. However, <span class="html-italic">Immp2l</span> deficiency in granulosa cells impairs the UPR<sup>mt</sup> and mitophagy, leading to excessive unfolded and misfolded protein aggregation in mitochondria, dysfunctional mitochondria aggregation in cells, and excessive aggregated dysfunctional mitochondria producing excessive ROS, thereby triggering granulosa cell senescence. Enocyanin treatment alleviates granulosa cell senescence and improves the function of UPR<sup>mt</sup> and mitophagy through STAT1/ATF4-mediated UPR<sup>mt</sup> and STAT1/(ATF4)/HIF1α/BNIP3-mediated mitophagy, restoring mitochondrial proteostasis and mitochondrial homeostasis in granulosa cells and reducing ROS production.</p>
Full article ">
22 pages, 3701 KiB  
Article
Physiological and Transcriptomic Analyses Reveal the Role of the Antioxidant System and Jasmonic Acid (JA) Signal Transduction in Mulberry (Morus alba L.) Response to Flooding Stress
by Xuejiao Bai, He Huang, Dan Li, Fei Yang, Xinyao Cong, Siqi Wu, Wenxu Zhu, Shengjin Qin and Yibo Wen
Horticulturae 2024, 10(10), 1100; https://doi.org/10.3390/horticulturae10101100 - 16 Oct 2024
Viewed by 355
Abstract
In recent decades, the frequency of flooding has increased as a result of global climate change. Flooding has become one of the major abiotic stresses that seriously affect the growth and development of plants. Mulberry (Morus alba L.) is an important economic [...] Read more.
In recent decades, the frequency of flooding has increased as a result of global climate change. Flooding has become one of the major abiotic stresses that seriously affect the growth and development of plants. Mulberry (Morus alba L.) is an important economic tree in China. Flooding stress is among the most severe abiotic stresses that affect the production of mulberry. However, the physiological and molecular biological mechanisms of mulberry responses to flooding stress are still unclear. In the present study, reactive oxygen species (ROS) metabolism, antioxidant mechanism, and plant hormones in mulberry associated with the response to flooding stress were investigated using physiological and transcriptomic analysis methods. The results showed significant increases in the production rate of superoxide anion (O2•−) and the content of hydrogen peroxide (H2O2) in leaves on the 5th day of flooding stress. This led to membrane lipid peroxidation and elevated malondialdehyde (MDA) levels. Antioxidant enzymes such as catalase (CAT), superoxide dismutase (SOD), and peroxidase (POD) exhibited enhanced activities initially, followed by fluctuations. The ascorbic acid–glutathione (AsA-GSH) cycle played a crucial role in scavenging ROS, promoting the reduction of oxidized glutathione (GSSG) to reduced glutathione (GSH). Transcriptomic analysis revealed the up-regulation of the gene-encoding antioxidant enzymes (APX, MDHAR, GPX, GR, GST) involved in ROS scavenging and stress tolerance mechanisms. Jasmonic acid (JA) levels and the expression of JA synthesis-related genes increased significantly in mulberry leaves under flooding stress. This activation of the JA signaling pathway contributed to the plant’s adaptability to flooding conditions. Proline (Pro) and soluble sugar (SS) contents increased notably in response to flooding stress. Proline helped maintain cell turgor and protected enzymes and membranes from damage, while soluble sugars supported anaerobic respiration and energy supply. However, soluble protein (SP) content decreased, suggesting inhibition of protein synthesis. The study provides insights into mulberry’s flooding tolerance mechanisms, guiding future molecular breeding efforts. This summary captures the key findings and implications of the study on mulberry’s response to flooding stress, focusing on physiological and molecular mechanisms identified in the research. Full article
Show Figures

Figure 1

Figure 1
<p>Production rate of O<sub>2</sub><sup>•−</sup> (<b>A</b>), H<sub>2</sub>O<sub>2</sub> (<b>B</b>) and MDA content (<b>C</b>), SOD activity (<b>D</b>), POD activity (<b>E</b>), CAT activity (<b>F</b>), and heatmaps of genes expression of ROS scavenging enzymes (<b>G</b>) in mulberry (<span class="html-italic">Morus alba</span> L.) leaves under flooding stress. Note: Heat map data are derived from the gene expression data of transcriptome analysis results, which are drawn according to the normalized gene expression amount under Ctl conditions. Under different waterlogging conditions, the gene expression amount higher than the average value under Ctl conditions is marked with red, and vice versa, and the gene expression amount lower than the average value is marked with blue. The color shading indicates the degree of difference between gene expression and Ctl. The darker the color, the more significant the difference. Data are represented as means ± SD of five replicate samples. Different lowercase letters indicating significant differences (<span class="html-italic">p</span> &lt; 0.05).</p>
Full article ">Figure 2
<p>Activity key enzymes, content of metabolic substance, and heatmaps of gene expression in ASA-GSH cycle in mulberry (<span class="html-italic">Morus alba</span> L.) leaves under flooding stress. The content of ascorbic acid (<b>A</b>), the content of dehydroascorbate (<b>B</b>), the ratio of the content of ascorbate to dehydroascorbate (<b>C</b>), the content of glutathione (<b>D</b>), the content of glutathiol (<b>E</b>), the ratio of the content of glutathione to glutathiol (<b>F</b>), the activity of ascorbate peroxidase (<b>G</b>), the activity of monodehydroascorbate reductase (<b>H</b>), the activity of dehydroascorbate reductase (<b>I</b>), the activity of glutathione peroxidase (<b>J</b>), the activity of glutathione reductase (<b>K</b>), the activity of glutathione S-transferase (<b>L</b>), heatmaps of the expression levels of relevant genes in the AsA-GSH cycle (<b>M</b>). APX: ascorbate peroxidase; MDHAR: monodehydroascorbate reductase; GPX: glutathione peroxidase; GR: gluathione reductase; DHAR: dehydroascorbate reductase; AsA: ascorbate; DHA: dehydroascorbate; GSH: glutathione; MDHA: monodehydroascorbate; GSSG: glutathiol. Note: Heat map data processing method is the same as above. Data are represented as means ± SD of five replicate samples. Different lowercase letters indicating significant differences (<span class="html-italic">p</span> &lt; 0.05).</p>
Full article ">Figure 3
<p>Activity key enzymes and heatmaps of genes expression in Trx-Prx pathway in mulberry (<span class="html-italic">Morus alba</span> L.) leaves under flooding stress. The activity of thioredoxin reductase (<b>A</b>), the activity of peroxiredoxin (<b>B</b>), heatmaps of the expression levels of related genes in the Trx-Prx pathway (<b>C</b>). Note: Data are represented as means ± SD of five replicate samples. Different lowercase letters indicating significant differences (<span class="html-italic">p</span> &lt; 0.05).</p>
Full article ">Figure 4
<p>JA content and heatmaps of gene expression in JA synthesis and JA signal in mulberry (<span class="html-italic">Morus alba</span> L.) leaves under flooding stress. The content of JA (<b>A</b>), heatmaps of the expression levels of related genes in JA synthesis and JA signal (<b>B</b>). Note: Data are represented as means ± SD of five replicate samples. Different lowercase letters indicating significant differences (<span class="html-italic">p</span> &lt; 0.05).</p>
Full article ">Figure 5
<p>Proline content (<b>A</b>), soluble sugar content (<b>B</b>), and soluble protein content (<b>C</b>) in mulberry (<span class="html-italic">Morus alba</span> L.) leaves under flooding stress. Note: Data are represented as means ± SD of five replicate samples. Different lowercase letters indicating significant differences (<span class="html-italic">p</span> &lt; 0.05).</p>
Full article ">Figure 6
<p>RT-qPCR verified transcript expression levels of DEGs in mulberry (<span class="html-italic">Morus alba</span> L.) leaves under flooding stress. Data are represented as means ± SD of five replicate samples. Different lowercase letters indicating significant differences (<span class="html-italic">p</span> &lt; 0.05).</p>
Full article ">
10 pages, 1197 KiB  
Article
Mitochondrial Creatine Kinase 2 (Ckmt2) as a Plasma-Based Biomarker for Evaluating Reperfusion Injury in Acute Myocardial Infarction
by Alexander Lang, Daniel Oehler, Marcel Benkhoff, Yvonne Reinders, Maike Barcik, Khatereh Shahrjerdi, Madlen Kaldirim, Albert Sickmann, Lisa Dannenberg, Amin Polzin, Susanne Pfeiler, Malte Kelm, Maria Grandoch, Christian Jung and Norbert Gerdes
Biomedicines 2024, 12(10), 2368; https://doi.org/10.3390/biomedicines12102368 - 16 Oct 2024
Viewed by 248
Abstract
Background/Objectives: Acute myocardial infarction (AMI), characterized by irreversible heart muscle damage and impaired cardiac function caused by myocardial ischemia, is a leading cause of global mortality. The damage associated with reperfusion, particularly mitochondrial dysfunction and reactive oxygen species (ROS) formation, has emerged as [...] Read more.
Background/Objectives: Acute myocardial infarction (AMI), characterized by irreversible heart muscle damage and impaired cardiac function caused by myocardial ischemia, is a leading cause of global mortality. The damage associated with reperfusion, particularly mitochondrial dysfunction and reactive oxygen species (ROS) formation, has emerged as a crucial factor in the pathogenesis of cardiac diseases, leading to the recognition of mitochondrial proteins as potential markers for myocardial damage. This study aimed to identify differentially expressed proteins based on the type of cardiac injury, in particular those with and without reperfusion. Methods: Male C57Bl/6J mice were either left untreated, sham-operated, received non-reperfused AMI, or reperfused AMI. Twenty-four hours after the procedures, left ventricular (LV) function and morphological changes including infarct size were determined using echocardiography and triphenyl tetrazolium chloride (TTC) staining, respectively. In addition, plasma was isolated and subjected to untargeted mass spectrometry and, further on, the ELISA-based validation of candidate proteins. Results: We identified mitochondrial creatine kinase 2 (Ckmt2) as a differentially regulated protein in plasma of mice with reperfused but not non-reperfused AMI. Elevated levels of Ckmt2 were significantly associated with infarct size and impaired LV function following reperfused AMI, suggesting a specific involvement in reperfusion damage. Conclusions: Our study highlights the potential of plasma Ckmt2 as a biomarker for assessing reperfusion injury and its impact on cardiac function and morphology in the acute phase of MI. Full article
(This article belongs to the Section Molecular and Translational Medicine)
Show Figures

Figure 1

Figure 1
<p>Protein abundance in the plasma of mice subjected to experimental AMI with and without reperfusion. A volcano plot illustrating the relative abundance (log2 fold changes) versus probability (−log10 <span class="html-italic">p</span>-values) in the plasma of mice that received either non-reperfused or reperfused AMI. Multiple <span class="html-italic">t</span>-test analyses with False Discovery Rate (FDR) correction were used to determine significance (indicated by red dots). Significant proteins are defined as those with a fold change &gt; | ±1.5| and a <span class="html-italic">p</span>-value &lt; 0.05. Two proteins are significantly upregulated in the non-reperfusion group (Mup3, Mup18), while six proteins are significantly upregulated in the reperfusion-operated mice (Ckmt2, Grn, Got1, Igfbp4, Got1, Actn3, Fermt3). <span class="html-italic">n</span> = 5 samples per group.</p>
Full article ">Figure 2
<p>The mass spectrometric quantification of eight proteins significantly altered in plasma between non-reperfusion- and reperfusion-operated mice and their abundance in the respective controls. Plasma was isolated 24 h after the induction of sham-operated, non-reperfused AMI, and reperfused AMI and in untreated animals and was analyzed by untargeted mass spectrometry. Each bar plot represents the mean value with standard deviation (SD), with individual data points overlaid to show the distribution of measurements. Statistical significance between groups was determined using paired one-way ANOVA for normally distributed data or Kruskal–Wallis test for non-normally distributed data, with a <span class="html-italic">p</span>-value &lt; 0.05 considered significant. Figure panels correspond to specific proteins: (<b>A</b>) Ckmt2, (<b>B</b>) Actn3, (<b>C</b>) Got1, (<b>D</b>) Igfbp4, (<b>E</b>) Grn, (<b>F</b>) Fermt3, (<b>G</b>) Mup3 and (<b>H</b>) Mup18. (<b>A</b>–<b>F</b>) show proteins that are upregulated in reperfusion-operated mice compared to non-reperfusion-operated mice, while (<b>G</b>,<b>H</b>) show proteins that are downregulated. Protein levels that were not detected are valued as 0. <span class="html-italic">n</span> = 5 per group.</p>
Full article ">Figure 3
<p>Ckmt2 correlates to infarct size and left ventricular dysfunction 24 h following reperfused myocardial infarction. (<b>A</b>) Ckmt2 plasma levels detected by ELISA show a significant positive correlation with infarct size (INF) relative to the area at risk (AAR), a negative relationship with (<b>B</b>) ejection fraction (EF) and a positive relationship with (<b>C</b>) end-systolic volume (ESV). (<b>D</b>) No significant correlation is observed between Ckmt2 levels and end-diastolic volume (EDV), (<b>E</b>) stroke volume (SV) or (<b>F</b>) heart rate (HR). Correlations are performed using Pearson’s correlation. Each dot represents an individual measurement from one mouse. Significance is defined as <span class="html-italic">p</span> &lt; 0.05, and correlations are displayed with R<sup>2</sup> values. <span class="html-italic">n</span> = 15.</p>
Full article ">
16 pages, 1106 KiB  
Article
Cryoprotective Potential of Theobromine in the Improvement of the Post-Thaw Quality of Bovine Spermatozoa
by Filip Benko, Štefan Baňas, Michal Ďuračka, Miroslava Kačániová and Eva Tvrdá
Cells 2024, 13(20), 1710; https://doi.org/10.3390/cells13201710 - 16 Oct 2024
Viewed by 183
Abstract
Theobromine (TBR) is a methylxanthine known for its bronchodilatory and stimulatory effects. This research evaluated the vitality, capacitation patterns, oxidative characteristics, microbial profile and expression of capacitation-associated proteins (CatSper1/2, sodium bicarbonate cotransporter [NBC], protein kinases A [PKA] and C [PKC] and adenylate cyclase [...] Read more.
Theobromine (TBR) is a methylxanthine known for its bronchodilatory and stimulatory effects. This research evaluated the vitality, capacitation patterns, oxidative characteristics, microbial profile and expression of capacitation-associated proteins (CatSper1/2, sodium bicarbonate cotransporter [NBC], protein kinases A [PKA] and C [PKC] and adenylate cyclase 10 [ADCY10]) in cryopreserved bovine spermatozoa (n = 30) in the absence (cryopreserved control [CtrlC]) or presence of different TBR concentrations (12.5, 25, and 50 µM) in egg yolk extender. Fresh ejaculate served as a negative control (CtrlN). Significant post-thaw maintenance of the sperm motility, membrane and DNA integrity and mitochondrial activity (p < 0.001) were recorded following the administration of 25 μM and 50 μM TBR, then compared to CtrlC. All groups supplemented with TBR exhibited a significantly lower percentage of prematurely capacitated spermatozoa (p < 0.001) than CtrlC. Significantly decreased levels of global reactive oxygen species (ROS), hydrogen peroxide and hydroxyl radicals were observed in the presence of 25 μM and 50 μM TBR (p < 0.01). Western blot analysis revealed that supplementation with 50 μM TBR significantly prevented the loss of NBC and ADCY10 (p < 0.01), while all TBR doses stabilized the levels of PKC (p < 0.05 at 50 μM TBR; p < 0.001 at 12.5 μM and 25 μM TBR). In summary, we suggest that TBR is effective in protecting the spermatozoa during the cryopreservation process through its potential to stimulate energy synthesis while preventing ROS overproduction and the loss of proteins involved in the sperm activation process. Full article
(This article belongs to the Section Reproductive Cells and Development)
Show Figures

Figure 1

Figure 1
<p>Protein levels of the cation channels of sperm isoforms 1 and 2 (CatSper1 and CatSper2), sodium bicarbonate cotransporter (NBC), protein kinase A (PKA), protein kinase C (PKC) and adenylyl cyclase 10 (ADCY10) in bovine spermatozoa in fresh state and cryopreserved in the absence or presence of selected theobromine (TBR) doses, assessed by Western blotting. Original photos of the gels and blots are available as <a href="#app1-cells-13-01710" class="html-app">Supplementary Materials</a>. Created with BioRender.com (accessed on 27 August 2024).</p>
Full article ">Figure 2
<p>Graphical representation of the relative quantification of the CatSper1 (<b>a</b>), CatSper2 (<b>b</b>), NBC (<b>c</b>), PKA (<b>d</b>), PKC (<b>e</b>) and ADCY 10 (<b>f</b>) proteins in bovine spermatozoa (n = 30) in fresh state (native control (Ctrl<sub>N</sub>)) and cryopreserved in the absence (cryopreserved control (Ctrl<sub>C</sub>)) or presence of selected theobromine (TBR) doses. Mean ± S.D. * <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 ">
20 pages, 10806 KiB  
Article
Advancing Mental Health Care: Intelligent Assessments and Automated Generation of Personalized Advice via M.I.N.I and RoBERTa
by Yuezhong Wu, Huan Xie, Lin Gu, Rongrong Chen, Shanshan Chen, Fanglan Wang, Yiwen Liu, Lingjiao Chen and Jinsong Tang
Appl. Sci. 2024, 14(20), 9447; https://doi.org/10.3390/app14209447 (registering DOI) - 16 Oct 2024
Viewed by 366
Abstract
As mental health issues become increasingly prominent, we are now facing challenges such as the severe unequal distribution of medical resources and low diagnostic efficiency. This paper integrates finite state machines, retrieval algorithms, semantic-matching models, and medical-knowledge graphs to design an innovative intelligent [...] Read more.
As mental health issues become increasingly prominent, we are now facing challenges such as the severe unequal distribution of medical resources and low diagnostic efficiency. This paper integrates finite state machines, retrieval algorithms, semantic-matching models, and medical-knowledge graphs to design an innovative intelligent auxiliary evaluation tool and a personalized medical-advice generation application, aiming to improve the efficiency of mental health assessments and the provision of personalized medical advice. The main contributions include the folowing: (1) Developing an auxiliary diagnostic tool that combines the Mini-International Neuropsychiatric Interview (M.I.N.I.) with finite state machines to systematically collect patient information for preliminary assessments; (2) Enhancing data processing by optimizing retrieval algorithms for efficient filtering and employing a fine-tuned RoBERTa model for deep semantic matching and analysis, ensuring accurate and personalized medical-advice generation; (3) Generating intelligent suggestions using NLP techniques; when semantic matching falls below a specific threshold, integrating medical-knowledge graphs to produce general medical advice. Experimental results show that this application achieves a semantic-matching degree of 0.9 and an accuracy of 0.87, significantly improving assessment accuracy and the ability to generate personalized medical advice. This optimizes the allocation of medical resources, enhances diagnostic efficiency, and provides a reference for advancing mental health care through artificial-intelligence technology. Full article
Show Figures

Figure 1

Figure 1
<p>Overall architecture.</p>
Full article ">Figure 2
<p>Auxiliary assessment effect.</p>
Full article ">Figure 3
<p>Working principle diagram of finite state machine.</p>
Full article ">Figure 4
<p>Program flowchart.</p>
Full article ">Figure 5
<p>Keyword-weight matching. Letters a–f represent different weight values.</p>
Full article ">Figure 6
<p>Cross-Encoder structure.</p>
Full article ">Figure 7
<p>RoBERTa model based on Cross-Encoder.</p>
Full article ">Figure 8
<p>Medical advice generation process flowchart. Letters a–e represent different keywords.</p>
Full article ">Figure 9
<p>Knowledge generation process in the mental and psychological health domain-knowledge graph. Different colors represent entity categories, and different letters represent subcategories of the same entity.</p>
Full article ">Figure 10
<p>Weight distribution.</p>
Full article ">Figure 11
<p>Weight-matching score.</p>
Full article ">Figure 12
<p>Manhattan distance score.</p>
Full article ">Figure 13
<p>Euclidean distance score.</p>
Full article ">Figure 14
<p>Cosine value score.</p>
Full article ">Figure 15
<p>RoBERTa vs. FT-RoBERTa.</p>
Full article ">Figure 16
<p>Mental and psychological health domain medical-knowledge graph.</p>
Full article ">Figure 17
<p>Entity-related info.</p>
Full article ">Figure 18
<p>Attribute-related info.</p>
Full article ">Figure 19
<p>Entity relationship-related Info.</p>
Full article ">Figure 20
<p>Depression knowledge generation.</p>
Full article ">Figure 21
<p>Mania knowledge generation.</p>
Full article ">
25 pages, 5904 KiB  
Article
In Vitro Evaluation of New 5-Nitroindazolin-3-one Derivatives as Promising Agents against Trypanosoma cruzi
by Josué Pozo-Martínez, Vicente J. Arán, Matías Zúñiga-Bustos, Sebastián Parra-Magna, Esteban Rocha-Valderrama, Ana Liempi, Christian Castillo, Claudio Olea-Azar and Mauricio Moncada-Basualto
Int. J. Mol. Sci. 2024, 25(20), 11107; https://doi.org/10.3390/ijms252011107 - 16 Oct 2024
Viewed by 261
Abstract
Chagas disease is a prevalent health problem in Latin America which has received insufficient attention worldwide. Current treatments for this disease, benznidazole and nifurtimox, have limited efficacy and may cause side effects. A recent study proposed investigating a wide range of nitroindazole and [...] Read more.
Chagas disease is a prevalent health problem in Latin America which has received insufficient attention worldwide. Current treatments for this disease, benznidazole and nifurtimox, have limited efficacy and may cause side effects. A recent study proposed investigating a wide range of nitroindazole and indazolone derivatives as feasible treatments. Therefore, it is proposed that adding a nitro group at the 5-position of the indazole and indazolone structure could enhance trypanocidal activity by inducing oxidative stress through activation of the nitro group by NTRs (nitroreductases). The study results indicate that the nitro group advances free radical production, as confirmed by several analyses. Compound 5a (5-nitro-2-picolyl-indazolin-3-one) shows the most favorable trypanocidal activity (1.1 ± 0.3 µM in epimastigotes and 5.4 ± 1.0 µM in trypomastigotes), with a selectivity index superior to nifurtimox. Analysis of the mechanism of action indicated that the nitro group at the 5-position of the indazole ring induces the generation of reactive oxygen species (ROS), which causes apoptosis in the parasites. Computational docking studies reveal how the compounds interact with critical residues of the NTR and FMNH2 (flavin mononucleotide reduced) in the binding site, which is also present in active ligands. The lipophilicity of the studied series was shown to influence their activity, and the nitro group was found to play a crucial role in generating free radicals. Further investigations are needed of derivatives with comparable lipophilic characteristics and the location of the nitro group in different positions of the base structure. Full article
(This article belongs to the Section Biochemistry)
Show Figures

Figure 1

Figure 1
<p>5-Nitroindazolin-3-ones with trypanocidal activity previously described [<a href="#B15-ijms-25-11107" class="html-bibr">15</a>,<a href="#B18-ijms-25-11107" class="html-bibr">18</a>] and structures of nifurtimox and benznidazole.</p>
Full article ">Figure 2
<p>5-Nitroindazolin-3-ones and other 5-nitroindazole derivatives studied in this work.</p>
Full article ">Figure 3
<p>Cyclic voltammogram of compound <b>3b</b>, recorded at different scan rates between 0.1 and 2.0 V/s.</p>
Full article ">Figure 4
<p>(<b>A</b>) Cyclic voltammogram of compound <b>19</b> with a potential sweep between −2.0 and 0.0 V and sweep speeds between 0.1 and 2.0 V/s. (<b>B</b>) Cyclic voltammogram using a potential range between −1.8 and −0.9 V and sweep speeds between 0.1 and 2.0 V/s. (<b>C</b>) Cyclic voltammogram using a speed of 2.0 V/s: (i) the first black line without NaOH; (ii) the red line in the presence of 30 mM NaOH.</p>
Full article ">Figure 5
<p>Experimental (black) and WimSIM (blue) simulated the ESR spectrum of (<b>A</b>) compound <b>5a</b>, (<b>B</b>) compound <b>10</b>, and (<b>C</b>) compound <b>23</b> at room temperature in DMSO.</p>
Full article ">Figure 5 Cont.
<p>Experimental (black) and WimSIM (blue) simulated the ESR spectrum of (<b>A</b>) compound <b>5a</b>, (<b>B</b>) compound <b>10</b>, and (<b>C</b>) compound <b>23</b> at room temperature in DMSO.</p>
Full article ">Figure 6
<p>(<b>A</b>). Percentage of SYTOX Green probe incorporation of the most active compounds on <span class="html-italic">T. cruzi</span> trypomastigotes. (<b>B</b>). Percentage of SYTOX Green probe incorporation of the most active compounds on RAW 264.7 cells. (<b>C</b>). Percentage of TMRM probe incorporation by variation of the mitochondrial membrane potential on <span class="html-italic">T. cruzi</span> trypomastigote. (<b>D</b>). Percentage of ATP levels in <span class="html-italic">T. cruzi</span> trypomastigotes by the effect of the compounds with the highest trypanocidal activity. The significant difference compared to the control (one-way ANOVA with Dunnett post-test, ****: <span class="html-italic">p</span> ≤ 0.001; **: <span class="html-italic">p</span> ≤ 0.05; *: <span class="html-italic">p</span> ≤ 0.1). ns: non significative variation.</p>
Full article ">Figure 7
<p>Increase in fluorescence as a function of time by generation of intracellular ROS in trypomastigotes of <span class="html-italic">T. cruzi</span>.</p>
Full article ">Figure 8
<p>Spectra of the spin adducts generated in trypomastigote forms of <span class="html-italic">T. cruzi</span> (Dm28c) at room temperature. (<b>A</b>) Spectrum recorded with trypomastigotes incubated with compound <b>5a</b> and DMPO, marked with (*), (↓), and (+), radicals centered on carbon, DMPOOX, and hydroxyl radical, respectively. (<b>B</b>) Spectrum recorded in trypomastigotes with the DMPO spin trap.</p>
Full article ">Figure 9
<p>(<b>A</b>) Structure of modeled <span class="html-italic">Tc</span>NTR receptor including two monomers (chain A in red and chain B in blue, respectively). (<b>B</b>) Superposition of FMN binding mode into ecNTR receptor and <span class="html-italic">Tc</span>NTR docking result. (<b>C</b>) Ligand interaction diagram of FMNH<sub>2</sub> and <span class="html-italic">Tc</span>NTR binding site.</p>
Full article ">Figure 10
<p>Ligand interaction maps of (<b>A</b>) <b>5a</b>, (<b>B</b>) <b>7</b>, (<b>C</b>) <b>1</b>, and active ligands (<b>D</b>) NFX, (<b>E</b>) <b>2a,</b> and (<b>F</b>) <b>2b</b> into the <span class="html-italic">Tc</span>NTR binding site predicted through docking calculations (carbons are in green). FMNH<sub>2</sub> is depicted in the figure with gray carbons, while chains A and B of <span class="html-italic">Tc</span>NTR are colored red and blue, respectively.</p>
Full article ">Figure 11
<p>Correlations between (<b>A</b>) lipophilicity and trypanocidal activity on the trypomastigote form of the 5-nitroindazolin-3-one series, the halogenated compounds that did not correlate are indicated in a circle, (<b>B</b>) reduction potentials (Epc) and trypanocidal activity on the trypomastigote form of the 5-nitroindazolin-3-one series, and (<b>C</b>) interaction energy with <span class="html-italic">Tc</span>NTR and trypanocidal activity on the trypomastigote form of the 5-nitroindazolin-3-one series.</p>
Full article ">Scheme 1
<p>Synthesis of 5-nitroindazolin-3-ones <b>7</b>–<b>18</b>. Reagents and conditions: Method A, for <b>7</b>–<b>18</b>: MeI, K<sub>2</sub>CO<sub>3</sub>, DMF, RT, overnight, 95–98%. Method B, for <b>8</b>–<b>12</b>, <b>14</b>, and <b>17</b>: substituted benzyl bromide, DMF, 150 °C, 4 h, 89–96%.</p>
Full article ">Scheme 2
<p>Proposed reduction mechanism for both 5-nitroindazole and 5-nitroindazolinone series with labile protons.</p>
Full article ">
20 pages, 6974 KiB  
Article
Targeting Ferroptosis with Small Molecule Atranorin (ATR) as a Novel Therapeutic Strategy and Providing New Insight into the Treatment of Breast Cancer
by Mine Ensoy and Demet Cansaran-Duman
Pharmaceuticals 2024, 17(10), 1380; https://doi.org/10.3390/ph17101380 - 16 Oct 2024
Viewed by 253
Abstract
Background/Objectives: Ferroptosis results from the accumulation of iron-dependent lipid peroxides and reactive oxygen species (ROS). Previous research has determined the effect of atranorin (ATR) on other cell death mechanisms, but its potential for a ferroptotic effect depending on ROS levels is unclear. This [...] Read more.
Background/Objectives: Ferroptosis results from the accumulation of iron-dependent lipid peroxides and reactive oxygen species (ROS). Previous research has determined the effect of atranorin (ATR) on other cell death mechanisms, but its potential for a ferroptotic effect depending on ROS levels is unclear. This study details the therapeutic role of small-molecule ATR through ferroptosis by suppressing MDA-MB-231, MCF-7, BT-474, and SK-BR-3 breast cancer cells. Methods: The anti-proliferative effect of ATR on cells was evaluated by xCELLigence analysis, and ferroptotic activity was evaluated by enzymatic assay kits. The changes in gene and protein expression levels of ATR were investigated by the qRT-PCR and western blot. In addition, mitochondrial changes were examined by transmission electron microscopy. Results: ATR was found to reduce cell viability in cancer cells in a dose- and time-dependent manner without showing cytotoxic effects on normal breast cells. In BT-474 and MDA-MB-231 cells, ATR, which had a higher anti-proliferative effect, increased iron, lipid peroxidation, and ROS levels in cells and decreased the T-GSH/GSSG ratio. The results revealed for the first time that small-molecule ATR exhibited anti-cancer activity by inducing the glutathione pathway and ferroptosis. Conclusions: This study highlights the potential of ATR as a drug candidate molecule that can be used in the development of new therapeutic strategies for the treatment of triple-negative and luminal-B breast cancer subtypes. Full article
(This article belongs to the Section Biopharmaceuticals)
Show Figures

Graphical abstract

Graphical abstract
Full article ">Figure 1
<p>Dose- and time-dependent anti-proliferative effects of ATR on (<b>A</b>) MCF-12A, (<b>B</b>) BT-474, (<b>C</b>) MDA-MB-231, (<b>D</b>) MCF-7, and (<b>E</b>) SK-BR-3 cells using MTT assays; (<b>F</b>) cell viability (%) after 48 h in cells treated with different concentrations of ATR compared to control. Data are represented as the mean ± SD error of three biological replicates. * <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 (compared to control).</p>
Full article ">Figure 2
<p>Dose- and time-dependent anti-proliferative effects of ATR on (<b>A</b>) BT-474, (<b>B</b>) MDA-MB-231, and (<b>C</b>) MCF-12A cells using the xCELLigence real-time cell analyzer. (** <span class="html-italic">p</span> &lt; 0.01; *** <span class="html-italic">p</span> &lt; 0.001).</p>
Full article ">Figure 3
<p>Dose- and time-dependent anti-proliferative effects of (<b>A</b>) erastin and (<b>B</b>) ferrostatin-1 on BT-474 cells and (<b>C</b>) erastin and (<b>D</b>) ferrostatin-1 on MDA-MB-231 cells using the xCELLigence real-time cell analyzer (** <span class="html-italic">p</span> &lt; 0.01).</p>
Full article ">Figure 4
<p>The effect of different combinations of ATR, erastin, and ferrostatin-1 molecules on cell viability and proliferation in (<b>A</b>) BT-474 and (<b>B</b>) MDA-MB-231 cell lines (*** <span class="html-italic">p</span> &lt; 0.001).</p>
Full article ">Figure 5
<p>(<b>A</b>) Iron ion level, (<b>B</b>) T-GSH/GSSG ratio, (<b>C</b>) MDA level, (<b>D</b>) ROS level in BT-474 cells, and (<b>E</b>) iron ion level, (<b>F</b>) T-GSH/GSSG ratio, (<b>G</b>) MDA level, (<b>H</b>) ROS level in MDA-MB-231 cells with and without ATR (IC<sub>50</sub> concentration); ferroptosis-related gene expression levels in (<b>I</b>) BT-474 and (<b>J</b>) MDA-MB-231 cells treated and untreated with ATR. <span class="html-italic">Gapdh</span> was used as the housekeeping gene. Data are represented as the mean ± SD error of three biological replicates; (<b>K</b>,<b>L</b>) ferroptosis-related protein levels in BT-474 and MDA-MB-231 cells with and without ATR. Results are normalized to β-Actin. The dash lines are used to show the change according to the control. (* <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 6
<p>Images of mitochondrial morphology under transmission electron microscopy in BT-474 cells of treated and non-treated ATR. Images were obtained with a magnification scale of 8000× and 15,000×. The red arrows indicate mitochondria.</p>
Full article ">Figure 7
<p>Images of mitochondrial morphology under transmission electron microscopy in MDA-MB-231 cells treated and non-treated ATR. Images were obtained with a magnification scale of 8000× and 15,000×. The red arrows indicate mitochondria.</p>
Full article ">
15 pages, 1419 KiB  
Article
Polydatin Prevents Electron Transport Chain Dysfunction and ROS Overproduction Paralleled by an Improvement in Lipid Peroxidation and Cardiolipin Levels in Iron-Overloaded Rat Liver Mitochondria
by Itzel Reyna-Bolaños, Elsa Paola Solís-García, Manuel Alejando Vargas-Vargas, Donovan J. Peña-Montes, Alfredo Saavedra-Molina, Christian Cortés-Rojo and Elizabeth Calderón-Cortés
Int. J. Mol. Sci. 2024, 25(20), 11104; https://doi.org/10.3390/ijms252011104 - 16 Oct 2024
Viewed by 342
Abstract
Increased intramitochondrial free iron is a key feature of various liver diseases, leading to oxidative stress, mitochondrial dysfunction, and liver damage. Polydatin is a polyphenol with a hepatoprotective effect, which has been attributed to its ability to enhance mitochondrial oxidative metabolism and antioxidant [...] Read more.
Increased intramitochondrial free iron is a key feature of various liver diseases, leading to oxidative stress, mitochondrial dysfunction, and liver damage. Polydatin is a polyphenol with a hepatoprotective effect, which has been attributed to its ability to enhance mitochondrial oxidative metabolism and antioxidant defenses, thereby inhibiting reactive oxygen species (ROS) dependent cellular damage processes and liver diseases. However, it has not been explored whether polydatin is able to exert its effects by protecting the phospholipid cardiolipin against damage from excess iron. Cardiolipin maintains the integrity and function of electron transport chain (ETC) complexes and keeps cytochrome c bound to mitochondria, avoiding uncontrolled apoptosis. Therefore, the effect of polydatin on oxidative lipid damage, ETC activity, cytochrome levels, and ROS production was explored in iron-exposed rat liver mitochondria. Fe2+ increased lipid peroxidation, decreased cardiolipin and cytochromes c + c1 and aa3 levels, inhibited ETC complex activities, and dramatically increased ROS production. Preincubation with polydatin prevented all these effects to a variable degree. These results suggest that the hepatoprotective mechanism of polydatin involves the attenuation of free radical production by iron, which enhances cardiolipin levels by counteracting membrane lipid peroxidation. This prevents the loss of cytochromes, improves ETC function, and decreases mitochondrial ROS production. Full article
(This article belongs to the Section Bioactives and Nutraceuticals)
Show Figures

Figure 1

Figure 1
<p>Effect of pretreatment with 100 µM polydatin (PD) or 5 µM butylated hydroxytoluene (BHT) on lipid peroxidation levels in mitochondria exposed to 50 µM ferrous iron (Fe). Results are expressed as the mean ± standard error of n ≥ 3. * <span class="html-italic">p</span> &lt; 0.05 vs. CTRL and <sup>§</sup> <span class="html-italic">p</span> &lt; 0.05 vs. Fe. (one-way ANOVA, Holm–Sidak post hoc test, <span class="html-italic">p</span> &lt; 0.05).</p>
Full article ">Figure 2
<p>Effect of pretreatment with 100 µM polydatin (PD) or 5 µM butylated hydroxytoluene (BHT) on cardiolipin levels in mitochondria exposed to 50 µM ferrous iron (Fe). Results are expressed as the mean ± standard error of n ≥ 4. * <span class="html-italic">p</span> &lt; 0.05 vs. CTRL, <sup>§</sup> <span class="html-italic">p</span> &lt; 0.05 vs. Fe and <sup>¶</sup> <span class="html-italic">p</span> &lt; 0.05 vs. PD + Fe (one-way ANOVA, Holm–Sidak post hoc test, <span class="html-italic">p</span> &lt; 0.05).</p>
Full article ">Figure 3
<p>Effect of pretreatment with 100 µM polydatin (PD) or 5 µM butylated hydroxytoluene (BHT) on cytochrome levels in mitochondria exposed to 50 µM ferrous iron (Fe): (<b>a</b>) representative cytochrome differential absorption spectra. Dotted lines indicate the absorption maximum for cytochromes <span class="html-italic">c</span> + <span class="html-italic">c</span><sub>1</sub> (550 nm) and cytochromes <span class="html-italic">aa</span><sub>3</sub> (600 nm); (<b>b</b>) quantification of cytochrome <span class="html-italic">c + c</span><sub>1</sub> levels; (<b>c</b>): quantification of cytochrome <span class="html-italic">aa</span><sub>3</sub> levels. Results are expressed in (<b>b</b>,<b>c</b>) as the mean ± standard error of n = 4. * <span class="html-italic">p</span> &lt; 0.05 vs. CTRL and <sup>§</sup> <span class="html-italic">p</span> &lt; 0.05 vs. Fe (one-way ANOVA, Holm–Sidak post hoc test, <span class="html-italic">p</span> &lt; 0.05).</p>
Full article ">Figure 4
<p>Effect of pretreatment with polydatin (PD) or butylated hydroxytoluene (BHT) on the activity of ETC complexes of mitochondria exposed to Fe<sup>2+</sup> (Fe): (<b>a</b>) complex I activity (50 µM PD, 2.5 µM BHT, 25 µM Fe<sup>2+</sup>); (<b>b</b>) complex II activity (100 µM PD, 5 µM BHT, 50 µM Fe<sup>2+</sup>); (<b>c</b>) complex III activity (100 µM PD, 5 µM BHT, 50 µM Fe<sup>2+</sup>); (<b>d</b>) complex IV activity (200 µM PD, 5 µM BHT, 100 µM Fe<sup>2+</sup>). Results are expressed as mean ± standard error of n ≥ 3. * <span class="html-italic">p</span> &lt; 0.05 vs. CTRL and <sup>§</sup> <span class="html-italic">p</span> &lt; 0.05 vs. Fe (one-way ANOVA, Holm–Sidak post hoc test, except for (<b>c</b>), where the Student–Newman–Keuls method was used, <span class="html-italic">p</span> &lt; 0.05).</p>
Full article ">Figure 5
<p>Effect of pretreatment with 50 µM polydatin (PD) on ROS levels of mitochondria exposed to 25 µM ferrous iron (Fe). Mitochondria were incubated with glutamate–malate (G + M; open bars) to stimulate ROS production and antimycin A (AA; diagonal striped bars) to stimulate maximal ROS production in the ETC. Results are expressed as the mean ± standard error of n ≥ 5. * <span class="html-italic">p</span> &lt; 0.05 vs. CTRL and <sup>§</sup> <span class="html-italic">p</span> &lt; 0.05 vs. Fe (one-way ANOVA, Holm–Sidak post hoc test, <span class="html-italic">p</span> &lt; 0.05).</p>
Full article ">
15 pages, 5526 KiB  
Article
Extracellular Vesicles Contribute to Oxidized LDL-Induced Stromal Cell Proliferation in Benign Prostatic Hyperplasia
by Franco F. Roldán Gallardo, Daniel E. Martínez Piñerez, Kevin F. Reinarz Torrado, Gabriela A. Berg, Jael D. Herzfeld, Vanina G. Da Ros, Manuel López Seoane, Cristina A. Maldonado and Amado A. Quintar
Biology 2024, 13(10), 827; https://doi.org/10.3390/biology13100827 - 16 Oct 2024
Viewed by 222
Abstract
Background: Clinical and experimental evidence has linked Benign Prostatic Hyperplasia (BPH) with dyslipidemic and hypercholesterolemic conditions, though the underlying cellular mechanisms remain unclear. This study investigates the impact of dyslipidemia, specifically oxidized LDL (OxLDL), on prostatic stromal cell proliferation and the release of [...] Read more.
Background: Clinical and experimental evidence has linked Benign Prostatic Hyperplasia (BPH) with dyslipidemic and hypercholesterolemic conditions, though the underlying cellular mechanisms remain unclear. This study investigates the impact of dyslipidemia, specifically oxidized LDL (OxLDL), on prostatic stromal cell proliferation and the release of extracellular vesicles (EVs). Methods: Mice were fed a high-fat diet, and human prostatic stromal cells (HPSCs) were treated with OxLDL. Proliferation assays and EV characterization were performed to assess the role of EVs in BPH progression. Results: Pro-atherogenic conditions significantly increased cell proliferation in both murine prostatic cells and HPSCs. Treatment with metformin effectively inhibited OxLDL-induced proliferation. Additionally, OxLDL stimulated the production and release of pro-proliferative EVs by HPSCs, which further promoted cellular proliferation. Conclusions: The findings suggest that dyslipidemia drives prostatic stromal cell proliferation and EV secretion, contributing to BPH progression. Metformin demonstrates potential as a therapeutic agent to mitigate these effects, offering insight into novel strategies for BPH management. This study highlights the complex interaction between dyslipidemia, cell proliferation, and extracellular communication in the context of BPH pathogenesis. Full article
(This article belongs to the Collection Extracellular Vesicles: From Biomarkers to Therapeutic Tools)
Show Figures

Figure 1

Figure 1
<p>Hypercholesterolemia induced by a high-fat diet promotes cell proliferation in the prostate gland. Mice were fed on a HFD for 12 weeks and their prostates were processed and analyzed by Ki-67 immunohistochemistry. (<b>A</b>) Representative images of prostate glands immunostained for the proliferation marker Ki-67 showing an increase in HFD-fed mice (brown nuclei, arrowheads). (<b>B</b>) Quantification of Ki-67-positive stromal cells in ventral prostates from chow- or HFD-fed mice (mean ± SEM; <span class="html-italic">n</span> = 10 per group; * <span class="html-italic">p</span> &lt; 0.05). (<b>C</b>) Correlative analysis showing serum cholesterol levels and prostatic Ki-67 positive counts (<span class="html-italic">n</span> = 20, Pearson’s correlation test).</p>
Full article ">Figure 2
<p>OxLDL at 20 µg/mL increases cell proliferation and viability of HPSCs. (<b>A</b>) HPSCs from patient HPSC-1 were used for the initial proliferation assay. Ki-67 immunocytochemistry analysis was performed to evaluate the effects of two types of OxLDL: moderately oxidized (OxLDL1) and highly oxidized (OxLDL2). Although OxLDL1 at 20 and 100 µg/mL and OxLDL2 at 100 µg/mL showed changes in cell proliferation (** <span class="html-italic">p</span> &lt; 0.01, * <span class="html-italic">p</span> &lt; 0.05, and ** <span class="html-italic">p</span> &lt; 0.01 vs. control, respectively), OxLDL at 20 µg/mL was the only concentration without signs of cell toxicity. Additionally, OxLDL at 20 µg/mL induced an increase in cell proliferation (** <span class="html-italic">p</span> &lt; 0.01 vs. control). (<b>B</b>) Quantification and representative merged images of DAPI/BrdU showing the proliferative effects of OxLDL1 at 20 µg/mL for 24 h vs. control (** <span class="html-italic">p</span> &lt; 0.01), which were evaluated in different primary cultures of HPSCs by BrdU incorporation (<span class="html-italic">n</span> = 7). (<b>C</b>) The same proliferative effect of OxLDL1 at 20 µg/mL for 24 h vs. control (** <span class="html-italic">p</span> &lt; 0.01) was evaluated in primary cultures of HPSCs by Ki-67 assay (<span class="html-italic">n</span> = 8). (<b>D</b>) Resazurin absorbance assay for HPSCs and THP-1 stimulated with OxLDL1 at 20 µg/mL for 24 h was carried out to assess cell viability. OxLDL1 at 20 µg/mL increased cell viability compared to control (** <span class="html-italic">p</span> &lt; 0.01). (<b>E</b>) Assays with OxLDL1 and OxLDL2 stimuli in THP-1 cells evaluating the number of total live and apoptotic cells. Neither OxLDL1 nor OxLDL2 exerted changes in the cell number of THP-1 cells. All tests were performed in two replicates, each in triplicate. Error bars represent mean ± SEM.</p>
Full article ">Figure 3
<p>OxLDL causes morphological and ultrastructural changes in HPSCs. (<b>A</b>) Ultrastructure of HPSCs derived from patient HPSC-1, treated with OxLDL1 at 20 µg/mL for 24 h, and evaluated by TEM. Cells stimulated by OxLDL display frequent actin filaments (FAs), dilated endoplasmic reticulum (ER), numerous mitochondria (M), Golgi apparatus (G), and prominent nucleoli (NLI). (<b>B</b>) Quantification of organelles observed by TEM for HPSCs from patient HPSC-1, control vs. treated with OxLDL1 at 20 µg/mL for 24 h. There was a significant increase in the number of mitochondria (*** <span class="html-italic">p</span> &lt; 0.001), nucleoli (** <span class="html-italic">p</span> &lt; 0.01), Golgi and endoplasmic reticulum (*** <span class="html-italic">p</span> &lt; 0.001), and actin myofibrils (* <span class="html-italic">p</span> &lt; 0.05) when stimulated by OxLDL1 compared to the control. All tests were performed in two replicates, each in triplicate. Bars represent mean ± SEM.</p>
Full article ">Figure 4
<p>Metformin inhibits the proliferative effect of OxLDL on HPSCs. Cells were stimulated with 20 µg/mL of OxLDL1, and atorvastatin, metformin, or a combination of both were used as co-stimuli, each at low and high doses for 24 h. (<b>A</b>) Cell proliferation assays by BrdU incorporation were carried out in all primary cultures of HPSCs. Images of cells are representative of patient HPSC-6. OxLDL1 enhanced cell proliferation compared to the control (** <span class="html-italic">p</span> &lt; 0.01), while metformin inhibited the proliferative effect induced by OxLDL1 (<sup>ΔΔ</sup> <span class="html-italic">p</span> &lt; 0.01 and <sup>ΔΔΔ</sup> <span class="html-italic">p</span> &lt; 0.001 vs. OxLDL1). (<b>B</b>) A similar effect was observed by Ki-67 immunocytochemistry analysis on HPSCs. Representative images of cells are from the same patient HPSC-6 (** <span class="html-italic">p</span> &lt; 0.01 for OxLDL1 vs. C, and <sup>ΔΔ</sup> <span class="html-italic">p</span> &lt; 0.01 for metformin vs. OxLDL1). (<b>C</b>) Determination of the total cell number was performed on cells from patient HPSC-5, HPSC-6 and HPSC-7. OxLDL1 increased the cell number compared to the control (*** <span class="html-italic">p</span> &lt; 0.001), while metformin reduced the cell number compared to OxLDL1 (<sup>Δ</sup> <span class="html-italic">p</span> &lt; 0.05 and <sup>ΔΔ</sup> <span class="html-italic">p</span> &lt; 0.01). (<b>D</b>) Cell viability assay evaluated by resazurin absorbance was performed on cells from patient HPSC-5, HPSC-6 and HPSC-7. All the tests were performed in two replicates, each in triplicate. Bars represent mean ± SEM.</p>
Full article ">Figure 5
<p>EVs derived from OxLDL-stimulated HPSCs exert a pro-proliferative effect on prostatic cells. (<b>A</b>) Images of HPSCs-derived EVs from patient HPSC-4, treated with OxLDL1 at 20 µg/mL for 24 h, isolated by ultracentrifugation and visualized by TEM. (<b>B</b>) Immunolabeling for CD63 in EVs derived from patient HPSC-4, using colloidal gold particles. (<b>C</b>) The histogram of size range distribution of EVs derived from HPSCs of patient HPSC-4 shows an increase in the number of EVs, mainly in the smaller size ranges (* <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 vs. control). EVs obtained from HPSCs stimulated with OxLDL1 at 20 µg/mL for 24 h (EVs OxLDL1) and the control (EVs control) were used as stimuli on HPSCs from the same patient (<b>D</b>) and on prostatic epithelial cells from the PC3 line (<b>E</b>). EVs from OxLDL-stimulated HPSCs induced a significant proliferative effect on HPSCs and PC3 cells in all assays (** <span class="html-italic">p</span> &lt; 0.01 and * <span class="html-italic">p</span> &lt; 0.05 vs. vehicle). The proliferative effect was evaluated by Ki-67 and all tests were performed in three replicates, each in triplicate. Bars represent mean ± SEM.</p>
Full article ">
15 pages, 1641 KiB  
Article
Expression of Myeloperoxidase in Patient-Derived Endothelial Colony-Forming Cells—Associations with Coronary Artery Disease and Mitochondrial Function
by Weiqian Eugene Lee, Elijah Genetzakis, Giannie Barsha, Joshua Vescovi, Carmen Mifsud, Stephen T. Vernon, Tung Viet Nguyen, Michael P. Gray, Stuart M. Grieve and Gemma A. Figtree
Biomolecules 2024, 14(10), 1308; https://doi.org/10.3390/biom14101308 (registering DOI) - 16 Oct 2024
Viewed by 294
Abstract
Background and Aims: Myeloperoxidase (MPO) plays a critical role in the innate immune response and has been suggested to be a surrogate marker of oxidative stress and inflammation, with elevated levels implicated in cardiovascular diseases, such as atherosclerosis and heart failure, as well [...] Read more.
Background and Aims: Myeloperoxidase (MPO) plays a critical role in the innate immune response and has been suggested to be a surrogate marker of oxidative stress and inflammation, with elevated levels implicated in cardiovascular diseases, such as atherosclerosis and heart failure, as well as in conditions like rheumatoid arthritis and cancer. While MPO is well-known in leukocytes, its expression and function in human endothelial cells remain unclear. This study investigates MPO expression in patient-derived endothelial colony-forming cells (ECFCs) and its potential association with CAD and mitochondrial function. Methods: ECFCs were cultured from the peripheral blood of 93 BioHEART-CT patients. MPO expression and associated functions were examined using qRT-PCR, immunochemistry, flow cytometry, and MPO activity assays. CAD presence was defined using CT coronary angiography (CACS > 0). Results: We report MPO presence in patient-derived ECFCs for the first time. MPO protein expression occurred in 70.7% of samples (n = 41) which had nuclear co-localisation, an atypical observation given its conventional localisation in the granules of neutrophils and monocytes. This suggests potential alternative roles for MPO in nuclear processes. MPO mRNA expression was detected in 66.23% of samples (n = 77). CAD patients had a lower proportion of MPO-positive ECFCs compared to non-CAD controls (57.45% vs. 80%, p = 0.04), a difference that persisted in the statin-naïve sub-cohort (53.85% vs. 84.62%, p = 0.02). Non-CAD patients with MPO expression showed upregulated mitochondrial-antioxidant genes (AIFM2, TXNRD1, CAT, PRDX3, PRDX6). In contrast, CAD patients with MPO gene expression had heightened mROS production and mitochondrial mass and decreased mitochondrial function compared to that of CAD patients without MPO gene expression. Conclusions: MPO is present in the nucleus of ECFCs. In non-CAD ECFCs, MPO expression is linked to upregulated mitochondrial-antioxidant genes, whereas in CAD ECFCs, it is associated with greater mitochondrial dysfunction. Full article
Show Figures

Figure 1

Figure 1
<p>MPO protein expression in patient-derived ECFCs. (<b>A</b>) Representative Western blot of MPO protein expression in corresponding patient-derived ECFCs using anti-MPO antibody and anti-β-actin for loading control. (<b>B</b>) MPO protein is co-localised to the nuclei of patient-derived ECFCs. Representative immunocytochemistry images of MPO protein expression co-localised to the nucleus of patient-derived ECFCs, showing nuclei (blue) and MPO granules (red) within the cell (20× magnification). The scale bar represents 100 μM. (<b>C</b>) Representative Western blot of subcellular expression of MPO in patient-derived ECFCs at the soluble nuclear and chromatin-bound nuclear subfractions. At least three biological replicates were used. Original images can be found in <a href="#app1-biomolecules-14-01308" class="html-app">Supplementary Materials</a> file.</p>
Full article ">Figure 2
<p>CAD patients were less likely to express MPO gene, as identified by qRT-PCR. Stacked bar plots showing the association between proportion in MPO gene expression and the presence of CAD in (<b>A</b>) all patients (<span class="html-italic">n</span> = 77; No CAD = 30, CAD = 47), (<b>B</b>) male patients (<span class="html-italic">n</span> = 40; No CAD = 14, CAD = 26) and (<b>C</b>) female patients (<span class="html-italic">n</span> = 36; No CAD = 16, CAD = 20). (<b>D</b>) Statin-naïve patients (<span class="html-italic">n</span> = 54; No CAD = 27, CAD = 27), (<b>E</b>) statin-naïve male patients (<span class="html-italic">n</span> = 27; No CAD = 12, CAD = 15) and (<b>F</b>) statin-naïve female patients (<span class="html-italic">n</span> = 26; No CAD = 14, CAD = 12). Statistical association was analysed using Pearson’s chi-square test (categorial variables). Categorical measurements are shown as percentages. * <span class="html-italic">p</span> &lt; 0.05, ** <span class="html-italic">p</span> &lt; 0.01.</p>
Full article ">Figure 2 Cont.
<p>CAD patients were less likely to express MPO gene, as identified by qRT-PCR. Stacked bar plots showing the association between proportion in MPO gene expression and the presence of CAD in (<b>A</b>) all patients (<span class="html-italic">n</span> = 77; No CAD = 30, CAD = 47), (<b>B</b>) male patients (<span class="html-italic">n</span> = 40; No CAD = 14, CAD = 26) and (<b>C</b>) female patients (<span class="html-italic">n</span> = 36; No CAD = 16, CAD = 20). (<b>D</b>) Statin-naïve patients (<span class="html-italic">n</span> = 54; No CAD = 27, CAD = 27), (<b>E</b>) statin-naïve male patients (<span class="html-italic">n</span> = 27; No CAD = 12, CAD = 15) and (<b>F</b>) statin-naïve female patients (<span class="html-italic">n</span> = 26; No CAD = 14, CAD = 12). Statistical association was analysed using Pearson’s chi-square test (categorial variables). Categorical measurements are shown as percentages. * <span class="html-italic">p</span> &lt; 0.05, ** <span class="html-italic">p</span> &lt; 0.01.</p>
Full article ">Figure 3
<p>MPO gene expression is associated with dysregulated mitochondrial function and dynamics. (<b>A</b>–<b>C</b>) CAD ECFCs with MPO gene expression at baseline had increased (<b>A</b>) mROS production and (<b>B</b>) mitochondrial mass and decreased (<b>C</b>) mitochondrial function. N = 8; no MPO gene expression: N = 3, MPO gene expression: N = 5. Data are represented as mean ± S.E.M. Welch’s <span class="html-italic">t</span>-test was performed.</p>
Full article ">Figure 4
<p>Antioxidant genes were upregulated with MPO gene expression in non-CAD patients. Relative fold gene expression was evaluated in ECFCs without CAD at baseline in (<b>A</b>) <span class="html-italic">AIFM2</span>, (<b>B</b>) <span class="html-italic">TXNRD1</span>, (<b>C</b>) <span class="html-italic">CAT</span>, (<b>D</b>) <span class="html-italic">PRDX3</span> and (<b>E</b>) <span class="html-italic">PRDX6</span>. Each sample was performed in triplicate. N = 30; no MPO gene expression: N = 6, MPO gene expression: N = 24. Data are represented as mean ± S.E.M. Student’s <span class="html-italic">t</span>-test was performed.</p>
Full article ">
Back to TopTop