[go: up one dir, main page]

 
 
Sign in to use this feature.

Years

Between: -

Subjects

remove_circle_outline
remove_circle_outline
remove_circle_outline
remove_circle_outline
remove_circle_outline
remove_circle_outline
remove_circle_outline
remove_circle_outline
remove_circle_outline

Journals

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

Article Types

Countries / Regions

remove_circle_outline
remove_circle_outline
remove_circle_outline
remove_circle_outline
remove_circle_outline
remove_circle_outline

Search Results (4,209)

Search Parameters:
Keywords = MMPs

Order results
Result details
Results per page
Select all
Export citation of selected articles as:
14 pages, 9322 KiB  
Article
Metabolic Reprogramming Induced by Aging Modifies the Tumor Microenvironment
by Xingyu Chen, Zihan Wang, Bo Zhu, Min Deng, Jiayue Qiu, Yunwen Feng, Ning Ding and Chen Huang
Cells 2024, 13(20), 1721; https://doi.org/10.3390/cells13201721 (registering DOI) - 17 Oct 2024
Abstract
Aging is an important risk factor for tumorigenesis. Metabolic reprogramming is a hallmark of both aging and tumor initiation. However, the manner in which the crosstalk between aging and metabolic reprogramming affects the tumor microenvironment (TME) to promote tumorigenesis was poorly explored. We [...] Read more.
Aging is an important risk factor for tumorigenesis. Metabolic reprogramming is a hallmark of both aging and tumor initiation. However, the manner in which the crosstalk between aging and metabolic reprogramming affects the tumor microenvironment (TME) to promote tumorigenesis was poorly explored. We utilized a computational approach proposed by our previous work, MMP3C (Modeling Metabolic Plasticity by Pathway Pairwise Comparison), to characterize aging-related metabolic plasticity events using pan-cancer bulk RNA-seq data. Our analysis revealed a high degree of metabolically organized heterogeneity across 17 aging-related cancer types. In particular, a higher degree of several energy generation pathways, i.e., glycolysis and impaired oxidative phosphorylation, was observed in older patients. Similar phenomena were also found via single-cell RNA-seq analysis. Furthermore, those energy generation pathways were found to be weakened in activated T cells and macrophages, whereas they increased in exhausted T cells, immunosuppressive macrophages, and Tregs in older patients. It was suggested that aging-induced metabolic switches alter glucose utilization, thereby influencing immune function and resulting in the remodeling of the TME. This work offers new insights into the associations between tumor metabolism and the TME mediated by aging, linking with novel strategies for cancer therapy. Full article
(This article belongs to the Section Cellular Metabolism)
19 pages, 5830 KiB  
Article
Exploration of the Anti-Photoaging Mechanisms of Lactiplantibacillus plantarum TWK10 in a UVB-Induced Mouse Model
by Te-Hua Liu, Wan-Jyun Lin, Meng-Chun Cheng, Yi-Chen Cheng, Chia-Chia Lee, Jin-Seng Lin and Tsung-Yu Tsai
Appl. Sci. 2024, 14(20), 9497; https://doi.org/10.3390/app14209497 - 17 Oct 2024
Abstract
Functional foods have shown promise in mitigating skin aging. This study aimed to evaluate the effects of Lactiplantibacillus plantarum TWK10 (LPTWK10) and its spray-dried supernatant powder on ultraviolet B (UVB)-induced skin photoaging in female BALB/c nude mice. Over a 13-week period of UVB [...] Read more.
Functional foods have shown promise in mitigating skin aging. This study aimed to evaluate the effects of Lactiplantibacillus plantarum TWK10 (LPTWK10) and its spray-dried supernatant powder on ultraviolet B (UVB)-induced skin photoaging in female BALB/c nude mice. Over a 13-week period of UVB exposure and concurrent administration of high doses of LPTWK10 or its spray-dried fermentation supernatant, significant improvements were observed, skin wrinkles were notably reduced, transepidermal water loss rate decreased by 68.94–70.77%, and stratum corneum hydration increased by 76.97–112.24%. Furthermore, LPTWK10 was effective in reducing erythema and inflammation while enhancing skin lightness. Histological assessments revealed substantial reductions in epidermal hyperplasia and collagen degradation. Additionally, LPTWK10 was found to influence critical mechanisms associated with collagen metabolism and proinflammatory cytokine production. In summary, LPTWK10 attenuates photoaging through modulation of collagen metabolism and reduction in inflammatory responses, suggesting its potential as a functional ingredient for delaying photoaging. Full article
(This article belongs to the Section Food Science and Technology)
Show Figures

Figure 1

Figure 1
<p>Effects of the administration of LPTWK10 or LPTWK10 spray-dried supernatant powders on dorsal wrinkles in UVB-induced nude mice. (<b>A</b>) NC (<b>B</b>) UVB (<b>C</b>) PC (<b>D</b>) LPLD (<b>E</b>) LPMD (<b>F</b>) LPHD (<b>G</b>) SPLD (<b>H</b>) SPMD (<b>I</b>) SPHD. Data are presented as mean ± SD (<span class="html-italic">n</span> = 7 or 8). UVB: ultraviolet B; NC: normal control; PC: positive control; LPLD, LPMD, and LPHD: administration of LPTWK10 at 2 × 10<sup>8</sup> CFU/kg bw/day, 2 × 10<sup>9</sup> CFU/kg bw/day and 2 × 10<sup>10</sup> CFU/kg bw/day, respectively; SPLD, SPMD and SPHD: administration of TWK10 spray-dried supernatant powders at 0.30 g/kg bw/day, 0.60 g/kg bw/day, and 1.20 g/kg bw/day, respectively. Red arrow: wrinkles.</p>
Full article ">Figure 2
<p>Effects of the administration of LPTWK10 or LPTWK10 spray-dried supernatant powders on (<b>A</b>) dorsal TEWL and (<b>B</b>) dorsal hydration of the stratum corneum in UVB-induced nude mice. Data are presented as mean ± SD (<span class="html-italic">n</span> = 7 or 8). Values with different uppercase letters were significantly different by Duncan’s multiple range test (<span class="html-italic">p</span> &lt; 0.05). TEWL: transepidermal water loss; UVB: ultraviolet B; NC: normal control; PC: positive control; LPLD, LPMD and LPHD: administration of LPTWK10 viable cells at 2 × 10<sup>8</sup> CFU/kg bw/day, 2 × 10<sup>9</sup> CFU/kg bw/day and 2 × 10<sup>10</sup> CFU/kg bw/day, respectively; SPLD, SPMD and SPHD: administration of TWK10 spray-dried supernatant powders at 0.30 g/kg bw/day, 0.60 g/kg bw/day, and 1.20 g/kg bw/day, respectively.</p>
Full article ">Figure 3
<p>Effects of the administration of LPTWK10 or LPTWK10 spray-dried supernatant powders on dorsal (<b>A</b>) erythema (<b>B</b>) a* value (<b>C</b>) melanin and (<b>D</b>) lightness in UVB-induced nude mice. Data are presented as mean ± SD (<span class="html-italic">n</span> = 7−8). Values with different uppercase letters were significantly different in the same weeks by Duncan’s multiple range test (<span class="html-italic">p</span> &lt; 0.05). UVB: ultraviolet B; NC: normal control; PC: positive control; LPLD, LPMD and LPHD: administration of LPTWK10 viable cells at 2 × 10<sup>8</sup> CFU/kg bw/day, 2 × 10<sup>9</sup> CFU/kg bw/day and 2 × 10<sup>10</sup> CFU/kg bw/day, respectively; SPLD, SPMD and SPHD: administration of LPTWK10 spray-dried supernatant powders at 0.30 g/kg bw/day, 0.60 g/kg bw/day, and 1.20 g/kg bw/day, respectively.</p>
Full article ">Figure 4
<p>Effects of the administration of LPTWK10 or LPTWK10 spray-dried supernatant powders on dorsal (<b>A</b>) TNF-α, (<b>B</b>) IL-6, and (<b>C</b>) IL-1β levels in UVB-induced nude mice. Data are presented as mean ± SD (<span class="html-italic">n</span> = 7 or 8). Values with different uppercase letters were significantly different by Duncan’s multiple range test (<span class="html-italic">p</span> &lt; 0.05). TNF-α: tumor necrosis factor-α; IL-6: interleukin-6; IL-1β: interleukin-1β; UVB: ultraviolet B; NC: normal control; PC: positive control; LPLD, LPMD and LPHD: administration of LPTWK10 viable cells at 2 × 10<sup>8</sup> CFU/kg bw/day, 2 × 10<sup>9</sup> CFU/kg bw/day and 2 × 10<sup>10</sup> CFU/kg bw/day, respectively; SPLD, SPMD and SPHD: administration of LPTWK10 spray-dried supernatant powders at 0.30 g/kg bw/day, 0.60 g/kg bw/day, and 1.20 g/kg bw/day, respectively.</p>
Full article ">Figure 5
<p>Effects of the administration of LPTWK10 or LPTWK10 spray-dried supernatant powders on dorsal (<b>A</b>) Type I procollagen, (<b>B</b>) Type III procollagen, (<b>C</b>) Type III procollagen/Type I procollagen, (<b>D</b>) MMP-1, (<b>E</b>) MMP-9 and (<b>F</b>) AP-1 levels in UVB-induced nude mice. Data are presented as mean ± SD (<span class="html-italic">n</span> = 7 or 8). Values with different uppercase letters were significantly different by Duncan’s multiple range test (<span class="html-italic">p</span> &lt; 0.05). UVB: ultraviolet B; NC: normal control; PC: positive control; LPLD, LPMD and LPHD: administration of LPTWK10 viable cells at 2 × 10<sup>8</sup> CFU/kg bw/day, 2 × 10<sup>9</sup> CFU/kg bw/day and 2 × 10<sup>10</sup> CFU/kg bw/day, respectively; SPLD, SPMD and SPHD: administration of LPTWK10 spray-dried supernatant powders at 0.30 g/kg bw/day, 0.60 g/kg bw/day, and 1.20 g/kg bw/day, respectively.</p>
Full article ">Figure 6
<p>Effects of the administration of LPTWK10 viable cells or LPTWK10 spray-dried supernatant powders on dorsal histopathological findings (<b>A</b>–<b>I</b>) H&amp;E stain, (<b>J</b>) epidermis thickness level and (<b>K</b>–<b>S</b>) MT stain in UVB-induced nude mice. (<b>A</b>,<b>K</b>) NC, (<b>B</b>,<b>L</b>) UVB, (<b>C</b>,<b>M</b>) PC, (<b>D</b>,<b>N</b>) LPLD, (<b>E</b>,<b>O</b>) LPMD, (<b>F</b>,<b>P</b>) LPHD, (<b>G</b>,<b>Q</b>) SPLD, (<b>H</b>,<b>R</b>) SPMD, and (<b>I</b>,<b>S</b>) SPHD. Original magnification, 400×; scale bar = 100 μm. Data are presented as mean ± SD (<span class="html-italic">n</span> = 7 or 8). H&amp;E: hematoxylin and eosin; UVB: ultraviolet B; NC: normal control; PC: positive control; LPLD, LPMD and LPHD: administration of LPTWK10 viable cells at 2 × 10<sup>8</sup> CFU/kg bw/day, 2 × 10<sup>9</sup> CFU/kg bw/day and 2 × 10<sup>10</sup> CFU/kg bw/day, respectively; SPLD, SPMD and SPHD: administration of LPTWK10 spray-dried supernatant powders at 0.30 g/kg bw/day, 0.60 g/kg bw/day, and 1.20 g/kg bw/day, respectively. Black arrow: epidermis thickness. Values with different uppercase letters were significantly different by Duncan’s multiple range test (<span class="html-italic">p</span> &lt; 0.05).</p>
Full article ">Figure 7
<p>Effects of the administration of LPTWK10 or LPTWK10 spray-dried supernatant powders on dorsal immunohistochemistry stain (claudin-1) in UVB-induced nude mice. (<b>A</b>) NC, (<b>B</b>) UVB, (<b>C</b>) PC, (<b>D</b>) LPLD, (<b>E</b>) LPMD, (<b>F</b>) LPHD, (<b>G</b>) SPLD, (<b>H</b>) SPMD, and (<b>I</b>) SPHD. Original magnification, 100×; scale bar = 200 μm. Data are presented as mean ± SD (<span class="html-italic">n</span> = 7 or 8). MT: Masson’s trichrome; UVB: ultraviolet B; NC: normal control; PC: positive control; LPLD, LPMD, and LPHD: administration of LPTWK10 viable cells at 2 × 10<sup>8</sup> CFU/kg bw/day, 2 × 10<sup>9</sup> CFU/kg bw/day and 2 × 10<sup>10</sup> CFU/kg bw/day, respectively; SPLD, SPMD and SPHD: administration of LPTWK10 spray-dried supernatant powders at 0.30 g/kg bw/day, 0.60 g/kg bw/day, and 1.20 g/kg bw/day, respectively. Black arrow: claudin-1 protein in the epidermis layer.</p>
Full article ">Figure 8
<p>Effects of the administration of LPTWK10 or LPTWK10 spray-dried supernatant powders on dorsal immunohistochemistry stain (laminin) in UVB-induced nude mice. (<b>A</b>) NC, (<b>B</b>) UVB, (<b>C</b>) PC, (<b>D</b>) LPLD, (<b>E</b>) LPMD, (<b>F</b>) LPHD, (<b>G</b>) SPLD, (<b>H</b>) SPMD, and (<b>I</b>) SPHD. Original magnification, 100×; scale bar = 200 μm. Data are presented as mean ± SD (<span class="html-italic">n</span> = 7 or 8). MT: Masson’s trichrome; UVB: ultraviolet B; NC: normal control; PC: positive control; LPLD, LPMD, and LPHD: administration of LPTWK10 viable cells at 2 × 10<sup>8</sup> CFU/kg bw/day, 2 × 10<sup>9</sup> CFU/kg bw/day and 2 × 10<sup>10</sup> CFU/kg bw/day, respectively; SPLD, SPMD and SPHD: administration of LPTWK10 spray-dried supernatant powders at 0.30 g/kg bw/day, 0.60 g/kg bw/day, and 1.20 g/kg bw/day, respectively. Black arrow: laminin protein in the basement membrane.</p>
Full article ">Figure 9
<p>Effects of the administration of LPTWK10 or LPTWK10 spray-dried supernatant powders on dorsal immunohistochemistry stain (type IV collagen) in UVB-induced nude mice. (<b>A</b>) NC, (<b>B</b>) UVB, (<b>C</b>) PC, (<b>D</b>) LPLD, (<b>E</b>) LPMD, (<b>F</b>) LPHD, (<b>G</b>) SPLD, (<b>H</b>) SPMD, and (<b>I</b>) SPHD. Original magnification, 100×; scale bar = 200 μm. Data are presented as mean ± SD (<span class="html-italic">n</span> = 7 or 8). MT: Masson’s trichrome; UVB: ultraviolet B; NC: normal control; PC: positive control; LPLD, LPMD and LPHD: administration of LPTWK10 viable cells at 2 × 10<sup>8</sup>, 2 × 10<sup>9</sup> and 2 × 10<sup>10</sup> CFU/kg bw/day, respectively; SPLD, SPMD and SPHD: administration of LPTWK10 spray-dried supernatant powders at 0.30, 0.60 and 1.20 g/kg bw/day, respectively. Black arrow: type IV collagen protein in the basement membrane.</p>
Full article ">
11 pages, 1009 KiB  
Article
Altered Expressions of IL-15, IFNG, and HPRT1 Genes in the Thin Endometria of Patients with Reproductive Disorders: A Prospective Comparative Study
by Almagul Kurmanova, Yeldar Ashirbekov, Gaukhar Kurmanova, Nagima Mamedaliyeva, Gaini Anartayeva, Gaukhar Moshkalova, Damilya Salimbayeva, Aidana Tulesheva and Zhamilya Zhankina
J. Clin. Med. 2024, 13(20), 6184; https://doi.org/10.3390/jcm13206184 - 17 Oct 2024
Abstract
Reproductive disorders are common events in modern reproductive medicine, occurring both in spontaneous and assisted pregnancies. Studies on the molecular mechanisms of implantation disorders in thin endometria, including the study of gene transcriptional activities, have shed light on the identification of the potential [...] Read more.
Reproductive disorders are common events in modern reproductive medicine, occurring both in spontaneous and assisted pregnancies. Studies on the molecular mechanisms of implantation disorders in thin endometria, including the study of gene transcriptional activities, have shed light on the identification of the potential biological markers of endometrial receptivity. Background/Objectives: The goal of this study was to reveal the significantly dysregulated selected gene expressions between RIF and RPL patients with thin endometria. Methods: Endometrial samples were collected from RIF patients (n = 20) and RPL patients (n = 19) during the implantation window days (LH + 7—LH + 10) of their natural menstrual cycles. Ten genes were chosen as the target genes regarding their possible relations with the implantation process. The total RNA was purified and reverse-transcribed, and gene expressions were quantified by RT-PCR. Results: The expressions of the IL-15, INFG, and HPRT1 genes were significantly decreased in the RIF patients with thin endometria compared to the RPL patients (log2 fold change = 0.92, p = 0.023 for IL-15; log2 fold change = 1.24, p = 0.046 for INFG; and log2 fold change = 0.579, p = 0.046 for HPRT1). There were no significant differences in the expressions of the CXCL8, CXCL1, MMP10, C4BPA, TNC, VEGFB, and HAND2 genes between the groups. Conclusions: Decreased expressions of the IL-15, INFG, and HPRT1 genes were found in patients with RIF with thin endometria compared to the endometria of women with RPL. This has practical significance for clinicians for the differentiated prescription of immunomodulatory therapy in patients undergoing ART programs. Full article
(This article belongs to the Special Issue Assisted Reproductive Technology: Clinical Advances and Challenges)
Show Figures

Figure 1

Figure 1
<p>Differences in gene expression levels (ΔCt values) in endometrial tissues between RIF and RPL patients: (<b>a</b>) <span class="html-italic">IL15</span> gene; (<b>b</b>) <span class="html-italic">IFNG</span> gene; (<b>c</b>) <span class="html-italic">HPRT1</span> gene.</p>
Full article ">Figure 2
<p>ROC curves for potential markers and combinations of markers in the discrimination of RIF and RPL in women. The ROC curves (red lines) allow for the evaluation of the quality of the binary classifications by displaying the relationship between sensitivity and specificity as the decision threshold is varied: (<b>a</b>) <span class="html-italic">IL15</span>; (<b>b</b>) <span class="html-italic">IFNG</span>; (<b>c</b>) <span class="html-italic">HPRT1</span>; (<b>d</b>) combination of <span class="html-italic">IL15</span> and <span class="html-italic">IFNG</span>; (<b>e</b>) combination of <span class="html-italic">IL15</span> and <span class="html-italic">HPRT1</span>; (<b>f</b>) combination of <span class="html-italic">IFNG</span> and <span class="html-italic">HPRT1</span>; (<b>g</b>) combination of <span class="html-italic">IL15</span>, <span class="html-italic">IFNG</span>, and <span class="html-italic">HPRT1</span>.</p>
Full article ">Figure 2 Cont.
<p>ROC curves for potential markers and combinations of markers in the discrimination of RIF and RPL in women. The ROC curves (red lines) allow for the evaluation of the quality of the binary classifications by displaying the relationship between sensitivity and specificity as the decision threshold is varied: (<b>a</b>) <span class="html-italic">IL15</span>; (<b>b</b>) <span class="html-italic">IFNG</span>; (<b>c</b>) <span class="html-italic">HPRT1</span>; (<b>d</b>) combination of <span class="html-italic">IL15</span> and <span class="html-italic">IFNG</span>; (<b>e</b>) combination of <span class="html-italic">IL15</span> and <span class="html-italic">HPRT1</span>; (<b>f</b>) combination of <span class="html-italic">IFNG</span> and <span class="html-italic">HPRT1</span>; (<b>g</b>) combination of <span class="html-italic">IL15</span>, <span class="html-italic">IFNG</span>, and <span class="html-italic">HPRT1</span>.</p>
Full article ">
18 pages, 901 KiB  
Systematic Review
Characterization of the Joint Microenvironment in Osteoarthritic Joints for In Vitro Strategies for MSC-Based Therapies: A Systematic Review
by Aline Silvestrini da Silva, Fernanda Campos Hertel, Fabrício Luciani Valente, Fabiana Azevedo Voorwald, Andrea Pacheco Batista Borges, Adriano de Paula Sabino, Rodrigo Viana Sepulveda and Emily Correna Carlo Reis
Appl. Biosci. 2024, 3(4), 450-467; https://doi.org/10.3390/applbiosci3040029 - 17 Oct 2024
Abstract
Osteoarthritis is a joint disease that causes pain, stiffness, and reduced joint function because the protective cushioning inside the joints, called cartilage, gradually wears away. This condition is caused by various factors and complex processes in the joint’s environment, involving different types of [...] Read more.
Osteoarthritis is a joint disease that causes pain, stiffness, and reduced joint function because the protective cushioning inside the joints, called cartilage, gradually wears away. This condition is caused by various factors and complex processes in the joint’s environment, involving different types of cells producing factors that can either maintain the joint health or contribute to osteoarthritis. This study aimed to understand the factors influencing both healthy and diseased joints in DDD strategies for the in vitro preconditioning of MSCs. An electronic search in the PubMed, Scopus, and Web of Science databases was carried out using the terms (cartilage OR chondr*) AND (repair OR regeneration OR healing) AND (niche OR microenvironment)) AND (“growth factor” OR GF OR cytokine). Researchers used various methods, including macroscopic examinations, histology, immunohistochemistry, and microCT. Molecules associated with joint inflammation were identified, like macrophage markers, MMP-13, TNF, apoptotic markers, and interleukins. Chondrogenesis-related factors such as aggrecan GAG, collagen type II, and TGF beta family were also identified. This study suggests that balancing certain molecules and ensuring the survival of joint chondrocytes could be crucial in improving the condition of osteoarthritic joints, emphasizing the importance of chondrocyte survival and activity. Future preconditioning methods for MSC- and EV-based therapies can find suitable strategies in the described microenvironments to explore co-culture systems and soluble or extracellular matrix factors. Full article
(This article belongs to the Special Issue Anatomy and Regenerative Medicine: From Methods to Applications)
Show Figures

Graphical abstract

Graphical abstract
Full article ">Figure 1
<p>Flowchart of the systematic review mechanism.</p>
Full article ">Figure 2
<p>Risk of bias in assessing the methodological quality of the 20 articles included in this systematic review.</p>
Full article ">
17 pages, 3970 KiB  
Article
Rats Exposed to Excess Sucrose During a Critical Period Develop Inflammation and Express a Secretory Phenotype of Vascular Smooth Muscle Cells
by Verónica Guarner-Lans, Elizabeth Soria-Castro, Agustina Cano-Martínez, María Esther Rubio-Ruiz, Gabriela Zarco, Elizabeth Carreón-Torres, Oscar Grimaldo, Vicente Castrejón-Téllez and Israel Pérez-Torres
Metabolites 2024, 14(10), 555; https://doi.org/10.3390/metabo14100555 - 17 Oct 2024
Viewed by 108
Abstract
Background: Neonatal rats that receive sucrose during a critical postnatal period (CP, days 12 to 28) develop hypertension by the time they reach adulthood. Inflammation might contribute to changes during this period and could be associated with variations in the vascular smooth muscle [...] Read more.
Background: Neonatal rats that receive sucrose during a critical postnatal period (CP, days 12 to 28) develop hypertension by the time they reach adulthood. Inflammation might contribute to changes during this period and could be associated with variations in the vascular smooth muscle (VSMC) phenotype. Objective: We studied changes in inflammatory pathways that could underlie the expression of the secretory phenotype in the VSMC in the thoracic aorta of rats that received sucrose during CP. Methods: We analyzed histological changes in the aorta and the expression of the COX-2, TLR4, iNOS, eNOS, MMP-2 and -9, and β- and α-actin, the quantities of TNF-α, IL-6, and IL-1β using ELISA, and the levels of fatty acids using gas chromatography. Results: The aortic wall presented disorganization, decellularization, and wavy elastic fibers and an increase in the lumen area. The α- and β-actin expressions were decreased, while COX-2, TLR4, TNF-α, and the activity of IL-6 were increased. Oleic acid was increased in CP in comparison to the control group. Conclusions: There is transient hypertension at the end of the CP that is accompanied by inflammation and a change in the phenotype of VSMC to the secretory phenotype. The inflammatory changes could act as epigenetic signals to determine the development of hypertension when animals reach adulthood. Full article
(This article belongs to the Special Issue Impact of Macronutrients on Metabolism)
Show Figures

Graphical abstract

Graphical abstract
Full article ">Figure 1
<p>Flow chart for the management of the experimental animals.</p>
Full article ">Figure 2
<p>Histological changes of the aortic rings from control rats receiving the normal diet (<b>A</b>,<b>C</b>) and rats exposed to sucrose in the drinking water during the CP (<b>B</b>,<b>D</b>). The pink-stained images were acquired from histological sections stained with HE using a bright field microscope coupled to a camera (see <a href="#sec2-metabolites-14-00555" class="html-sec">Section 2</a>). From the same sections with HE staining, images in a gray tone were acquired with the relief phase color channel for close-ups in which the differences in the structure of the aortic wall between the two groups are distinguished in more detail. In the lower-right gray image, the locations of undulations (black arrow) and aneurysms (white arrows) are indicated. On the right side of the image, the graphs of the comparison of the measurements of the thickness (<b>E</b>) and the total area of the wall (<b>F</b>), as well as the total area of the aortic lumen (<b>G</b>), are shown. Values represent the mean ± standard error (<span class="html-italic">n</span> = 6), * <span class="html-italic">p</span> = 0.01, *** <span class="html-italic">p</span> = 0.001 (<a href="#metabolites-14-00555-f001" class="html-fig">Figure 1</a>).</p>
Full article ">Figure 3
<p>Immunohistochemistry for COX-2, iNOS, and eNOS in control aortas and aortas from rats that received sucrose during the CP. Panels (<b>A</b>,<b>C</b>,<b>E</b>) (COX-2, iNOS, and eNOS in the C group, respectively) and panels (<b>B</b>,<b>D</b>,<b>F</b>) (COX-2, iNOS and eNOS in the CP group, respectively). *** <span class="html-italic">p</span> &lt; 0.001. Values represent the mean ± SE, <span class="html-italic">n</span> = 10, per group. Abbreviations: EF = elastic fibers, E = endothelium, MZ = muscular zone.</p>
Full article ">Figure 4
<p>Immunohistochemistry for TLR4 in control aortas (<b>A</b>) and aortas from rats that received sucrose during the CP (<b>B</b>). * <span class="html-italic">p</span> &lt; 0.03 Values represent the mean ± SE (<span class="html-italic">n</span> = 10 per group). Abbreviations: EF = elastic fibers, E = endothelium, MZ = muscular zone.</p>
Full article ">Figure 5
<p>Concentrations of interleukins IL-1β (<b>A</b>), IL-6 (<b>B</b>), and TNF-α (<b>C</b>) in serum from control and CP rats.. Values represent the mean ± SE (n = 8 per group), * <span class="html-italic">p</span> &lt; 0.05.</p>
Full article ">Figure 6
<p>Changes in the expression of SMA (<b>A</b>) and β-actin (<b>B</b>) in thoracic aortas from C rats and CP rats. Values represent the mean ± SE (<span class="html-italic">n</span> = 8 animals per group). * <span class="html-italic">p</span> &lt; 0.05. Representative western blot images are included in the lower panel. AU refers to arbitrary units, which are determined as the relative density of the band of the protein of interest in relation to the control of charge protein (GAPDH).</p>
Full article ">Figure 7
<p>Changes in the expression of MMP-2 (<b>A</b>) and -9 (<b>B</b>) in thoracic aortas from control rats receiving the normal diet (C) and rats receiving sucrose during the critical period (CP). Values represent the mean ± SE, <span class="html-italic">n</span> = 8 animals per group. * <span class="html-italic">p</span> &lt; 0.05. Representative western blot images are included in the lower panel. AU refers to arbitrary units which are determined as the relative density of the band of the protein of interest in relation to the control of charge protein (GAPDH).</p>
Full article ">Figure 8
<p>Excess consumption of sucrose in drinking water during the CP of vessel development in rats results in transitory hypertension accompanied by a change in the phenotype of VSMCs to the secretory type. There is a decrease in smooth muscle of β- and α-actin, which could act as an epigenetic cue to determine the development of hypertension when the animals reach adulthood. This change in the phenotype might be induced by increased inflammation characterized by increased levels of TNF-α, IL-6, and an increase in the expression of COX-2, TLR-4, and MMP-9, arrow-up = increase, arrow-down= decrease.</p>
Full article ">
30 pages, 5645 KiB  
Article
Exploring the Antidiabetic Potential of Salvia officinalis Using Network Pharmacology, Molecular Docking and ADME/Drug-Likeness Predictions
by Chimaobi J. Ononamadu and Veronique Seidel
Plants 2024, 13(20), 2892; https://doi.org/10.3390/plants13202892 (registering DOI) - 16 Oct 2024
Viewed by 323
Abstract
A combination of network pharmacology, molecular docking and ADME/drug-likeness predictions was employed to explore the potential of Salvia officinalis compounds to interact with key targets involved in the pathogenesis of T2DM. These were predicted using the SwissTargetPrediction, Similarity Ensemble Approach and BindingDB databases. [...] Read more.
A combination of network pharmacology, molecular docking and ADME/drug-likeness predictions was employed to explore the potential of Salvia officinalis compounds to interact with key targets involved in the pathogenesis of T2DM. These were predicted using the SwissTargetPrediction, Similarity Ensemble Approach and BindingDB databases. Networks were constructed using the STRING online tool and Cytoscape (v.3.9.1) software. Gene Ontology (GO), Kyoto Encyclopedia of Genes and Genomes (KEGG) pathways analysis and molecular docking were performed using DAVID, SHINEGO 0.77 and MOE suite, respectively. ADME/drug-likeness parameters were computed using SwissADME and Molsoft L.L.C. The top-ranking targets were CTNNB1, JUN, ESR1, RELA, NR3C1, CREB1, PPARG, PTGS2, CYP3A4, MMP9, UGT2B7, CYP2C19, SLCO1B1, AR, CYP19A1, PARP1, CYP1A2, CYP1B1, HSD17B1, and GSK3B. Apigenin, caffeic acid, oleanolic acid, rosmarinic acid, hispidulin, and salvianolic acid B showed the highest degree of connections in the compound-target network. Gene enrichment analysis identified pathways involved in insulin resistance, adherens junctions, metabolic processes, IL-17, TNF-α, cAMP, relaxin, and AGE-RAGE in diabetic complications. Rosmarinic acid, caffeic acid, and salvianolic acid B showed the most promising interactions with PTGS2, DPP4, AMY1A, PTB1B, PPARG, GSK3B and RELA. Overall, this study enhances understanding of the antidiabetic activity of S. officinalis and provides further insights for future drug discovery purposes. Full article
(This article belongs to the Section Phytochemistry)
Show Figures

Figure 1

Figure 1
<p>Venn diagram depiction of the biological targets common (intersection) to <span class="html-italic">S. officinalis</span> compounds and T2DM.</p>
Full article ">Figure 2
<p>(<b>A</b>) Protein–Protein Interaction (PPI) network and (<b>B</b>) top-ranked 20 (Hub Genes) T2DM targets of <span class="html-italic">S. officinalis</span> compounds. The sizes of the nodes in (<b>A</b>) are proportional to the magnitude of the degree.</p>
Full article ">Figure 3
<p>Compound-target network. The sizes of the nodes are proportional to the magnitude of the degree. Red circles = T2DM targets and blue circles = <span class="html-italic">S. officinalis</span> compounds.</p>
Full article ">Figure 4
<p>Top-ranked 10 (core) <span class="html-italic">S. officinalis</span> compounds predicted to interact with T2DM targets.</p>
Full article ">Figure 5
<p>Target-pathway (TP) network of enriched T2DM-related KEGG pathways for the identified targets. Orange = KEGG Pathways, Green = Targets.</p>
Full article ">Figure 6
<p>GO and KEGG pathway analysis showing the top-ranked enrichments of (<b>A</b>) molecular function and (<b>B</b>) biological process for the common targets/genes associated with <span class="html-italic">S. officinalis</span> compounds and T2DM.</p>
Full article ">Figure 7
<p>GO and KEGG pathway analysis showing the top-ranked enrichments of (<b>A</b>) cellular compartment (<b>B</b>) KEGG pathways for the common targets/genes associated with <span class="html-italic">S. officinalis</span> compounds and T2DM.</p>
Full article ">Figure 8
<p>Docked pose of rosmarinic acid with (<b>A</b>) PGTS2 (<b>B</b>) AMY1A and (<b>C</b>) DDP4.</p>
Full article ">Figure 9
<p>Docked pose of salvianolic acid B acid with (<b>A</b>) PPARG, (<b>B</b>) GSK3B and (<b>C</b>) RELA.</p>
Full article ">Figure 10
<p>Docked pose of caffeic acid with PTP1B.</p>
Full article ">
20 pages, 5554 KiB  
Article
Syn-Propanethial S-Oxide as an Available Natural Building Block for the Preparation of Nitro-Functionalized, Sulfur-Containing Five-Membered Heterocycles: An MEDT Study
by Mikołaj Sadowski, Ewa Dresler, Karolina Zawadzińska, Aneta Wróblewska and Radomir Jasiński
Molecules 2024, 29(20), 4892; https://doi.org/10.3390/molecules29204892 (registering DOI) - 15 Oct 2024
Viewed by 392
Abstract
The regio- and stereoselectivity and the molecular mechanisms of the [3 + 2] cycloaddition reactions between Syn-propanethial S-oxide and selected conjugated nitroalkenes were explored theoretically in the framework of the Molecular Electron Density Theory. It was found that cycloadditions with the participation [...] Read more.
The regio- and stereoselectivity and the molecular mechanisms of the [3 + 2] cycloaddition reactions between Syn-propanethial S-oxide and selected conjugated nitroalkenes were explored theoretically in the framework of the Molecular Electron Density Theory. It was found that cycloadditions with the participation of nitroethene as well as its methyl- and chloro-substituted analogs can be realized via a single-step mechanism. On the other hand, [3 + 2] cycloaddition reactions between Syn-propanethial S-oxide and 1,1-dinitroethene can proceed according to a stepwise mechanism with a zwitterionic intermediate. Finally, we evaluated the affinity of model reaction products for several target proteins: cytochrome P450 14α-sterol demethylase CYP51 (RSCB Database PDB ID: 1EA1), metalloproteinase gelatinase B (MMP-9; PDB ID: 4XCT), and the inhibitors of cyclooxygenase COX-1 (PDB:3KK6) and COX-2 (PDB:5KIR). Full article
(This article belongs to the Special Issue Heterocyclic Compounds: Synthesis, Application and Theoretical Study)
Show Figures

Figure 1

Figure 1
<p>Topology of molecule <b>1</b> ELF, as rendered at an isovalue of 0.8. Core basins are given in magenta, protonated basins in cyan, disynaptic basins in green, and monosynaptic basins in red.</p>
Full article ">Figure 2
<p>Positions and populations of ELF attractors of unprotonated basins in molecule <b>1</b>.</p>
Full article ">Figure 3
<p>Views of critical structures for the [3 + 2] cycloaddition between <span class="html-italic">Syn</span>-propanethial S-oxide (<b>1</b>) and nitroethene <b>2a</b> in toluene solution according to wB97XD/6-31G(d) (PCM) calculations.</p>
Full article ">Figure 4
<p>Visualization of ELF basins of system <b>IA</b> of reaction <b>1</b> + <b>2d</b> at an ELF isovalue of 0.8, as computed by wB97XD/6-31G(d) in toluene (PCM). Core basins are given in magenta, protonated basins in cyan, disynaptic basins in green, and monosynaptic basins in red.</p>
Full article ">Figure 5
<p>Populations of ELF attractors in intermediate <b>IA</b> of reaction <b>1</b> + <b>2d</b> and populations of significant basins, as computed by wB97XD/6-31G(d)) in toluene (PCM).</p>
Full article ">Figure 6
<p>The natural charges of the atoms in intermediate <b>IA</b> of reaction <b>1</b> + <b>2d</b> (WB97XD/6-31G(d) in toluene (PCM)). Charges greater than 0.2 are given in red, those less than −0.2 are given in blue, and those between or equal to −0.2 and 0.2 are given in black.</p>
Full article ">Figure 7
<p>Views of critical structures for path A of the [3 + 2] cycloaddition between <span class="html-italic">Syn</span>-propanethial S-oxide (<b>1</b>) and 1,1-dinitroethene <b>2d</b> in toluene solution, according to wB97XD/6-31G(d) (PCM) calculations.</p>
Full article ">Figure 8
<p>Visualization of hydrogen bond formation between ligand and protein for (<b>A</b>) <b>4a</b>, (<b>B</b>) <b>3a</b>, and COX-2 (PDB:5kir).</p>
Full article ">Figure 9
<p>Visualization of hydrogen bond formation between ligand and protein for (<b>A</b>) <b>6a</b>, (<b>B</b>) <b>5a</b>, and CYP51 (PDB:1ea1).</p>
Full article ">Figure 10
<p>Visualization of (<b>A</b>) electrostatic and (<b>B</b>) hydrophobic–hydrophilic potential of CYP51 (PDB:1EA1) with <b>6a</b> in the pocket.</p>
Full article ">Scheme 1
<p>Examples of bioactive 1,3-oxathiolane derivatives.</p>
Full article ">Scheme 2
<p>Single-step and stepwise mechanisms of [3 + 2] cycloaddition reactions.</p>
Full article ">Scheme 3
<p>Theoretically possible regio- and stereoisomeric paths of the [3 + 2] cycloaddition between <span class="html-italic">Syn</span>-propanethial S-oxide (<b>1</b>) and 1-R-1-nitroethenes (<b>2a</b>–<b>d</b>).</p>
Full article ">Scheme 4
<p>Geometry of the CSO group of oxide <b>1</b> calculated at the ground state.</p>
Full article ">Scheme 5
<p>Natural charges of atoms (in electrons) in the S-oxide Xet, as computed at the ground state in the gaseous phase (B3LYP/6-31G(d)); charges &gt; 0.2 e are given in red, and charges &lt; −0.2 are given in blue.</p>
Full article ">Scheme 6
<p>Parr function values for radical anion and radical cation of molecule <b>1</b>.</p>
Full article ">
11 pages, 1336 KiB  
Article
Therapeutic Potential of Intermittent Hypoxia in Atrial Fibrillation
by Hyewon Park, Bokyeong Park, Kyu-sung Kim, Young Hoon Son, Sung Jin Park, Kichang Lee, Hyelim Park and Junbeom Park
Int. J. Mol. Sci. 2024, 25(20), 11085; https://doi.org/10.3390/ijms252011085 (registering DOI) - 15 Oct 2024
Viewed by 203
Abstract
Intermittent hypoxia (IH) has been extensively studied in recent years, demonstrating adverse and beneficial effects on several physiological systems. However, the precise mechanism underlying its cardiac effects on the heart remains unclear. This study aims to explore the effect of treatment on atrial [...] Read more.
Intermittent hypoxia (IH) has been extensively studied in recent years, demonstrating adverse and beneficial effects on several physiological systems. However, the precise mechanism underlying its cardiac effects on the heart remains unclear. This study aims to explore the effect of treatment on atrial fibrillation under IH conditions, providing data that can potentially be used in the treatment of heart disease. An atrial fibrillation (AF) model was induced by injecting monocrotaline (MCT, 60 mg/kg) into rats. The study included 32 rats divided into four groups: Control, Control + IH, AF, and AF + IH. We evaluated molecular changes associated with AF using ELISA and Western blot and performed electrophysiological experiments to evaluate AF. Arrhythmia-related calcium and fibrosis markers were investigated. Phosphorylation levels of CaMKII, Phospholamban, and RyR2 all increased in the AF group but decreased in the IH-exposed group. Additionally, fibrosis marker expressions such as SMA, MMP2, MMP9, and TGF-β increased in the AF group but were significantly downregulated with IH treatment. Connexin 43 and AQP4 expression were restored in the IH-treated group. These findings suggest that IH may prevent AF by downregulating the expression of calcium-handling proteins and fibrosis-associated proteins in an AF-induced rat model. Full article
(This article belongs to the Special Issue Therapeutic Target in Cardiovascular Disease)
Show Figures

Figure 1

Figure 1
<p>Effect of intermittent hypoxia (IH) exposure on atrial fibrillation (AF) in rats. (<b>A</b>) Schematic representation of the design of the animal study and intermittent hypoxia chamber image. (<b>B</b>) Representative EKG of a rat at baseline. (<b>C</b>) Percentage of lethal arrhythmias observed. (<b>D</b>) AF incidence in each group. (<b>E</b>) Representative electrocardiogram recording. The underline indicates AF. Control <span class="html-italic">n</span> = 8; Control + IH <span class="html-italic">n</span> = 8; AF <span class="html-italic">n</span> = 8; AF + IH <span class="html-italic">n</span> = 8. Statistical analyses: χ2 test for incidence, Student <span class="html-italic">t</span>-test, or one-way ANOVA with Bonferroni’s post hoc test for other comparisons.</p>
Full article ">Figure 2
<p>IH effect on HRV. (<b>A</b>,<b>B</b>) ELISA analysis of serum levels of Norepinephrine and Epinephrine levels across four groups. (<b>C</b>) Representative power spectra indicating autonomic controls in cardiac regulation across 5 min segments. (<b>D</b>,<b>E</b>) Heart rate variability-low frequency (LF) to high frequency (HF) power ratio, denoting cardiac sympathetic activity. Data expressed as mean ± SEM. The data were analyzed using a one-way analysis of variance. * <span class="html-italic">p</span> &lt; 0.01 vs. CTL.</p>
Full article ">Figure 3
<p>Intermittent hypoxia (IH) suppresses inflammation and fibrosis in atrial fibrillation (AF)-induced rat atrial tissues. (<b>A</b>–<b>D</b>) Levels of Interleukin-6 (IL-6), Tumor necrosis factor-alpha (TNF-α), Transforming Growth Factor-beta (TGF-B), MMP2. Control <span class="html-italic">n</span> = 4; Control + IH <span class="html-italic">n</span> = 4; AF <span class="html-italic">n</span> = 4; AF + IH <span class="html-italic">n</span> = 4. * <span class="html-italic">p</span> &lt; 0.01 vs. control group or AF + IH group. Data expressed as mean ± SEM. The data were analyzed using a one-way analysis of variance. * <span class="html-italic">p</span> &lt; 0.01 vs. CTL.</p>
Full article ">Figure 4
<p>Comparison of Ca<sup>2+</sup>/calmodulin-dependent protein kinase II (CaMKII) and p-CaMKII expression and consequent phosphorylation of downstream targets. Each figure panel shows a (<b>A</b>) representative Western blot. (<b>B</b>–<b>E</b>) Relative expression of phosphorylated and total CaMKII (<b>B</b>), Phospholamban (PLB) (<b>C</b>), Ryanodine receptor (RyR2) (<b>D</b>,<b>E</b>). No significant difference was observed in CaMKII, PLB, RyR2 expression, or auto-phosphorylation in the Control and Control + IH groups. Conversely, these factors were increased in the AF group but recovered in the AF + IH group. Data are presented as mean ± SEM of the ratio of normalized phosphorylated protein to total protein extracted from rat hearts (n = 3–4 each). The data were analyzed using a one-way analysis of variance. * <span class="html-italic">p</span> &lt; 0.01 vs. CTL. # <span class="html-italic">p</span> &lt; 0.01 vs. AF.</p>
Full article ">Figure 5
<p>Effect of intermittent hypoxia on fibrosis and water channel protein marker levels in hearts. (<b>A</b>) Western blot analysis depicting fibrosis markers α-SMA, MMP2, MMP9, TGF-β, Connexin 43, AQP4, and endogenous control GAPDH. (<b>B</b>–<b>G</b>) Quantification graphs for each group. Statistical data represent results from four rats in each group. The experimental results are expressed as means ± standard deviations. The data were analyzed using a one-way analysis of variance. * <span class="html-italic">p</span> &lt; 0.01 vs. CTL. # <span class="html-italic">p</span> &lt; 0.01 vs. AF.</p>
Full article ">
12 pages, 5458 KiB  
Article
Anti-Photoaging Effects of Antioxidant Peptide from Seahorse (Hippocampus abdominalis) in In Vivo and In Vitro Models
by Fengqi Yang, Yang Yang, Dandan Xiao, Poongho Kim, Jihee Lee, You-Jin Jeon and Lei Wang
Mar. Drugs 2024, 22(10), 471; https://doi.org/10.3390/md22100471 - 14 Oct 2024
Viewed by 369
Abstract
Overexposure to ultraviolet (UV) radiation can lead to photoaging, which contributes to skin damage. The objective of this study was to evaluate the effects of an antioxidant peptide (SHP2) purified from seahorse (Hippocampus abdominalis) alcalase hydrolysate on UVB-irradiated skin damage in [...] Read more.
Overexposure to ultraviolet (UV) radiation can lead to photoaging, which contributes to skin damage. The objective of this study was to evaluate the effects of an antioxidant peptide (SHP2) purified from seahorse (Hippocampus abdominalis) alcalase hydrolysate on UVB-irradiated skin damage in human keratinocyte (HaCaT) and human dermal fibroblast (HDF) cells and a zebrafish model. The data revealed that SHP2 significantly enhanced cell viability by attenuating apoptosis through the reduction of intracellular reactive oxygen species (ROS) levels in UVB-stimulated HaCaT cells. Moreover, SHP2 effectively inhibited ROS, improved collagen synthesis, and suppressed the secretion of matrix metalloproteinases (MMPs) in UVB-irradiated HDF cells. SHP2 restored the protein levels of HO-1, Nrf2, and SOD, while decreasing Keap1 expression in UVB-treated HDF, indicating stimulation of the Keap1/Nrf2/HO-1 signaling pathway. Furthermore, an in vivo study conducted in zebrafish confirmed that SHP2 inhibited photoaging by reducing cell death through the suppression of ROS generation and lipid peroxidation. Particularly, 200 µg/mL of SHP2 exerted a remarkable anti-photoaging effect on both in vitro and in vivo models. These results demonstrate that SHP2 possesses antioxidant properties and regulates skin photoaging activities, suggesting that SHP2 may have the potential for use in the development of cosmetic products. Full article
(This article belongs to the Special Issue Marine Anti-inflammatory and Antioxidant Agents 4.0)
Show Figures

Figure 1

Figure 1
<p>SHP2 protects HaCaT cells against UVB-induced photodamage. (<b>A</b>) Cytotoxicity of SHP2 on HaCaT cells; (<b>B</b>) intracellular ROS scavenging effect of SHP2 in UVB-irradiated HaCaT cells; (<b>C</b>) protective effect of SHP2 against UVB-induced cell death in HaCaT cells. The experiments were conducted in triplicate and the data are expressed as the mean ± SE. * <span class="html-italic">p</span> &lt; 0.05, ** <span class="html-italic">p</span> &lt; 0.01 as compared to the UVB-irradiated group and ## <span class="html-italic">p</span> &lt; 0.01 as compared to the control group.</p>
Full article ">Figure 2
<p>SHP2 protects HaCaT cells against UVB-induced apoptosis. (<b>A</b>) Morphology of the normal cells; (<b>B</b>) morphology of the cells irradiated by UVB; (<b>C</b>) morphology of the cells treated with 50 μg/mL SHP2 and irradiated by UVB; (<b>D</b>) morphology of the cells treated with 100 μg/mL SHP2 and irradiated by UVB; (<b>E</b>) morphology of the cells treated with 200 μg/mL SHP2 and irradiated by UVB; (<b>F</b>) quantification of apoptotic cells. The apoptotic body formation was evaluated by Hoechst 33342 staining assay. White arrows indicate apoptotic bodies. **** <span class="html-italic">p</span> &lt; 0.0001 as compared to the UVB-irradiated group and #### <span class="html-italic">p</span> &lt; 0.0001 as compared to the control group.</p>
Full article ">Figure 3
<p>SHP2 protects HDF cells against UVB-induced damage. (<b>A</b>) Cytotoxicity of SHP2 on HDF cells; (<b>B</b>) intracellular ROS scavenging effect of SHP2 in UVB-irradiated HDF cells; (<b>C</b>) protective effect of SHP2 against UVB-induced cell death in HDF cells. The experiments were conducted in triplicate and the data are expressed as the mean ± SE. * <span class="html-italic">p</span> &lt; 0.05, ** <span class="html-italic">p</span> &lt; 0.01 as compared to the UVB-irradiated group and ## <span class="html-italic">p</span> &lt; 0.01 as compared to the control group.</p>
Full article ">Figure 4
<p>SHP2 improves collagen synthesis and inhibits MMPs’ expression in UVB-irradiated HDF cells. (<b>A</b>) Collagen contents in UVB-irradiated HDF cells; (<b>B</b>) MMP-1 expression levels in UVB-irradiated HDF cells; (<b>C</b>) MMP-2 expression levels in UVB-irradiated HDF cells; (<b>D</b>) MMP-8 expression levels in UVB-irradiated HDF cells; (<b>E</b>) MMP-9 expression levels in UVB-irradiated HDF cells; (<b>F</b>) MMP-13 expression levels in UVB-irradiated HDF cells. The experiments were conducted in triplicate and the data are expressed as the mean ± SE. * <span class="html-italic">p</span> &lt; 0.05, ** <span class="html-italic">p</span> &lt; 0.01 as compared to the UVB-irradiated group and ## <span class="html-italic">p</span> &lt; 0.01 as compared to the control group.</p>
Full article ">Figure 5
<p>Protective effects of SHP2 on Keap1/Nrf2/HO-1 signaling pathway in UVB-irradiated HDF cells. (<b>A</b>) Western blot bands; (<b>B</b>) HO-1 protein expression; (<b>C</b>) Keap1 protein expression; (<b>D</b>) Nrf2 protein expression; (<b>E</b>) SOD protein expression. The experiments were conducted in triplicate and the data are expressed as the mean ± SE. **** <span class="html-italic">p</span> &lt; 0.0001 as compared to the UVB-irradiated group and ## <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 as compared to the control group.</p>
Full article ">Figure 6
<p>SHP2 protects zebrafish against UVB-induced oxidative stress in the zebrafish. (<b>A</b>) ROS generation in zebrafish; (<b>B</b>) cell death level in zebrafish; (<b>C</b>) lipid peroxidation level in zebrafish. ROS, cell death, and lipid peroxidation levels were measured by Image J software (v.1.8.0). The data are expressed as means ± SE. ** <span class="html-italic">p</span> &lt; 0.01 as compared to the UVB-treated group and ## <span class="html-italic">p</span> &lt; 0.01 as compared to the control group.</p>
Full article ">
13 pages, 4334 KiB  
Article
Limited Alleviation of Lysosomal Acid Lipase Deficiency by Deletion of Matrix Metalloproteinase 12
by Martin Buerger, Melina Amor, Alena Akhmetshina, Valentina Bianco, Bianca Perfler, Armin Zebisch, Thomas Weichhart and Dagmar Kratky
Int. J. Mol. Sci. 2024, 25(20), 11001; https://doi.org/10.3390/ijms252011001 - 13 Oct 2024
Viewed by 400
Abstract
Lysosomal acid lipase (LAL) is the only known enzyme that degrades cholesteryl esters and triglycerides at an acidic pH. In LAL deficiency (LAL-D), dysregulated expression of matrix metalloproteinase 12 (MMP-12) has been described. The overexpression of MMP-12 in myeloid lineage cells causes an [...] Read more.
Lysosomal acid lipase (LAL) is the only known enzyme that degrades cholesteryl esters and triglycerides at an acidic pH. In LAL deficiency (LAL-D), dysregulated expression of matrix metalloproteinase 12 (MMP-12) has been described. The overexpression of MMP-12 in myeloid lineage cells causes an immune cell dysfunction resembling that of Lal knockout (Lal KO) mice. Both models develop progressive lymphocyte dysfunction and expansion of myeloid-derived suppressor (CD11b+ Gr-1+) cells. To study whether MMP-12 might be a detrimental contributor to the pathology of LAL-D, we have generated Lal/Mmp12 double knockout (DKO) mice. The phenotype of Lal/Mmp12 DKO mice closely resembled that of Lal KO mice, while the weight and morphology of the thymus were improved in Lal/Mmp12 DKO mice. Cytological examination of blood smears showed a mildly reversed lymphoid-to-myeloid shift in DKO mice. Despite significant decreases in CD11b+ Ly6G+ cells in the peripheral blood, bone marrow, and spleen of Lal/Mmp12 DKO mice, the hematopoietic bone marrow progenitor compartment and markers for neutrophil chemotaxis were unchanged. Since the overall severity of LAL-D remains unaffected by the deletion of Mmp12, we conclude that MMP-12 does not represent a viable target for treating the inflammatory pathology in LAL-D. Full article
(This article belongs to the Special Issue Peroxisome and Lysosome in Health and Disease)
Show Figures

Figure 1

Figure 1
<p>Phenotype of <span class="html-italic">Lal/Mmp12</span> DKO mice. Male 30-week-old chow-diet-fed WT, <span class="html-italic">Lal</span> KO, and <span class="html-italic">Lal/Mmp12</span> DKO mice were fasted for 6 h. (<b>A</b>) Body weight. (<b>B</b>) Weight of subcutaneous white adipose tissue (sWAT) and brown adipose tissue (BAT), normalized to brain weight; WT mice were arbitrarily set to 1. (<b>C</b>) Mean diameter of adipocytes from sWAT paraffin sections. (<b>D</b>) Liver weight normalized to brain weight; WT mice were arbitrarily set to 1. (<b>E</b>) Hepatic mRNA expression of macrophage markers (<span class="html-italic">Emr1</span> and <span class="html-italic">Cd68</span>) relative to <span class="html-italic">Hprt</span> expression. (<b>F</b>) Hepatic lipid parameters. (<b>G</b>) Representative images of H&amp;E-stained liver sections. Scale bars, 100 µm. Asterisks indicate the granuloma-like accumulation of lipid-laden macrophages. Hepatic mRNA expression of (<b>H</b>) <span class="html-italic">Mmp12</span> and (<b>I</b>) inflammation/chemotactic markers (<span class="html-italic">Tnf</span>, <span class="html-italic">Il1b</span>, and <span class="html-italic">Ccl2</span>) relative to <span class="html-italic">Hprt</span> expression. Plasma concentrations of (<b>J</b>) aspartate aminotransferase (AST), alanine aminotransferase (ALT), and (<b>K</b>) serum amyloid A (SAA). Data are shown as means (n = 4–8) + SD. <sup>###</sup> <span class="html-italic">p</span> ≤ 0.001 for the comparison between <span class="html-italic">Lal</span> KO and <span class="html-italic">Lal/Mmp12</span> DKO mice.</p>
Full article ">Figure 2
<p>Minor changes in total blood cell counts but ameliorated thymus weight and morphology in <span class="html-italic">Lal/Mmp12</span> DKO mice. Complete blood counts in peripheral blood collected from 28–32-week-old chow-diet-fed male WT, <span class="html-italic">Mmp12</span>, <span class="html-italic">Lal</span> KO, and <span class="html-italic">Lal/Mmp12</span> DKO mice for (<b>A</b>) total leukocytes and (<b>B</b>) thrombocytes. (<b>C</b>) Hemoglobin concentrations. (<b>D</b>) The frequency of myeloid and lymphoid cells determined by cytological examination of Giemsa-stained blood smears. (<b>E</b>) Thymus weight normalized to brain weight; WT mice were arbitrarily set to 1. (<b>F</b>) Representative images of H&amp;E-stained thymus sections. Scale bars, 100 µm. Red arrows indicate the accumulation of lipid-laden macrophages. Data are shown as means (n = 4–9) + SD. <sup>†</sup> <span class="html-italic">p</span> &lt; 0.05 for the comparison between WT and <span class="html-italic">Mmp12</span> KO mice; <sup>#</sup> <span class="html-italic">p</span> &lt; 0.05 and <sup>##</sup> <span class="html-italic">p</span> ≤ 0.01 for the comparison between <span class="html-italic">Lal</span> KO and <span class="html-italic">Lal/Mmp12</span> DKO mice.</p>
Full article ">Figure 3
<p>Reduced CD11b+ Ly6G+ counts in blood, bone marrow, and spleen, but no changes in the hematopoietic progenitor compartment of <span class="html-italic">Lal/Mmp12</span> DKO mice. (<b>A</b>) Representative flow cytometric plots and their quantification, depicting a decreased CD11b+ Ly6G+ fraction in peripheral blood, bone marrow, and spleen of <span class="html-italic">Lal/Mmp12</span> DKO compared to <span class="html-italic">Lal</span> KO mice. (<b>B</b>) Spleen weight normalized to brain weight; WT mice were arbitrarily set to 1. (<b>C</b>) The content of Lin- Sca1+ c-Kit+ (LSK) and Lin- c-Kit+ (LK) and (<b>D</b>) progenitor cells [common myeloid progenitors (CMPs), granulocyte–macrophage progenitors (GMPs), megakaryocyte–erythrocyte progenitor cells (MEPs)] shown as % of 7-AAD lineage-negative bone marrow cells. Data are shown as means (n = 4–10) + SD. ** <span class="html-italic">p</span> &lt; 0.01 and *** <span class="html-italic">p</span> ≤ 0.001 for the comparison between WT and Lal KO mice; <sup>#</sup> <span class="html-italic">p</span> ≤ 0.05 and <sup>##</sup> <span class="html-italic">p</span> ≤ 0.01 for the comparison between <span class="html-italic">Lal</span> KO and <span class="html-italic">Lal/Mmp12</span> DKO mice.</p>
Full article ">Figure 4
<p>Altered spleen morphology but unchanged neutrophil chemotaxis and lymphocyte markers in <span class="html-italic">Lal/Mmp12</span> DKO mice. (<b>A</b>) Representative images of H&amp;E-stained spleen sections. Green arrows indicate white pulps; red arrows indicate red pulps. Scale bars, 100 µm. Gene expression of (<b>B</b>) the macrophage marker <span class="html-italic">Cd68</span>, (<b>C</b>) <span class="html-italic">Mmp12</span>, (<b>D</b>,<b>E</b>) neutrophil markers (<span class="html-italic">Ly6g</span> and <span class="html-italic">Elane</span>), (<b>F</b>–<b>H</b>) chemokine receptors (<span class="html-italic">Cxcr1</span>, <span class="html-italic">Cxcr2</span>, and <span class="html-italic">Cxcr4</span>), (<b>I</b>–<b>K</b>) chemokine ligands (<span class="html-italic">Cxcl1</span>, <span class="html-italic">Cxcl2</span>, and <span class="html-italic">Cxcl5</span>), and (<b>L</b>,<b>M</b>) lymphocyte markers (<span class="html-italic">Cd3e</span> and <span class="html-italic">B220</span>) relative to <span class="html-italic">Ppia</span> expression. (<b>N</b>) Immunoblot analysis showing the absence of <span class="html-italic">CD3e</span> expression in the spleen of <span class="html-italic">Lal</span> KO (LKO) and <span class="html-italic">Lal/Mmp12</span> DKO mice. Calnexin was used as loading control. Data are shown as means (n = 3–5) + SD. * <span class="html-italic">p</span> ≤ 0.05, ** <span class="html-italic">p</span> ≤ 0.01, and *** <span class="html-italic">p</span> ≤ 0.001 for the comparison between WT and <span class="html-italic">Lal</span> KO mice; <sup>###</sup> <span class="html-italic">p</span> ≤ 0.001 for the comparison between <span class="html-italic">Lal</span> KO and <span class="html-italic">Lal/Mmp12</span> DKO mice.</p>
Full article ">
12 pages, 2901 KiB  
Article
Development of a Bioactive Titanium Surface via Alkalinization and Naringenin Coating for Peri-Implant Repair: In Vitro Study
by Isabela Massaro Ribeiro, Lais Medeiros Cardoso, Taisa Nogueira Pansani, Ana Carolina Chagas, Carlos Alberto de Souza Costa and Fernanda Gonçalves Basso
Coatings 2024, 14(10), 1303; https://doi.org/10.3390/coatings14101303 - 12 Oct 2024
Viewed by 292
Abstract
This study assessed the effects of titanium (Ti) surface modification with sodium hydroxide (NaOH) associated or not with Naringenin (NA) citrus flavonoid-coating on osteoblastic-like cells (Ob) metabolism. Ti discs were submitted to alkalinization by NaOH solution (5 M, 60 °C) for 24 h; [...] Read more.
This study assessed the effects of titanium (Ti) surface modification with sodium hydroxide (NaOH) associated or not with Naringenin (NA) citrus flavonoid-coating on osteoblastic-like cells (Ob) metabolism. Ti discs were submitted to alkalinization by NaOH solution (5 M, 60 °C) for 24 h; then, the discs were impregnated or not with 100 µg/mL of NA and dried for 1 h at room temperature. The chemical composition, surface topography, and NA release were evaluated. For the biological assays, the discs were placed on 24-well cell culture plates and Ob (Saos-2; ATCC HTB-85) was seeded onto the discs. After different periods, cell adhesion and viability, alkaline phosphatase activity (ALP), and mineralized nodules deposition (MND) were assessed. In addition, cells stimulated with tumor necrosis factor-alpha (TNF-α) were submitted to matrix metalloproteinase (MMP)-2 synthesis and ALP gene expression assessment. Since data presented normal distribution and homogeneity (Shapiro-Wilk; Levene), Student’s t-test or one-way ANOVA/post-hoc tests were selected for data analysis (α = 0.05). Higher roughness was observed on Ti discs submitted to NaOH treatment, while the chemical and NA release evaluations indicated the successful adsorption of NA to alkali-treated Ti surface. Higher cell adhesion, cell viability (after 7 days of culture), ALP activity, and MND were observed on Ti NaOH coated with NA compared to the control group (Ti NaOH) (p < 0.05). Moreover, NA coating also promoted decreased MMP-2 synthesis and increased ALP gene expression in the presence of the inflammatory stimulus TNF-α (p < 0.05). The modification of Ti disks with NaOH associated with NA-coating enhanced bone cell metabolism, suggesting that this type of surface modification has a promising potential to accelerate bone repair and formation around dental implants. Full article
(This article belongs to the Special Issue Synthesis and Applications of Bioactive Coatings)
Show Figures

Figure 1

Figure 1
<p>Qualitative analysis of surface topography of Ti and NaOH-treated Ti disks by scanning electron microscopy (SEM) and ImageJ software and quantitative analysis of surface roughness (Ra) for both surfaces.</p>
Full article ">Figure 2
<p>(<b>A</b>) Displays Fourier Transform Infrared (FTIR) spectra for Naringenin (NA) powder and Titanium (Ti) discs submitted to alkali treatment (NaOH) and coated or not with NA. (<b>B</b>) NA release from the alkalinized Ti-discs coated with NA over time after incubation at 37 °C in ultrapure water (<span class="html-italic">n</span> = 6). Data points are mean values and error bars denote 95% confidence intervals (α = 5%).</p>
Full article ">Figure 3
<p>(<b>A</b>) Cell viability (% of control Ti NaOH group) after 24 h, 48 h, and 7 days of cell culture on modified Ti surfaces (<span class="html-italic">n</span> = 6). Columns represent mean values and error bars represent standard deviations. Statistical notation is indicated by * (Student <span class="html-italic">t</span>-test, <span class="html-italic">p</span> &lt; 0.05). (<b>B</b>) Representative images of fluorescence microscopy for cell adhesion and spreading on modified Ti surfaces after 24 h and 48 h of culture (<span class="html-italic">n</span> = 4). Cell nuclei are stained in blue (Hoechst) and actin filaments are stained in red (ActinRed). Scale bar: 250 μm—10× magnification; 125 μm—20× magnification).</p>
Full article ">Figure 4
<p>(<b>A</b>) ALP activity and (<b>B</b>) mineralized nodule formation (% of control Ti NaOH group) after 7 days of cell culture on modified Ti surfaces (<span class="html-italic">n</span> = 6). Columns represent mean values and error bars represent standard deviations. Statistical notation is indicated by * (Student <span class="html-italic">t</span>-test, <span class="html-italic">p</span> &lt; 0.05).</p>
Full article ">Figure 5
<p>(<b>A</b>) Synthesis of MMP-2 (% of control Ti NaOH group) and (<b>B</b>) ALP gene expression (mRNA fold change) by osteoblasts exposed (+TNF) or not (−TNF) to tumor necrosis factor alpha (TNF-α) (<span class="html-italic">n</span> = 6). Columns represent mean values and error bars represent standard deviations. Groups identified by different symbols (*, ** or ***) are statistically different from each other (one-way ANOVA, followed by Tukey’s post-hoc test, <span class="html-italic">p</span> &lt; 0.05).</p>
Full article ">
12 pages, 4398 KiB  
Article
Impacts of Hyperglycemia on Epigenetic Modifications in Human Gingival Fibroblasts and Gingiva in Diabetic Rats
by Kento Kojima, Nobuhisa Nakamura, Airi Hayashi, Shun Kondo, Megumi Miyabe, Takeshi Kikuchi, Noritaka Sawada, Tomokazu Saiki, Tomomi Minato, Reina Ozaki, Sachiko Sasajima, Akio Mitani and Keiko Naruse
Int. J. Mol. Sci. 2024, 25(20), 10979; https://doi.org/10.3390/ijms252010979 - 12 Oct 2024
Viewed by 321
Abstract
Periodontal disease is considered one of the diabetic complications with high morbidity and severity. Recent studies demonstrated the involvement of the epigenome on diabetic complications. Histone modifications change chromatin architecture and gene activation. Histone modifications have been reported to alter chromatin structure and [...] Read more.
Periodontal disease is considered one of the diabetic complications with high morbidity and severity. Recent studies demonstrated the involvement of the epigenome on diabetic complications. Histone modifications change chromatin architecture and gene activation. Histone modifications have been reported to alter chromatin structure and regulate gene transcription. In this study, we investigated the impacts of H3 lysine 4 trimethylation (H3K4me3) and specific histone methyltransferases of H3K4 methylation, su(var)3-9, enhancer-of-zeste, and trithorax domain 1A (SETD1A) on periodontal tissue affected by the diabetic condition. We observed the increase in H3K4me3 and SETD1A in gingival tissue of diabetic rats compared with the normal rats. Cultured human fibroblasts (hGFs) confirmed a high glucose-induced increase in H3K4me3 and SETD1A. We further demonstrated that high glucose increased the gene expression of matrix metalloproteinase (MMP) 1 and MMP13, which were canceled by sinefungin, an SETD1A inhibitor. Our investigation suggests that diabetes triggers histone modifications in the gingival tissue, resulting in gingival inflammation. Histone modifications may play crucial roles in the development of periodontal disease in diabetes. Full article
Show Figures

Figure 1

Figure 1
<p>Comparison of alveolar bone between control group and diabetic group. (<b>A</b>) Three-dimensional reconstructed image of alveolar bone using microcomputed tomography (µCT). Scale = 3000 µm. (<b>B</b>) Alveolar bone was assessed by µCT of the distance from the cement–enamel junction (CEJ) to the alveolar bone crest (ABC). The results are presented as the mean ± S.E.M. (<span class="html-italic">n</span> = 6).</p>
Full article ">Figure 2
<p>Expression of H3K4me3 and SETD1A in periodontal tissue in the control and diabetes groups. The expression of local H3K4me3 and SETD1A in gingival tissues was visualized by immunohistochemical staining. The immunofluorescence staining of nuclei stained with 4′,6-diamidino-2-phenylindole (DAPI, blue) and H3K4me3 (red), SETD1A (red) was performed. (<b>A</b>) Low-power images of H3K4me3 in the periapical maxillary second molar gingiva of control and diabetic rats. Scale bar = 50 µm (×20). (<b>B</b>) High-magnification images of H3K4me3. Scale bar = 50 µm (×40). (<b>C</b>) Quantitative measurement of H3K4me3 expression per cell was performed using Image J (<span class="html-italic">n</span> = 4). (<b>D</b>) Low-power image of SETD1A in the periapical maxillary second molar gingiva of normal and diabetes rats. Scale bar = 50 µm (×20). (<b>E</b>) High-magnification images of SETD1A. Scale bar = 50 µm (×40). (<b>F</b>) Quantitative measurement of SETD1A expression per cell was performed using Image J. * <span class="html-italic">p</span> &lt; 0.05.</p>
Full article ">Figure 3
<p>Immunohistochemical staining revealed changes in the expression of histone-methylated proteins (H3K4me3) and histone methyltransferases (SETD1A) owing to high glucose levels. The immunofluorescence staining of the nuclei with DAPI (blue), H3K4me3 (red), and SETD1A (red) was performed. (<b>A</b>) H3K4me3 stained images of control, high glucose, and high glucose with sinefungin. Scale bar = 50 µm (×40). (<b>B</b>) ImageJ software (<span class="html-italic">n</span> = 6) was used to quantitatively measure the H3K4me3 expression per cell. (<b>C</b>) SETD1A stained images of control, high glucose, and high glucose with sinefungin. Scale bar = 50 µm (×40). (<b>D</b>) ImageJ software (<span class="html-italic">n</span> = 6) was used to quantitatively measure the SETD1A expression per cell. ** <span class="html-italic">p</span> &lt; 0.01.</p>
Full article ">Figure 4
<p>The expression of MMPs and TIMPs in cells cultured with D-glucose (27.5 mM) was inhibited by sinefungin. The hGFs were cultured with D-glucose (27.5 mM) and inhibited by sinefungin. The mRNA expression levels of MMP1, MMP13, and TIMP1 were analyzed using real-time polymerase chain reaction. Results are presented as the mean ± S.E.M. (<span class="html-italic">n</span> = 6). * <span class="html-italic">p</span> &lt; 0.05. The ratio was calculated and analyzed using comparative quantitative values of MMP and TIMP (<span class="html-italic">n</span> = 6). * <span class="html-italic">p</span> &lt; 0.05, ** <span class="html-italic">p</span> &lt; 0.01.</p>
Full article ">Figure 5
<p>Schematic of mechanisms of increasing susceptibility to periodontal disease in diabetes. Hyperglycemia-induced upregulation of SETD1A increases H3K4me3 levels, which in turn leads to elevated expression of MMPs.</p>
Full article ">
14 pages, 3836 KiB  
Article
Mitochondria-Targeted Antioxidant MitoQ Improves In Vitro Maturation and Subsequent Embryonic Development from Culled Cows
by Zhihao Feng, Junsong Shi, Jiajie Ren, Lvhua Luo, Dewu Liu, Yongqing Guo, Baoli Sun, Guangbin Liu, Ming Deng and Yaokun Li
Animals 2024, 14(20), 2929; https://doi.org/10.3390/ani14202929 - 11 Oct 2024
Viewed by 292
Abstract
The purpose of this study was to investigate the effects and mechanisms of MitoQ on the IVM of culled bovine oocytes and subsequent embryonic development. The results revealed that in comparison to the control group (0 µmol/L), the IVM rate (p < [...] Read more.
The purpose of this study was to investigate the effects and mechanisms of MitoQ on the IVM of culled bovine oocytes and subsequent embryonic development. The results revealed that in comparison to the control group (0 µmol/L), the IVM rate (p < 0.05) and subsequent blastocyst rate (p < 0.05) of the low-concentration 1 and 5 µmol/L MitoQ treatment group were increased. The level of ROS (p < 0.05) in the MitoQ treatment group was decreased in comparison to the control group. Additionally, the level of GSH, MMP, ATP, and mt-DNA in the MitoQ treatment group was increased (p < 0.05) in comparison to the control group. The expression level of BAX was decreased (p < 0.05) in the MitoQ treatment group, and the BCL2, DNM1, Mfn2, SOD, and CAT were increased (p < 0.05). In conclusion, MitoQ improved mitochondrial dysfunction, increased mitochondrial activity during IVM, and reduced oxidative stress, resulting in increased IVM rates and subsequent embryonic development from culled cows. Full article
(This article belongs to the Section Cattle)
Show Figures

Figure 1

Figure 1
<p>Influences of MitoQ on subsequent embryonic development and quality. Representative images of blastocyst development on day 7 for control group and (1, 5, 10 µmol/L) MitoQ treatment groups. Scale bar: 200 µm.</p>
Full article ">Figure 2
<p>Influences of MitoQ on the ROS level in matured bovine oocytes. (<b>A</b>) Representative images of ROS (green) staining of oocytes in control group and MitoQ treatment group. Scale bar: 200 µm. (<b>B</b>) Analysis of ROS fluorescent intensity data. All experiments were performed in at least three replications, and the data statistics were represented by mean ± SEM. Values (a, b) with different superscripts in the columns are significantly different (<span class="html-italic">p</span> &lt; 0.05).</p>
Full article ">Figure 3
<p>Influences of MitoQ on the GSH level in matured bovine oocytes. (<b>A</b>) Representative images of GSH (blue) staining of oocytes in control group and MitoQ treatment group. Scale bar: 200 µm. (<b>B</b>) Analysis of GSH fluorescent intensity data. All experiments were performed in at least three replications, and the data statistics were represented by mean ± SEM. Values (a, b) with different superscripts in the columns are significantly different (<span class="html-italic">p</span> &lt; 0.05).</p>
Full article ">Figure 4
<p>Influences of MitoQ on the MMP level in matured bovine oocytes. (<b>A</b>) Representative images of MMP (red) staining of oocytes in control group and MitoQ treatment group. Scale bar: 200 µm. (<b>B</b>) Analysis of MMP fluorescent intensity data. All experiments were performed in at least three replications, and the data statistics were represented by mean ± SEM. Values (a, b) with different superscripts in the columns are significantly different (<span class="html-italic">p</span> &lt; 0.05).</p>
Full article ">Figure 5
<p>Influences of MitoQ on mitochondrial function in matured bovine oocytes. (<b>A</b>) Quantitative analysis of ATP content level. (<b>B</b>) Quantitative analysis of mt-DNA copies number level. All experiments were performed in at least three replications, and the data statistics were represented by mean ± SEM. Values (a, b) with different superscripts in the columns are significantly different (<span class="html-italic">p</span> &lt; 0.05).</p>
Full article ">Figure 6
<p>Influences of MitoQ on antioxidant-related and mitochondria-related genes in matured bovine oocytes. qPCR quantitative analysis of mRNA expression of apoptosis-related genes (BAX, BCL2), antioxidant-related genes (CAT, SOD), and mitochondria-related genes (DNM1, MFN2). All experiments were performed in three replications, and the data statistics were represented by mean ± SEM. Values (a, b) with different superscripts in the columns are significantly different (<span class="html-italic">p</span> &lt; 0.05).</p>
Full article ">
18 pages, 3426 KiB  
Article
Effect of Gossypol on Gene Expression in Swine Granulosa Cells
by Min-Wook Hong, Hun Kim, So-Young Choi, Neelesh Sharma and Sung-Jin Lee
Toxins 2024, 16(10), 436; https://doi.org/10.3390/toxins16100436 - 10 Oct 2024
Viewed by 345
Abstract
Gossypol (GP), a polyphenolic compound in cottonseed, has notable effects on female reproduction and the respiratory system in pigs. This study aimed to discern the alterations in gene expression within swine granulosa cells (GCs) when treated with two concentrations of GP (6.25 and [...] Read more.
Gossypol (GP), a polyphenolic compound in cottonseed, has notable effects on female reproduction and the respiratory system in pigs. This study aimed to discern the alterations in gene expression within swine granulosa cells (GCs) when treated with two concentrations of GP (6.25 and 12.5 µM) for 72 h, in vitro. The analysis revealed significant changes in the expression of numerous genes in the GP-treated groups. A Gene Ontology analysis highlighted that the differentially expressed genes (DEGs) primarily pertained to processes such as the mitotic cell cycle, chromosome organization, centromeric region, and protein binding. Pathway analysis using the Kyoto Encyclopedia of Genes and Genomes (KEGG) indicated distinct impacts on various pathways in response to different GP concentrations. Specifically, in the GP6.25 group, pathways related to the cycle oocyte meiosis, progesterone-mediated oocyte maturation, and p53 signaling were prominently affected. Meanwhile, in the GP12.5 group, pathways associated with PI3K-Akt signaling, focal adhesion, HIF-1 signaling, cell cycle, and ECM–receptor interaction showed significant alterations. Notably, genes linked to female reproductive function (CDK1, CCNB1, CPEB1, MMP3), cellular component organization (BIRC5, CYP1A1, TGFB3, COL1A2), and oxidation–reduction processes (PRDX6, MGST1, SOD3) exhibited differential expression in GP-treated groups. These findings offer valuable insights into the changes in GC gene expression in pigs exposed to GP. Full article
Show Figures

Figure 1

Figure 1
<p>Effect of GP on the viability of swine GCs <span class="html-italic">in vitro</span>. Swine GCs were cultured with 0, 6.25, 12.5, and 25 µM GP for 72 h. (<b>a</b>) GC viability at different concentrations (6.25–25 µM) of GP. Data are expressed as mean ± SEM (n = 3) *** <span class="html-italic">p</span> &lt; 0.001. (<b>b</b>–<b>d</b>) present the morphology of cultured GCs treated with GP. Large pictures were captured at 40× magnification, while small pictures were captured at 100× magnification. (<b>b</b>) GCs cultured in 0 µM GP (non-treated). (<b>c</b>) GCs cultured in 6.25 µM GP. (<b>d</b>) GCs cultured in 12.5 µM GP. GP: gossypol; GCs: granulosa cells; Ctrl: GP-untreated GCs.</p>
Full article ">Figure 2
<p>Summary of RNA−seq mapping data. (<b>a</b>) Comparative expression of selected transcriptomes across the GP−treated and Ctrl groups. Heatmap for the expression values in log10 (FPKM) units of the selected DEGs. Red indicates upregulated, while green indicates downregulated gene expression (<span class="html-italic">p</span> &lt; 0.05). (<b>b</b>) Density plot for log10 (FPKM) values to compare gene expression between each sample. (<b>c</b>) Venn diagram showing the number of differently expressed genes between the GP6.25 and GP12.5 groups (<span class="html-italic">p</span> &lt; 0.05). Red color: upregulated, blue color: downregulated; Ctrl: GP-untreated GCs, GP6.25: GCs treated with 6.25 µM GP; GP12.5: GCs treated with 12.5 µM GP; GP: gossypol; GCs: granulosa cells.</p>
Full article ">Figure 3
<p>GO functional analysis of DEGs. GO term enrichment analysis results based on the different GP treatments were retrieved and compared with the Ctrl group using g:Profiler (<a href="https://biit.cs.ut.ee/gprofiler" target="_blank">https://biit.cs.ut.ee/gprofiler</a>). The 10 most significantly (<span class="html-italic">p</span> &lt; 0.05) enriched GO terms related to biological processes, cellular components, and molecular functions are shown. All adjusted statistically significant values of the terms were −log10 converted. (<b>a</b>) GO term enrichment result of the DEGs in the GP6.25 group. (<b>b</b>) GO term enrichment result of the DEGs in the GP12.5 group. (<b>c</b>) GO term enrichment result of the DEGs in the IPC group. GO: Gene Ontology; GP: gossypol; Ctrl: GP-untreated GCs; GP6.25: GCs treated with 6.25 µM GP; GP12.5: GCs treated with 12.5 µM GP; IPC: intersection portion of both concentrations; DEGs: differentially expressed genes.</p>
Full article ">Figure 4
<p>A scatter plot of the significantly enriched KEGG pathways (<span class="html-italic">p</span> &lt; 0.05, <a href="http://www.kegg.jp/kegg/kegg1.html" target="_blank">www.kegg.jp/kegg/kegg1.html</a>) of the DEGs based on transcriptome sequencing analysis of each GP-treated group compared with the Ctrl group. The y-axis represents the name of the pathway, while the x-axis represents the gene ratio. Dot size represents the number of genes and the color indicates the <span class="html-italic">p</span>-value. (<b>a</b>) Enriched KEGG pathways of the DEGs in the GP6.25 group. (<b>b</b>) Enriched KEGG pathways of the DEGs in the GP12.5 group. (<b>c</b>) Enriched KEGG pathways of the DEGs in the IPC group. KEGG: Kyoto Encyclopedia of Genes and Genomes; DEGs: differentially expressed genes; GP6.25: GCs treated with 6.25 µM GP; GP12.5: GCs treated with 12.5 µM GP; IPC: intersection portion of both concentrations.</p>
Full article ">Figure 5
<p>Quantification of the mRNA profile of GP-treated GCs (6.25 and 12.5 µM) using RNA-seq and qRT-PCR. GCs incubated with each GP concentration for 72 h. GP exposure affected the mRNA abundance of certain genes in the cells. (<b>a</b>) The cellular component organization of related genes (<span class="html-italic">BIRC5</span>, <span class="html-italic">CYP1A1</span>, <span class="html-italic">COL1A2</span>, and <span class="html-italic">TGFβ3</span>) was measured and presented. (<b>b</b>) The female reproductive function-related genes (<span class="html-italic">CDK1</span>, <span class="html-italic">CCNB1</span>, <span class="html-italic">CPEB1</span>, and <span class="html-italic">MMP3</span>) measured and presented. (<b>c</b>) Oxidation–reduction process-related genes (<span class="html-italic">PRDX6</span>, <span class="html-italic">MGST1</span>, and <span class="html-italic">SOD3</span>) were measured and presented. The mRNA levels of all genes were normalized to the swine <span class="html-italic">β-actin</span> gene. Results are presented as mean ± SEM. All experiments were repeated at least three times. Ctrl: GP-untreated GCs; GP: gossypol; GCs: granulosa cells; * <span class="html-italic">p</span> &lt; 0.05, ** <span class="html-italic">p</span> &lt; 0.01, *** <span class="html-italic">p</span> &lt; 0.001.</p>
Full article ">Figure 6
<p>Western blot analysis of CCNB1, MMP3, MGST1, and PRDX6 proteins with different concentrations of GP (6.25 and 12.5 µM) in swine GCs. The protein levels were normalized to β-actin. Data are expressed as the mean ± SEM. All experiments were repeated at least three times. Full-length images of the above cropped images are presented in <a href="#app1-toxins-16-00436" class="html-app">Supplementary Figure S1</a>. Ctrl: GP-untreated GCs; GP: gossypol; GCs: granulosa cells; *: <span class="html-italic">p</span> &lt; 0.05, **: <span class="html-italic">p</span> &lt; 0.01, ***: <span class="html-italic">p</span> &lt; 0.001.</p>
Full article ">
14 pages, 3038 KiB  
Article
Role of IL3RA in a Family with Lumbar Spinal Stenosis
by Kai-Ming Liu, Chi-Fan Yang, Weng-Siong H’ng, Hui-Ping Chuang, Eunice Han Xian Khor, Pei-Chun Tsai, Vivia Khosasih, Liang-Suei Lu, Erh-Chan Yeh, Wan-Jia Lin, Feng-Jen Hsieh, Chien-Hsiun Chen, Shiuh-Lin Hwang and Jer-Yuarn Wu
Int. J. Mol. Sci. 2024, 25(20), 10915; https://doi.org/10.3390/ijms252010915 - 10 Oct 2024
Viewed by 253
Abstract
Lumbar spinal stenosis (LSS) is a degenerative spinal condition characterized by the narrowing of the spinal canal, resulting in low back pain (LBP) and limited leg mobility. Twin and family studies have suggested that genetics contributes to disease progression. However, the genetic causes [...] Read more.
Lumbar spinal stenosis (LSS) is a degenerative spinal condition characterized by the narrowing of the spinal canal, resulting in low back pain (LBP) and limited leg mobility. Twin and family studies have suggested that genetics contributes to disease progression. However, the genetic causes of familial LSS remain unclear. We performed whole-exome and direct sequencing on seven female patients from a Han Chinese family with LBP, among whom four developed LSS. Based on our genetic findings, we performed gene knockdown studies in human chondrocytes to study possible pathological mechanisms underlying LSS. We found a novel nonsense mutation, c.417C > G (NM_002183, p.Y139X), in IL3RA, shared by all the LBP/LSS cases. Knockdown of IL3RA led to a reduction in the total collagen content of 81.6% in female chondrocytes and 21% in male chondrocytes. The expression of MMP-1, -3, and/or -10 significantly increased, with a more pronounced effect observed in females than in males. Furthermore, EsRb expression significantly decreased following IL3RA knockdown. Moreover, the knockdown of EsRb resulted in increased MMP-1 and -10 expression in chondrocytes from females. We speculate that IL3RA deficiency could lead to a reduction in collagen content and intervertebral disk (IVD) strength, particularly in females, thereby accelerating IVD degeneration and promoting LSS occurrence. Our results illustrate, for the first time, the association between IL3RA and estrogen receptor beta, highlighting their importance and impact on MMPs and collagen in degenerative spines in women. Full article
Show Figures

Figure 1

Figure 1
<p>The pedigree of the family with low back pain and lumbar stenosis. The dark circles represent the patients with low back pain. Among them, II-3, III-2, III-7, and III-9 were diagnosed with lumbar stenosis and/or underwent spinal surgery.</p>
Full article ">Figure 2
<p>The expression of IL3RA (green), MMP-1 (red) COL11A2 (red), and DAPI (blue) in human intervertebral disk (<b>A</b>) and human chondrocyte cells (<b>B</b>). Bar = 20 μm.</p>
Full article ">Figure 3
<p>The effects of <span class="html-italic">IL3RA</span> knockdown on the total amount of collagen in chondrocyte lines from females and males. The difference between the <span class="html-italic">IL3RA</span> knockdown and control groups is presented as fold change. For each group, six individual experiments were performed. <span class="html-italic">n</span> = 6 each group. ** <span class="html-italic">p</span> &lt; 0.01, **** <span class="html-italic">p</span> &lt; 0.0001. Independent <span class="html-italic">t</span>-test was used for statistical analysis. The values are shown as mean with a 95% confidence interval (CI).</p>
Full article ">Figure 4
<p>The effects of <span class="html-italic">IL3RA</span> knockdown on mRNA expression of estrogen receptors (<span class="html-italic">EsRa</span> and <span class="html-italic">EsRb</span>), <span class="html-italic">MMPs</span>, and <span class="html-italic">COL11A2</span> in chondrocytes from human females. The fold change in gene expression in each knockdown (KD) group compared with that in the non-KD control was determined using qPCR. For each group, the assay was performed in triplicate, and chondrocytes were treated with 10 nM of IL-3. The effects are presented for different concentrations of estrogen or testosterone. <span class="html-italic">n</span> = 3 each group. * <span class="html-italic">p</span> &lt; 0.05, ** <span class="html-italic">p</span> &lt; 0.01. Independent <span class="html-italic">t</span>-test was used for statistical analysis. The values are shown as mean with a 95% confidence interval (CI).</p>
Full article ">Figure 5
<p>The effects of <span class="html-italic">IL3RA</span> knockdown on mRNA expression of <span class="html-italic">MMPs</span> and <span class="html-italic">COL11A2</span> in chondrocytes from human males. Fold change in gene expression in each knockdown (KD) group compared with that in the non-KD control was determined using qPCR. For each group, the assay was performed in triplicate, and chondrocytes were treated with 10 nM of IL-3. The effects are presented for different concentrations of testosterone groups. <span class="html-italic">n</span> = 3 each group. * <span class="html-italic">p</span> &lt; 0.05, ** <span class="html-italic">p</span> &lt; 0.01, and *** <span class="html-italic">p</span> &lt; 0.001. Independent <span class="html-italic">t</span>-test was used for statistical analysis. The values are shown as mean with a 95% confidence interval (CI).</p>
Full article ">Figure 6
<p>The effects of estrogen receptor beta (<span class="html-italic">EsRb</span>) gene knockdown on mRNA expression of <span class="html-italic">IL3RA</span>, estrogen receptor alpha (<span class="html-italic">EsRa</span>), <span class="html-italic">MMPs</span>, and <span class="html-italic">COL11A2</span> in chondrocytes from human females. Fold change in gene expression in each knockdown (KD) group compared with that in the non-KD control was determined using qPCR. For each group, the assay was performed in triplicate, and chondrocytes were treated with 10 nM of IL-3. The effects are presented for the 10 nM estrogen groups. <span class="html-italic">n</span> = 3 each group. * <span class="html-italic">p</span> &lt; 0.05, ** <span class="html-italic">p</span> &lt; 0.01. Independent <span class="html-italic">t</span>-test was used for statistical analysis. The values are shown as mean with a 95% confidence interval (CI).</p>
Full article ">
Back to TopTop