www.fgks.org   »   [go: up one dir, main page]

You seem to have javascript disabled. Please note that many of the page functionalities won't work as expected without javascript enabled.
 
 
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

Article Types

Countries / Regions

remove_circle_outline
remove_circle_outline
remove_circle_outline
remove_circle_outline

Search Results (371)

Search Parameters:
Keywords = nod like receptor

Order results
Result details
Results per page
Select all
Export citation of selected articles as:
16 pages, 4135 KiB  
Article
Wound Healing Efficacy of Cucurbitaceae Seed Oils in Rats: Comprehensive Phytochemical, Pharmacological, and Histological Studies Tackling AGE/RAGE and Nrf2/Ho-1 Cue
by Ayat M. Emad, Engy A. Mahrous, Dalia M. Rasheed, Fatma Alzahraa M. Gomaa, Ahmed Mohsen Elsaid Hamdan, Heba Mohammed Refat M. Selim, Einas M. Yousef, Hagar B. Abo-Zalam, Amira A. El-Gazar and Ghada M. Ragab
Pharmaceuticals 2024, 17(6), 733; https://doi.org/10.3390/ph17060733 - 5 Jun 2024
Viewed by 304
Abstract
The Cucurbitaceae family includes several edible species that are consumed globally as fruits and vegetables. These species produce high volumes of seeds that are often discarded as waste. In this study, we investigate the chemical composition and biological activity of three seed oils [...] Read more.
The Cucurbitaceae family includes several edible species that are consumed globally as fruits and vegetables. These species produce high volumes of seeds that are often discarded as waste. In this study, we investigate the chemical composition and biological activity of three seed oils from Cucurbitaceae plants, namely, cantaloupe, honeydew, and zucchini, in comparison to the widely used pumpkin seed oil for their ability to enhance and accelerate wound healing in rats. Our results showed that honeydew seed oil (HSO) was effective in accelerating wound closure and enhancing tissue repair, as indicated by macroscopic, histological, and biochemical analyses, as compared with pumpkin seed oil (PSO). This effect was mediated by down-regulation of the advanced glycation end products (AGE) and its receptor (RAGE) cue, activating the cytoprotective enzymes nuclear factor erythroid 2 (Nrf2) and heme oxygenase-1 (HO-1), suppressing the inflammatory mediators tumor necrosis factor (TNF)-α, nuclear factor kappa B (NF-κB), and nod-like receptor protein 3 (NLRP3), and reducing the levels of the skin integral signaling protein connexin (CX)-43. Furthermore, immunohistochemical staining for epidermal growth factor (EGF) showed the lowest expression in the skin after treatment with HSO, indicating a well-organized and complete healing process. Other seed oils from cantaloupe and zucchini exhibited favorable activity when compared with untreated rats; however, their efficacy was comparatively lower than that of PSO and HSO. Gas chromatographic analysis of the derivatized oils warranted the superior activity of HSO to its high nutraceutical content of linoleic acid, which represented 65.9% of the fatty acid content. This study’s findings validate the use of honeydew seeds as a wound-healing fixed oil and encourage further investigation into the potential of Cucurbitaceae seeds as sources of medicinally valuable plant oils. Full article
(This article belongs to the Section Medicinal Chemistry)
Show Figures

Figure 1

Figure 1
<p>Macroscopic assessment of the morphological appearance and progression of wound healing (%) on days 4, 8, 10, and 14 after wound induction. (<b>A</b>): Representative photo-macrographs of wound morphology, (<b>B</b>): progression of wound healing (%) determined using Image J, an image analyzer, and compared using Mixed model ANOVA, (<b>C</b>): progression of wound healing (%) determined using Image J, an image analyzer, and compared using one-way ANOVA and subsequent multiple comparisons using Tukey’s test. Data are expressed as the mean ± SD [n = 9], and the number of asterisks “*” above the columns indicates the strength of significance as follows: ** <span class="html-italic">p</span> &lt; 0.01, *** <span class="html-italic">p</span> &lt; 0.001, and **** <span class="html-italic">p</span> &lt; 0.0001, while (ns) stands for non-significant. CSO; cantaloupe seed oil, HSO; honeydew melon seed oil, PSO; pumpkin seed oil, ST; standard, ZSO; zucchini seed oil.</p>
Full article ">Figure 2
<p>Representative photomicrographs of histopathological alterations after wound induction (hematoxylin and eosin stain; scale bar 50 μm, 200×). Photomicrographs of skin tissue from the normal control group (<b>I</b>) with the normal histological structure of the skin, the wound injury group (<b>II</b>) showing the formation of a large scab (S) covering some newly formed epidermal layers (E) with fibrous connective tissue formation (F), the PSO/ST-treated group (<b>III</b>) showing migration of epidermal cells (EM) and formation of some layers of epidermis (E) and well-organized fibrous connective tissue (F), the HSO-treated group (<b>IV</b>) presenting migration of epidermal cells (EM) and formation of some layers of epidermis (E), the CSO-treated group (<b>V</b>) showing well-organized fibrous connective tissue (F) and an epidermal layer (E) at the site of the defect, and finally, the ZSO-treated group (<b>VI</b>) showing migration of epidermal cells (EM) and formation of some layers of epidermis (E) with the presence of ill-organized fibrous connective tissue (F).</p>
Full article ">Figure 3
<p>Effect of different oils on the skin tissue protein expressions of (<b>A</b>) AGE and (<b>B</b>) RAGE as signaling cue/genes in wound-injured rats. Statistical analysis was achieved with one-way ANOVA and subsequent multiple comparisons using Tukey’s test. Data are expressed as the mean ± SD [n = 3] with <span class="html-italic">p</span>-value &lt; 0.05 as compared with the normal control group (α), wound injury (β), wound injury + HSO (φ), and wound injury + CSO (ε). [F-value of AGE = 60.62 and F-value of RAGE = 224.1.] AGE; advanced glycation end products, CSO; cantaloupe seed oil, HSO; honeydew melon seed oil, PSO; pumpkin seed oil, RAGE; receptor of advanced glycation end products, ST; standard, ZSO; zucchini seed oil.</p>
Full article ">Figure 4
<p>Effect of different oils on the skin tissue contents of (<b>A</b>) Nrf2 and (<b>B</b>) HO-1 as antioxidant signaling molecules in wound-injured rats. Statistical analysis was performed using one-way ANOVA and subsequent multiple comparisons using Tukey’s test. Data are expressed as the mean ± SD [n = 6] with <span class="html-italic">p</span>-value &lt; 0.05 as compared with the normal control group (α) and the wound injury (β), wound injury + PSO/ST (γ), wound injury + HSO (φ), and wound injury + CSO (ε) groups. [F-value of Nrf2 = 127.9 and F-value of HO-1 = 53.09.] CSO; cantaloupe seed oil, HO-1; heme oxygenase-1, HSO; honeydew melon seed oil, Nrf2, nuclear factor erythropoietin-2-related factor 2, PSO; pumpkin seed oil, ST; standard, ZSO; zucchini seed oil.</p>
Full article ">Figure 5
<p>Effect of different oils on the skin tissue contents of (<b>A</b>) TNF-α, (<b>B</b>) NF-κB, and (<b>C</b>) NLRP3 as inflammatory markers in addition to serum levels of (<b>D</b>) CX-43 as a skin integrity marker in wound-injured rats. Statistical analysis was achieved with one-way ANOVA and subsequent multiple comparisons using Tukey’s test. Data are expressed as the mean ± SD [n = 6] with <span class="html-italic">p</span>-value &lt; 0.05 as compared with the normal control group (α) and the wound injury (β), wound injury + PSO/ST (γ), wound injury + HSO (φ), and wound injury + CSO (ε) groups. [F-value of TNF-α = 198.9, F-value of NF-κB = 139.8, F-value of NLRP3 = 253.5, and F-value of CX-43 = 423.8.] CX-43; connexin-43, CSO; Cantaloupe seed oil, HSO; Honeydew melon seed oil, NLRP3; NLR family pyrin domain containing 3, NF-κB; nuclear transcription factor kappa B, PSO; pumpkin seed oil, ST; standard, TNF-α; tumor necrosis factor-alpha, ZSO; Zucchini seed oil.</p>
Full article ">Figure 6
<p>Representative photomicrographs of the immunohistochemistry expression of epidermal growth factor (EGF) after 14 days of wound induction. Panel [<b>A</b>]: skin sections of the newly formed epidermal layer showing areas stained for EGF expression in the normal skin of healthy rats (I), wound injury group (II), PSO/ST-treated rats (III), rats treated with HSO (IV), CSO (V), and ZSO (VI). Panel [<b>B</b>] represents EGF area% expression (the mean of 10 microscopic fields ± SD) in all treated groups. One-way ANOVA and subsequent multiple comparisons using Tukey’s test as compared with the normal control group (α) and the wound injury (β), wound injury + PSO/ST (γ), and wound injury + HSO (φ), groups [F-value of EGF% = 197.3]. CSO; cantaloupe seed oil; EGF; epidermal growth factor; IHC; immunohistochemistry; HSO; honeydew melon seed oil; PSO; pumpkin seed oil; ST; standard; ZSO; zucchini seed oil.</p>
Full article ">Figure 7
<p>Diagram of the experimental protocol timeline.</p>
Full article ">
14 pages, 4631 KiB  
Article
TBHQ Alleviates Particulate Matter-Induced Pyroptosis in Human Nasal Epithelial Cells
by Ji-Sun Kim, Hyunsu Choi, Jeong-Min Oh, Sung Won Kim, Soo Whan Kim, Byung Guk Kim, Jin Hee Cho, Joohyung Lee and Dong Chang Lee
Toxics 2024, 12(6), 407; https://doi.org/10.3390/toxics12060407 - 3 Jun 2024
Viewed by 119
Abstract
Pyroptosis represents a type of cell death mechanism notable for its cell membrane disruption and the subsequent release of proinflammatory cytokines. The Nod-like receptor family pyrin domain containing inflammasome 3 (NLRP3) plays a critical role in the pyroptosis mechanism associated with various diseases [...] Read more.
Pyroptosis represents a type of cell death mechanism notable for its cell membrane disruption and the subsequent release of proinflammatory cytokines. The Nod-like receptor family pyrin domain containing inflammasome 3 (NLRP3) plays a critical role in the pyroptosis mechanism associated with various diseases resulting from particulate matter (PM) exposure. Tert-butylhydroquinone (tBHQ) is a synthetic antioxidant commonly used in a variety of foods and products. The aim of this study is to examine the potential of tBHQ as a therapeutic agent for managing sinonasal diseases induced by PM exposure. The occurrence of NLRP3 inflammasome-dependent pyroptosis in RPMI 2650 cells treated with PM < 4 µm in size was confirmed using Western blot analysis and enzyme-linked immunosorbent assay results for the pyroptosis metabolites IL-1β and IL-18. In addition, the inhibitory effect of tBHQ on PM-induced pyroptosis was confirmed using Western blot and immunofluorescence techniques. The inhibition of tBHQ-mediated pyroptosis was abolished upon nuclear factor erythroid 2-related factor 2 (Nrf2) knockdown, indicating its involvement in the antioxidant mechanism. tBHQ showed potential as a therapeutic agent for sinonasal diseases induced by PM because NLRP3 inflammasome activation was effectively suppressed via the Nrf2 pathway. Full article
(This article belongs to the Special Issue Air Pollutant Exposure and Respiratory Diseases)
Show Figures

Graphical abstract

Graphical abstract
Full article ">
25 pages, 1605 KiB  
Review
NLRP3 Inflammasome Inhibitors for Antiepileptogenic Drug Discovery and Development
by Inamul Haque, Pritam Thapa, Douglas M. Burns, Jianping Zhou, Mukut Sharma, Ram Sharma and Vikas Singh
Int. J. Mol. Sci. 2024, 25(11), 6078; https://doi.org/10.3390/ijms25116078 - 31 May 2024
Viewed by 265
Abstract
Epilepsy is one of the most prevalent and serious brain disorders and affects over 70 million people globally. Antiseizure medications (ASMs) relieve symptoms and prevent the occurrence of future seizures in epileptic patients but have a limited effect on epileptogenesis. Addressing the multifaceted [...] Read more.
Epilepsy is one of the most prevalent and serious brain disorders and affects over 70 million people globally. Antiseizure medications (ASMs) relieve symptoms and prevent the occurrence of future seizures in epileptic patients but have a limited effect on epileptogenesis. Addressing the multifaceted nature of epileptogenesis and its association with the Nod-like receptor family pyrin domain containing 3 (NLRP3) inflammasome-mediated neuroinflammation requires a comprehensive understanding of the underlying mechanisms of these medications for the development of targeted therapeutic strategies beyond conventional antiseizure treatments. Several types of NLRP3 inhibitors have been developed and their effect has been validated both in in vitro and in vivo models of epileptogenesis. In this review, we discuss the advances in understanding the regulatory mechanisms of NLRP3 activation as well as progress made, and challenges faced in the development of NLRP3 inhibitors for the treatment of epilepsy. Full article
Show Figures

Figure 1

Figure 1
<p>Schematic illustration of the mechanism regulating NLRP3 inflammasome activation in canonical, non-canonical, and alternate pathway. Optimal activation of NLRP3 requires two steps. The first step is priming, which is initiated by extracellular PAMPs and endogenous cytokines by the PRRs, which then upregulates the NF-κB-mediated transcription of NLRP3, pro-IL-1β, and pro-IL-18. The second step is activation, which includes canonical and non-canonical pathways. The canonical pathway is triggered by multiple pathogens and inflammatory agents through a combination of important and related events such as activation of the purinergic receptor P2X7 by ATP, cathepsin release following lysozyme rupture, opening of Ca2+ channels to allow ion flux, mitochondrial dysfunction, ROS formation, Golgi apparatus disassembly, and endoplasmic reticulum stress. Once activated, oligomerization of the NLRP3 inflammasome is thought to induce conformational changes that generate active caspase-1, which converts pro-pro-IL-1β and pro-IL-18 to mature bioactive IL-1β and IL-18. Additionally, Casp1 cleaves the protein gasdermin D to generate N-terminal gasdermin D to form pores, allowing IL-1β and IL-18 to leave the cell and effectively execute a highly inflammatory form of cell death that is termed pyroptosis. In the non-canonical pathway, intracellular LPS is directly recognized by the CARD domain of caspase-11 in mice and caspase-4/5 in humans, ultimately leading to IL-1β and IL-18 release through the activation of the NLRP3-ASC-Casp1 pathway. The alternative pathway of activation is caused by TLR4 agonists like LPS, which activates the TLR4-TRIF-RIPK1-FADD-Casp8 signaling pathway. Casp8 activates the NLRP3 inflammasome but lacks ASC speck formation, pyroptosis induction, or K+ efflux.</p>
Full article ">Figure 2
<p>Chemical structures of clinically approved antiseizure medications discussed in this review.</p>
Full article ">Figure 3
<p>Chemical structure of different NLRP3 inflammasome inhibitors discussed in this review.</p>
Full article ">
12 pages, 2056 KiB  
Communication
NLRX1 Mediates the Disruption of Intestinal Mucosal Function Caused by Porcine Astrovirus Infection via the Extracellular Regulated Protein Kinases/Myosin Light–Chain Kinase (ERK/MLCK) Pathway
by Jie Tao, Jinghua Cheng, Ying Shi, Benqiang Li, Pan Tang, Jiajie Jiao and Huili Liu
Cells 2024, 13(11), 913; https://doi.org/10.3390/cells13110913 - 25 May 2024
Viewed by 319
Abstract
Porcine astrovirus (PAstV) has a potential zoonotic risk, with a high proportion of co-infection occurring with porcine epidemic diarrhea virus (PEDV) and other diarrheal pathogens. Despite its high prevalence, the cellular mechanism of PAstV pathogenesis is ill–defined. Previous proteomics [...] Read more.
Porcine astrovirus (PAstV) has a potential zoonotic risk, with a high proportion of co-infection occurring with porcine epidemic diarrhea virus (PEDV) and other diarrheal pathogens. Despite its high prevalence, the cellular mechanism of PAstV pathogenesis is ill–defined. Previous proteomics analyses have revealed that the differentially expressed protein NOD–like receptor X1 (NLRX1) located in the mitochondria participates in several important antiviral signaling pathways in PAstV–4 infection, which are closely related to mitophagy. In this study, we confirmed that PAstV–4 infection significantly up-regulated NLRX1 and mitophagy in Caco–2 cells, while the silencing of NLRX1 or the treatment of mitophagy inhibitor 3–MA inhibited PAstV–4 replication. Additionally, PAstV–4 infection triggered the activation of the extracellular regulated protein kinases/ myosin light-chain kinase (ERK/MLCK) pathway, followed by the down-regulation of tight–junction proteins (occludin and ZO–1) as well as MUC–2 expression. The silencing of NLRX1 or the treatment of 3–MA inhibited myosin light-chain (MLC) phosphorylation and up-regulated occludin and ZO–1 proteins. Treatment of the ERK inhibitor PD98059 also inhibited MLC phosphorylation, while MLCK inhibitor ML-7 mitigated the down-regulation of mucosa-related protein expression induced by PAstV–4 infection. Yet, adding PD98059 or ML–7 did not affect NLRX1 expression. In summary, this study preliminarily explains that NLRX1 plays an important role in the disruption of intestinal mucosal function triggered by PAstV–4 infection via the ERK/MLC pathway. It will be helpful for further antiviral drug target screening and disease therapy. Full article
(This article belongs to the Special Issue Charming Micro-Insights into Health and Diseases)
Show Figures

Graphical abstract

Graphical abstract
Full article ">Figure 1
<p>Down-regulation of tight junctions in Caco−2 cells upon <span class="html-italic">PAstV</span> infection. (<b>A</b>) <span class="html-italic">PAstV</span> propagation on Caco−2 cells with 15 μg/mL Pancreatin resulted in specific red fluorescence detected in the cytoplasm, as demonstrated by an indirect immunofluorescence assay using a <span class="html-italic">PAstV</span> monoclonal antibody. (<b>B</b>) Following inoculation of Caco-2 cells with 1MOI of <span class="html-italic">PAstV</span>, the expression of tight-junction proteins was detected 24 h later through Western blotting. (<b>C</b>) The influence of <span class="html-italic">PAstV</span> infection at various MOIs (0.1, 0.5, and 1.0) on the transcription of MUC−2, occludin, and ZO−1 was determined by relative fluorescence quantification PCR, utilizing <span class="html-italic">β-actin</span> as the intrinsic reference protein. Fold changes were calculated using the 2<sup>−ΔΔCt</sup> method. (“**” <span class="html-italic">p</span> &lt; 0.01, ”*” <span class="html-italic">p</span> &lt; 0.05).</p>
Full article ">Figure 2
<p>Reduction in <span class="html-italic">PAstV</span> replication through NLRX1 knockdown in Caco−2 cells. (<b>A</b>) Caco−2 cells were infected at an MOI of 1.0 with either <span class="html-italic">PAstV</span> or UV−inactivated <span class="html-italic">PAstV</span> for 24 h. Real−time PCR was employed to determine the relative NLRX1 levels, normalized to <span class="html-italic">β−actin</span>. Asterisks denote significant differences from uninfected cells(“**” <span class="html-italic">p</span> &lt; 0.01, ”*” <span class="html-italic">p</span> &lt; 0.05). “ns” indicates no significant difference. (<b>B</b>) Cell lysates were collected at 24 h and 48 h, respectively, following infection, and NLRX1 expression was assessed through Western blotting using the indicated antibodies. (<b>C</b>,<b>D</b>) Caco−2 cells were transfected with negative control siRNA or three different siRNA duplexes targeting NLRX1 for 24 h. Subsequently, the relative NLRX1 and expression levels were detected, with <span class="html-italic">β−actin</span> as an internal control. (<b>E</b>) Cells were transfected with negative control siRNA or siRNA/NLRX1−2 for 24 h and then infected with 1.0 MOI PAstV. Virus titers were determined through a TCID<sub>50</sub> assay at 60 h post-infection.</p>
Full article ">Figure 3
<p>Induction of mitophagy by <span class="html-italic">PAstV</span> infection via up-regulation of NLRX1 protein. (<b>A</b>) Caco-2 cells were transfected with negative control siRNA or siRNA/NLRX1−2 for 24 h, followed by infection with 1MOI <span class="html-italic">PAstV</span>. Cell lysates were blotted with anti-LC3I/II, anti-NLRX1, and anti-<span class="html-italic">β−actin</span> antibodies. (<b>B</b>) Cells treated with the mitophagy inhibitor 3−MA for 6 h were subsequently infected with 1.0MOI <span class="html-italic">PAstV</span>. Virus titers were determined through a TCID<sub>50</sub> assay 60 h post-infection. (<b>C</b>,<b>D</b>) show that cells subjected to treatment were infected with 1.0 MOI <span class="html-italic">PAstV</span>, and the relative and expression levels of NLRX1 were assessed 24 h later. (“**” <span class="html-italic">p</span> &lt; 0.01, ”*” <span class="html-italic">p</span> &lt; 0.05). “ns” indicates no significant difference.</p>
Full article ">Figure 4
<p>Reduction in tight-junction protein expression by <span class="html-italic">PAstV</span> infection via the ERK/MLC pathway. (<b>A</b>) Caco-2 cells were transfected with negative control siRNA or siRNA/NLRX1-2 for 24 h, followed by <span class="html-italic">PAstV</span> infection at an MOI of 1.0. After 24 h, cell lysates were blotted with the indicated antibodies. (<b>B</b>) Cells were treated with 20 μM PD98059 (ERK inhibitor) for 6 h, followed by infection with 1MOI <span class="html-italic">PAstV</span>. Subsequent Western blotting using the indicated antibodies was conducted as described above. (<b>C</b>) Cells were treated with 20 μM 3-MA (mitophagy inhibitor), and the subsequent steps were the same as in (<b>B</b>).</p>
Full article ">Figure 5
<p>Essential role of NLRX1 in impaired intestinal barrier function induced by <span class="html-italic">PAstV</span> infection. (<b>A</b>) Caco−2 cells were transfected with negative control siRNA or siRNA/NLRX1−2 for 24 h, followed by <span class="html-italic">PAstV</span> infection at an MOI of 1.0. Cell lysates were then blotted with the indicated antibodies 24 h later. Further, cells were treated with 25 μM ML−7 (MLCK inhibitor) for 6 h and infected with 1.0 MOI <span class="html-italic">PAstV</span>. (<b>B</b>) The relative MUC−2, occludin, ZO−1, and NLRX1 levels were detected after treatment with ML−7 (MLCK inhibitor), using <span class="html-italic">β−actin</span> as an internal reference. (<b>C</b>) MUC−2, occludin, ZO−1, and NLRX1 expression levels were assessed after treatment with ML−7 (MLCK inhibitor), with <span class="html-italic">β−actin</span> as an internal control. (“**” <span class="html-italic">p</span> &lt; 0.01, ”*” <span class="html-italic">p</span> &lt; 0.05). “ns” indicates no significant difference.</p>
Full article ">
16 pages, 3819 KiB  
Article
Single-Cell Transcriptomic Profiling Unveils Dynamic Immune Cell Responses during Haemonchus contortus Infection
by Wenxuan Wang, Zhe Jin, Mei Kong, Zhuofan Yan, Liangliang Fu and Xiaoyong Du
Cells 2024, 13(10), 842; https://doi.org/10.3390/cells13100842 - 15 May 2024
Viewed by 526
Abstract
Background: Haemonchus contortus is a parasite widely distributed in tropical, subtropical, and warm temperate regions, causing significant economic losses in the livestock industry worldwide. However, little is known about the genetics of H. contortus resistance in livestock. In this study, we monitor the [...] Read more.
Background: Haemonchus contortus is a parasite widely distributed in tropical, subtropical, and warm temperate regions, causing significant economic losses in the livestock industry worldwide. However, little is known about the genetics of H. contortus resistance in livestock. In this study, we monitor the dynamic immune cell responses in diverse peripheral blood mononuclear cells (PBMCs) during H. contortus infection in goats through single-cell RNA sequencing (scRNA-Seq) analysis. Methods and Results: A total of four Boer goats, two goats with oral infection with the L3 larvae of H. contortus and two healthy goats as controls, were used in the animal test. The infection model in goats was established and validated by the fecal egg count (FEC) test and qPCR analysis of the gene expression of IL-5 and IL-6. Using scRNA-Seq, we identified seven cell types, including T cells, monocytes, natural killer cells, B cells, and dendritic cells with distinct gene expression signatures. After identifying cell subpopulations of differentially expressed genes (DEGs) in the case and control groups, we observed the upregulation of multiple inflammation-associated genes, including NFKBIA and NFKBID. Kyoto Encyclopedia of the Genome (KEGG) enrichment analysis revealed significant enrichment of NOD-like receptor pathways and Th1/Th2 cell differentiation signaling pathways in CD4 T cells DEGs. Furthermore, the analysis of ligand–receptor interaction networks showed a more active state of cellular communication in the PBMCs from the case group, and the inflammatory response associated MIF–(CD74 + CXCR4) ligand receptor complex was significantly more activated in the case group, suggesting a potential inflammatory response. Conclusions: Our study preliminarily revealed transcriptomic profiling characterizing the cell type specific mechanisms in host PBMCs at the single-cell level during H. contortus infection. Full article
Show Figures

Figure 1

Figure 1
<p>The change in EPG and immune gene expression. (<b>A</b>) Pipeline for case and control group processing and analysis; (<b>B</b>) EPG in goats infected with <span class="html-italic">H. contortus</span> at different times; (<b>C</b>) qPCR detection of immune gene expression in goat PBMCs at various time points postinfection with <span class="html-italic">H. contortus</span> (*, <span class="html-italic">p</span> &lt; 0.05; **, <span class="html-italic">p</span> &lt; 0.01).</p>
Full article ">Figure 2
<p>Annotation of cell subpopulations. (<b>A</b>) The clustered grouping plot of integrated samples; (<b>B</b>) UMAP plot annotated with SingleR; (<b>C</b>) UMAP plots illustrating the seven primary subpopulations of PBMCs; (<b>D</b>) bubble plots showing the expression and distribution of marker genes in each cell subpopulation of PBMCs; (<b>E</b>) UMAP plots depicting the 10 major subpopulations of T cells; (<b>F</b>) bubble plots depicting the expression and distribution of marker genes in each cell subpopulation of T cells.</p>
Full article ">Figure 3
<p>Differentially expressed genes in cell subpopulations of PBMCs from case and control groups. (<b>A</b>) The top 10 upregulated and downregulated genes in CD8 T cells; (<b>B</b>) the top 10 upregulated and downregulated genes in NK cells; (<b>C</b>) the top 10 upregulated and downregulated genes in CD4 T cells; (<b>D</b>) the top 10 upregulated and downregulated genes in monocytes; (<b>E</b>) the top 10 upregulated and downregulated genes in B cells; (<b>F</b>) the top 10 upregulated and downregulated genes in DCs; (<b>G</b>) GO enrichment analysis of DEGs; (<b>H</b>) KEGG enrichment analysis of DEGs.</p>
Full article ">Figure 4
<p>Cell communication among subpopulations of goat PBMCs. (<b>A</b>) Number of interactions among subpopulations of PBMCs in the control group of goats; (<b>B</b>) number of interactions among subpopulations of PBMCs in the case group of goats; (<b>C</b>) the strength of interaction among subpopulations of goat PBMCs in the control group calculated by summing the probability values; (<b>D</b>) the strength of interactions among subpopulations in the case group of goat PBMCs; (<b>E</b>) cell communication among subpopulations of goat PBMCs in the control group: the <span class="html-italic">x</span>-axis represents cell combinations, and the <span class="html-italic">y</span>-axis represents ligand–receptor pairs that participate in cell–cell interactions in the communication network; (<b>F</b>) goat PBMC subpopulation cell communication in the case group; (<b>G</b>) hierarchical plot showing inferred cell interactions in the control group: solid and open circles refer to the source and target, respectively, while lines represent intercellular interactions, their thickness is proportional to the communication probability of the cell interaction, and; line colors correspond to the signaling source; (<b>H</b>) hierarchical plot showing inferred cell interactions in the case group.</p>
Full article ">Figure 5
<p>Comparative analysis of transcription factor regulation. (<b>A</b>) The top 15 transcription factors with the greatest differential upregulation or downregulation between the control group and the case group of T cells; (<b>B</b>) the top 15 transcription factors with the greatest differential upregulation and downregulation between the control group and the case group of monocytes.</p>
Full article ">
19 pages, 2800 KiB  
Article
Nucleotide-Binding Oligomerization Domain 1 (NOD1) Agonists Prevent SARS-CoV-2 Infection in Human Lung Epithelial Cells through Harnessing the Innate Immune Response
by Edurne Garcia-Vidal, Ignasi Calba, Eva Riveira-Muñoz, Elisabet García, Bonaventura Clotet, Pere Serra-Mitjà, Cecilia Cabrera, Ester Ballana and Roger Badia
Int. J. Mol. Sci. 2024, 25(10), 5318; https://doi.org/10.3390/ijms25105318 - 13 May 2024
Viewed by 581
Abstract
The lung is prone to infections from respiratory viruses such as Severe Acute Respiratory Syndrome Coronavirus 2 (SARS-CoV-2). A challenge in combating these infections is the difficulty in targeting antiviral activity directly at the lung mucosal tract. Boosting the capability of the respiratory [...] Read more.
The lung is prone to infections from respiratory viruses such as Severe Acute Respiratory Syndrome Coronavirus 2 (SARS-CoV-2). A challenge in combating these infections is the difficulty in targeting antiviral activity directly at the lung mucosal tract. Boosting the capability of the respiratory mucosa to trigger a potent immune response at the onset of infection could serve as a potential strategy for managing respiratory infections. This study focused on screening immunomodulators to enhance innate immune response in lung epithelial and immune cell models. Through testing various subfamilies and pathways of pattern recognition receptors (PRRs), the nucleotide-binding and oligomerization domain (NOD)-like receptor (NLR) family was found to selectively activate innate immunity in lung epithelial cells. Activation of NOD1 and dual NOD1/2 by the agonists TriDAP and M-TriDAP, respectively, increased the number of IL-8+ cells by engaging the NF-κB and interferon response pathways. Lung epithelial cells showed a stronger response to NOD1 and dual NOD1/2 agonists compared to control. Interestingly, a less-pronounced response to NOD1 agonists was noted in PBMCs, indicating a tissue-specific effect of NOD1 in lung epithelial cells without inducing widespread systemic activation. The specificity of the NOD agonist pathway was confirmed through gene silencing of NOD1 (siRNA) and selective NOD1 and dual NOD1/2 inhibitors in lung epithelial cells. Ultimately, activation induced by NOD1 and dual NOD1/2 agonists created an antiviral environment that hindered SARS-CoV-2 replication in vitro in lung epithelial cells. Full article
(This article belongs to the Special Issue Viral and Host Targets to Fight RNA Viruses)
Show Figures

Figure 1

Figure 1
<p>NLR agonists induce innate immune activation in in vitro lung epithelial and myeloid models. (<b>A</b>) Workflow to screen for potential immunomodulators of the innate immune system in A549 lung epithelial and THP-1 myeloid cell lines. (<b>B</b>) Library classification of tested compounds according to their reported target. (<b>C</b>) Heatmap illustrates the immune activation induced by immunomodulators targeting PRR subfamilies in lung epithelial A549 and myeloid THP-1 cells, as determined by the intracellular staining of IL-8 by flow cytometry. (<b>D</b>) Representative dot-plots showing IL-8+ intracellular staining of lung epithelial A549 (<b>left panel</b>) and myeloid THP-1 (<b>right panel</b>) cells upon treatment with NLR agonists, as determined by flow cytometry compared to untreated (UN) cells.</p>
Full article ">Figure 2
<p>Cytokine response is preferentially triggered by NOD1 and dual NOD1/2 agonists in lung epithelial cells. (<b>A</b>) Cytokine response to NLR agonists triggered by NOD1-, NOD1/2- and NOD2-specific agonists in lung epithelial A549-Dual cells. Immune response was determined by the percentage of intracellular IL-8+ (<b>left</b>) and TNFα+ (<b>right</b>) cell quantification by flow cytometry after 24 h of treatment, using LPS (1 µg/mL, yellow bar) non-treated condition (UN, black bar) as controls. (<b>B</b>) Induction of the proinflammatory response upon treatment with increasing concentrations of NOD1, NOD1/2 and NOD2 ligands in lung epithelial A549-Dual cells after 24 h of treatment. The intracellular stainings of IL-8 and TNFα were determined by flow cytometry as subrogate representative markers of the proinflammatory response, using LPS and UN as controls. Mean ± SD of three independent experiments is shown. * <span class="html-italic">p</span> &lt; 0.05; *** <span class="html-italic">p</span> &lt; 0.001.</p>
Full article ">Figure 3
<p>NOD1 and NOD1/2 agonists induce innate immune activation in vitro in lung epithelial through the NF-κB and ISRE pathways. (<b>A</b>) Induction of the NF-κB activity triggered by NLR agonists upon recognition by the NOD1, NOD1/2 and NOD2 receptors in lung epithelial A549-Dual cells after 24 h of treatment. LPS (yellow bar) and Poly(I:C) (grey bar) were used as controls for NF-κB activation. (<b>B</b>) Assessment of NLR agonist activity on type I IFN response signaling by the quantification of interferon-stimulated response element (ISRE)-dependent gene expression in lung epithelial A549-Dual cells after 24 h of treatment. Values were relativized to the untreated (ND, black bar) condition. (<b>C</b>) Relative mRNA expression of IL-8, CXCL10 and ISG15 in A549-Dual treated cells with 50 µM of selected NOD agonists for 8 h measured by qPCR (normalized to GAPDH expression). Mean ± SD of three independent experiments is shown. * <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 4
<p>Activity of NOD1 and dual NOD1/2 agonists is specific in lung epithelial cells. (<b>A</b>) Gene expression of NOD1 receptor in A549-Dual cells transiently silenced with siRNA targeting NOD1 (siNOD1). Mock and non-specific siRNA (siNT) were used as controls. (<b>B</b>) Cell viability of A549-Dual cells treated with siNOD1 and siNT, using mock condition as control. Cell viability was determined by LIVE/DEAD staining and measured by flow cytometry. (<b>C</b>) Activity of NOD1 agonists (TriDAP and C12-iE-DAP), dual NOD1/2 (M-TriDAP) and NOD2 (MDP) in A549 cells treated with siNOD1. Intracellular staining of proinflammatory IL-8+ cells was determined by flow cytometry using the mock and siNT conditions, respectively. (<b>D</b>) Induction of the NF-κB and ISRE (<b>E</b>) activation pathways in A549-Dual cells treated with NOD1, dual NOD1/2 or NOD2 agonists with 50 µM of selective NOD1 inhibitor ML130, 50 µM of NOD1/2 inhibitor NOD-IN-1 or untreated (UNT), respectively. Red dotted line indicates the basal NF-κB (<b>left</b>) or ISRE activity (<b>right</b>) in A549-Dual cells. Mean ± SD of three independent experiments is shown. * <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 5
<p>NLR agonist-induced cytokine response is preferentially triggered by NOD2 in PBMCs. (<b>A</b>) Assessment of the cytokine response to NLR agonists triggered by specific NOD1, dual NOD1/2 and NOD2 agonists in PBMCs. The percentages of intracellular IL-1β+, TNFα+ and IL-6+ cells were measured as representative markers of the proinflammatory response. Values were relativized to the non-treated condition (ND, black bar). LPS (1 µg/mL, yellow bar) and PMA (50 ng/mL) + ionomycin (1 µM) were used as positive controls. (<b>B</b>) Dose-response induction of proinflammatory cytokines IL-1β+, TNFα+ and IL-6+ in PBMCs treated with increasing concentrations of TriDAP (NOD1), M-TriDAP (dual NOD1/2) and MDP (NOD2) agonists. Mean ± SD of three independent experiments is shown. * <span class="html-italic">p</span> &lt; 0.05; ** <span class="html-italic">p</span> &lt; 0.01.</p>
Full article ">Figure 6
<p>NOD1 and dual NOD1/2 agonists impair SARS-CoV-2 replication in lung epithelial cells. Pretreatment of lung epithelial A549-Dual cells for 3 h with increasing concentrations of NOD1 and dual NOD1/2 agonists preferentially inhibits SARS-CoV-2 replication. (<b>A</b>) Representative dot-plots of infected cells treated with NOD agonists as measured by flow cytometry. (<b>B</b>) Quantification of viral replication measured as the percentage of SARS-CoV-2-GFP+ cells determined by flow cytometry after 48 h of infection. Values were relativized to the untreated condition (INF, black bar). TLR3 agonist Poly(I:C) (light gray bars) was used as control for the induction of the innate immune response. Mean ± SD of three independent experiments is shown. * <span class="html-italic">p</span> &lt; 0.05; ** <span class="html-italic">p</span> &lt; 0.01; EC<sub>50</sub>: half maximal effective concentration.</p>
Full article ">
17 pages, 3518 KiB  
Article
Diagnosis of Canine Tumours and the Value of Combined Detection of VEGF, P53, SF and NLRP3 for the Early Diagnosis of Canine Mammary Carcinoma
by Ning-Yu Yang, Hui-Hua Zheng, Chao Yu, Yan Ye and Guang-Hong Xie
Animals 2024, 14(9), 1272; https://doi.org/10.3390/ani14091272 - 24 Apr 2024
Viewed by 707
Abstract
The average life of a dog is generally maintained at ten to fifteen years, and tumours are the predominant reason that leads to the death of dogs, especially canine mammary carcinoma. Therefore, early diagnosis of tumours is very important. In this study, tumor [...] Read more.
The average life of a dog is generally maintained at ten to fifteen years, and tumours are the predominant reason that leads to the death of dogs, especially canine mammary carcinoma. Therefore, early diagnosis of tumours is very important. In this study, tumor size, morphology, and texture could be seen through general clinical examination, tumor metastasis could be seen through imaging examination, inflammatory reactions could be seen through hematological examination, and abnormal cell morphology could be seen through cytological and histopathological examination. In the 269 malignant cases and 179 benign cases, we randomly selected 30 cases each, and an additional 30 healthy dogs were selected for the experiment (healthy dogs: dogs in good physical condition without any tumor or other diseases). We used RT-qPCR and ELISA to determine the relative expression of vascular endothelial growth factor (VEGF), tumor protein P53 (P53), serum ferritin (SF), and NOD-like receptor protein 3 (NLRP3) in 30 healthy dogs, 30 dogs with benign mammary tumours, and 30 dogs with malignant mammary tumours. In the results, the same expression trend was obtained both in serum and tissues, and the expression of the four markers was the highest in malignant mammary tumours, with highly significant differences compared with the benign and healthy/paracancerous groups. By plotting the ROC curves, it was found that the results of combined tests were better than a single test and the combination of the four markers was the best for the early diagnosis. In conclusion, this can assist the clinical early diagnosis to a certain extent, and also provides some references and assistance for the development of tumor detection kits in clinical practice. Full article
(This article belongs to the Special Issue Companion Animals’ Molecular Oncology)
Show Figures

Figure 1

Figure 1
<p>Pictures of clinical signs in dogs with tumours.</p>
Full article ">Figure 2
<p>Canine tumour imaging results.</p>
Full article ">Figure 3
<p>Cytological results in canine tumours.</p>
Full article ">Figure 4
<p>(<b>A</b>–<b>L</b>) show the histopathological results in canine tumours.</p>
Full article ">Figure 5
<p>(<b>A</b>–<b>D</b>) show the gel electrophoresis results of SF, NLRP3, VEGF and P53.</p>
Full article ">Figure 6
<p>The relative expression levels of markers NLRP3, P53, SF and VEGF mRNA. ① * represents a significant difference (<span class="html-italic">p</span> &lt; 0.05); ** represents a highly significant difference (<span class="html-italic">p</span> &lt; 0.01), and *** represents a highly significant difference (<span class="html-italic">p</span> &lt; 0.0001). ② ## represents a highly significant difference (<span class="html-italic">p</span> &lt; 0.01), and ### represents a highly significant difference (<span class="html-italic">p</span> &lt; 0.0001).</p>
Full article ">Figure 7
<p>The results of SF, P53, VEGF and NLRP3 determination in the serum of the three groups. ① * represents a significant difference (<span class="html-italic">p</span> &lt; 0.05); ** represents a highly significant difference (<span class="html-italic">p</span> &lt; 0.01), and *** represents a highly significant difference (<span class="html-italic">p</span> &lt; 0.0001). ② ## represents a highly significant difference (<span class="html-italic">p</span> &lt; 0.01), and ### represents a highly significant difference (<span class="html-italic">p</span> &lt; 0.0001).</p>
Full article ">Figure 8
<p>ROC curves of single and combined tumor markers for canine mammary tumor diagnosis.</p>
Full article ">
16 pages, 2222 KiB  
Article
NOD2 Responds to Dengue Virus Type 2 Infection in Macrophage-like Cells Interacting with MAVS Adaptor and Affecting IFN-α Production and Virus Titers
by Diana Alhelí Domínguez-Martínez, Mayra Silvia Pérez-Flores, Daniel Núñez-Avellaneda, Jesús M. Torres-Flores, Gloria León-Avila, Blanca Estela García-Pérez and Ma Isabel Salazar
Pathogens 2024, 13(4), 306; https://doi.org/10.3390/pathogens13040306 - 10 Apr 2024
Viewed by 1621
Abstract
In pathogen recognition, the nucleotide-binding domain (NBD) and leucine rich repeat receptors (NLRs) have noteworthy functions in the activation of the innate immune response. These receptors respond to several viral infections, among them NOD2, a very dynamic NLR, whose role in dengue virus [...] Read more.
In pathogen recognition, the nucleotide-binding domain (NBD) and leucine rich repeat receptors (NLRs) have noteworthy functions in the activation of the innate immune response. These receptors respond to several viral infections, among them NOD2, a very dynamic NLR, whose role in dengue virus (DENV) infection remains unclear. This research aimed to determine the role of human NOD2 in THP-1 macrophage-like cells during DENV-2 infection. NOD2 levels in DENV-2 infected THP-1 macrophage-like cells was evaluated by RT-PCR and Western blot, and an increase was observed at both mRNA and protein levels. We observed using confocal microscopy and co-immunoprecipitation assays that NOD2 interacts with the effector protein MAVS (mitochondrial antiviral signaling protein), an adaptor protein promoting antiviral activity, this occurring mainly at 12 h into the infection. After silencing NOD2, we detected increased viral loads of DENV-2 and lower levels of IFN-α in supernatants from THP-1 macrophage-like cells with NOD2 knock-down and further infected with DENV-2, compared with mock-control or cells transfected with Scramble-siRNA. Thus, NOD2 is activated in response to DENV-2 in THP-1 macrophage-like cells and participates in IFN-α production, in addition to limiting virus replication at the examined time points. Full article
(This article belongs to the Special Issue Emerging Arboviruses: Epidemiology, Vector Dynamics, and Pathogenesis)
Show Figures

Graphical abstract

Graphical abstract
Full article ">Figure 1
<p>Detection of NOD2 in THP-1 macrophage-like cells stimulated with L18-MDP or infected with DENV-2. (<b>A</b>). NOD2 protein was detected as early as 3 hpi in cells infected by DENV-2 (MOI of 10) and in the positive control with LPS (1 µg/mL). (<b>B</b>,<b>C</b>). NOD2 expression was assessed at different time points in THP-1 macrophage-like cells under different treatments: mock (white bars), positive control transfected with L18-MDP (pink bars) or DENV-2-infected at an MOI of 10 (dark blue bars). (<b>B</b>). <span class="html-italic">NOD2</span> mRNA was measured after 3, 6, and 12 h using end point RT-PCR using specific primers for human <span class="html-italic">NOD2</span> and <span class="html-italic">GAPDH</span>. Bar graph indicates fold change calculated by the ratio <span class="html-italic">NOD2</span>/<span class="html-italic">GAPDH.</span> (<b>C</b>). NOD2 protein expression was evaluated by Western blotting of the whole-cell lysates using an anti-NOD2 antibody and an anti-β-actin antibody as a loading control. Bar graph indicates the fold change calculated by the ratio NOD2/β-actin. All experiments were performed three times. Statistical differences were assessed with one-way ANOVA, where *** <span class="html-italic">p</span> &lt; 0.001, ** <span class="html-italic">p</span> &lt; 0.01, and * <span class="html-italic">p</span> &lt; 0.05. (<b>D</b>) Confocal micrograph of THP-1 macrophage-like cells that have undergone the treatments indicated in the images for 12 h, followed by immunostaining for NOD2 (green), with the nuclei contrasted (blue). Images show the percentage of NOD2-positive cells which were found after counting 100 cells in each condition. Pictures were cropped to improve presentation, the digital magnification is 2100×, and the scale bar indicates 10 µM.</p>
Full article ">Figure 2
<p>DENV-2 induces NOD2–MAVS colocalization in THP-1 macrophage-like cells. (<b>A</b>,<b>B</b>) Representative confocal micrographs shown in (<b>A</b>) THP-1 macrophage-like cells with a mock treatment and (<b>B</b>) DENV-2-infected cells at 6, 12, and 24 h. Fluorescence images show the merge fields of NOD2 (green), MAVS (red), and nuclei (blue). Arrows indicate the cells where colocalization was observed. The scale bars indicate 20 µm, magnification 630×. (<b>C</b>–<b>K</b>) Representative confocal micrographs of the colocalization (<b>C</b>–<b>E</b>) of NOD2 (green) and (<b>F</b>–<b>H</b>) MAVS (red) in THP-1 macrophage-like cells in the following groups: mock, DENV-2-infected, and positive control (PIC) at 12 h. (<b>I</b>–<b>K</b>) On the far right, the images of NOD2 and MAVS are merged, with a nuclear contrast (blue). Individual cells displayed in the square were selected for observation in greater detail. A digital magnification of 630× is illustrated on the left and one of 2100× on the right side of each column. The arrows indicate points of NOD2–MAVS colocalization. Each condition was repeated three times.</p>
Full article ">Figure 3
<p>DENV-2 induces the NOD2–MAVS interaction in THP-1 macrophage-like cells after 12 hpi. NOD2 interaction with MAVS was evaluated via immunoprecipitation (IP) and Co-IP assays in whole protein cell lysates of THP-1 macrophage-like cells. THP-1 macrophage-like cells were grouped as follows: mock, DENV-2-infected, and the positive control (cells transfected with PIC). (<b>A</b>) Western blot image of the IP with an anti-NOD2 antibody, and IB with the same antibody shows an increase in NOD2 production in the positive control (PIC), and in cells infected with DENV-2 at 12 h. (<b>B</b>). Western blot image of the Co-IP shows the interaction of NOD2–MAVS in the positive control (PIC) and in cells infected with DENV-2 (asterisk) at 12 hpi. Proteins were IP with an anti-NOD2 antibody and IB with an anti-MAVS antibody. The images show one representative experiment. Uncropped membranes are presented in <a href="#app1-pathogens-13-00306" class="html-app">Supplementary Figure S3A</a>).</p>
Full article ">Figure 4
<p>Analysis of THP-1 macrophage-like cells knocked down in <span class="html-italic">NOD2</span> by specific NOD2-siRNAs shows a diminished production of IFN-α in THP-1 macrophage-like cells infected with DENV-2. (<b>A</b>) Quantification of cellular viability in THP-1 macrophage-like cells were performed via MTT assay. Bar graphs represent the mean ± SEM of the percentage of cell viability in THP-1 macrophage-like cells transfected with a Scramble-siRNA or NOD2-siRNAs and after 24 h of stimulation with the agonist PIC or untreated. A paired <span class="html-italic">t</span>-test was performed for statistical analysis to compare UT vs. PIC in each group; no differences were observed. N = 3. (<b>B</b>) Western blot of the silencing assay shows that NOD2 was downregulated in THP-1 macrophage-like cells transfected with specific NOD2-siRNAs. Bar graphs shown the percentage of the ratio NOD2/β-actin. (<b>C</b>) Western blot of THP-1 macrophage-like cells transfected with specific NOD2-siRNAs and stimulated with the agonist PIC produce lower levels of NOD2 protein than the cells transfected with the Scramble-siRNA and stimulated with the agonist PIC. (<b>D</b>) THP-1 macrophage-like cells were silenced with a mix of specific human NOD2-siRNAs or with a Scramble-siRNA (negative control); earlier, the cells were stimulated with PIC and infected with DENV-2; after 24 h, the cell supernatants were collected, and IFN-α levels were measured by ELISA. Bar graphs shows the IFN-α levels (pg/mL) mean ± SEM of three independent experiments. Statistical differences were evaluated by one-way ANOVA, where *** <span class="html-italic">p</span> &lt; 0.001, and ** <span class="html-italic">p</span> &lt; 0.01.</p>
Full article ">Figure 5
<p>Silencing <span class="html-italic">NOD</span>2 in THP-1 macrophage-like cells infected with DENV-2 increases viral titers and affects IFNα levels. (<b>A</b>) DENV-2 growth curve. THP-1 macrophage-like cells were untreated (UT) or transfected with a Scramble-siRNA or with an NOD2-siRNAs after all groups were infected with DENV-2 for 12, 24, and 48 h. The graph shows the viral titers of the three experimental groups expressed as FFU/mL mean ± SEM of three independent experiments at 12, 24, and 48 h. The DENV-2 growth curve shows comparisons between the viral loads in THP-1 macrophage-like cells treated with Scramble-siRNA or NOD2-siRNAs vs. untreated cells at each time point. The statistical differences to each time point of the curve were analyzed with one-way ANOVA. *** <span class="html-italic">p</span> &lt; 0.001, ** <span class="html-italic">p</span> &lt; 0.01, and * <span class="html-italic">p</span> &lt; 0.05. The viral progeny was quantified by focus-forming assay in C6/36 cells. (<b>B</b>) IFN-α levels were quantified in supernatants of THP-1 macrophage-like cells at 24 h post-infection with DENV-2 in mock cells or transfected with Scramble-siRNA or NOD2-siRNAs. Bar graph shows the IFN-α levels (pg/mL) mean ± SEM of three independent experiments. Statistical differences were evaluated by paired <span class="html-italic">t</span>-test, where ** <span class="html-italic">p</span> &lt; 0.01.</p>
Full article ">
13 pages, 1896 KiB  
Article
Vitamin D Mitigates Hepatic Fat Accumulation and Inflammation and Increases SIRT1/AMPK Expression in AML-12 Hepatocytes
by Eugene Chang
Molecules 2024, 29(6), 1401; https://doi.org/10.3390/molecules29061401 - 21 Mar 2024
Viewed by 1054
Abstract
Emerging evidence has demonstrated a strong correlation between vitamin D status and fatty liver disease. Aberrant hepatic fat infiltration contributes to oxidant overproduction, promoting metabolic dysfunction, and inflammatory responses. Vitamin D supplementation might be a good strategy for reducing hepatic lipid accumulation and [...] Read more.
Emerging evidence has demonstrated a strong correlation between vitamin D status and fatty liver disease. Aberrant hepatic fat infiltration contributes to oxidant overproduction, promoting metabolic dysfunction, and inflammatory responses. Vitamin D supplementation might be a good strategy for reducing hepatic lipid accumulation and inflammation in non-alcoholic fatty liver disease and its associated diseases. This study aimed to investigate the role of the most biologically active form of vitamin D, 1,25-dihydroxyvitamin D (1,25(OH)2D), in hepatic fat accumulation and inflammation in palmitic acid (PA)-treated AML-12 hepatocytes. The results indicated that treatment with 1,25(OH)2D significantly decreased triglyceride contents, lipid peroxidation, and cellular damage. In addition, mRNA levels of apoptosis-associated speck-like CARD-domain protein (ASC), thioredoxin-interacting protein (TXNIP), NOD-like receptor family pyrin domain-containing 3 (NLRP3), and interleukin-1β (IL-1β) involved in the NLRP3 inflammasome accompanied by caspase-1 activity and IL-1β expression were significantly suppressed by 1,25(OH)2D in PA-treated hepatocytes. Moreover, upon PA exposure, 1,25(OH)2D-incubated AML-12 hepatocytes showed higher sirtulin 1 (SIRT1) expression and adenosine monophosphate-activated protein kinase (AMPK) phosphorylation. A SIRT1 inhibitor alleviated the beneficial effects of 1,25(OH)2D on PA-induced hepatic fat deposition, IL-1β expression, and caspase-1 activity. These results suggest that the favorable effects of 1,25(OH)2D on hepatic fat accumulation and inflammation may be, at least in part, associated with the SIRT1. Full article
Show Figures

Figure 1

Figure 1
<p>Effects of 1,25-dihydroxyvitamin D (1,25(OH)2D) on cell viability and lipid accumulation. Murine AML-12 hepatocytes were co-treated with palmitic acid (PA, 0.5 mM, 24 h) together with 1,25(OH)2D (0, 1, 10, or 100 nM for 24 h). The conversion to blue formazan (<b>A</b>) and intracellular triglyceride levels (<b>B</b>) were expressed as fold-change compared to vehicle control. The results are expressed as mean ± standard error of the mean. Experiments represent at least two or three independent experiments (<span class="html-italic">n</span> = 6–9 per group). ** <span class="html-italic">p</span> &lt; 0.01 compared to vehicle control. # <span class="html-italic">p</span> &lt; 0.05; ## <span class="html-italic">p</span> &lt; 0.01 compared to PA control.</p>
Full article ">Figure 2
<p>Influence of 1,25-dihydroxyvitamin D (1,25(OH)2D) on oxidative stress, lipid peroxidation, and intracellular damage. Murine AML-12 hepatocytes were co-treated with palmitic acid (PA, 0.5 mM for 24 h) together with 1,25(OH)2D (0, 1, 10, or 100 nM for 24 h). Reactive oxygen species (<b>A</b>), lipid peroxidation (<b>B</b>), and lactate dehydrogenase (<b>C</b>) were measured and expressed as fold-change compared to vehicle control. The results are expressed as mean ± standard error of the mean. Experiments represent at least two or three independent experiments (<span class="html-italic">n</span> = 6–9 per group). ** <span class="html-italic">p</span> &lt; 0.01 compared to the vehicle control. # <span class="html-italic">p</span> &lt; 0.05; ## <span class="html-italic">p</span> &lt; 0.01 compared to PA control.</p>
Full article ">Figure 3
<p>Effects of 1,25-dihydroxyvitamin D (1,25(OH)2D) on gene expression involved in NLRP3 inflammasome components (<b>A</b>), caspase-1 activity (<b>B</b>), and IL-1β contents (<b>C</b>) in palmitic acid (PA)-treated murine AML-12 hepatocytes. mRNA levels were determined by RT-PCR, normalized for all samples to β-actin, and presented as fold-change compared to vehicle control (Con). Caspase-1 activity and IL-1β levels were measured using commercial colorimetric ELISA kits, normalized to their respective protein levels, and expressed as fold-change compared to vehicle control (Con). Values are presented as mean ± standard error of the mean. Experiments represent two independent experiments (<span class="html-italic">n</span> = 8 per group). * <span class="html-italic">p</span> &lt; 0.05, ** <span class="html-italic">p</span> &lt; 0.01 compared to vehicle control (Con). # <span class="html-italic">p</span> &lt; 0.05, ## <span class="html-italic">p</span> &lt; 0.01 compared to PA control (PA). ASC, apoptosis-associated speck-like protein containing a CARD; IL, interleukin; NLRP3, NOD-, LRR- and pyrin domain-containing protein 3; TXNIP, thioredoxin interacting protein.</p>
Full article ">Figure 4
<p>Effects of 1,25(OH)2D on AMPK phosphorylation and SIRT1 protein abundance in palmitic acid (PA)-treated hepatocytes. AML-12 mouse hepatocytes were treated with 0.5 mM PA (lanes 3 and 4) together with 100 nM 1,25(OH)2D (lanes 2 and 4) for 24 h. The density of the signal was quantified, normalized by GAPDH or AMPK, and expressed as fold-change compared to vehicle control (Con, lane 1) (<b>A</b>). Representative Western blots for p-AMPK, AMPK, SIRT1, and GAPDH (<b>B</b>). Experiments were repeated at least twice (<span class="html-italic">n</span> = 4–6 per group). Values are presented as mean ± standard error of the mean. Bars with asterisks indicate significant differences compared to vehicle control (Con) or PA control (PA). ** <span class="html-italic">p</span> &lt; 0.01 compared to vehicle control (Con, lane 1). # <span class="html-italic">p</span> &lt; 0.05, ## <span class="html-italic">p</span> &lt; 0.01 compared to PA control (PA, lane 3). AMPK, adenosine monophosphate-activated protein kinase; GAPDH, glyceraldehyde-3-phosphate dehydrogenase; SIRT1, sirtuin1.</p>
Full article ">Figure 5
<p>Effects of 1,25(OH)2D on fat deposition and inflammation through SIRT1 in palmitic acid (PA)-treated AML-12 mouse hepatocytes. AML-12 cells were incubated with PA (0.5 mM. 24 h; lanes 3, 4, 7, and 8) to induce hepatic steatosis. Hepatocytes were co-treated with 1,25(OH)2D (100 nM, 24 h; lanes 2, 4, 6, and 8), combined with EX-527 (10 μM, 24 h; lanes 5, 6, 7, and 8). (<b>A</b>) Triglyceride concentration, (<b>B</b>) caspase-1 activity, and (<b>C</b>) IL-1β levels were measured using commercial colorimetric ELISA kits, normalized to their respective protein levels, and expressed as fold-change compared to vehicle control in the absence of SIRT1 inhibitor (lane 1). Values are expressed as means ± standard error of the mean (<span class="html-italic">n</span> = 6) of two independent experiments. ** <span class="html-italic">p</span> &lt; 0.01 compared to vehicle control in the absence of EX-527 (lane 1). # <span class="html-italic">p</span> &lt; 0.05, ## <span class="html-italic">p</span> &lt; 0.01 compared to PA control in the absence of EX-527 (lane 3). <span>$</span> <span class="html-italic">p</span> &lt; 0.05, <span>$</span><span>$</span> <span class="html-italic">p</span> &lt; 0.01 compared to vehicle control in the presence of EX-527 (lane 5). Different letters (a, b, c) indicate the statistically different effects of 1,25(OH)2D, determined by one-way analysis of variance (ANOVA), followed by Student–Newman–Keuls multiple comparison post hoc test.</p>
Full article ">
14 pages, 5245 KiB  
Article
Ameliorative Effect of Areca Nut Polyphenols on Adverse Effects Induced by Lipopolysaccharides in RAW264.7 Cells
by Luyan Zou, Shuhan Yi and Yuanliang Wang
Molecules 2024, 29(6), 1329; https://doi.org/10.3390/molecules29061329 - 16 Mar 2024
Viewed by 789
Abstract
In Asian regions, areca nuts are tropical fruits that are extensively consumed. The areca nut contains a lot of polyphenols and its safety is unknown. In this research, we investigated the effects of lipopolysaccharides (LPS) and areca nut polyphenols (ANP) on normal RAW264.7 [...] Read more.
In Asian regions, areca nuts are tropical fruits that are extensively consumed. The areca nut contains a lot of polyphenols and its safety is unknown. In this research, we investigated the effects of lipopolysaccharides (LPS) and areca nut polyphenols (ANP) on normal RAW264.7 cells. The results showed that LPS stimulated adverse effects in normal cells by affecting cytokine production. The GO analysis results mainly affected DNA repair, cell division, and enzyme activities. In the KEGG analysis results, the NOD-like receptor signaling pathway, which is related to NF-κB, MAPK, and the pro-inflammatory cytokines, is the most significant. In the protein–protein interaction network (PPI) results, significant sub-networks in all three groups were shown to be related to cytokine–cytokine receptor interaction. Collectively, our findings showed a comprehensive understanding of LPS-induced toxicity and the protective effects of ANP by RNA sequencing. Full article
Show Figures

Figure 1

Figure 1
<p>(<b>a</b>) The chart of gene expression distribution and (<b>b</b>) the heat map of sample correlation.</p>
Full article ">Figure 2
<p>The chart of principal component analysis.</p>
Full article ">Figure 3
<p>The differential gene volcano plot: (<b>a</b>) LPS/Con, (<b>b</b>) P160/Con, and (<b>c</b>) P320/Con.</p>
Full article ">Figure 4
<p>The GO ontology analysis: (<b>a</b>) LPS/Con, (<b>b</b>) LPS/Con, (<b>c</b>) P160/Con, (<b>d</b>) P160/Con, (<b>e</b>) P320/Con, and (<b>f</b>) P320/Con.</p>
Full article ">Figure 5
<p>The PPI results of LPS/Con. The darker the color, the higher the degree of the gene.</p>
Full article ">Figure 6
<p>The PPI results of P160/Con. The darker the color, the higher the degree of the gene.</p>
Full article ">Figure 7
<p>The PPI results of P320/Con. The darker the color, the higher the degree of the gene.</p>
Full article ">
20 pages, 6009 KiB  
Article
The Mechanism of Elizabethkingia miricola Infection of the Black Spotted Frog as Revealed by Multi-Omics Analysis
by Qingcong Wei, Dan Wang, Kaijin Wei, Bin Xu and Jin Xu
Fishes 2024, 9(3), 91; https://doi.org/10.3390/fishes9030091 - 28 Feb 2024
Cited by 1 | Viewed by 1268
Abstract
Elizabethkingia miricola (E. miricola) is a significant pathogen that causes the crooked head disease in black spotted frogs. This disease has plagued numerous frog farms in China and has resulted in substantial losses to the frog farming industry. Nonetheless, the exact [...] Read more.
Elizabethkingia miricola (E. miricola) is a significant pathogen that causes the crooked head disease in black spotted frogs. This disease has plagued numerous frog farms in China and has resulted in substantial losses to the frog farming industry. Nonetheless, the exact mechanism that causes the disease in frogs remains unknown. In this study, transcriptomic and microbiomic analyses were conducted to analyze frog samples infected with E. miricola to reveal the infection mechanism of the pathogen. Liver transcriptomic analysis indicated that the livers of infected frogs had 1469 differentially expressed genes when compared with an uninfected group. These DEGs are mainly involved in immunity and metabolism, including neutrophil extracellular trap formation, the NOD-like receptor signaling pathway, leukocyte transendothelial migration, chemokine signaling pathway, Fc gamma R-mediated phagocytosis, and “metabolism”-related pathways such as the pentose phosphate pathway, carbon metabolism, glycerophospholipid metabolism, and glycerolipid metabolism. Similarly, 4737 DEGs were found in the kidney of infected frogs. These DEGs are mainly involved in immunity, including neutrophil extracellular trap formation, the NOD-like receptor signaling pathway, B cell receptor signaling pathway, C-type lectin receptor signaling pathway, complement and coagulation cascade, and Toll-like receptor signaling pathway. Ten immune-associated DEGs were screened in liver and kidney DEGs, respectively. And it was hypothesized that E. miricola infection could influence the host immune response. Microbiome analysis results showed that some opportunistic pathogens such as Citrobacter, Shigella, and Providencia were significantly elevated (p < 0.05) in infected frogs. Additionally, functional prediction confirmed that most of the microbiota in infected frogs were linked to metabolism-related KEGG pathways. In this study, the screened genes linked to immunity showed an association with the gut microbiome. The majority of these genes were found to be linked with the abundance of opportunistic pathogens. The results showed that E. miricola infection led to the downregulation of immune and metabolic-related genes, which led to the inhibition of immune function and metabolic disorder, and then increased the abundance of opportunistic pathogens in the gut microbiota. The findings of this study offer a preliminary foundation for comprehending the pathogenic processes of E. miricola infection in black spotted frogs. Full article
(This article belongs to the Special Issue Prevention and Treatment of Aquaculture Animal Diseases)
Show Figures

Figure 1

Figure 1
<p>“Volcano plot” of DEGs in liver. The red dots represent genes that are significantly upregulated. The yellow dots represent genes that are significantly downregulated. The blue dots represent genes that are no significant differences.</p>
Full article ">Figure 2
<p>“Volcano plot” of DEGs in kidney. The red dots represent genes that are significantly upregulated. The yellow dots represent genes that are significantly downregulated. The blue dots represent genes that are no significant differences.</p>
Full article ">Figure 3
<p>GO enrichment analysis of DEGs in liver.</p>
Full article ">Figure 4
<p>KEGG pathway analysis of DEGs in liver.</p>
Full article ">Figure 5
<p>GO enrichment analysis of DEGs in kidney.</p>
Full article ">Figure 6
<p>KEGG pathway enrichment analysis of DEGs in kidney.</p>
Full article ">Figure 7
<p>Microbial composition in the gut of black spotted frogs, (<b>a</b>,<b>b</b>) represent the dominant phyla and genera in groups.</p>
Full article ">Figure 8
<p>Bacterial taxa with significant differences (LDA score &gt; 2.0) in the relative abundance identified by Lefse in uninfected and infected groups (G24 vs. GE24).</p>
Full article ">Figure 9
<p>Bacterial taxa with significant differences (LDA score &gt; 2.0) in the relative abundance identified by Lefse in uninfected and infected groups (G-48 vs. GE-48).</p>
Full article ">Figure 10
<p>Function prediction of gut bacteria in black spotted frogs. G-48 and GE-48 represent the uninfected and infected groups, respectively.</p>
Full article ">Figure 11
<p>Co-occurrence network (FDR &lt; 0.01) between the genus level (top 30) of intestinal microbiota and immune-related DEGs in liver. The blue line shows the negative correlation between genes and genera; the red line shows the positive correlation between genes and genera.</p>
Full article ">Figure 12
<p>Co-occurrence network (FDR &lt; 0.01) between the genus level (top 30) of intestinal microbiota and immune-related DEGs in kidney. The blue line shows the negative correlation between genes and genera; the red line shows the positive correlation between genes and genera.</p>
Full article ">
17 pages, 5670 KiB  
Article
The Peptide LLTRAGL Derived from Rapana venosa Exerts Protective Effect against Inflammatory Bowel Disease in Zebrafish Model by Regulating Multi-Pathways
by Yongna Cao, Fenghua Xu, Qing Xia, Kechun Liu, Houwen Lin, Shanshan Zhang and Yun Zhang
Mar. Drugs 2024, 22(3), 100; https://doi.org/10.3390/md22030100 - 22 Feb 2024
Viewed by 1325
Abstract
Inflammatory bowel disease (IBD) is a chronic inflammatory bowel disease with unknown pathogenesis which has been gradually considered a public health challenge worldwide. Peptides derived from Rapana venosa have been shown to have an anti-inflammatory effect. In this study, peptide LLTRAGL derived from [...] Read more.
Inflammatory bowel disease (IBD) is a chronic inflammatory bowel disease with unknown pathogenesis which has been gradually considered a public health challenge worldwide. Peptides derived from Rapana venosa have been shown to have an anti-inflammatory effect. In this study, peptide LLTRAGL derived from Rapana venosa was prepared by a solid phase synthesis technique. The protective effects of LLTRAGL were studied in a 2,4,6-trinitrobenzene sulfonic acid (TNBS)-induced zebrafish colitis model. The underlying mechanisms of LLTRAGL were predicted and validated by transcriptome, real-time quantitative PCR assays and molecular docking. The results showed that LLTRAGL reduced the number of macrophages migrating to the intestine, enhanced the frequency and rate of intestinal peristalsis and improved intestinal inflammatory damage. Furthermore, transcriptome analysis indicated the key pathways (NOD-like receptor signal pathway and necroptosis pathway) that link the underlying protective effects of LLTRAGL’s molecular mechanisms. In addition, the related genes in these pathways exhibited different expressions after TNBS treatment. Finally, molecular docking techniques further verified the RNA-sequencing results. In summary, LLTRAGL exerted protective effects in the model of TNBS-induced colitis zebrafish. Our findings provide valuable information for the future application of LLTRAGL in IBD. Full article
Show Figures

Figure 1

Figure 1
<p>Effect of peptide LLTRAGL on TNBS-induced migration of zebrafish to intestinal macrophages. (<b>A</b>) Typical fluorescence images of Tg (<span class="html-italic">zlyz</span>: <span class="html-italic">EGFP</span>) macrophage migration in transgenic zebrafish juveniles. (<b>B</b>) Box chart showing the number of macrophages migrating from zebrafish larvae to the intestine. (<b>A</b>) the image on the right is an enlarged image of the white box on the left. Compared with the blank group, ## <span class="html-italic">p</span> ≤ 0.01; compared with the TNBS group, * <span class="html-italic">p</span> ≤ 0.05 and ** <span class="html-italic">p</span> ≤ 0.01.</p>
Full article ">Figure 2
<p>LLTRAGL improves TNBS-induced intestinal peristalsis damage. (<b>A</b>) Typical fluorescence images of intestinal efflux rate (IEE) of wild zebrafish juveniles. The blue box shows that after staining with calcein for 1.5 h, the drug was administered for 16 h. (<b>B</b>) Box chart showing the intestinal efflux rate of zebrafish juveniles. (<b>C</b>) Violin chart showing the number of intestinal peristalses in zebrafish juveniles. Compared with the blank group, ## <span class="html-italic">p</span> ≤ 0.01; compared with the TNBS group, * <span class="html-italic">p</span> ≤ 0.05 and ** <span class="html-italic">p</span> ≤ 0.01.</p>
Full article ">Figure 3
<p>Effects of the peptide on TNBS-induced intestinal tissue pathology and ultrastructure of zebrafish. (<b>A</b>) H&amp;E staining of intestinal tissues. Black arrow: sparse arrangement of cells; Blue arrow: Mucosal layer necrosis, cell lysis, enhanced cytoplasmic basophilia, disappearance of intestinal folds. Scale bar is 100 µm. (<b>B</b>) Alcian blue (AB) staining of intestinal tissues. Scale bar is 100 µm. (<b>C</b>) Electron microscopy of intestinal ultrastructure. Red arrow: intestinal microvilli status; Mv represents microvilli; Green arrow: goblet cell state; GC stands for goblet cell. Scale bar is 10.0 µm.</p>
Full article ">Figure 4
<p>The differentially expressed genes among the control group, TNBS group and peptide treatment group in RNA-Seq. (<b>A</b>) Statistical histogram of differentially expressed genes in TNBS vs. Control and LLTRAGL vs. TNBS. (<b>B</b>) TNBS vs. Control differential gene grouping cluster diagram. (<b>C</b>) LLTRAGL vs. TNBS differential gene grouping cluster diagram. (<b>D</b>) Venn diagram of common and unique differentially expressed genes between the TNBS vs. Control and LLTRAGL vs. TNBS comparative groups. The red color in the figure represents genes encoding relatively high expression proteins, while the blue color represents genes encoding relatively low expression proteins.</p>
Full article ">Figure 5
<p>GO enrichment analysis of DEGs and GSEA of all tested genes. (<b>A</b>) Circle plots of the distribution of DEGs in different GO categories. The circle chart shows the distribution of DEGs in the molecular function (MF), biological process (BP), and cellular composition (CC) categories. (<b>B</b>) Grouping clustering diagram of inflammatory response in the GSEA analysis. (<b>C</b>) Analysis Results of the Inflammatory Response Gene Set, mainly including the distribution map of enrichment score (ES), gene distribution map of gene set, and distribution map of the measurement and control bar sorting matrix.</p>
Full article ">Figure 6
<p>KEGG pathway enrichment analysis of DEGs. (<b>A</b>) KEGG-enriched top 10 chord diagram of TNBS vs. Control. (<b>B</b>) KEGG-enriched top 10 chord diagram of LLTRAGL-vs-TNBS. KEGG, Kyoto Encyclopedia of Genes and Genomes.</p>
Full article ">Figure 7
<p>Effect of LLTRAG on the inflammation-related gene expression in zebrafish, determined by RT-PCR. Compared with the blank group, # <span class="html-italic">p</span> ≤ 0.05 and ## <span class="html-italic">p</span> ≤ 0.01; compared with the TNBS group, * <span class="html-italic">p</span> ≤ 0.05 and ** <span class="html-italic">p</span> ≤ 0.01.</p>
Full article ">Figure 8
<p>Effect of LLTRAG on gene expression in TNBS-induced colitis in zebrafish. (<b>A</b>) NOD-like receptor signal pathway-related genes. (<b>B</b>) Inflammatory factors-related genes. (<b>C</b>) necroptosis signal pathway-related gene. Compared with the blank group, # <span class="html-italic">p</span> ≤ 0.05 and ## <span class="html-italic">p</span> ≤ 0.01; compared with the TNBS group, * <span class="html-italic">p</span> ≤ 0.05 and ** <span class="html-italic">p</span> ≤ 0.01.</p>
Full article ">
19 pages, 4171 KiB  
Article
Hesperitin-Copper(II) Complex Regulates the NLRP3 Pathway and Attenuates Hyperuricemia and Renal Inflammation
by Xi Peng, Kai Liu, Xing Hu, Deming Gong and Guowen Zhang
Foods 2024, 13(4), 591; https://doi.org/10.3390/foods13040591 - 15 Feb 2024
Viewed by 1060
Abstract
Background: Hyperuricaemia (HUA) is a disorder of purine metabolism in the body. We previously synthesized a hesperitin (Hsp)-Cu(II) complex and found that the complex possessed strong uric acid (UA)-reducing activity in vitro. In this study we further explored the complex’s UA-lowering and nephroprotective [...] Read more.
Background: Hyperuricaemia (HUA) is a disorder of purine metabolism in the body. We previously synthesized a hesperitin (Hsp)-Cu(II) complex and found that the complex possessed strong uric acid (UA)-reducing activity in vitro. In this study we further explored the complex’s UA-lowering and nephroprotective effects in vivo. Methods: A mouse with HUA was used to investigate the complex’s hypouricemic and nephroprotective effects via biochemical analysis, RT-PCR, and Western blot. Results: Hsp-Cu(II) complex markedly decreased the serum UA level and restored kidney tissue damage to normal in HUA mice. Meanwhile, the complex inhibited liver adenosine deaminase (ADA) and xanthine oxidase (XO) activities to reduce UA synthesis and modulated the protein expression of urate transporters to promote UA excretion. Hsp-Cu(II) treatment significantly suppressed oxidative stress and inflammatory in the kidney, reduced the contents of cytokines and inhibited the activation of the nucleotide-binding oligomerization domain (NOD)-like receptor thermal protein domain associated protein 3 (NLRP3) inflammatory pathway. Conclusions: Hsp-Cu(II) complex reduced serum UA and protected kidneys from renal inflammatory damage and oxidative stress by modulating the NLRP3 pathway. Hsp-Cu(II) complex may be a promising dietary supplement or nutraceutical for the therapy of hyperuricemia. Full article
(This article belongs to the Special Issue Polyphenols and Health Benefits)
Show Figures

Graphical abstract

Graphical abstract
Full article ">Figure 1
<p>Effect of Hsp-Cu(II) complex on (<b>A</b>) body weight, (<b>B</b>) kidney index, and (<b>C</b>) liver index in the HUA mice. Results are expressed as mean ± SD of 10 samples per group. Values with different superscript letters are significantly different (<span class="html-italic">p</span> &lt; 0.05).</p>
Full article ">Figure 2
<p>Effect of Hsp-Cu(II) complex on serum physicochemical indexes of (<b>A</b>) UA, (<b>B</b>) BUN, and (<b>C</b>) Cr in the HUA mice. Results are means ± standard deviation, <span class="html-italic">n</span> = 10 per group. The different letters represent a significant difference (<span class="html-italic">p</span> &lt; 0.05).</p>
Full article ">Figure 3
<p>Effect of Hsp-Cu(II) complex on the histopathology of kidney tissue of HUA mice. (<b>A</b>) Normal, (<b>B</b>) Diseased, (<b>C</b>) Alp, (<b>D</b>) Hsp, (<b>E</b>) LHC, (<b>F</b>) MHC, and (<b>G</b>) HHC. The red arrows pointed to the glomerular atrophy and necrosis; the blue arrows pointed to the tubular edema and fibrosis.</p>
Full article ">Figure 4
<p>Effect of Hsp-Cu(II) complex on the activity of (<b>A</b>) XO and (<b>B</b>) ADA in the liver of HUA mice. Results are means ± standard deviation, <span class="html-italic">n</span> = 10 per group. The different letters represent a significant difference (<span class="html-italic">p</span> &lt; 0.05).</p>
Full article ">Figure 5
<p>Effect of Hsp-Cu(II) complex on protein mRNA expression of the UA transporters in HUA mice. (<b>A</b>) Western blot analysis, the expression level of (<b>B</b>) URAT1, (<b>C</b>) GLUT9, and (<b>D</b>) OAT1. The mRNA expression level of (<b>E</b>) URAT1, (<b>F</b>) GLUT9, and (<b>G</b>) OAT1. Data are presented as the mean ± SD of three independent experiments. The different letters represent a significant difference (<span class="html-italic">p</span> &lt; 0.05).</p>
Full article ">Figure 6
<p>Effect of Hsp-Cu(II) complex on the activities of (<b>A</b>) T-SOD, (<b>B</b>) CAT, and (<b>C</b>) MDA level in HUA mice. Results are means ± standard deviation, <span class="html-italic">n</span> = 8 per group. The different letters represent a significant difference (<span class="html-italic">p</span> &lt; 0.05).</p>
Full article ">Figure 7
<p>Effect of Hsp-Cu(II) complex on the expression of the proteins involved in NLRP3 inflammatory pathway in the kidneys of HUA mice. (<b>A</b>) Western blot analysis, the content of (<b>B</b>) NLRP3, (<b>C</b>) Caspase-1, and (<b>D</b>) ASC. Data are presented as the mean ± SD of three independent experiments. The different letters represent a significant difference (<span class="html-italic">p</span> &lt; 0.05).</p>
Full article ">Figure 8
<p>Effect of the Hsp-Cu(II) complex on the level of (<b>A</b>) IL-1β, (<b>B</b>) IL-6, (<b>C</b>) TGF-β, and (<b>D</b>) TNF-α cytokines in the kidneys of HUA mice. Results are means ± standard deviation, <span class="html-italic">n</span> = 6 per group. The different letters represent a significant difference (<span class="html-italic">p</span> &lt; 0.05).</p>
Full article ">Figure 9
<p>Hsp-Cu(II) complex ameliorates potassium oxalate and hypoxanthine-induced hyperuricemia and kidney inflammation in mice through inhibiting XO and ADA activities and modulating the NLRP3 pathway.</p>
Full article ">
14 pages, 1828 KiB  
Article
Imeglimin Exhibits Novel Anti-Inflammatory Effects on High-Glucose-Stimulated Mouse Microglia through ULK1-Mediated Suppression of the TXNIP–NLRP3 Axis
by Hisashi Kato, Kaori Iwashita, Masayo Iwasa, Sayaka Kato, Hajime Yamakage, Takayoshi Suganami, Masashi Tanaka and Noriko Satoh-Asahara
Cells 2024, 13(3), 284; https://doi.org/10.3390/cells13030284 - 5 Feb 2024
Cited by 2 | Viewed by 1694
Abstract
Type 2 diabetes mellitus (T2DM) is an epidemiological risk factor for dementia and has been implicated in multifactorial pathologies, including neuroinflammation. In the present study, we aimed to elucidate the potential anti-inflammatory effects of imeglimin, a novel antidiabetic agent, on high-glucose (HG)-stimulated microglia. [...] Read more.
Type 2 diabetes mellitus (T2DM) is an epidemiological risk factor for dementia and has been implicated in multifactorial pathologies, including neuroinflammation. In the present study, we aimed to elucidate the potential anti-inflammatory effects of imeglimin, a novel antidiabetic agent, on high-glucose (HG)-stimulated microglia. Mouse microglial BV2 cells were stimulated with HG in the presence or absence of imeglimin. We examined the effects of imeglimin on the levels of proinflammatory cytokines, intracellular reactive oxygen species (ROS), mitochondrial integrity, and components related to the inflammasome or autophagy pathways in these cells. Our results showed that imeglimin suppressed the HG-induced production of interleukin-1beta (IL-1β) by reducing the intracellular ROS levels, ameliorating mitochondrial dysfunction, and inhibiting the activation of the thioredoxin-interacting protein (TXNIP)–NOD-like receptor family pyrin domain containing 3 (NLRP3) axis. Moreover, the inhibitory effects of imeglimin on the TXNIP–NLRP3 axis depended on the imeglimin-induced activation of ULK1, which also exhibited novel anti-inflammatory effects without autophagy induction. These findings suggest that imeglimin exerted novel suppressive effects on HG-stimulated microglia through the ULK1–TXNIP–NLRP3 axis, and may, thereby, contribute to the development of innovative strategies to prevent T2DM-associated cognitive impairment. Full article
Show Figures

Figure 1

Figure 1
<p>Effects of imeglimin on inflammatory mediators in high glucose (HG)-stimulated BV2 cells. Cells were pretreated with 500 μM imeglimin or the vehicle control for 30 min and then incubated for 24 h under low glucose or HG conditions. Expression levels of genes related to inflammatory mediators <span class="html-italic">Il-1β</span> (<b>A</b>), <span class="html-italic">Tnf-α</span> (<b>B</b>), and <span class="html-italic">Hmgb1</span> (<b>C</b>) were analyzed using real-time RT-PCR with 18S rRNA used as an internal control. The amounts of proinflammatory cytokines IL-1β (<b>D</b>) and TNF-α (<b>E</b>) in the culture supernatant were quantified using ELISA. HMGB1 protein levels in the microglia were determined through Western blot (<b>upper panel</b>: representative images) and densitometry (<b>lower panel</b>) (<b>F</b>). The amount of HMGB1 was normalized to that of β-actin. Expression levels are displayed relative to vehicle controls (1.0) (<b>A</b>–<b>C</b>,<b>F</b>). Data are presented as the mean ± SEM from three independent experiments (<span class="html-italic">n</span> = 3). * <span class="html-italic">p</span> &lt; 0.05, ** <span class="html-italic">p</span> &lt; 0.01, SEM: standard error of the mean. LG: low glucose, HG: high glucose, IMG: imeglimin, IL-1β: interleukin-1β, TNF-α: tumor necrosis factor alpha, HMGB1: high mobility group box 1.</p>
Full article ">Figure 2
<p>Effects of imeglimin on reactive oxygen species (ROS) generation, mitochondrial membrane potential (MMP), mitophagy, and apoptosis in high glucose (HG)-stimulated BV2 cells. Cells were pretreated with 500 μM imeglimin or the vehicle control for 30 min and then incubated for 24 h under low glucose or HG conditions. Intercellular ROS levels were examined using the DCFH-DA assay; fluorescence images (<b>A</b>) were obtained using a fluorescence microscope (scale bar, 50 μm), and the ROS levels (<b>B</b>) were quantified by measuring the fluorescence intensities. MMP levels were examined using the JC-1 assay; fluorescence images (<b>C</b>) were acquired (scale bar, 50 μm), and the MMP levels (<b>D</b>) were quantified by measuring the fluorescence intensities. Mitophagy components PINK1 and Parkin were analyzed using Western blot (<b>E</b>) and densitometry (<b>F</b>,<b>G</b>). Apoptosis-related proteins, cleaved cas-3 and PARP, were examined using Western blot (<b>H</b>) and densitometry (<b>I</b>,<b>J</b>). The amount of protein of interest was normalized to that of β-actin (<b>F</b>,<b>G</b>,<b>J</b>) or cas-3 (<b>I</b>). Fold changes are displayed relative to vehicle controls (1.0) (<b>B</b>,<b>D</b>,<b>F</b>,<b>G</b>,<b>J</b>). Data are presented as the mean ± SEM from three independent experiments (<span class="html-italic">n</span> = 3). * <span class="html-italic">p</span> &lt; 0.05, ** <span class="html-italic">p</span> &lt; 0.01, SEM: standard error of the mean. LG: low glucose, HG: high glucose, IMG: imeglimin, PINK1: phosphatase and tensin homolog (PTEN)-induced putative kinase 1, cleaved cas-3: cleaved caspase-3, cas-3: caspase-3, PARP: poly(ADP-ribose) polymerase, 2′,7′-DCFH-DA: dichloro-dihydro-fluorescein diacetate.</p>
Full article ">Figure 3
<p>Effects of imeglimin on the activation of the TXNIP–NLRP3 axis in high glucose (HG)-stimulated BV2 cells. Cells were pretreated with 500 μM imeglimin or the vehicle control for 30 min and then incubated for 24 h under low glucose or HG conditions. Proteins of interest were analyzed using Western blot (<b>A</b>). The levels of TXNIP (<b>B</b>), NLRP3 (<b>C</b>), or ASC (<b>D</b>) relative to β-actin and the levels of cleaved cas-1 (<b>E</b>) relative to cas-1 were determined using densitometry. Expression levels are displayed relative to vehicle controls (1.0) (<b>B</b>–<b>D</b>). Data are presented as the mean ± SEM from three independent experiments (<span class="html-italic">n</span> = 3). * <span class="html-italic">p</span> &lt; 0.05, ** <span class="html-italic">p</span> &lt; 0.01, SEM: standard error of the mean. LG: low glucose, HG: high glucose, IMG: imeglimin, TXNIP: thioredoxin-interacting protein, NLRP3: NOD-like receptor family pyrin domain containing 3, ASC: apoptosis-associated speck-like protein containing a caspase recruitment domain, cleaved cas-1: cleaved caspase-1, cas-1: caspase-1.</p>
Full article ">Figure 4
<p>Effects of imeglimin on autophagy component levels in high glucose (HG)-stimulated BV2 cells. Cells were pretreated with 500 μM imeglimin or the vehicle control for 30 min and then incubated for 24 h under low glucose or HG conditions. Expression levels of the proteins of interest were analyzed using Western blot (<b>A</b>) and densitometry (<b>B</b>–<b>F</b>); the protein levels of (<b>B</b>) ATG7 relative to β-actin; (<b>C</b>) LC3-II relative to LC3-I; (<b>D</b>) p62 relative to β-actin; (<b>E</b>) p-AMPKα relative to total AMPKα; (<b>F</b>) p-ULK1 relative to total ULK1. Protein levels are displayed relative to vehicle controls (1.0) (<b>B</b>,<b>D</b>). Data are presented as the mean ± SEM from three independent experiments (<span class="html-italic">n</span> = 3). * <span class="html-italic">p</span> &lt; 0.05, ** <span class="html-italic">p</span> &lt; 0.01, SEM: standard error of the mean. LG: low glucose, HG: high glucose, IMG: imeglimin, ATG: autophagy-related, LC3: microtubule-associated protein light chain 3, AMPK: adenosine monophosphate-activated protein kinase, t-AMPK: total AMPK, ULK1: UNC-51 like autophagy activating kinase 1, t-ULK1: total ULK1.</p>
Full article ">Figure 5
<p>Effects of imeglimin on inflammatory responses of high glucose (HG)-stimulated BV2 cells with or without ULK1 silencing. BV2 cells were transfected with ULK1 siRNA or control siRNA for 24 h and then subjected to treatment with 500 μM imeglimin or the vehicle control for 30 min, followed by incubation for 24 h under low glucose or HG conditions in the continued presence of imeglimin or the vehicle control. <span class="html-italic">Ulk1</span> mRNA levels 24 h after siRNA transfection were examined using quantitative RT-PCR and normalized to those of 18S rRNA (<b>A</b>). Protein levels of interest were determined using Western blot (<b>B</b>) and densitometry (<b>C</b>,<b>D</b>); (<b>C</b>) the amount of ULK1 relative to GAPDH; (<b>D</b>) the amount of p-ULK1 relative to total ULK1. <span class="html-italic">Il-1β</span> mRNA levels were examined using quantitative RT-PCR and normalized to those of 18S rRNA (<b>E</b>). Protein levels of TXNIP, NLRP3, and GAPDH were determined using Western blot (<b>F</b>). The amount of TXNIP (<b>G</b>) or NLRP3 (<b>H</b>) relative to GAPDH was obtained through densitometry. Fold changes are displayed relative to vehicle controls (1.0) (<b>A</b>,<b>C</b>,<b>E</b>,<b>G</b>,<b>H</b>). Data are presented as the mean ± SEM from three independent experiments (<span class="html-italic">n</span> = 3). * <span class="html-italic">p</span> &lt; 0.05, ** <span class="html-italic">p</span> &lt; 0.01, n.s., not significant, SEM: standard error of the mean. CON: control, ULK1: UNC-51 like autophagy activating kinase 1, t-ULK1: total ULK1, LG: low glucose, HG: high glucose, IMG: imeglimin, TXNIP: thioredoxin-interacting protein, NLRP3: NOD-like receptor family pyrin domain containing 3, GAPDH: glyceraldehyde-3-phosphate dehydrogenase.</p>
Full article ">
13 pages, 1090 KiB  
Review
Contribution of Nucleotide-Binding Oligomerization Domain-like (NOD) Receptors to the Immune and Metabolic Health
by César Jeri Apaza, Marisol Días, Aurora García Tejedor, Lisardo Boscá and José Moisés Laparra Llopis
Biomedicines 2024, 12(2), 341; https://doi.org/10.3390/biomedicines12020341 - 1 Feb 2024
Viewed by 1130
Abstract
Nucleotide-binding oligomerization domain-like (NOD) receptors rely on the interface between immunity and metabolism. Dietary factors constitute critical players in the activation of innate immunity and modulation of the gut microbiota. The latter have been involved in worsening or improving the control and promotion [...] Read more.
Nucleotide-binding oligomerization domain-like (NOD) receptors rely on the interface between immunity and metabolism. Dietary factors constitute critical players in the activation of innate immunity and modulation of the gut microbiota. The latter have been involved in worsening or improving the control and promotion of diseases such as obesity, type 2 diabetes, metabolic syndrome, diseases known as non-communicable metabolic diseases (NCDs), and the risk of developing cancer. Intracellular NODs play key coordinated actions with innate immune ‘Toll-like’ receptors leading to a diverse array of gene expressions that initiate inflammatory and immune responses. There has been an improvement in the understanding of the molecular and genetic implications of these receptors in, among others, such aspects as resting energy expenditure, insulin resistance, and cell proliferation. Genetic factors and polymorphisms of the receptors are determinants of the risk and severity of NCDs and cancer, and it is conceivable that dietary factors may have significant differential consequences depending on them. Host factors are difficult to influence, while environmental factors are predominant and approachable with a preventive and/or therapeutic intention in obesity, T2D, and cancer. However, beyond the recognition of the activation of NODs by peptidoglycan as its prototypical agonist, the underlying molecular response(s) and its consequences on these diseases remain ill-defined. Metabolic (re)programming is a hallmark of NCDs and cancer in which nutritional strategies might play a key role in preventing the unprecedented expansion of these diseases. A better understanding of the participation and effects of immunonutritional dietary ingredients can boost integrative knowledge fostering interdisciplinary science between nutritional precision and personalized medicine against cancer. This review summarizes the current evidence concerning the relationship(s) and consequences of NODs on immune and metabolic health. Full article
(This article belongs to the Section Immunology and Immunotherapy)
Show Figures

Figure 1

Figure 1
<p>Schematic summary of the participation of immunonutritional agonists in interaction with the nucleotide-binding oligomerization domain-like (NODs) and ‘Toll-like’ (TLRs) receptors and their implication for non-communicable diseases (NCDs) and cancer.</p>
Full article ">Figure 2
<p>Schematic diagram summarizing the synergies between nucleotide-binding oligomerization domain-containing proteins (NOD) and Toll-like receptors in human immune and metabolic systems and their main agonists. meso-DAP, γ-D-glutamyl-meso-diaminopimelic acid; MDP, muramyl dipeptide; TAB1, mitogen-activated protein kinase kinase 7-interacting protein-1; Rip2, receptor-interacting protein 2.</p>
Full article ">
Back to TopTop