[go: up one dir, main page]
More Web Proxy on the site http://driver.im/
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 (604)

Search Parameters:
Keywords = pyroptosis

Order results
Result details
Results per page
Select all
Export citation of selected articles as:
15 pages, 1290 KiB  
Review
Dysregulated Signaling Pathways in Canine Mammary Tumor and Human Triple Negative Breast Cancer: Advances and Potential Therapeutic Targets
by Chen Mei, Ying Liu, Zhenyi Liu, Yan Zhi, Zhaoling Jiang, Xueze Lyu and Hongjun Wang
Int. J. Mol. Sci. 2025, 26(1), 145; https://doi.org/10.3390/ijms26010145 (registering DOI) - 27 Dec 2024
Abstract
In 2022, human breast cancer (HBC) and canine mammary tumors (CMTs) remained the most prevalent malignant tumors worldwide, with high recurrence and lethality rates, posing a significant threat to human and dog health. The development of breast cancer involves multiple signaling pathways, highlighting [...] Read more.
In 2022, human breast cancer (HBC) and canine mammary tumors (CMTs) remained the most prevalent malignant tumors worldwide, with high recurrence and lethality rates, posing a significant threat to human and dog health. The development of breast cancer involves multiple signaling pathways, highlighting the need for effective inhibitory drugs that target key proteins in these pathways. This article reviews the dysregulation of the EGFR, PI3K/AKT/mTOR, Hippo, pyroptosis, and PD-1/PD-L1 signaling pathways in HBC and CMT, as well as the corresponding drugs used to inhibit tumor growth, with the aim of providing theoretical support for the development of more efficient drugs. Full article
Show Figures

Figure 1

Figure 1
<p>EGFR signaling pathway inhibitors and therapeutic monoclonal antibodies.</p>
Full article ">Figure 2
<p>PI3K-AKT-mTOR signaling pathway inhibitors in HBC and CMT.</p>
Full article ">Figure 3
<p>Antibody–drug conjugates for the treatment of mammary cancer.</p>
Full article ">
31 pages, 1122 KiB  
Review
Therapeutic Significance of NLRP3 Inflammasome in Cancer: Friend or Foe?
by Aliea M. Jalali, Kenyon J. Mitchell, Christian Pompoco, Sudeep Poludasu, Sabrina Tran and Kota V. Ramana
Int. J. Mol. Sci. 2024, 25(24), 13689; https://doi.org/10.3390/ijms252413689 - 21 Dec 2024
Viewed by 481
Abstract
Besides various infectious and inflammatory complications, recent studies also indicated the significance of NLRP3 inflammasome in cancer progression and therapy. NLRP3-mediated immune response and pyroptosis could be helpful or harmful in the progression of cancer, and also depend on the nature of the [...] Read more.
Besides various infectious and inflammatory complications, recent studies also indicated the significance of NLRP3 inflammasome in cancer progression and therapy. NLRP3-mediated immune response and pyroptosis could be helpful or harmful in the progression of cancer, and also depend on the nature of the tumor microenvironment. The activation of NLRP3 inflammasome could increase immune surveillance and the efficacy of immunotherapy. It can also lead to the removal of tumor cells by the recruitment of phagocytic macrophages, T-lymphocytes, and other immune cells to the tumor site. On the other hand, NLRP3 activation can also be harmful, as chronic inflammation driven by NLRP3 supports tumor progression by creating an environment that facilitates cancer cell proliferation, migration, invasion, and metastasis. The release of pro-inflammatory cytokines such as IL-1β and IL-18 can promote tumor growth and angiogenesis, while sustained inflammation may lead to immune suppression, hindering effective anti-tumor responses. In this review article, we discuss the role of NLRP3 inflammasome-mediated inflammatory response in the pathophysiology of various cancer types; understanding this role is essential for the development of innovative therapeutic strategies for cancer growth and spread. Full article
(This article belongs to the Section Molecular Pathology, Diagnostics, and Therapeutics)
Show Figures

Figure 1

Figure 1
<p>Canonical and non-canonical activation of NLRP3 inflammasome. The canonical pathway involves the activation of NLRP3 inflammasomes through signals such as mitochondrial ROS, calcium influx, and potassium efflux, leading to NF-κB activation and the production of pro-inflammatory cytokines (IL-1β and IL-18). This pathway ultimately activates caspase-1, resulting in cytokine release and pyroptosis. The non-canonical pathway involving LPS from Gram-negative bacteria triggers caspase-11, which indirectly activates NLRP3, leading to similar inflammasome responses, cytokine production, and pyroptosis. Both pathways generally play critical roles in innate immunity and inflammation.</p>
Full article ">Figure 2
<p>Significance of NLRP3 inflammasome in the melanoma progression. Various factors such as UV radiation, tumor microenvironment, and melanoma cells contribute to oxidative stress and cause DNA damage, immune cell activation, and cytokine release. Oxidative stress, in turn, triggers reactive oxygen species (ROS), mitochondrial DNA damage, potassium efflux, and NF-κB activation, which influence the activation of the NLRP3 inflammasome. NLRP3 activation promotes melanoma progression, metastasis, immune evasion, and therapy resistance. In contrast, inhibition of NLRP3 could enhance immunotherapy, inhibit tumor growth, and reduce metastasis.</p>
Full article ">Figure 3
<p>Significance of NLRP3 inflammasome in leukemias. Oxidative stress and mitochondrial dysfunction in leukemia cells could activate NLRP3 inflammasomes through mitochondrial ROS, potassium efflux, and NF-κB signaling pathways. NLRP3 activation leads to the generation of active IL-1β and IL-18 cytokines, and could cause pyroptosis. Further, NLRP3 activation plays various roles in different leukemias. For example, in Acute Myeloid Leukemia (AML), NLRP3 promotes immune evasion and survival, while inhibition reduces the disease burden. In Chronic Myeloid Leukemia (CML), NLRP3 is linked to KRAS mutations and therapy resistance. In Acute Lymphoblastic Leukemia (ALL), NLRP3 activation is correlated with glucocorticoid resistance, and in Chronic Lymphocytic Leukemia (CLL), P2X7R overexpression leads to increased NLRP3.</p>
Full article ">Figure 4
<p>Role of NLRP3 inflammasome in breast cancer growth and spread. Several factors such as reactive oxygen species (ROS)-induced mitochondrial damage, BRCA1-associated genetic mutations causing mitochondrial dysfunction, extracellular ATP leading to P2X7R overexpression, and inflammatory cytokines that activate NF-κB-mediated inflammasome components could lead to activation of NLRP3 inflammasome. NLRP3-mediated release of IL-1β and IL-18 promotes cancer cell proliferation, survival, migration, immune evasion, and resistance to therapy. Further, the outcomes also include increased tumor growth, metastasis, compromised immune surveillance, and drug resistance.</p>
Full article ">Figure 5
<p>Significance of NLRP3 inflammasome activation in cigarette smoke and COPD-induced lung cancer development. Cigarette smoke leads to reactive oxygen species (ROS), mitochondrial damage, and tissue dysfunction, activating the NLRP3 inflammasome. The release of pro-inflammatory cytokines IL-1β and IL-18 contributes to prolonged inflammation. COPD-induced inflammation and immune cell recruitment, such as macrophages, further amplify this process. The persistent inflammation and oxidative stress promote DNA damage, genetic mutations, and genomic instability, ultimately leading to lung carcinogenesis.</p>
Full article ">Figure 6
<p>Role of NLRP3 inflammasome activation in promoting colon cancer development. During inflammatory bowel disease (IBD), such as Crohn’s disease and ulcerative colitis, the inflammation of the gut lining activates immune cells, causing oxidative stress, mitochondrial damage, and the release of DAMPs. Gut microbiome imbalance (dysbiosis) leads to pathogenic bacterial growth and loss of gut-barrier integrity, allowing pathogen and toxin leakage, further driving oxidative stress. These pathways cause NLRP3 activation and trigger IL-1β and IL-18 release. Increased inflammasome response results in immune system imbalance, epithelial-barrier dysfunction, and neoplasia initiation, ultimately contributing to colon cancer growth and metastasis.</p>
Full article ">
28 pages, 3756 KiB  
Review
Unveiling the Emerging Role of Extracellular Vesicle–Inflammasomes in Hyperoxia-Induced Neonatal Lung and Brain Injury
by Karen Young, Merline Benny, Augusto Schmidt and Shu Wu
Cells 2024, 13(24), 2094; https://doi.org/10.3390/cells13242094 - 18 Dec 2024
Viewed by 572
Abstract
Extremely premature infants are at significant risk for developing bronchopulmonary dysplasia (BPD) and neurodevelopmental impairment (NDI). Although BPD is a predictor of poor neurodevelopmental outcomes, it is currently unknown how BPD contributes to brain injury and long-term NDI in pre-term infants. Extracellular vesicles [...] Read more.
Extremely premature infants are at significant risk for developing bronchopulmonary dysplasia (BPD) and neurodevelopmental impairment (NDI). Although BPD is a predictor of poor neurodevelopmental outcomes, it is currently unknown how BPD contributes to brain injury and long-term NDI in pre-term infants. Extracellular vesicles (EVs) are small, membrane-bound structures released from cells into the surrounding environment. EVs are involved in inter-organ communication in diverse pathological processes. Inflammasomes are large, multiprotein complexes that are part of the innate immune system and are responsible for triggering inflammatory responses and cell death. Apoptosis-associated speck-like protein containing a caspase recruitment domain (ASC) is pivotal in inflammasome assembly and activating inflammatory caspase-1. Activated caspase-1 cleaves gasdermin D (GSDMD) to release a 30 kD N-terminal domain that can form membrane pores, leading to lytic cell death, also known as pyroptosis. Activated caspase-1 can also cleave pro-IL-1β and pro-IL-18 to their active forms, which can be rapidly released through the GSDMD pores to induce inflammation. Recent evidence has emerged that activation of inflammasomes is associated with neonatal lung and brain injury, and inhibition of inflammasomes reduces hyperoxia-induced neonatal lung and brain injury. Additionally, multiple studies have demonstrated that hyperoxia stimulates the release of lung-derived EVs that contain inflammasome cargos. Adoptive transfer of these EVs into the circulation of normal neonatal mice and rats induces brain inflammatory injury. This review focuses on EV–inflammasomes’ roles in mediating lung-to-brain crosstalk via EV-dependent and EV-independent mechanisms critical in BPD, brain injury, and NDI pathogenesis. EV–inflammasomes will be discussed as potential therapeutic targets for neonatal lung and brain injury. Full article
(This article belongs to the Special Issue Perinatal Brain Injury—from Pathophysiology to Therapy)
Show Figures

Figure 1

Figure 1
<p>Mouse model of hyperoxia-induced BPD. Newborn mice were exposed to room air (RA) or 85% O<sub>2</sub> from P1 to P14. (<b>A</b>,<b>B</b>): Lung morphology was assessed by mean linear intercept (MLI) (<b>C</b>) and radial alveolar count (RAC) (<b>D</b>). Hyperoxia exposure increased MLI and decreased RAC, suggesting impaired alveolarization. (<b>E</b>,<b>F</b>): immunostaining for vWF (white arrows) and alpha-smooth muscle actin (α-SMA) (red arrows). Hyperoxia exposure reduced vWF+ vessel counts (<b>G</b>) and increased α-SMA positive vessels (<b>H</b>). Magnification: 20×. Scale bar: 50 μm. *** <span class="html-italic">p</span> &lt; 0.001. **** <span class="html-italic">p</span> &lt; 0.0001. Ref. [<a href="#B55-cells-13-02094" class="html-bibr">55</a>] with permission.</p>
Full article ">Figure 2
<p>Hyperoxia-induced neonatal brain injury in mice. Hyperoxia increases neuroinflammation and oxidative stress and decreases neurotrophins. These lead to brain cell death and impaired differentiation, altered brain microvascular development, and impaired myelination and axonal development. NGF: nerve growth factor. BDNF: brain-derived neurotrophic factor. VEGF: vascular endothelial growth factor. NT-3: neurotrophin-3. NT-4: neurotrophin-4. Ref. [<a href="#B68-cells-13-02094" class="html-bibr">68</a>] with permission.</p>
Full article ">Figure 3
<p>Structure, cargo, and function of extracellular vesicles. Extracellular vesicles (EVs) are composed of a lipid bilayer containing transmembrane proteins with a cargo consisting of proteins, mRNA, miRNA, DNA, and lipids. EVs can be isolated from various body fluids and have a diverse range of sizes ranging from 100 to 1000 nm. EVs isolated from the lung fluids and peripheral blood can be used as biomarkers for neonatal lung diseases. EVs have also been linked to the mediation of neonatal lung disease-associated brain injury.</p>
Full article ">Figure 4
<p>Characterization of EVs. (<b>A</b>,<b>B</b>): Nanoparticle tracking demonstrates that sizes and concentrations of EVs isolated from the plasma of pre-term infants on lower O<sub>2</sub> (LO<sub>2</sub>, &lt;30%) and higher O<sub>2</sub> (HO<sub>2</sub>, &gt;30%) on the seventh day of life. (<b>C</b>–<b>E</b>): transmission electron microscopy (TEM) shows the EV particles are smaller in size in the HO<sub>2</sub> group (red arrows). (<b>F</b>–<b>H</b>): Western blot analysis detects EV surface markers, CD9, CD63, and CD81. The CD63 and CD81 expression levels of the HO<sub>2</sub> group are lower than the LO<sub>2</sub> group. * <span class="html-italic">p</span> &lt; 0.05, ** <span class="html-italic">p</span> &lt; 0.01. Scale bar: 200 nm. Ref. [<a href="#B106-cells-13-02094" class="html-bibr">106</a>] with permission.</p>
Full article ">Figure 5
<p>Activation of NLRP3 inflammasome. NLRP3 inflammasome activation requires 2 steps: priming and activation. During the priming step, expression of NLRP3 and other inflammasome components is increased by activation of NF-κB upon PAMPs or DAMPs interaction with toll-like receptors (TLR). In the activation step, NLRP3 is activated by diverse stimuli and the formation of NLRP3 inflammasome that relies on homotypic interaction between the pyrin domain (PYD) and caspase-recruitment domain (CARD). ASC is recruited to cluster PRDs of oligomerized NLRP3 molecules, creating a platform for recruiting the effector caspase-1. The CARD of the procaspase-1 can interact with the aggregated CARDs of ASC, resulting in autolytic cleavage of pro-caspase-1 to P20 and P10 subunits that lead to caspase-1 activation. The activated caspase-1 can cleave and activate GSDMD to release an N-terminal domain that forms membrane pores and leads to pyroptosis. The caspase-1 can also cleave pro-IL-1β and pro-IL-18 to their active forms, IL-1β and IL-18, which can be rapidly released via the GSDMD pores into the extracellular space, leading to inflammation.</p>
Full article ">Figure 6
<p>GSDMD-KO reduces hyperoxia modulation of inflammatory, tissue remodeling, and developmental pathways in the neonatal lung. (<b>A</b>). Over-representation analysis using Toppcluster to identify similarities and dissimilarities of Gene Ontology terms and pathways modulated by hyperoxia in WT and GSDMD-KO lungs. Bars represent the log p-value, and the number of genes associated with each term is displayed at the end of the bar. In GSDMD-KO lungs, genes induced by hyperoxia were more strongly associated with TNF superfamily cytokine production, cellular extravasation, and cellular response to IFN-g, while suppressed genes in GSDMD-KO were uniquely associated with lobar bronchus epithelium development and B cell receptor signaling pathways. n = 3 animals/group. qRT-PCR validation of differentially expressed genes between hyperoxia-exposed WT and hyperoxia-exposed GSDMD-KO lungs included <span class="html-italic">Alas2</span> (<b>B</b>), <span class="html-italic">Scl4a1</span> (<b>C</b>), <span class="html-italic">Edn1</span> (<b>D</b>), <span class="html-italic">Mif</span> (<b>E</b>), <span class="html-italic">Pik3cg</span> (<b>F</b>), and <span class="html-italic">Trem2</span> (<b>G</b>). n = 4/group. * <span class="html-italic">p</span> &lt; 0.05, ** <span class="html-italic">p</span> &lt; 0.01, *** <span class="html-italic">p</span> &lt; 0.001, WT-O<sub>2</sub> vs. WT-RA. <sup>#</sup> <span class="html-italic">p</span> &lt; 0.05, <sup>##</sup> <span class="html-italic">p</span> &lt; 0.01, WT-O<sub>2</sub> vs. KO-O<sub>2</sub>. Ref. [<a href="#B55-cells-13-02094" class="html-bibr">55</a>] with permission.</p>
Full article ">Figure 7
<p>Adoptively transferred circulating EVs track to the lung and brain of normal neonatal rats. EVs isolated from the plasma of room air-maintained (RA) or hyperoxia-exposed (O<sub>2</sub>) rats were labeled with Exo-Glow and adoptively transferred into normal neonatal rats by intravenous injection. As illustrated in (<b>A</b>–<b>C</b>), both Exo-Glow labeled RA-EVs and O<sub>2</sub>-EVs rapidly distributed throughout the body and were localized in both lung and brain at 1 and 4 h after tail vein injection. Their homing to the brain (<b>D</b>) and lung (<b>E</b>) tissues was further confirmed by ex vivo imaging after dissection at 4 h. To determine if circulating EVs can cross BBB and are present in brain tissue for longer than 4 h, Dil-dye labeled EVs were similarly injected via tail veins, and at 24 h, EVs were isolated from the CSF and examined in brain tissues. The Dil signals were detected in the brain tissue sections of rats injected with both RA-EVs and O<sub>2</sub>-EVs but not from sham animals (<b>F</b>–<b>H</b>). Magnification: 20×. Scale bars: 50 μm. In addition, high concentrations of CSF EV particles were detected in animals that received either RA-EVs ((<b>J</b>), 17.55 ± 1.9 × 10<sup>7</sup>, n = 2 pooled of 3 CSF, <span class="html-italic">p</span> &lt; 0.01) or O<sub>2</sub>-EVs ((<b>K</b>), 26.7 ± 16.6 × 10<sup>7</sup>, n = 2 pooled of 3 CSF, <span class="html-italic">p</span> &lt; 0.05), compared to the sham animals ((<b>I</b>), 4.95 ± 0.57 × 10<sup>7</sup>, n = 4 pooled of 3 CSF). Overall, these results confirm that hyperoxia-induced circulating EVs can cross BBB and be taken up by brain cells. Ref. [<a href="#B56-cells-13-02094" class="html-bibr">56</a>] with permission.</p>
Full article ">Figure 8
<p>EV-inflammasome mediated lung-brain axis. The alveolar macrophages (AMs) in early injured lungs release EVs that contain an increased cargo of ASC. These EVs contribute to BPD pathogenesis by inducing lung inflammation and inhibiting alveolarization and vascularization. These EVs are released to the circulation, cross the BBB, and are taken up by neural cells. The ASC cargo can activate GSDMD in specific neural cells and result in brain injury by activating microglial cells and inducing cell death, possibly through pyroptosis mechanisms. The molecular and cellular changes can lead to long-term neurodevelopmental impairment (NDI). Ref. [<a href="#B106-cells-13-02094" class="html-bibr">106</a>] with permission.</p>
Full article ">
17 pages, 894 KiB  
Review
Mechanisms and Therapeutic Potential of Multiple Forms of Cell Death in Myocardial Ischemia–Reperfusion Injury
by Shinya Tsurusaki and Eddy Kizana
Int. J. Mol. Sci. 2024, 25(24), 13492; https://doi.org/10.3390/ijms252413492 - 17 Dec 2024
Viewed by 293
Abstract
Programmed cell death, especially programmed necrosis such as necroptosis, ferroptosis, and pyroptosis, has attracted significant attention recently. Traditionally, necrosis was thought to occur accidentally without signaling pathways, but recent discoveries have revealed that molecular pathways regulate certain forms of necrosis, similar to apoptosis. [...] Read more.
Programmed cell death, especially programmed necrosis such as necroptosis, ferroptosis, and pyroptosis, has attracted significant attention recently. Traditionally, necrosis was thought to occur accidentally without signaling pathways, but recent discoveries have revealed that molecular pathways regulate certain forms of necrosis, similar to apoptosis. Accumulating evidence indicates that programmed necrosis is involved in the development of various diseases, including myocardial ischemia–reperfusion injury (MIRI). MIRI occurs when blood flow and oxygen return to an ischemic area, causing excessive production of reactive oxygen species. While this reperfusion is critical for treating myocardial infarction, it inevitably causes cellular damage via oxidative stress. Furthermore, this cellular damage triggers multiple forms of cardiomyocyte death, which is the primary cause of inflammation, cardiac tissue remodeling, and ensuing heart failure. Therefore, understanding the molecular mechanisms of various forms of cell death in MIRI is crucial for therapeutic target discovery. Developing therapeutic strategies to inhibit multiple cell death pathways simultaneously could provide effective protection against MIRI. In this paper, we review the fundamental molecular pathways and MIRI-specific mechanisms of apoptosis, necroptosis, ferroptosis, and pyroptosis. Additionally, we suggest that the simultaneous suppression of multiple cell death pathways could be an effective therapy and identify potential therapeutic targets for implementing this strategy. Full article
(This article belongs to the Special Issue Advances in Cardiac Disease)
Show Figures

Figure 1

Figure 1
<p>The involvement of multiple cell death pathways in cardiomyocyte loss in MIRI. The blockage of the coronary artery by cholesterol plaque and blood clots causes hypoxia and damages cardiomyocytes. In addition, restoring blood supply via revascularization results in a further loss of cardiomyocytes through massive production of reactive oxygen species, causing reperfusion injury. In this process, various types of cell death are induced and involved in cardiomyocyte loss.</p>
Full article ">
19 pages, 31681 KiB  
Article
Comparison of Endoplasmic Reticulum Stress and Pyroptosis Induced by Pathogenic Calcium Oxalate Monohydrate and Physiologic Calcium Oxalate Dihydrate Crystals in HK-2 Cells: Insights into Kidney Stone Formation
by Wei-Jian Nong, Xin-Yi Tong and Jian-Ming Ouyang
Cells 2024, 13(24), 2070; https://doi.org/10.3390/cells13242070 - 15 Dec 2024
Viewed by 443
Abstract
Endoplasmic reticulum stress (ERS) can activate pyroptosis through CHOP and TXNIP; however, the correlation between this process and the formation of kidney stones has not been reported. The purpose is to investigate the effects of calcium oxalate monohydrate (COM) and calcium oxalate dihydrate [...] Read more.
Endoplasmic reticulum stress (ERS) can activate pyroptosis through CHOP and TXNIP; however, the correlation between this process and the formation of kidney stones has not been reported. The purpose is to investigate the effects of calcium oxalate monohydrate (COM) and calcium oxalate dihydrate (COD) on ERS and pyroptosis in HK-2 cells and to explore the formation mechanism of calcium oxalate stones. HK-2 cells were injured by 3 μm COM and COD. COM and COD significantly upregulated the expression levels of GRP78, CHOP, TXNIP, and pyroptosis-related proteins (NLRP3, caspase-1, GSDMD-N, and IL-1β). Fluorescence colocalization revealed that COM induced pyroptosis by inducing the interaction between TXNIP and NLRP3. Both COM and COD crystals can induce ERS and pyroptosis in HK-2 cells. COM induces the interaction with NLRP3 by the upregulation of CHOP and TXNIP and then promotes pyroptosis, while COD only promotes pyroptosis by the upregulation of CHOP. The cytotoxicity and the ability of COM to promote crystal adhesion and aggregation are higher than COD, suggesting that COM is more dangerous for calcium oxalate kidney stone formation. Full article
(This article belongs to the Collection The Role of NLRP3 in Health and Disease)
Show Figures

Figure 1

Figure 1
<p>Synthesis and characterization of COM and COD. (<b>A</b>) SEM; (<b>B</b>) the particle size distributions fitted to normal distribution curves (The red curve is a normal fitting distribution); (<b>C</b>) crystal XRD pattern; (<b>D</b>) zeta potential. Calcium oxalate monohydrate, COM. Calcium oxalate dihydrate, COD. Scanning electron microscope, SEM. X-ray diffraction, XRD. Data were extracted from independent samples, and experiments were performed in triplicate.</p>
Full article ">Figure 2
<p>Cytotoxicity of COM and COD and their differences in adhesion to HK-2 cells. (<b>A</b>) Cell viability was measured by CCK8; (<b>B</b>) microscope images of crystal adhesion after 1 h and 48 h exposure to HK-2 cells. Control: normal control group; COM: 3 μm COM with a concentration of 300 μg/mL; COD: 3 μm COD with a concentration of 300 μg/mL; comparison among different groups, * <span class="html-italic">p</span> &lt; 0.05; ** <span class="html-italic">p</span> &lt; 0.01; *** <span class="html-italic">p</span> &lt; 0.001. Scale bar: 20 μm. Calcium oxalate monohydrate, COM. Calcium oxalate dihydrate, COD. Data were extracted from independent samples, and experiments were performed in triplicate. The white box is the enlarged area, and the images pointed by the arrow is the enlarged images in the white box area.</p>
Full article ">Figure 3
<p>ERS induced by COM and COD. (<b>A</b>) The expression of GRP78 was observed by immunofluorescence (scale: 20 μm); (<b>B</b>) semi-quantitative analysis of GRP78 fluorescence images; (<b>C</b>,<b>G</b>) Western blot analysis of endoplasmic reticulum stress-related proteins; (<b>D</b>–<b>F</b>,<b>H</b>) semi-quantitative analysis histograms of IRE1α, ATF6, CHOP, and P-PERK, respectively. Control: normal control group; COM: 3 μm COM with a concentration of 300 μg/mL; COD: 3 μm COD with a concentration of 300 μg/mL; comparison among different groups, * <span class="html-italic">p</span> &lt; 0.05; ** <span class="html-italic">p</span> &lt; 0.01; and *** <span class="html-italic">p</span> &lt; 0.001. Calcium oxalate monohydrate, COM. Calcium oxalate dihydrate, COD. Glucose-regulated protein 78, GRP78. 4,6-diamino-2-phenylindole, DAPI. Inositol requiring enzyme 1α, IRE1α. Activating transcription factor-6, ATF6. C/EBP homologous protein, CHOP. Phosphorylated PERK, p-PERK. Data were extracted from independent samples, and experiments were performed in triplicate.</p>
Full article ">Figure 4
<p>COM- and COD-induced pyroptosis and their differences. (<b>A</b>) Double staining flow quantitative analysis of caspase-1/PI; (<b>B</b>) quantitative statistical histogram of pyroptosis; (<b>C</b>) caspase-1/PI double dye confocal observation, scale: 50 μm; (<b>D</b>) semi-quantitative analysis of IL-18 in supernatant after cell injury by Elisa. (<b>E</b>,<b>H</b>) Western blot analysis of pyroptosis related pathway proteins. (<b>F</b>,<b>G</b>,<b>I</b>,<b>J</b>) semi-quantitative histograms of NLRP3, pro-caspase-1, GSDMD-N, and Pro-IL-1β, respectively. Control: normal control group; COM: 3 μm COM with a concentration of 300 μg/mL; COD: 3 μm COD with a concentration of 300 μg/mL; comparison among different groups, * <span class="html-italic">p</span> &lt; 0.05; ** <span class="html-italic">p</span> &lt; 0.01; *** <span class="html-italic">p</span> &lt; 0.001. Calcium oxalate monohydrate, COM. Calcium oxalate dihydrate, COD. Propidium iodide, PI. N-terminal cleavage product of GSDMD, GSDMD-N. Interleukin-1β, IL-1β. NOD-like receptor thermal protein domain associated protein 3, NLRP3. Data were extracted from independent samples, and experiments were performed in triplicate. The FLICA-YVAD probe binds to caspase-1 and is excited as green fluorescence. PI binds to the nuclei of the cells with membrane rupture and was excited as red fluorescence. DAPI bound to the nuclei of all cells and was excited as blue fluorescence. More intense green and red fluorescence represents more intense pyroptosis.</p>
Full article ">Figure 5
<p>Activation effects of COM and COD on TXNIP. (<b>A</b>) Western blot analysis of TXNIP; (<b>B</b>) semi-quantitative analysis histogram of TXNIP; (<b>C</b>) visualization of the colocalization of NLRP3 and TXNIP in HK-2 cells by laser confocal microscopy, scale: 10 μm; (<b>D</b>) copositioning curve analysis diagram for the white line region of figure (<b>C</b>). Control: normal control group; COM: 3 μm COM with a concentration of 300 μg/mL; COD: 3 μm COD with a concentration of 300 μg/mL; comparison among different groups, *** <span class="html-italic">p</span> &lt; 0.001. Calcium oxalate monohydrate, COM. Calcium oxalate dihydrate, COD. Thioredoxin-interacting protein, TXNIP. NOD-like receptor thermal protein domain associated protein 3, NLRP3. Data were extracted from independent samples, and experiments were performed in triplicate. TXNIP is observed as red fluorescence. NLRP3 is observed as green fluorescence. DAPI binding nuclei is observed as blue fluorescence. The image on the far right is a magnified view of the red box.</p>
Full article ">Figure 6
<p>The mechanism of COM and COD damages HK-2 cells through the ERS–NLRP3 pyroptosis pathway and promotes the formation of kidney stones (by Figdraw). Calcium oxalate monohydrate, COM. Calcium oxalate dihydrate, COD. Glucose-regulated protein 78, GRP78. Endoplasmic reticulum stress, ERS. Activating transcription factor-6, ATF6. Inositol requiring enzyme 1α, IRE1α. C/EBP homologous protein, CHOP. Thioredoxin-interacting protein, TXNIP. NOD-like receptor thermal protein domain associated protein 3, NLRP3. N-terminal cleavage product of GSDMD, GSDMD-N. Interleukin-18, IL-18. Interleukin-1β, IL-1β. Arrows indicate activation or upregulation effects.</p>
Full article ">
20 pages, 1292 KiB  
Review
Unmasking the Invisible Threat: Biological Impacts and Mechanisms of Polystyrene Nanoplastics on Cells
by Wenxia Bu, Ye Cui, Yueyuan Jin, Xuehai Wang, Mengna Jiang, Ruiyao Huang, JohnPaul Otuomasiri Egbobe, Xinyuan Zhao and Juan Tang
Toxics 2024, 12(12), 908; https://doi.org/10.3390/toxics12120908 - 14 Dec 2024
Viewed by 488
Abstract
Polystyrene nanoplastics (PS-NPs), a pervasive component of plastic pollution, have emerged as a significant environmental and health threat due to their microscopic size and bioaccumulative properties. This review systematically explores the biological effects and mechanisms of PS-NPs on cellular systems, encompassing oxidative stress, [...] Read more.
Polystyrene nanoplastics (PS-NPs), a pervasive component of plastic pollution, have emerged as a significant environmental and health threat due to their microscopic size and bioaccumulative properties. This review systematically explores the biological effects and mechanisms of PS-NPs on cellular systems, encompassing oxidative stress, mitochondrial dysfunction, DNA damage, inflammation, and disruptions in autophagy. Notably, PS-NPs induce multiple forms of cell death, including apoptosis, ferroptosis, necroptosis, and pyroptosis, mediated through distinct yet interconnected molecular pathways. The review also highlights various factors that influence the cytotoxicity of PS-NPs, such as particle size, surface modifications, co-exposure with other pollutants, and protein corona formation. These complex interactions underscore the extensive and potentially hazardous impacts of PS-NPs on cellular health. The findings presented here emphasize the need for continued research on the mechanisms underlying PS-NP toxicity and the development of effective strategies for mitigating their effects, thereby informing regulatory frameworks aimed at minimizing environmental and biological risks. Full article
(This article belongs to the Section Exposome Analysis and Risk Assessment)
Show Figures

Figure 1

Figure 1
<p>Mechanistic overview of PS-NP-induced cellular impacts. This figure illustrates the key biological impacts of PS-NPs on cellular systems, including oxidative stress, mitochondrial dysfunction, DNA damage, inflammation, autophagy disruption, and various forms of cell death.</p>
Full article ">Figure 2
<p>Venn diagram of key cell death pathways induced by PS-NPs. Venn diagram illustrating the key pathways involved in different forms of cell death induced by PS-NPs. The diagram shows the overlapping mechanisms of apoptosis, ferroptosis, necroptosis, and pyroptosis.</p>
Full article ">
16 pages, 4966 KiB  
Article
Quercetin Ameliorates Acute Lethal Sepsis in Mice by Inhibiting Caspase-11 Noncanonical Inflammasome in Macrophages
by Eojin Kim, Deok-Hyeong Choi and Young-Su Yi
Molecules 2024, 29(24), 5900; https://doi.org/10.3390/molecules29245900 - 13 Dec 2024
Viewed by 334
Abstract
Quercetin is a natural polyphenolic flavonoid widely found in plants, fruits, and vegetables, and has been reported to play pharmacological roles in numerous pathogenic conditions. The anti-inflammatory effects of quercetin in various inflammatory conditions and diseases have been well-documented. However, its regulatory role [...] Read more.
Quercetin is a natural polyphenolic flavonoid widely found in plants, fruits, and vegetables, and has been reported to play pharmacological roles in numerous pathogenic conditions. The anti-inflammatory effects of quercetin in various inflammatory conditions and diseases have been well-documented. However, its regulatory role in noncanonical inflammasome activation has not yet been demonstrated. This study investigated the anti-inflammatory effects of quercetin in caspase-11 noncanonical inflammasome-activated inflammatory responses in macrophages and a mouse model of acute lethal sepsis. Quercetin protected J774A.1 macrophages from lipopolysaccharide (LPS)-induced cell death and caspase-11 noncanonical inflammasome-induced pyroptosis. It significantly decreased the production and mRNA expression of pro-inflammatory cytokines, such as interleukin (IL)-1β, IL-18, and IL-6, but not tumor necrosis factor (TNF)-α, and inflammatory molecules, such as nitric oxide (NO) and inducible NO synthase in caspase-11 noncanonical inflammasome-activated J774A.1 cells. Mechanistically, quercetin strongly suppressed the autoproteolysis and secretion of caspase-11 and the proteolysis of gasdermin D in caspase-11 noncanonical inflammasome-activated J774A.1 cells. However, quercetin did not inhibit the direct binding of caspase-11 to LPS. In vivo, the study revealed that quercetin increased the survival rate of mice with acute lethal sepsis and decreased serum levels of pro-inflammatory cytokines without causing significant toxicity. In conclusion, this study highlights quercetin-mediated anti-inflammatory action in inflammatory responses and acute lethal sepsis through a novel mechanism that targets the caspase-11 noncanonical inflammasome in macrophages, suggesting quercetin as a promising anti-inflammatory agent in natural medicine. Full article
(This article belongs to the Special Issue Natural Polyphenols in Human Health (Volume II))
Show Figures

Graphical abstract

Graphical abstract
Full article ">Figure 1
<p>Quercetin prevents pyroptosis induced by caspase-11 noncanonical inflammasome activation in macrophages. (<b>A</b>) Chemical structure of quercetin. (<b>B</b>) J774A.1 cells were treated with quercetin (5, 10, 20, 25, 50, and 100 μM) for 24 h, and cell viability was determined using an MTT assay. (<b>C</b>) J774A.1 cells pretreated with quercetin (5, 10, 20, 25, 50, and 100 μM) for 1 h were then treated with LPS (1 μg/mL) for 24 h, and cell viability was determined using an MTT assay. (<b>D</b>) J774A.1 cells pretreated with quercetin (5, 10, 15, 20, 25, 50, and 100 μM) for 1 h were treated with Pam3CSK4 (1 μg/mL) for 4 h and subsequently transfected with LPS (2.5 μg/mL) for 20 h. LDH levels released from the J774A.1 cells were determined. J774A.1 cells pretreated with either quercetin (10 and 15 μM) or diclofenac (15 μM) for 1 h were treated with Pam3CSK4 (1 μg/mL) for 4 h and transfected with LPS (2.5 μg/mL) for 20 h. (<b>E</b>) The cells were stained with PI, and the PI-stained cells were analyzed by flow cytometry. Numbers indicate % of pyroptotic cell death. (<b>F</b>) Cell morphology was observed and photographed under a light microscope. Red arrows indicate pyroptotic cells. (<b>G</b>) LDH levels released from J774A.1 cells were measured. (<b>H</b>) Cell viability was determined using an MTT assay. ** <span class="html-italic">p</span> &lt; 0.01 compared to a (−) control (<b>B</b>). <sup>##</sup> <span class="html-italic">p</span> &lt; 0.01 compared to a (−) control, * <span class="html-italic">p</span> &lt; 0.05 and ** <span class="html-italic">p</span> &lt; 0.01 compared to an LPS-treated control (<b>C</b>). <sup>##</sup> <span class="html-italic">p</span> &lt; 0.01 compared to Pam3CSK4-treated controls, and ** <span class="html-italic">p</span> &lt; 0.01 compared to LPS-transfected controls (<b>D</b>,<b>F</b>,<b>G</b>). Arrows indicate pyroptotic death of J774A.1 cells.</p>
Full article ">Figure 1 Cont.
<p>Quercetin prevents pyroptosis induced by caspase-11 noncanonical inflammasome activation in macrophages. (<b>A</b>) Chemical structure of quercetin. (<b>B</b>) J774A.1 cells were treated with quercetin (5, 10, 20, 25, 50, and 100 μM) for 24 h, and cell viability was determined using an MTT assay. (<b>C</b>) J774A.1 cells pretreated with quercetin (5, 10, 20, 25, 50, and 100 μM) for 1 h were then treated with LPS (1 μg/mL) for 24 h, and cell viability was determined using an MTT assay. (<b>D</b>) J774A.1 cells pretreated with quercetin (5, 10, 15, 20, 25, 50, and 100 μM) for 1 h were treated with Pam3CSK4 (1 μg/mL) for 4 h and subsequently transfected with LPS (2.5 μg/mL) for 20 h. LDH levels released from the J774A.1 cells were determined. J774A.1 cells pretreated with either quercetin (10 and 15 μM) or diclofenac (15 μM) for 1 h were treated with Pam3CSK4 (1 μg/mL) for 4 h and transfected with LPS (2.5 μg/mL) for 20 h. (<b>E</b>) The cells were stained with PI, and the PI-stained cells were analyzed by flow cytometry. Numbers indicate % of pyroptotic cell death. (<b>F</b>) Cell morphology was observed and photographed under a light microscope. Red arrows indicate pyroptotic cells. (<b>G</b>) LDH levels released from J774A.1 cells were measured. (<b>H</b>) Cell viability was determined using an MTT assay. ** <span class="html-italic">p</span> &lt; 0.01 compared to a (−) control (<b>B</b>). <sup>##</sup> <span class="html-italic">p</span> &lt; 0.01 compared to a (−) control, * <span class="html-italic">p</span> &lt; 0.05 and ** <span class="html-italic">p</span> &lt; 0.01 compared to an LPS-treated control (<b>C</b>). <sup>##</sup> <span class="html-italic">p</span> &lt; 0.01 compared to Pam3CSK4-treated controls, and ** <span class="html-italic">p</span> &lt; 0.01 compared to LPS-transfected controls (<b>D</b>,<b>F</b>,<b>G</b>). Arrows indicate pyroptotic death of J774A.1 cells.</p>
Full article ">Figure 2
<p>Quercetin suppresses the production of inflammatory mediators induced by caspase-11 noncanonical inflammasome activation in macrophages. J774A.1 cells pretreated with either quercetin (10 and 15 μM) or diclofenac (15 μM) for 1 h were treated with Pam3CSK4 (1 μg/mL) for 4 h and transfected with LPS (2.5 μg/mL) for 20 h. (<b>A</b>) IL-1β, (<b>B</b>) IL-18, (<b>C</b>) IL-6, and (<b>D</b>) TNF-α released in the cell culture media were quantified by ELISA. mRNA expressions of (<b>E</b>) IL-1β, (<b>F</b>) IL-18, (<b>G</b>) IL-6, (<b>H</b>) TNF-α, and (<b>K</b>) iNOS were determined by qPCR. (<b>I</b>) J774A.1 cells pretreated with quercetin (5, 10, 20, 25, 50, and 100 μM) for 1 h were treated with LPS (1 μg/mL) for 24 h, and NO levels in the cell culture media were determined by a Griess assay. (<b>J</b>) J774A.1 cells pretreated with either quercetin (10 and 15 μM) or diclofenac (15 μM) for 1 h were treated with Pam3CSK4 (1 μg/mL) for 4 h and transfected with LPS (2.5 μg/mL) for 20 h, and NO levels in the cell culture media were determined by a Griess assay. <sup>##</sup> <span class="html-italic">p</span> &lt; 0.01 compared to a (−) control, and ** <span class="html-italic">p</span> &lt; 0.01 compared to an LPS-treated control (<b>I</b>). <sup>##</sup> <span class="html-italic">p</span> &lt; 0.01 compared to Pam3CSK4-treated controls, and ** <span class="html-italic">p</span> &lt; 0.01 compared to LPS-transfected controls (<b>A</b>–<b>H</b>,<b>J</b>,<b>K</b>).</p>
Full article ">Figure 2 Cont.
<p>Quercetin suppresses the production of inflammatory mediators induced by caspase-11 noncanonical inflammasome activation in macrophages. J774A.1 cells pretreated with either quercetin (10 and 15 μM) or diclofenac (15 μM) for 1 h were treated with Pam3CSK4 (1 μg/mL) for 4 h and transfected with LPS (2.5 μg/mL) for 20 h. (<b>A</b>) IL-1β, (<b>B</b>) IL-18, (<b>C</b>) IL-6, and (<b>D</b>) TNF-α released in the cell culture media were quantified by ELISA. mRNA expressions of (<b>E</b>) IL-1β, (<b>F</b>) IL-18, (<b>G</b>) IL-6, (<b>H</b>) TNF-α, and (<b>K</b>) iNOS were determined by qPCR. (<b>I</b>) J774A.1 cells pretreated with quercetin (5, 10, 20, 25, 50, and 100 μM) for 1 h were treated with LPS (1 μg/mL) for 24 h, and NO levels in the cell culture media were determined by a Griess assay. (<b>J</b>) J774A.1 cells pretreated with either quercetin (10 and 15 μM) or diclofenac (15 μM) for 1 h were treated with Pam3CSK4 (1 μg/mL) for 4 h and transfected with LPS (2.5 μg/mL) for 20 h, and NO levels in the cell culture media were determined by a Griess assay. <sup>##</sup> <span class="html-italic">p</span> &lt; 0.01 compared to a (−) control, and ** <span class="html-italic">p</span> &lt; 0.01 compared to an LPS-treated control (<b>I</b>). <sup>##</sup> <span class="html-italic">p</span> &lt; 0.01 compared to Pam3CSK4-treated controls, and ** <span class="html-italic">p</span> &lt; 0.01 compared to LPS-transfected controls (<b>A</b>–<b>H</b>,<b>J</b>,<b>K</b>).</p>
Full article ">Figure 3
<p>Quercetin-mediated inhibitory mechanisms of caspase-11 noncanonical inflammasome in macrophages. J774A.1 cells pretreated with either quercetin (10 and 15 μM) or diclofenac (15 μM) for 1 h were treated with Pam3CSK4 (1 μg/mL) for 4 h and transfected with LPS (2.5 μg/mL) for 20 h. (<b>A</b>) Uncleaved and cleaved forms of caspase-11 in the whole cell lysates and cell culture media detected by Western blot analysis. (<b>B</b>) Uncleaved and cleaved forms of GSDMD in the whole cell lysates detected by Western blot analysis. (<b>C</b>) Direct interaction between Flag-caspase-11 expressed in HEK293 cells and biotin conjugated LPS (1 μg) in the absence or presence of quercetin (50, 100, and 150 μg) was detected by Western blot analysis. The bands were quantified and plotted using an ImageJ software (version 1.54k).</p>
Full article ">Figure 4
<p>Quercetin ameliorates LPS-induced acute lethal sepsis in mice. (<b>A</b>) Experimental design and schedule. C57BL/6 mice were intraperitoneally injected with quercetin (25 and 50 mg/kg) every 24 h for 7 days, followed by an intraperitoneal injection of LPS (30 mg/kg). (<b>B</b>) The survival rates, and (<b>D</b>) the body weights measured for 72 h. (<b>C</b>) Serum levels of IL-1β and IL-18 measured at 72 h in the mice. * <span class="html-italic">p</span> &lt; 0.05 compared to vehicle-injected controls (<b>B</b>). <sup>##</sup> <span class="html-italic">p</span> &lt; 0.01 compared to vehicle-injected controls, and ** <span class="html-italic">p</span> &lt; 0.01 compared to LPS-injected controls (<b>C</b>).</p>
Full article ">Figure 5
<p>Graphical summary of quercetin-mediated anti-inflammatory action through targeting caspase-11 noncanonical inflammasome in macrophages. Pro-caspase-11 senses intracellular LPS through direct interaction, forming LPS-pro-caspase-11 complexes, which then undergo oligomerization to generate the caspase-11 non-canonical inflammasome. The caspase-11 noncanonnical inflammasome is activate by the autoproteolysis of caspase-11, leading to: (1) the proteolytic activation of GSDMD and GSDMD pore-induced pyroptosis, (2) iNOS-catalyzed NO production, and (3) the caspase-1-mediated proteolytic maturation and release of pro-inflammatory cytokines through the GSDMD pores in macrophages. Quercetin inhibits the autoproteolytic activation of the caspase-11 noncanonical inflammasome (red box), leading to the suppression of pyroptosis and the release of pro-inflammatory cytokines and NO in macrophages. Additionally, it ameliorates LPS-induced acute lethal sepsis in mice. Black arrows indicate the activation pathways.</p>
Full article ">
22 pages, 16663 KiB  
Article
Gene-Silencing Therapeutic Approaches Targeting PI3K/Akt/mTOR Signaling in Degenerative Intervertebral Disk Cells: An In Vitro Comparative Study Between RNA Interference and CRISPR–Cas9
by Masao Ryu, Takashi Yurube, Yoshiki Takeoka, Yutaro Kanda, Takeru Tsujimoto, Kunihiko Miyazaki, Hiroki Ohnishi, Tomoya Matsuo, Naotoshi Kumagai, Kohei Kuroshima, Yoshiaki Hiranaka, Ryosuke Kuroda and Kenichiro Kakutani
Cells 2024, 13(23), 2030; https://doi.org/10.3390/cells13232030 - 9 Dec 2024
Viewed by 692
Abstract
The mammalian target of rapamycin (mTOR), a serine/threonine kinase, promotes cell growth and inhibits autophagy. The following two complexes contain mTOR: mTORC1 with the regulatory associated protein of mTOR (RAPTOR) and mTORC2 with the rapamycin-insensitive companion of mTOR (RICTOR). The phosphatidylinositol 3-kinase (PI3K)/Akt/mTOR [...] Read more.
The mammalian target of rapamycin (mTOR), a serine/threonine kinase, promotes cell growth and inhibits autophagy. The following two complexes contain mTOR: mTORC1 with the regulatory associated protein of mTOR (RAPTOR) and mTORC2 with the rapamycin-insensitive companion of mTOR (RICTOR). The phosphatidylinositol 3-kinase (PI3K)/Akt/mTOR signaling pathway is important in the intervertebral disk, which is the largest avascular, hypoxic, low-nutrient organ in the body. To examine gene-silencing therapeutic approaches targeting PI3K/Akt/mTOR signaling in degenerative disk cells, an in vitro comparative study was designed between small interfering RNA (siRNA)-mediated RNA interference (RNAi) and clustered regularly interspaced short palindromic repeat (CRISPR)–CRISPR-associated protein 9 (Cas9) gene editing. Surgically obtained human disk nucleus pulposus cells were transfected with a siRNA or CRISPR–Cas9 plasmid targeting mTOR, RAPTOR, or RICTOR. Both of the approaches specifically suppressed target protein expression; however, the 24-h transfection efficiency differed by 53.8–60.3% for RNAi and 88.1–89.3% for CRISPR–Cas9 (p < 0.0001). Targeting mTOR, RAPTOR, and RICTOR all induced autophagy and inhibited apoptosis, senescence, pyroptosis, and matrix catabolism, with the most prominent effects observed with RAPTOR CRISPR–Cas9. In the time-course analysis, the 168-h suppression ratio of RAPTOR protein expression was 83.2% by CRISPR–Cas9 but only 8.8% by RNAi. While RNAi facilitates transient gene knockdown, CRISPR–Cas9 provides extensive gene knockout. Our findings suggest that RAPTOR/mTORC1 is a potential therapeutic target for degenerative disk disease. Full article
Show Figures

Figure 1

Figure 1
<p>Schematic illustration of human disk intracellular PI3K/Akt/mTOR signaling pathway. The mTOR is a serine/threonine kinase that integrates nutrient signals to promote drive cell growth and division. It operates within the following two primary complexes: mTORC1 and mTORC2, which include RAPTOR and RICTOR, respectively. The downstream effectors of mTORC1, such as p70/S6K, are involved in controlling cell proliferation, mRNA translation, and protein synthesis, also associated with senescence and matrix catabolism. Autophagy is tightly suppressed by mTORC1 as well. The regulation of mTORC1 is mediated by the upstream class-I PI3K, with Akt serving as a crucial pro-survival mediator that prevents apoptosis. Furthermore, the negative feedback loop between p70/S6K and the class-I PI3K exists. To analyze the cascade-dependent functions of PI3K/Akt/mTOR signaling, gene suppression was performed using both siRNA-mediated RNAi-based and CRISPR–Cas9-based methods to target <span class="html-italic">mTOR</span> for both mTORC1 and mTORC2, <span class="html-italic">RAPTOR</span> for mTORC1, and <span class="html-italic">RICTOR</span> for mTORC2.</p>
Full article ">Figure 2
<p>Schematic illustration of the in vitro study design. Human degenerative intervertebral disk NP cells were surgically collected from patients who underwent lumbar discectomy or interbody fusion surgery. To retain the phenotype and replicate the physiologically hypoxic intervertebral disk environment, first-passage cells were cultured under 2% O<sub>2</sub> until they reached ~80% confluence. Gene knockdown and knockout targeting <span class="html-italic">mTOR</span>, <span class="html-italic">RAPTOR</span>, and <span class="html-italic">RICTOR</span> were performed using both siRNA-mediated RNAi and CRISPR–Cas9, respectively. After the cells were transfected for 24 h, the suppression of mTOR, RAPTOR, and RICTOR and autophagy were evaluated by Western blotting. The cell number was counted. Cell viability was measured using the CCK-8 assay to evaluate the toxicity associated with RNAi and CRISPR–Cas9. Additionally, to mimic the clinically relevant low-nutrient and inflammatory disease conditions, following siRNA or CRISPR–Cas9 treatment for 24 h, the cells were stimulated with pro-inflammatory IL-1β in serum-free DMEM for an additional 24 h. Subsequent analyses included evaluating the apoptosis, pyroptosis, senescence, and matrix metabolism using Western blotting, TUNEL staining for apoptosis, and SA-β-gal staining for senescence.</p>
Full article ">Figure 3
<p>RNAi and CRISPR–Cas9 enhance the selective suppression of mTOR, RAPTOR, and RICTOR in human disk NP cells. (<b>A</b>) Western blot analysis for brachyury, CD24, and tubulin in the total protein extracts from five different batches of human disk NP cells in DMEM with 10% FBS. (<b>B</b>) Western blot analysis for mTOR, RAPTOR, RICTOR, and tubulin in the total protein extracts of human disk NP cells 24 h post-transfection with <span class="html-italic">mTOR</span>, <span class="html-italic">RAPTOR</span>, <span class="html-italic">RICTOR</span>, or control siRNA with each of two different sequences (Seq. 1 and Seq. 2) in DMEM with 10% FBS to assess the expression levels of the target protein relative to tubulin. (<b>C</b>) Western blot analysis for mTOR, RAPTOR, RICTOR, and tubulin in the total protein extracts of human disk NP cells 24 h after transfection with <span class="html-italic">mTOR</span>, <span class="html-italic">RAPTOR</span>, <span class="html-italic">RICTOR</span>, or control CRISPR–Cas9 plasmid with each of the three different guide RNA sequences (Seq. 1, Seq. 2, and Seq. 3) in DMEM with 10% FBS to assess the expression levels of the target protein relative to tubulin. (<b>D</b>) Fluorescence for phase contrast (gray), GFP (green), DAPI (blue), and merged signals in human disk NP cells 24 h post-transfection with <span class="html-italic">mTOR</span>, <span class="html-italic">RAPTOR</span>, <span class="html-italic">RICTOR</span>, or control siRNA containing a GFP sequence in DMEM with 10% FBS to assess the transfection efficiency of the GFP-positive cells relative to the total DAPI-positive cells. (<b>E</b>) Morphological appearance of human disk NP cells 24 h post-transfection with <span class="html-italic">RAPTOR</span> siRNA or <span class="html-italic">RAPTOR</span> CRISPR–Cas9 plasmid in DMEM with 10% FBS to assess the number of adherent cells treated relative to the control. (<b>F</b>) CCK-8 assay in human disk NP cells 24 h post-transfection with control siRNA, control CRISPR–Cas9 plasmid, lipofection only, <span class="html-italic">RAPTOR</span> siRNA, or <span class="html-italic">RAPTOR</span> CRISPR–Cas9 plasmid in DMEM with 10% FBS to assess the viability of the cells treated relative to the control. Cells were counted in duplicated five random low-power fields (100×). Statistical analysis was performed using one-way repeated measures ANOVA with the Tukey–Kramer post hoc test. Data are presented with dot and box plots (<span class="html-italic">n</span> = 6). In (<b>A</b>), the immunoblots shown are all results from experiments with similar outcomes (<span class="html-italic">n</span> = 5). In (<b>B</b>–<b>E</b>), the immunoblots and cellular images shown represent typical results from the experiments with similar outcomes (<span class="html-italic">n</span> = 6).</p>
Full article ">Figure 4
<p>Selective suppression of RAPTOR/mTORC1 inhibits autophagy and p70/S6K but differentially induces Akt activation in human disk NP cells. (<b>A</b>) Western blot analysis for mTOR, RAPTOR, RICTOR, and tubulin in total protein extracts of human disk NP cells 24 h post-transfection with <span class="html-italic">mTOR</span>, <span class="html-italic">RAPTOR</span>, <span class="html-italic">RICTOR</span>, or control siRNA with the sequence showing the highest suppression efficiency in DMEM with 10% FBS to assess the expression levels of the target protein relative to tubulin. (<b>B</b>) Western blot analysis for mTOR, RAPTOR, RICTOR, and tubulin in total protein extracts of human disk NP cells 24 h after transfection with <span class="html-italic">mTOR</span>, <span class="html-italic">RAPTOR</span>, <span class="html-italic">RICTOR</span>, or control CRISPR–Cas9 plasmid with the sequence presenting the highest suppression efficiency in DMEM with 10% FBS to assess the expression levels of the target protein relative to tubulin. (<b>C</b>) Western blot analysis for Akt, phosphorylated Akt (p-Akt), p70/S6K, phosphorylated p70/S6K (p-p70/S6K), and tubulin in the total protein extracts of human disk NP cells 24 h post-transfection with <span class="html-italic">mTOR</span>, <span class="html-italic">RAPTOR</span>, <span class="html-italic">RICTOR</span>, or control siRNA in DMEM with 10% FBS. (<b>D</b>) Western blot analysis for Akt, p-Akt, p70/S6K, p-p70/S6K, and tubulin in the total protein extracts of human disk NP cells 24 h post-transfection with <span class="html-italic">mTOR</span>, <span class="html-italic">RAPTOR</span>, <span class="html-italic">RICTOR</span>, or control CRISPR–Cas9 plasmid in DMEM with 10% FBS. (<b>E</b>) Western blot analysis for LC3, p62/SQSTM1, and tubulin in the total protein extracts of human disk NP cells 24 h post-transfection with <span class="html-italic">mTOR</span>, <span class="html-italic">RAPTOR</span>, <span class="html-italic">RICTOR</span>, or control siRNA in DMEM with 10% FBS to assess the expression levels of the target protein relative to tubulin. (<b>F</b>) Western blot analysis for LC3, p62/SQSTM1, and tubulin in the total protein extracts of human disk NP cells 24 h after transfection with <span class="html-italic">mTOR</span>, <span class="html-italic">RAPTOR</span>, <span class="html-italic">RICTOR</span>, or control CRISPR–Cas9 plasmid in DMEM with 10% FBS to assess the expression levels of the target protein relative to tubulin. Statistical analysis was performed using the paired <span class="html-italic">t</span>-test or one-way repeated measures ANOVA with the Tukey–Kramer post hoc test. Data are presented with dot and box plots (<span class="html-italic">n</span> = 6). The immunoblots shown represent the typical results from experiments with similar outcomes (<span class="html-italic">n</span> = 6).</p>
Full article ">Figure 5
<p>Selective suppression of RAPTOR/mTORC1 inhibits apoptosis in human disk NP cells. (<b>A</b>) Western blot analysis for PARP, cleaved PARP, cleaved caspase-9, and tubulin in the total protein extracts of human disk NP cells 24 h post-transfection with <span class="html-italic">mTOR</span>, <span class="html-italic">RAPTOR</span>, <span class="html-italic">RICTOR</span>, or control siRNA in 10 ng/mL IL-1β-supplemented DMEM with 0% FBS. (<b>B</b>) Western blot analysis for PARP, cleaved PARP, cleaved caspase-9, and tubulin in the total protein extracts of human disk NP cells 24 h after transfection with <span class="html-italic">mTOR</span>, <span class="html-italic">RAPTOR</span>, <span class="html-italic">RICTOR</span>, or control CRISPR–Cas9 plasmid in 10 ng/mL IL-1β-supplemented DMEM with 0% FBS. (<b>C</b>) Fluorescence for TUNEL (green), DAPI (blue), and merged signals in human disk NP cells 24 h post-transfection with <span class="html-italic">mTOR</span>, <span class="html-italic">RAPTOR</span>, <span class="html-italic">RICTOR</span>, or control siRNA in 10 ng/mL IL-1β-supplemented DMEM with 0% FBS to assess the ratio of TUNEL-positive cells relative to the total DAPI-positive cells. (<b>D</b>) Fluorescence for TUNEL (green), DAPI (blue), and merged signals in human disk NP cells 24 h after transfection with <span class="html-italic">mTOR</span>, <span class="html-italic">RAPTOR</span>, <span class="html-italic">RICTOR</span>, or control CRISPR–Cas9 plasmid in 10 ng/mL IL-1β-supplemented DMEM with 0% FBS to assess the ratio of TUNEL-positive cells relative to the total DAPI-positive cells. Cells were counted in duplicated five random low-power fields (100×). Statistical analysis was performed using one-way repeated measures ANOVA with the Tukey–Kramer post hoc test. Data are presented with dot and box plots (<span class="html-italic">n</span> = 6). The immunoblots and cellular images shown represent typical results from the experiments with similar outcomes (<span class="html-italic">n</span> = 6).</p>
Full article ">Figure 6
<p>Selective suppression of RAPTOR/mTORC1 inhibits pyroptosis in human disk NP cells. (<b>A</b>) Western blot analysis for caspase-1, cleaved caspase-1, GSDMD, N-terminal GSDMD, and tubulin in the total protein extracts of human disk NP cells 24 h post-transfection with <span class="html-italic">mTOR</span>, <span class="html-italic">RAPTOR</span>, <span class="html-italic">RICTOR</span>, or control siRNA in 10 ng/mL IL-1β-supplemented DMEM with 0% FBS to assess the expression levels of the target protein relative to tubulin. (<b>B</b>) Western blot analysis for caspase-1, cleaved caspase-1, GSDMD, N-terminal GSDMD, and tubulin in the total protein extracts of human disk NP cells 24 h post-transfection with <span class="html-italic">mTOR</span>, <span class="html-italic">RAPTOR</span>, <span class="html-italic">RICTOR</span>, or control CRISPR–Cas9 plasmid in 10 ng/mL IL-1β-supplemented DMEM with 0% FBS to assess the expression levels of the target protein relative to tubulin. Statistical analysis was performed using one-way repeated measures ANOVA with the Tukey–Kramer post hoc test. Data are presented with dot and box plots (<span class="html-italic">n</span> = 6). The immunoblots shown represent typical results from the experiments with similar outcomes (<span class="html-italic">n</span> = 6).</p>
Full article ">Figure 7
<p>Selective suppression of RAPTOR/mTORC1 inhibits senescence in human disk NP cells. (<b>A</b>) Western blot analysis for p16/INK4A, p21/WAF1/CIP1, p53, and tubulin in the total protein extracts of human disk NP cells 24 h post-transfection with <span class="html-italic">mTOR</span>, <span class="html-italic">RAPTOR</span>, <span class="html-italic">RICTOR</span>, or control siRNA in 10 ng/mL IL-1β-supplemented DMEM with 0% FBS. (<b>B</b>) Western blot analysis for p16/INK4A, p21/WAF1/CIP1, p53, and tubulin in the total protein extracts of human disk NP cells 24 h after transfection with <span class="html-italic">mTOR</span>, <span class="html-italic">RAPTOR</span>, <span class="html-italic">RICTOR</span>, or control CRISPR–Cas9 plasmid in 10 ng/mL IL-1β-supplemented DMEM with 0% FBS. (<b>C</b>) Colorimetric assay for the SA-β-gal signals (blue, indicated by black arrowheads) in human disk NP cells 24 h post-transfection with <span class="html-italic">mTOR</span>, <span class="html-italic">RAPTOR</span>, <span class="html-italic">RICTOR</span>, or control siRNA in 10 ng/mL IL-1β-supplemented DMEM with 0% FBS to assess the ratio of SA-β-gal-positive cells relative to the total cells. (<b>D</b>) Colorimetric assay for the SA-β-gal signals (blue, indicated by black arrowheads) in human disk NP cells 24 h post-transfection with <span class="html-italic">mTOR</span>, <span class="html-italic">RAPTOR</span>, <span class="html-italic">RICTOR</span>, or control CRISPR–Cas9 plasmid in 10 ng/mL IL-1β-supplemented DMEM with 0% FBS to assess the ratio of SA-β-gal-positive cells relative to the total cells. Cells were counted in duplicated five random low-power fields (100×). Statistical analysis was performed using one-way repeated measures ANOVA with the Tukey–Kramer post hoc test. Data are presented with dot and box plots (<span class="html-italic">n</span> = 6). The immunoblots and cellular images shown represent typical results from the experiments with similar outcomes (<span class="html-italic">n</span> = 6).</p>
Full article ">Figure 8
<p>Selective suppression of RAPTOR/mTORC1 increases matrix anabolism through decreased catabolic enzymes in human disk NP cells. (<b>A</b>) Western blot analysis for aggrecan, COL2A1, and tubulin in the total protein extracts of human disk NP cells 24 h post-transfection with <span class="html-italic">mTOR</span>, <span class="html-italic">RAPTOR</span>, <span class="html-italic">RICTOR</span>, or control siRNA in 10 ng/mL IL-1β-supplemented DMEM with 0% FBS. (<b>B</b>) Western blot analysis for aggrecan, COL2A1, and tubulin in the total protein extracts of human disk NP cells 24 h post-transfection with <span class="html-italic">mTOR</span>, <span class="html-italic">RAPTOR</span>, <span class="html-italic">RICTOR</span>, or control CRISPR–Cas9 plasmid in 10 ng/mL IL-1β-supplemented DMEM with 0% FBS. (<b>C</b>) Western blot analysis for MMP-3, MMP-13, TIMP-1, and TIMP-2 in the supernatant protein extracts of human disk NP cells 24 h post-transfection with <span class="html-italic">mTOR</span>, <span class="html-italic">RAPTOR</span>, <span class="html-italic">RICTOR</span>, or control siRNA in 10 ng/mL IL-1β-supplemented DMEM with 0% FBS. (<b>D</b>) Western blot analysis for MMP-3, MMP-13, TIMP-1, and TIMP-2 in the supernatant protein extracts of human disk NP cells 24 h post-transfection with <span class="html-italic">mTOR</span>, <span class="html-italic">RAPTOR</span>, <span class="html-italic">RICTOR</span>, or control CRISPR–Cas9 plasmid in 10 ng/mL IL-1β-supplemented DMEM with 0% FBS. The immunoblots shown represent typical results from the experiments with similar outcomes (<span class="html-italic">n</span> = 6).</p>
Full article ">Figure 9
<p>RNAi facilitates transient <span class="html-italic">RAPTOR</span> gene knockdown but CRISPR–Cas9 provides extensive <span class="html-italic">RAPTOR</span> gene knockout in human disk NP cells. Western blot analysis for RAPTOR and tubulin in the total protein extracts from five different batches of human disk NP cells at 0, 24, 48, 72, 120, and 168 h post-transfection with <span class="html-italic">RAPTOR</span> siRNA or CRISPR–Cas9 plasmid in 10% FBS-supplemented DMEM with a media change every 48 h to assess the time-course expression levels of the RAPTOR protein relative to tubulin. Statistical analysis was performed using two-way repeated measures ANOVA with the Tukey–Kramer post hoc test. Data are represented as the mean ± standard deviation (<span class="html-italic">n</span> = 5). The immunoblots shown are all results from the experiments with similar outcomes (<span class="html-italic">n</span> = 5).</p>
Full article ">
16 pages, 2423 KiB  
Article
DENV-1 Infection of Macrophages Induces Pyroptosis and Causes Changes in MicroRNA Expression Profiles
by Qinyi Zhang, Sicong Yu, Zhangnv Yang, Xingxing Wang, Jianhua Li, Lingxuan Su, Huijun Zhang, Xiuyu Lou, Haiyan Mao, Yi Sun, Lei Fang, Hao Yan and Yanjun Zhang
Biomedicines 2024, 12(12), 2752; https://doi.org/10.3390/biomedicines12122752 - 30 Nov 2024
Viewed by 586
Abstract
Background: Dengue virus (DENV) is the most widespread mosquito-borne virus, which can cause dengue fever with mild symptoms, or progress to fatal dengue hemorrhagic fever and dengue shock syndrome. As the main target cells of DENV, macrophages are responsible for the innate immune [...] Read more.
Background: Dengue virus (DENV) is the most widespread mosquito-borne virus, which can cause dengue fever with mild symptoms, or progress to fatal dengue hemorrhagic fever and dengue shock syndrome. As the main target cells of DENV, macrophages are responsible for the innate immune response against the virus. Methods: In this study, we investigated the role of pyroptosis in the pathogenic mechanism of dengue fever by examining the level of pyroptosis in DENV-1-infected macrophages and further screened differentially expressed microRNAs by high-throughput sequencing to predict microRNAs that could affect the pyroptosis of the macrophage. Results: Macrophages infected with DENV-1 were induced with decreased cell viability, decreased release of lactate dehydrogenase and IL-1β, activation of NLRP3 inflammasome and caspase-1, cleavage of GSDMD to produce an N-terminal fragment bound to cell membrane, and finally induced macrophage pyroptosis. MicroRNA expression profiles were obtained by sequencing macrophages from all periods of DENV-1 infection and comparing with the negative control. Sixty-three microRNAs differentially expressed in both the early and later stages of infection were also identified. In particular, miR-223-3p, miR-148a-3p, miR-125a-5p, miR-146a-5p and miR-34a-5p were recognized as small molecules that may be involved in the regulation of inflammation. Conclusions: In summary, this study aimed to understand the pathogenic mechanism of DENV through relevant molecular mechanisms and provide new targets for dengue-specific therapy. Full article
(This article belongs to the Special Issue Pathogenic Mechanism and Biosafety of Pathogenic Microorganisms)
Show Figures

Figure 1

Figure 1
<p>Results of macrophage cell viability and LDH release. (<b>A</b>,<b>B</b>) Macrophage cell viability was detected using CCK-8; (<b>C</b>,<b>D</b>) LDH release results were detected by collecting cell supernatants in different culture states. ** <span class="html-italic">p</span> &lt; 0.01, *** <span class="html-italic">p</span> &lt; 0.001 compared with the negative control group; <sup>##</sup> <span class="html-italic">p</span> &lt; 0.01, <sup>###</sup> <span class="html-italic">p</span> &lt; 0.001 compared between the two groups.</p>
Full article ">Figure 2
<p>Results of macrophage IL-1β release and mRNA and protein expression level. (<b>A</b>,<b>B</b>) Supernatants were collected under different culture conditions, and IL-1β content was detected using an ELISA kit; (<b>C</b>,<b>D</b>) Relative quantification of <span class="html-italic">IL-1β</span> mRNA expression level by RT-qPCR. (<b>E</b>) IL-1β protein expression levels were detected using Western blot, and β-actin was used as an internal reference control. * <span class="html-italic">p</span> &lt; 0.05, ** <span class="html-italic">p</span> &lt; 0.01, *** <span class="html-italic">p</span> &lt; 0.001 compared with the negative control group; <sup>##</sup> <span class="html-italic">p</span> &lt; 0.01, <sup>###</sup> <span class="html-italic">p</span> &lt; 0.001 compared between the two groups.</p>
Full article ">Figure 3
<p>Results of the detection of macrophage pyroptosis-related gene mRNA and its protein expression level. (<b>A</b>–<b>F</b>) Detection results of the expression levels of pyroptosis-related gene mRNA under different culture conditions; (<b>G</b>) The expression levels of pyroptosis-related proteins in macrophages infected with DENV-1 (MOI = 1) for 24 h, 48 h and 72 h, where β-actin was used as an internal reference control. * <span class="html-italic">p</span> &lt; 0.05, ** <span class="html-italic">p</span> &lt; 0.01, *** <span class="html-italic">p</span> &lt; 0.001 compared with the negative control group; <sup>#</sup> <span class="html-italic">p</span> &lt; 0.05, <sup>###</sup> <span class="html-italic">p</span> &lt; 0.001 compared between the two groups.</p>
Full article ">Figure 4
<p>Transcriptome sequencing results and analysis after DENV-1 infection in macrophages. (<b>A</b>) Differentially expressed miRNAs in the infected group and the negative control group at different infection times, where blue indicates downregulation and red indicates upregulation; (<b>B</b>) GO enrichment analysis of differentially expressed miRNA target genes at h.p.i = 72 h, showing the top 10 entries of difference in each classification; (<b>C</b>) KEGG enrichment analysis of differentially expressed miRNA target genes at h.p.i = 72 h, showing the entries in the top 20 differences.</p>
Full article ">Figure 5
<p>Analysis and validation of differentially expressed miRNAs after DENV-1 infection of macrophages. (<b>A</b>) Venn diagram of miRNAs differentially expressed at 24 h, 48 h and 72 h of infection; (<b>B</b>) Heatmap of miRNA clustering for miRNAs that were differentially expressed at all infection periods, with the red colored-frame indicating upregulation and blue colored-frame indicating downregulation; (<b>C</b>) Detection of the expression levels of <span class="html-italic">miR-223-3p</span>, <span class="html-italic">miR-148a-3p</span>, <span class="html-italic">miR-125a-5p</span>, <span class="html-italic">miR-146a-5p</span> and <span class="html-italic">miR-34a-5p</span> by RT-qPCR in macrophages at all infection periods. ** <span class="html-italic">p</span> &lt; 0.01, *** <span class="html-italic">p</span> &lt; 0.001 compared with the negative control group.</p>
Full article ">
15 pages, 2544 KiB  
Article
Ganglioside GM1 Alleviates Propofol-Induced Pyroptosis in the Hippocampus of Developing Rats via the PI3K/AKT/NF-κB Signaling Cascade
by Zhiheng Zhang, Shan Du, Xinzhang Chen, Di Qiu, Siyao Li, Lin Han, Hui Bai and Ruifeng Gao
Int. J. Mol. Sci. 2024, 25(23), 12662; https://doi.org/10.3390/ijms252312662 - 25 Nov 2024
Viewed by 442
Abstract
In pediatric and intensive care units, propofol is widely used for general anesthesia and sedation procedures as a short-acting anesthetic. Multiple studies have revealed that propofol causes hippocampal injury and cognitive dysfunction in developing animals. As is known, GM1, a type of ganglioside, [...] Read more.
In pediatric and intensive care units, propofol is widely used for general anesthesia and sedation procedures as a short-acting anesthetic. Multiple studies have revealed that propofol causes hippocampal injury and cognitive dysfunction in developing animals. As is known, GM1, a type of ganglioside, plays a crucial role in promoting nervous system development. Consequently, this study explored whether GM1 mitigated neurological injury caused by propofol during developmental stages and investigated its underlying mechanisms. Seven-day-old SD rats or PC12 cells were used in this study for histopathological analyses, a Morris water maze test, a lactate dehydrogenase release assay, Western blotting, and an ELISA. Furthermore, LY294002 was employed to explore the potential neuroprotective effect of GM1 via the PI3K/AKT signaling cascade. The results indicated that GM1 exerted a protective effect against hippocampal morphological damage and pyroptosis as well as behavioral abnormalities following propofol exposure by increasing p-PI3K and p-AKT expression while decreasing p-p65 expression in developing rats. Nevertheless, the inhibitor LY294002, which targets the PI3K/AKT cascade, attenuated the beneficial effects of GM1. Our study provides evidence that GM1 confers neuroprotection and attenuates propofol-induced developmental neurotoxicity, potentially involving the PI3K/AKT/NF-κB signaling cascade. Full article
(This article belongs to the Section Molecular Neurobiology)
Show Figures

Graphical abstract

Graphical abstract
Full article ">Figure 1
<p>Effect of GM1 on hippocampus morphological injury induced by propofol in young rats. (<b>A</b>) The hippocampal sections underwent histological analysis using Nissl staining. The whole hippocampal region was examined at 40× magnification and the CA1 and CA3 regions were observed at 400× magnification. (<b>B</b>) Cell counting was conducted in the hippocampal CA1 and CA3 regions (n = 3), where 9 randomly selected fields of view (10<sup>4</sup> μm<sup>2</sup> each) were subjected to semi-quantitative analysis to determine cell density (n = 3). ** <span class="html-italic">p</span> &lt; 0.01 vs. Con, <sup>##</sup> <span class="html-italic">p</span> &lt; 0.01 vs. PPF, <sup>▲</sup> <span class="html-italic">p</span> &lt; 0.05, and <sup>▲▲</sup> <span class="html-italic">p</span> &lt; 0.01 vs. PPF + GM1.</p>
Full article ">Figure 2
<p>GM1 ameliorates cognitive deficits caused by propofol in young rats. (<b>A</b>) Escape latency to reach a platform. (<b>B</b>) Mean swimming trail. (<b>C</b>) Travel time to target region. (<b>D</b>) Time to the target platform (n = 10). ** <span class="html-italic">p</span> &lt; 0.01 vs. Con, <sup>##</sup> <span class="html-italic">p</span> &lt; 0.01 vs. PPF, and <sup>▲▲</sup> <span class="html-italic">p</span> &lt; 0.01 vs. PPF + GM1.</p>
Full article ">Figure 3
<p>GM1 modulates the PI3K/AKT/NF-κB signaling pathway in the hippocampus of young rats following exposure to propofol. (<b>A</b>) Immunoblots and representative images of the PI3K/AKT/NF-κB pathway. (<b>B</b>,<b>C</b>) Corresponding quantification analysis of PI3K, p-PI3K, AKT, p-AKT, p65, p-p65, p-PI3K/PI3K, p-AKT/AKT, and p-p65/p65 in the hippocampus (n = 3). * <span class="html-italic">p</span> &lt; 0.05, ** <span class="html-italic">p</span> &lt; 0.01 vs. Con, <sup>#</sup> <span class="html-italic">p</span> &lt; 0.05, <sup>##</sup> <span class="html-italic">p</span> &lt; 0.01 vs. PPF, and <sup>▲▲</sup> <span class="html-italic">p</span> &lt; 0.01 vs. PPF + GM1.</p>
Full article ">Figure 4
<p>GM1 attenuates propofol-induced pyroptosis and inflammation in the hippocampus of young rats. (<b>A</b>) Immunofluorescence images of the CA1 region (NF-κB/NLRP3/caspase-1 triple labeling); scale bar: 20 μm. (<b>B</b>–<b>E</b>) Immunoblot images and corresponding quantification of NLRP3, p20, and GSDMD-N protein in the hippocampus. (<b>F</b>,<b>G</b>) Content of inflammatory cytokines in the hippocampus ((<b>B</b>–<b>E</b>), n = 3; (<b>F</b>–<b>G</b>), n = 6). * <span class="html-italic">p</span> &lt; 0.05, ** <span class="html-italic">p</span> &lt; 0.01 vs. Con, <sup>#</sup> <span class="html-italic">p</span> &lt; 0.05, <sup>##</sup> <span class="html-italic">p</span> &lt; 0.01 vs. PPF, <sup>▲</sup> <span class="html-italic">p</span> &lt; 0.05, and <sup>▲▲</sup> <span class="html-italic">p</span> &lt; 0.01 vs. PPF + GM1.</p>
Full article ">Figure 5
<p>Impact of GM1 on propofol-induced cell viability in PC12 cells (n = 6). ** <span class="html-italic">p</span> &lt; 0.01 vs. Con, <sup>##</sup> <span class="html-italic">p</span> &lt; 0.01 vs. PPF, and <sup>▲▲</sup> <span class="html-italic">p</span> &lt; 0.01 vs. PPF + GM1.</p>
Full article ">Figure 6
<p>Regulation of the PI3K/AKT/NF-κB pathway by GM1 in PC12 cells following propofol exposure. (<b>A</b>) Immunoblots and representative bands of PI3K/AKT/NF-κB proteins. (<b>B</b>,<b>C</b>) Corresponding quantification analyses of PI3K, p-PI3K, AKT, p-AKT, p65, p-p65, p-PI3K/PI3K, p-AKT/AKT, and p-p65/p65 in the hippocampus (n = 3). * <span class="html-italic">p</span> &lt; 0.05, ** <span class="html-italic">p</span> &lt; 0.01 vs. Con, <sup>#</sup> <span class="html-italic">p</span> &lt; 0.05, <sup>##</sup> <span class="html-italic">p</span> &lt; 0.01 vs. PPF, <sup>▲</sup> <span class="html-italic">p</span> &lt; 0.05, and <sup>▲▲</sup> <span class="html-italic">p</span> &lt; 0.01 vs. PPF + GM1.</p>
Full article ">Figure 7
<p>GM1 downregulates propofol-induced pyroptosis and inflammation in vitro. (<b>A</b>) Immunoblots and representative bands of NLRP3, p20, and GSDMD-N. (<b>B</b>) Concentrations of LDH release in PC12 cells. Corresponding quantification of NLRP3 (<b>C</b>), p20 (<b>D</b>), and GSDMD-N (<b>E</b>). (<b>F</b>,<b>G</b>) Pro-inflammatory cytokine levels in PC12 cells ((<b>A</b>–<b>E</b>), n = 3; (<b>F</b>,<b>G</b>), n = 6). * <span class="html-italic">p</span> &lt; 0.05, ** <span class="html-italic">p</span> &lt; 0.01 vs. Con, <sup>#</sup> <span class="html-italic">p</span> &lt; 0.05, <sup>##</sup> <span class="html-italic">p</span> &lt; 0.01 vs. PPF, <sup>▲</sup> <span class="html-italic">p</span> &lt; 0.05, and <sup>▲▲</sup> <span class="html-italic">p</span> &lt; 0.01 vs. PPF + GM1.</p>
Full article ">
20 pages, 428 KiB  
Review
Cytomegalovirus Biology Viewed Through a Cell Death Suppression Lens
by Edward S. Mocarski
Viruses 2024, 16(12), 1820; https://doi.org/10.3390/v16121820 - 23 Nov 2024
Viewed by 704
Abstract
Cytomegaloviruses, species-specific members of the betaherpesviruses, encode an impressive array of immune evasion strategies committed to the manipulation of the host immune system enabling these viruses to remain for life in a stand-off with host innate and adaptive immune mechanisms. Even though they [...] Read more.
Cytomegaloviruses, species-specific members of the betaherpesviruses, encode an impressive array of immune evasion strategies committed to the manipulation of the host immune system enabling these viruses to remain for life in a stand-off with host innate and adaptive immune mechanisms. Even though they are species-restricted, cytomegaloviruses are distributed across a wide range of different mammalian species in which they cause systemic infection involving many different cell types. Regulated, or programmed cell death has a recognized potential to eliminate infected cells prior to completion of viral replication and release of progeny. Cell death also naturally terminates replication during the final stages of replication. Over the past two decades, the host defense potential of known programmed cell death pathways (apoptosis, necroptosis, and pyroptosis), as well as a novel mitochondrial serine protease pathway have been defined through studies of cytomegalovirus-encoded cell death suppressors. Such virus-encoded inhibitors prevent virus-induced, cytokine-induced, and stress-induced death of infected cells while also moderating inflammation. By evading cell death and consequent inflammation as well as innate and adaptive immune clearance, cytomegaloviruses represent successful pathogens that become a critical disease threat when the host immune system is compromised. This review will discuss cell death programs acquired for mammalian host defense against cytomegaloviruses and enumerate the range of modulatory strategies this type of virus employs to balance host defense in favor of lifelong persistence. Full article
(This article belongs to the Special Issue Immune Modulation by Human Cytomegalovirus)
30 pages, 2650 KiB  
Review
Neuroinflammation in Age-Related Neurodegenerative Diseases: Role of Mitochondrial Oxidative Stress
by Xenia Abadin, Cristina de Dios, Marlene Zubillaga, Elia Ivars, Margalida Puigròs, Montserrat Marí, Albert Morales, Marisa Vizuete, Javier Vitorica, Ramon Trullas, Anna Colell and Vicente Roca-Agujetas
Antioxidants 2024, 13(12), 1440; https://doi.org/10.3390/antiox13121440 - 22 Nov 2024
Viewed by 839
Abstract
A shared hallmark of age-related neurodegenerative diseases is the chronic activation of innate immune cells, which actively contributes to the neurodegenerative process. In Alzheimer’s disease, this inflammatory milieu exacerbates both amyloid and tau pathology. A similar abnormal inflammatory response has been reported in [...] Read more.
A shared hallmark of age-related neurodegenerative diseases is the chronic activation of innate immune cells, which actively contributes to the neurodegenerative process. In Alzheimer’s disease, this inflammatory milieu exacerbates both amyloid and tau pathology. A similar abnormal inflammatory response has been reported in Parkinson’s disease, with elevated levels of cytokines and other inflammatory intermediates derived from activated glial cells, which promote the progressive loss of nigral dopaminergic neurons. Understanding the causes that support this aberrant inflammatory response has become a topic of growing interest and research in neurodegeneration, with high translational potential. It has been postulated that the phenotypic shift of immune cells towards a proinflammatory state combined with the presence of immunogenic cell death fuels a vicious cycle in which mitochondrial dysfunction plays a central role. Mitochondria and mitochondria-generated reactive oxygen species are downstream effectors of different inflammatory signaling pathways, including inflammasomes. Dysfunctional mitochondria are also recognized as important producers of damage-associated molecular patterns, which can amplify the immune response. Here, we review the major findings highlighting the role of mitochondria as a checkpoint of neuroinflammation and immunogenic cell deaths in neurodegenerative diseases. The knowledge of these processes may help to find new druggable targets to modulate the inflammatory response. Full article
(This article belongs to the Special Issue Mitochondrial Oxidative Stress in Aging and Disease—2nd Edition)
Show Figures

Figure 1

Figure 1
<p>Activation of the cGAS–STING signaling pathway by cytosolic mtDNA. Mitochondrial oxidative stress favors the release of mtDNA. In the cytosol, mtDNA is recognized by cGAS, which then assembles into dimers and catalyzes the synthesis of cGAMP from ATP and GTP. cGAMP initiates the signaling cascade by binding to the ER-associated adaptor protein STING, which translocates to the Golgi apparatus, where it recruits and activates TBK1. TBK1 phosphorylates IRF3, which, in turn, targets the nucleus and transcribes interferon-stimulated genes. TBK1 can also relieve the inhibition of NF-κB, triggering the transcription of gene-encoding proinflammatory cytokines [<a href="#B72-antioxidants-13-01440" class="html-bibr">72</a>]. Abbreviations: ER, endoplasmic reticulum; cGAMP, cyclic 2′,3′-cyclic guanosine monophosphate; cGAS, cyclic GMP–AMP synthase; IRF3, interferon regulatory factor 3; NF-κB, nuclear factor kappa B; STING, stimulator of interferon genes; TBK1, TANK-binding kinase.</p>
Full article ">Figure 2
<p>Inflammasome induction. Inflammasome assembly can be triggered by a repertoire of PPRs that detect and bind microbial ligands (PAMPs) and endogenous molecules exposed under stress conditions (DAMPs). The activated sensor then oligomerizes with the adapter protein ASC through PYD interactions and recruits the CASP1 zymogen, which undergoes auto-proteolysis to generate catalytically active CASP1. Mature CASP1 then cleaves and activates the inflammatory cytokines pro-IL-1β and pro-IL-18 and the pore-forming protein GSDMD. Abbreviations: ASC, associated speck-like protein containing a CARD; CASP1, caspase 1; DAMPs, damage-associated molecular patterns; GSDMD, gasdermin D; PAMPs, pathogen-associated molecular patterns; PPRs, pattern recognition receptors; PYD, pyrin domain.</p>
Full article ">Figure 3
<p>Inflammasome components and their domain architecture. Members of the NLR receptor family contain NBD and LRR motifs. NLR family members may be further subdivided into NLRP (with a PYD motif) and NLRC (with a CARD motif). The AIM2 receptor is composed of a PYD and a dsDNA-binding Hin-200/HIN domain. Finally, the inflammasome scaffold protein pyrin consists of PYD, BBOX, and CC domains that are followed by a carboxy-terminal B30.2 domain in the human but not its murine ortholog. NLRP3, human (h)NLRP1, AIM2, and pyrin, containing a PYD, bind to the PYD of ASC, allowing the ASC to activate CASP1 by interacting with the CARD of pro-CASP1. Card-containing sensors such as mouse (m)NLRP1 and NLRC4 activate CASP1 by directly binding the CARD of pro-CASP1 without ASC or binding the paired ASC scaffold. Abbreviations: AIM2, absent in melanoma 2; ASC, associated speck-like protein containing a CARD; BBOX, B-Box-type zinc finger; CASP1, caspase 1; CARD, caspase recruitment domain; CC, coiled coil; HIN/Hin-200, hematopoietic interferon-inducible nuclear protein with a 200 amino acid repeat; NBD, nucleotide-binding domain; NLR, nucleotide-binding domain and leucine-rich repeat; LRR, leucine-rich repeat; PYD, pyrin domain.</p>
Full article ">Figure 4
<p>NLRP3 inflammasome induction by mtROS. The mtROS generated in complexes I and III of the electron transport chain confer the priming signal necessary for the NLRP3 inflammasome activation. The cytosolic spread of mtROS activates redox-sensitive transcription factors, such as NF-κB, which stimulate the expression of inflammasome-related genes. Additionally, excessive production of mtROS causes oxidative modifications of mitochondrial proteins, mtDNA, and membrane lipids such as cardiolipin, which favor inflammasome assembly. ox-mtDNA release requires a TLR-dependent priming signal that triggers mtDNA replication through IRF1-mediated induction of CMPK2, an enzyme responsible for dNTPs supply. Dysfunctional mitochondria also engage the NRF2-ARE signaling pathway and mitophagy, which inhibit inflammasome assembly. Abbreviations: ARE, antioxidant response element; CMPK2, cytidine/uridine monophosphate kinase 2; IRF1, interferon regulatory factor 1; NF-κB, nuclear factor kappa B; NRF2, nuclear factor-erythroid 2 (NFE2) p45-related factor 2; dNTPs, deoxynucleotide triphosphates; TLR, toll-like receptor.</p>
Full article ">Figure 5
<p>TNFR1-mediated survival and cell death pathways. Cell death signaling engaged by TNFα initiates with the adaptor TRADD protein that binds TNFR1 by its death domain and, in turn, serves as a signal to recruit RIPK1. Then, RIPK1 is polyubiquitinated by the combined action of TRAFs with c-IAPs and triggers the translocation of different ubiquitin-associated proteins and kinases, including OPTN, TAK1, and the serine/threonine-protein kinase TBK1. This signaling cascade promotes the stabilization of RIPK1 in complex I and the activation of pro-survival NF-κB and MAPK pathways. Conversely, inhibition of c-IAPs permits the removal of polyubiquitin chains of RIPK1 by the ubiquitin carboxyl-terminal hydrolase CYLD. RIPK1 is released from complex I and binds FADD and pro-CASP8 in complex IIa. TRAIL/TNFSF10 and FASL-mediated signaling can also participate in the assembly of complex IIa. Then, activated CASP8 can cleave the kinase domain of RIPK1, thereby triggering apoptosis; however, when CASP8 is deficient, RIPK1 remains active, and the signaling pathway is switched from apoptosis to necroptosis. Under necroptotic permissive conditions, RIPK1 interacts with RIPK3 through their RHIMs motifs, leading to the formation of a signaling platform, named necrosome (complex IIb). Upon necrosome assembly, RIPK3 forms homodimers, which leads to activating cis-autophosphorylation of T231/S232 in mouse and S227 in human RIPK3. Then, the kinase domain of RIPK3 binds to the C-terminal domain of MLKL and phosphorylates S345 in mice and T357/S358 in human MLKL, allowing the assembly of MLKL into oligomers. Ultimately, these MLKL polymers form pores in the plasma membrane that drive lytic cell death [<a href="#B156-antioxidants-13-01440" class="html-bibr">156</a>]. Abbreviations: CASP8, caspase 8; c-IAPs, cellular inhibitors of apoptosis; CYLD, CYLD lysine 63 deubiquitinase; FADD, Fas-associated death domain; FASL, Fas ligand; MAPK, mitogen-activated protein kinases; MLKL, mixed lineage kinase domain-like; NF-κB, nuclear factor kappa B; OPTN, optineurin; RIPK, receptor-interacting protein kinase; TAK1, transforming growth factor-β-activated kinase 1; TBK1, TANK-binding kinase; TRADD, TNFR1-associated death domain protein; TRAF, TNFR-associated factor; TRAIL/TNFSF10, TNF superfamily member 10; RHIM, RIP homotypic interaction motifs.</p>
Full article ">Figure 6
<p>Core mechanisms of ferroptosis. TF and LCN2 transport extracellular iron into the cell through their respective receptors. Meanwhile, ferroportin is responsible for exporting intracellular iron to maintain iron balance. Inside the cells, HMOX1 breaks down heme to release free Fe<sup>2+</sup>, while ferritin, the iron storage protein, regulates intracellular iron levels to prevent iron overload. Ferrous iron (Fe<sup>2+</sup>) converts peroxides into free radicals through the Fenton reaction. This process leads to excessive lipid peroxidation and ultimately triggers ferroptosis. Ferroptosis can be prevented by two main antioxidant systems. One involves GPx4, which reduces lipid peroxides in a GSH-dependent reaction. The other is mediated by FSP1, which regenerates ubiquinone (CoQ10) to act as a trap for lipid peroxyl radicals. The cystine–glutamate antiporter SLC7A11 (xCT system) mediates the uptake of cystine, which is used for GSH synthesis. Abbreviations: CoQ10, coenzyme Q10; FSP1, ferroptosis suppressor protein 1; GPX4, glutathione peroxidase 4; HMOX1, heme oxygenase 1; LCN2, lipocalin 2; (PL-PUFA) phospholipids PUFA; SLC7A11, solute carrier family 7 member 1; TF, transferrin.</p>
Full article ">Figure 7
<p>Molecular mechanisms of Cuproptosis. Extracellular copper can be transported within the cell via a copper ionophore (elesclomol, disulfiram, etc.). FDX1 reduces Cu<sup>2+</sup> to Cu<sup>+</sup>, facilitating the release of copper ions in the mitochondrial matrix. FDX1 also binds to LIAS to promote lipoyl moiety generation. Then, LIPT1 transfers the LA to target proteins, such as DLAT. Lipoylated mitochondrial proteins (specifically DLAT) exhibit a high affinity for Cu<sup>+</sup> binding, which causes their aggregation and loss of function. Moreover, excess of Cu<sup>+</sup> reduces Fe-S cluster biosynthesis, destabilizes Fe-S cluster-containing proteins and contributes to overload Cu-mediated mitochondrial dysfunction and cell death. Copper importer SLC31A1 and exporters ATP7A/B can regulate the degree of cuproptosis sensitivity by controlling the intracellular concentration of copper ions. Also, GSH can serve as an endogenous copper chelator and protect against cuproptosis. Abbreviations: ATP7A/B, ATPase copper transporting alpha and beta; DLAT, dihydrolipoyl transacetylase; FDX1, ferredoxin 1; GSH, glutathione; LA, lipoic acid; LIAS, lipoic acid synthase; LIPT1, lipoyltransferase 1; SLC31A1, solute carrier family 31 member 1.</p>
Full article ">
21 pages, 5705 KiB  
Article
Pyroptosis in Endothelial Cells and Extracellular Vesicle Release in Atherosclerosis via NF-κB-Caspase-4/5-GSDM-D Pathway
by Salman Shamas, Razia Rashid Rahil, Laveena Kaushal, Vinod Kumar Sharma, Nissar Ahmad Wani, Shabir H. Qureshi, Sheikh F. Ahmad, Sabry M. Attia, Mohammad Afzal Zargar, Abid Hamid and Owais Mohmad Bhat
Pharmaceuticals 2024, 17(12), 1568; https://doi.org/10.3390/ph17121568 - 22 Nov 2024
Viewed by 602
Abstract
Background: Pyroptosis, an inflammatory cell death, is involved in the progression of atherosclerosis. Pyroptosis in endothelial cells (ECs) and its underlying mechanisms in atherosclerosis are poorly understood. Here, we investigated the role of a caspase-4/5-NF-κB pathway in pyroptosis in palmitic acid (PA)-stimulated [...] Read more.
Background: Pyroptosis, an inflammatory cell death, is involved in the progression of atherosclerosis. Pyroptosis in endothelial cells (ECs) and its underlying mechanisms in atherosclerosis are poorly understood. Here, we investigated the role of a caspase-4/5-NF-κB pathway in pyroptosis in palmitic acid (PA)-stimulated ECs and EVs as players in pyroptosis. Methods: Human umbilical vein endothelial cells (HUVECs) were cultured in an endothelial cell medium, treated with Ox-LDL, PA, caspase-4/5 inhibitor, NF-κB inhibitor, and sEV release inhibitor for 24 h, respectively. The cytotoxicity of PA was determined using an MTT assay, cell migration using a scratch-wound-healing assay, cell morphology using bright field microscopy, and lipid deposition using oil red O staining. The mRNA and protein expression of GSDM-D, CASP4, CASP5, NF-κB, NLRP3, IL-1β, and IL-18 were determined with RT-PCR and Western blot. Immunofluorescence was used to determine NLRP3 and ICAM-1 expressions. Extracellular vesicles (EVs) were isolated using an exosome isolation kit and were characterized by Western blot and scanning electron microscopy. Results: PA stimulation significantly changed the morphology of the HUVECs characterized by cell swelling, plasma membrane rupture, and increased LDH release, which are features of pyroptosis. PA significantly increased lipid accumulation and reduced cell migration. PA also triggered inflammation and endothelial dysfunction, as evidenced by NLRP3 activation, upregulation of ICAM-1 (endothelial activation marker), and pyroptotic markers (NLRP3, GSDM-D, IL-1β, IL-18). Inhibition of caspase-4/5 (Ac-FLTD-CMK) and NF-κB (trifluoroacetate salt (TFA)) resulted in a significant reduction in LDH release and expression of caspase-4/5, NF-κB, and gasdermin D (GSDM-D) in PA-treated HUVECs. Furthermore, GW4869, an exosome release inhibitor, markedly reduced LDH release in PA-stimulated HUVECs. EVs derived from PA-treated HUVECs exacerbated pyroptosis, as indicated by significantly increased LDH release and augmented expression of GSDM-D, NF-κB. Conclusions: The present study revealed that inflammatory, non-canonical caspase-4/5-NF-κB signaling may be one of the crucial mechanistic pathways associated with pyroptosis in ECs, and pyroptotic EVs facilitated pyroptosis in normal ECs during atherosclerosis. Full article
Show Figures

Graphical abstract

Graphical abstract
Full article ">Figure 1
<p>Cell viability and pyroptosis in palmitate-stimulated HUVECs. (<b>A</b>) Dose-dependent effect of PA on HUVECs viability compared to untreated control cells by the MTT assay. (<b>B</b>) The cellular supernatant LDH level was evaluated with a cytotoxicity detection LDH kit. (<b>C</b>) Representative images of HUVECs after their treatment with different concentrations of PA for 24 h. Hematoxylin was used for nuclear staining, and nuclei appear blue. Pyroptotic cells are indicated with square icon and marked structures in second panel indicates blubbed membrane and a swollen structure (Scale bar, 100 μm). <span class="html-italic">n</span> = 3. Results are presented as mean ± standard deviation. * <span class="html-italic">p</span> &lt; 0.05 vs. the control group.</p>
Full article ">Figure 2
<p>Effect of palmitic acid on cell migration and lipotoxicity in HUVECs. (<b>A</b>) The wound-healing assay was performed in the presence of BSA or PA (200 μM) for 24 h (Scale bar: 500 nm) <span class="html-italic">n</span> = 3. (<b>B</b>) Summarized bar graph showed quantification of cell migration. (<b>C</b>) Cells were stained with the oil red O stain, and lipid accumulation was visualized under a microscope after 24 h treatment. Second panel is enlarged portion of square icons in the first panel, and indicates the cells with lipid deposition. Scale bar (100 μm). (<b>D</b>) Quantification of the stained lipid droplets was performed using the eluted oil red O stain by measuring absorbance at 495 nm. <span class="html-italic">n</span> = 3. Results are presented as mean ± standard deviation. * <span class="html-italic">p</span> &lt; 0.05 vs. the control group, <span>$</span> &lt; 0.05 vs. the PA group, # &lt; 0.05 vs. the Ox-LDL group.</p>
Full article ">Figure 2 Cont.
<p>Effect of palmitic acid on cell migration and lipotoxicity in HUVECs. (<b>A</b>) The wound-healing assay was performed in the presence of BSA or PA (200 μM) for 24 h (Scale bar: 500 nm) <span class="html-italic">n</span> = 3. (<b>B</b>) Summarized bar graph showed quantification of cell migration. (<b>C</b>) Cells were stained with the oil red O stain, and lipid accumulation was visualized under a microscope after 24 h treatment. Second panel is enlarged portion of square icons in the first panel, and indicates the cells with lipid deposition. Scale bar (100 μm). (<b>D</b>) Quantification of the stained lipid droplets was performed using the eluted oil red O stain by measuring absorbance at 495 nm. <span class="html-italic">n</span> = 3. Results are presented as mean ± standard deviation. * <span class="html-italic">p</span> &lt; 0.05 vs. the control group, <span>$</span> &lt; 0.05 vs. the PA group, # &lt; 0.05 vs. the Ox-LDL group.</p>
Full article ">Figure 3
<p>Effect of palmitic acid on inflammatory caspases and pyroptotic markers. Summarized bar graphs showed mRNA expression of (<b>A</b>) caspase-4, (<b>B</b>) caspase-5, (<b>C</b>) IL-1β, (<b>D</b>) IL-18, and (<b>E</b>) GSDM-D determined by RT-PCR. GAPDH was used as internal control. <span class="html-italic">n</span> = 4–5. Results are presented as mean ± standard deviation. * <span class="html-italic">p</span> &lt; 0.05 vs. control group.</p>
Full article ">Figure 4
<p>Effect of palmitic acid on endothelial dysfunction. Representative photomicrographs depicted (<b>A</b>) NLRP3 (green) and (<b>B</b>) ICAM-1 (red) (Scale bar, 100 μm). Summarized bar graph showed mRNA expression of (<b>C</b>) NLRP3 and (<b>D</b>) ICAM-1 determined by RT-PCR. <span class="html-italic">n</span> = 3. GAPDH was used as internal control. Results are presented as mean ± standard deviation. * <span class="html-italic">p</span> &lt; 0.05 vs. control group.</p>
Full article ">Figure 4 Cont.
<p>Effect of palmitic acid on endothelial dysfunction. Representative photomicrographs depicted (<b>A</b>) NLRP3 (green) and (<b>B</b>) ICAM-1 (red) (Scale bar, 100 μm). Summarized bar graph showed mRNA expression of (<b>C</b>) NLRP3 and (<b>D</b>) ICAM-1 determined by RT-PCR. <span class="html-italic">n</span> = 3. GAPDH was used as internal control. Results are presented as mean ± standard deviation. * <span class="html-italic">p</span> &lt; 0.05 vs. control group.</p>
Full article ">Figure 5
<p>Inhibition of non-canonical caspase-4/5 blocked mRNA expression of pyroptotic pathway markers in PA-induced HUVECs. HUVECs were pre-incubated with caspase-4/5 inhibitor AC-FLTD-CMK (10 µm/mL) for 2 h, then exposed to PA (200 µm) for 24 h. Summarized bar graph showed mRNA expression of (<b>A</b>) caspase-4, (<b>B</b>) caspase-5, (<b>C</b>) NF-κB, (<b>D</b>) GSDM-D, (<b>E</b>) IL-1β, and (<b>F</b>) IL-18, determined via RT-PCR. <span class="html-italic">n</span> = 3. GAPDH was used as internal control. <span class="html-italic">n</span> = 3–4. Results are presented as mean ± standard deviation. * <span class="html-italic">p</span> &lt; 0.05 vs. control group, <span>$</span> &lt; 0.05 vs. PA group.</p>
Full article ">Figure 5 Cont.
<p>Inhibition of non-canonical caspase-4/5 blocked mRNA expression of pyroptotic pathway markers in PA-induced HUVECs. HUVECs were pre-incubated with caspase-4/5 inhibitor AC-FLTD-CMK (10 µm/mL) for 2 h, then exposed to PA (200 µm) for 24 h. Summarized bar graph showed mRNA expression of (<b>A</b>) caspase-4, (<b>B</b>) caspase-5, (<b>C</b>) NF-κB, (<b>D</b>) GSDM-D, (<b>E</b>) IL-1β, and (<b>F</b>) IL-18, determined via RT-PCR. <span class="html-italic">n</span> = 3. GAPDH was used as internal control. <span class="html-italic">n</span> = 3–4. Results are presented as mean ± standard deviation. * <span class="html-italic">p</span> &lt; 0.05 vs. control group, <span>$</span> &lt; 0.05 vs. PA group.</p>
Full article ">Figure 6
<p>Inhibition of non-canonical caspase-4/5 blocked protein levels of pyroptotic pathway markers in PA-induced HUVECs. HUVECs were pre-incubated with caspase-4/5 inhibitor, AC-FLTD-CMK (10 µm/mL) for 2 h, then exposed to PA (200 µm) for 24 h. (<b>A</b>) Lactate dehydrogenase release (LDH). (<b>B</b>) Representative Western blot analysis showing the effects of caspase-4/5 inhibition on caspase-4, caspase-5, GSDM-D, and NF-κB protein expression. Summarized data showed the changes in the protein expression of (<b>C</b>) caspase-4, (<b>D</b>) caspase-5, (<b>E</b>) NF-κB, and (<b>F</b>) GSDM-D. β-actin was used as an internal control. <span class="html-italic">n</span> = 3–4. Results are presented as mean ± standard deviation. * <span class="html-italic">p</span> &lt; 0.05 vs. the control group, <span>$</span> &lt; 0.05 vs. the PA group.</p>
Full article ">Figure 6 Cont.
<p>Inhibition of non-canonical caspase-4/5 blocked protein levels of pyroptotic pathway markers in PA-induced HUVECs. HUVECs were pre-incubated with caspase-4/5 inhibitor, AC-FLTD-CMK (10 µm/mL) for 2 h, then exposed to PA (200 µm) for 24 h. (<b>A</b>) Lactate dehydrogenase release (LDH). (<b>B</b>) Representative Western blot analysis showing the effects of caspase-4/5 inhibition on caspase-4, caspase-5, GSDM-D, and NF-κB protein expression. Summarized data showed the changes in the protein expression of (<b>C</b>) caspase-4, (<b>D</b>) caspase-5, (<b>E</b>) NF-κB, and (<b>F</b>) GSDM-D. β-actin was used as an internal control. <span class="html-italic">n</span> = 3–4. Results are presented as mean ± standard deviation. * <span class="html-italic">p</span> &lt; 0.05 vs. the control group, <span>$</span> &lt; 0.05 vs. the PA group.</p>
Full article ">Figure 7
<p>Inhibition of NF-κB blocked mRNA expression of pyroptotic pathway markers in PA-induced HUVECs. HUVECs were pre-incubated with NF-κB inhibitor, trifluoroacetate (TFA) (20 µm/mL) for 2 h, and then exposed to PA (200 µm) for 24 h. Summarized bar graph showed mRNA expression for (<b>A</b>) caspase-4, (<b>B</b>) caspase-5, (<b>C</b>) NF-κB, (<b>D</b>) GSDM-D, (<b>E</b>) IL-1β, and (<b>F</b>) IL-18, determined via RT-PCR. GAPDH was used as internal control. <span class="html-italic">n</span> = 3. Results are presented as mean ± standard deviation. * <span class="html-italic">p</span> &lt; 0.05 vs. control group, <span>$</span> &lt; 0.05 vs. PA group.</p>
Full article ">Figure 7 Cont.
<p>Inhibition of NF-κB blocked mRNA expression of pyroptotic pathway markers in PA-induced HUVECs. HUVECs were pre-incubated with NF-κB inhibitor, trifluoroacetate (TFA) (20 µm/mL) for 2 h, and then exposed to PA (200 µm) for 24 h. Summarized bar graph showed mRNA expression for (<b>A</b>) caspase-4, (<b>B</b>) caspase-5, (<b>C</b>) NF-κB, (<b>D</b>) GSDM-D, (<b>E</b>) IL-1β, and (<b>F</b>) IL-18, determined via RT-PCR. GAPDH was used as internal control. <span class="html-italic">n</span> = 3. Results are presented as mean ± standard deviation. * <span class="html-italic">p</span> &lt; 0.05 vs. control group, <span>$</span> &lt; 0.05 vs. PA group.</p>
Full article ">Figure 8
<p>Inhibition of NF-κB blocked protein levels of pyroptotic pathway markers in PA-induced HUVECs. HUVECs were pre-incubated with NF-κB inhibitor, trifluoroacetate (TFA) (20 µm/mL) for 2 h, and then exposed to PA (200 µm) for 24 h. (<b>A</b>) Lactate dehydrogenase (LDH) release. (<b>B</b>) Representative Western blot analysis showing the effects of NF-κB inhibition on caspase-4, caspase-5, GSDM-D, and NF-κB protein expression. Summarized data show the changes in the protein expression of (<b>C</b>) caspase-4, (<b>D</b>) caspase-5, (<b>E</b>) NF-κB, and (<b>F</b>) GSDM-D. β-actin was used as internal control. <span class="html-italic">n</span> = 3. Results are presented as mean ± standard deviation. * <span class="html-italic">p</span> &lt; 0.05 vs. control group, <span>$</span> &lt; 0.05 vs. PA group.</p>
Full article ">Figure 8 Cont.
<p>Inhibition of NF-κB blocked protein levels of pyroptotic pathway markers in PA-induced HUVECs. HUVECs were pre-incubated with NF-κB inhibitor, trifluoroacetate (TFA) (20 µm/mL) for 2 h, and then exposed to PA (200 µm) for 24 h. (<b>A</b>) Lactate dehydrogenase (LDH) release. (<b>B</b>) Representative Western blot analysis showing the effects of NF-κB inhibition on caspase-4, caspase-5, GSDM-D, and NF-κB protein expression. Summarized data show the changes in the protein expression of (<b>C</b>) caspase-4, (<b>D</b>) caspase-5, (<b>E</b>) NF-κB, and (<b>F</b>) GSDM-D. β-actin was used as internal control. <span class="html-italic">n</span> = 3. Results are presented as mean ± standard deviation. * <span class="html-italic">p</span> &lt; 0.05 vs. control group, <span>$</span> &lt; 0.05 vs. PA group.</p>
Full article ">Figure 9
<p>Characterization of small extracellular vesicles/exosomes and their role in pyroptosis in PA-induced HUVECs. (<b>A</b>) Representative scanning electron micrographs of sEVs/exosomes isolated from culture medium of HUVECs (Scale bar, 100 nm). (<b>B</b>) Representative Western blot analysis of different exosomal protein markers (Alix TSG 101) and cell lysate proteins (Calnexin and β-Actin) collected from HUVECs. (<b>C</b>) LDH release was measured in HUVECs, pre-incubated with exosome release inhibitor (GW4869) (5 µm/mL) for 2 h, and then exposed to PA (200 µm) for 24 h. (<b>D</b>) HUVECs were co-cultured with EVs isolated from PA-induced HUVECs, and lactate dehydrogenase release (LDH) was measured. (<b>E</b>) Representative Western blot analysis showing effects of EVs isolated from PA-induced HUVECs on GSDM-D and NF-κB protein expression. Summarized data showed changes in protein expression of (<b>F</b>) GSDM-D and (<b>G</b>) NF-κB. <span class="html-italic">n</span> = 3. Results are presented as mean ± standard deviation. Data from two groups were analyzed using two-tailed Student’s <span class="html-italic">t</span>-test. * <span class="html-italic">p</span> &lt; 0.05 vs. control group, <span>$</span> &lt; 0.05 vs. PA group.</p>
Full article ">
18 pages, 20350 KiB  
Article
Paeoniflorin Inhibits the Activation of Microglia and Alleviates Depressive Behavior by Regulating SIRT1-NF-kB-NLRP3/Pyroptosis Pathway
by Xue Wang, Lili Su, Silu Liu, Zhongmei He, Jianming Li, Ying Zong, Weijia Chen and Rui Du
Int. J. Mol. Sci. 2024, 25(23), 12543; https://doi.org/10.3390/ijms252312543 - 22 Nov 2024
Viewed by 465
Abstract
Inflammation assumes a vital role in the pathogenesis of depression and in antidepressant treatment. Paeoniflorin (PF), a monoterpene glycoside analog possessing anti-inflammatory attributes, exhibits therapeutic efficacy on depression-like behavior in mice. The objective of this study was to evaluate the antidepressant effects of [...] Read more.
Inflammation assumes a vital role in the pathogenesis of depression and in antidepressant treatment. Paeoniflorin (PF), a monoterpene glycoside analog possessing anti-inflammatory attributes, exhibits therapeutic efficacy on depression-like behavior in mice. The objective of this study was to evaluate the antidepressant effects of PF on depression elicited by the chronic unpredictable mild stress (CUMS) model and the precise neural sequence associated with the inflammatory process. In this study, we established an in vivo mouse model induced by CUMS and an in vitro BV2 cell model induced by LPS+ATP. The mechanism of PF for depression was assessed by the SIRT1 selective inhibitor EX-527. The findings demonstrated that PF significantly alleviated the damage of BV2 cells treated with LPS and ATP, inhibited the generation of ROS, up-regulated the expression of SIRT1 mRNA, and down-regulated the expression of nuclear NF-κB, p65, NLRP3, Caspase-1 and GSDMD-N in vitro. In vivo, PF mitigated the depressive-like behavior induced by CUMS, reduced the number of neurons, and decreased the secretion of pro-inflammatory factors IL-1β, IL-6, and TNF-α in the hippocampus. Immunohistochemical results indicated that PF attenuated CUMS-induced hyperactivation of microglia. Moreover, the expression level of SIRT1 in the hippocampus was augmented, while the protein levels of NF-κB, p65, NLRP3, Caspase-1, IL-1β and GSDMD-N were diminished after PF treatment. Additionally, the selective inhibition of SIRT1 attenuated the therapeutic effect of PF on depression. These results imply that PF possesses antidepressant properties that rely on SIRT1 signaling to regulate NLRP3 inflammasome inactivation. Full article
Show Figures

Graphical abstract

Graphical abstract
Full article ">Figure 1
<p>PF inhibited the activation of apoptosis, inflammation, pyroptosis and ROS of BV2 cells induced by LPS+ATP. (<b>A</b>) Cell viability of BV2. (<b>B</b>–<b>E</b>) The content of LDH, Caspase-1, IL-1β and IL-18. (<b>F</b>) Hoechst/PI double staining (×400). (<b>G</b>) ROS staining (×400). (<b>H</b>) Quantitative analysis of PI/Hoechst and DCFH-DA. Data are presented as mean ± SEM. ## <span class="html-italic">p</span> &lt; 0.01, compared to control group; ** <span class="html-italic">p</span> &lt; 0.01, compared to LPS+ATP group.</p>
Full article ">Figure 2
<p>Effects of PF on protein levels of SIRT1, NF-κB and NLRP3 inflammasome in LPS+ATP treated BV2 cells. The relative mRNA expression level of (<b>A</b>) SIRT1, (<b>B</b>) NF-κB, (<b>C</b>) NLRP3, (<b>D</b>) ASC, (<b>E</b>) Caspase-1, (<b>F</b>) IL-1β, (<b>G</b>) GSDMD-N. Data are presented as mean ± SEM. ## <span class="html-italic">p</span> &lt; 0.01, compared to control group; * <span class="html-italic">p</span> &lt; 0.05, ** <span class="html-italic">p</span> &lt; 0.01, compared to LPS+ATP group, <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 PF+EX-527 group.</p>
Full article ">Figure 3
<p>PF improved depressive-like behaviors in CUMS mice. (<b>A</b>) Experimental procedures. (<b>B</b>) SPT, (<b>C</b>) TST, (<b>D</b>) FST, (<b>E</b>,<b>G</b>) OFT and (<b>F</b>,<b>H</b>) EPM. Data are presented as mean ± SEM. ## <span class="html-italic">p</span> &lt; 0.01, compared to control group; * <span class="html-italic">p</span> &lt; 0.05, ** <span class="html-italic">p</span> &lt; 0.01, compared to CUMS group.</p>
Full article ">Figure 4
<p>PF alleviated neuroinflammation and neuronal damage in CUMS mice. (<b>A</b>) The levels of IL-6, IL-1β, TNF-α. (<b>B</b>) Quantitative analysis of Nissl bodies positive cells. (<b>C</b>) The positive signal intensity of IBA1. (<b>D</b>) Nissl staining image (×200). (<b>E</b>) Immunohistochemical image of IBA1 (×200). Data are presented as mean ± SEM. ## <span class="html-italic">p</span> &lt; 0.01, compared to control group; * <span class="html-italic">p</span> &lt; 0.05, ** <span class="html-italic">p</span> &lt; 0.01, compared to CUMS group.</p>
Full article ">Figure 5
<p>PF inhibited the activation of NF-κB by activating the expression of SIRT1. (<b>A</b>) Representative Western blots. (<b>B</b>) Quantification of SIRT1 and NF-κB P65 expression levels and fluorescence intensity. (<b>C</b>) The fluorescence intensity of SIRT1 (×400). (<b>D</b>) The fluorescence intensity of NF-κB P65 in CA3 area of hippocampus (×400). Data are presented as mean ± SEM. ## <span class="html-italic">p</span> &lt; 0.01, compared to control group; * <span class="html-italic">p</span> &lt; 0.05, ** <span class="html-italic">p</span> &lt; 0.01, compared to CUMS group.</p>
Full article ">Figure 6
<p>PF inhibited the activation of NLRP3 inflammasome and its mediated pyroptosis. (<b>A</b>) Representative Western blots. (<b>B</b>) The quantitative analysis of the expression level of NLRP3, ASC, Caspase-1, IL-1β and GSDMD-N. (<b>C</b>) The positive signal intensity of NLRP3 (×200). (<b>D</b>) IHC quantitative analysis of NLRP3. (<b>E</b>) The fluorescence intensity of GSDMD-N/DAPI in CA3 area of hippocampus (×400). Data are presented as mean ± SEM. ## <span class="html-italic">p</span> &lt; 0.01, ### <span class="html-italic">p</span> &lt; 0.001, compared to control group; * <span class="html-italic">p</span> &lt; 0.05, ** <span class="html-italic">p</span> &lt; 0.01, *** <span class="html-italic">p</span> &lt; 0.001 compared to CUMS group.</p>
Full article ">Figure 7
<p>EX-527 blocked the inhibitory effect of PF on the activation of NF-κB, NLRP3 inflammasome and pyroptosis in CUMS model mice. (<b>A</b>) Representative Western blots. (<b>B</b>) Quantification of SIRT1, NF-kB-P65, NLRP3, Caspase-1 and GSDMD-N expression levels and quantification of SIRT1 and GSDMD-N fluorescence intensity. (<b>C</b>) The fluorescence intensity of SIRT1/DAPI in CA3 area of hippocampus (×400). (<b>D</b>) The fluorescence intensity of GSDMD-N/DAPI in CA3 area of hippocampus (×400). Data are presented as mean ± SEM. ## <span class="html-italic">p</span> &lt; 0.01, compared to control group; * <span class="html-italic">p</span> &lt; 0.05, ** <span class="html-italic">p</span> &lt; 0.01, compared to CUMS group, <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 PF+EX-527 group.</p>
Full article ">
12 pages, 4663 KiB  
Article
Microplastics Exposure Aggravates Synovitis and Pyroptosis in SLE by Activating NF-κB and NRF2/KEAP1 Signaling
by Wenxiang Zeng, Shiqiao He, Ying Zhao, Minjian Jiang, Wenla Wang, Limeng Yang, Weibin Du and Wei Zhuang
Toxics 2024, 12(12), 840; https://doi.org/10.3390/toxics12120840 - 22 Nov 2024
Viewed by 530
Abstract
Microplastics (MPs) represent an emerging pollutant capable of entering the human body through the respiratory and digestive systems, thereby posing significant health risks. Systemic lupus erythematosus (SLE) is a complex autoimmune disease that affects multiple organ systems, often presenting with polyarticular joint manifestations. [...] Read more.
Microplastics (MPs) represent an emerging pollutant capable of entering the human body through the respiratory and digestive systems, thereby posing significant health risks. Systemic lupus erythematosus (SLE) is a complex autoimmune disease that affects multiple organ systems, often presenting with polyarticular joint manifestations. Despite its relevance, there is currently limited research on the impact of MPs on lupus arthritis. This study aims to investigate the effects of MPs on joint inflammation in SLE. MRL/lpr mice exhibit SLE similar to that of humans. We administered either 0.5 mg/kg or 5 mg/kg of MPs to 8-week-old female MRL/lpr mice via oral ingestion. Our findings indicate that exposure to MPs can lead to synovial damage, adversely affecting the morphology and function of the knee joint, along with increased oxidative stress, apoptosis, synovial fibrosis, and the secretion of inflammatory cytokines. Notably, MPs significantly enhanced synovial cell pyroptosis by upregulating the expression of NLRP3, CASPASE-1, GSDMD, IL-1β, and IL-18. Mechanistic analyses further demonstrated that MPs exposure activates the NF-κB and NRF2/KEAP1 signaling pathways. Overall, our in vivo findings suggest that MPs exposure promotes synovial cell pyroptosis through increased oxidative stress and NF-κB signaling, thereby disrupting the structure and function of synovial tissue. This research provides new insights into the synovial damage associated with MPs exposure. Full article
(This article belongs to the Section Reproductive and Developmental Toxicity)
Show Figures

Graphical abstract

Graphical abstract
Full article ">Figure 1
<p>The MP structure damages the synovial structure and aggravate synovitis in SLE mice: (<b>A</b>) flow chart of the experimental design; (<b>B</b>) HE staining of mouse knee synovium, demonstrating thickening of synovial layers (Black arrow) and increased vasculature (Blue arrow); (<b>C</b>) histological scoring of the synovium based on HE staining; (<b>D</b>) MASSON’s trichrome staining of mouse knee synovium, indicating increased fiber content (Black arrow); (<b>E</b>–<b>H</b>) immunohistochemical staining illustrating the expression of MMP-13 and MMP-19 in the synovium of the knee joint of mice, along with quantification of their expression levels. Red arrowheads denote positively stained cells. Data are presented as mean ± standard deviation (SD). ns indicates no statistical significance. * <span class="html-italic">p</span> &lt; 0.05, ** <span class="html-italic">p</span> &lt; 0.01 (vs. 0 mg/kg group); ## <span class="html-italic">p</span> &lt; 0.01 (vs. C57BL/6J group), with <span class="html-italic">n</span> = 6 per group.</p>
Full article ">Figure 2
<p>MPs promote synovial cell apoptosis in SLE mice: (<b>A</b>,<b>C</b>) immunofluorescence staining demonstrating the expression of CASPASE-3 and BCL-2 in the synovium of the knee joints of mice, along with (<b>B</b>,<b>D</b>) quantification of expression; (<b>E</b>) TUNEL staining and (<b>F</b>) quantification of the rate of TUNEL-positive cells. DAPI stains the nuclei blue, and white arrowheads indicate positively stained cells. Data are presented as mean ± SD. ns indicates no statistical significance. ** <span class="html-italic">p</span> &lt; 0.01 (vs. 0 mg/kg group); ## <span class="html-italic">p</span> &lt; 0.01 (vs. C57BL/6J group), <span class="html-italic">n</span> = 6 per group.</p>
Full article ">Figure 3
<p>MPs compound synovial inflammation in SLE mice. (<b>A</b>–<b>H</b>) Immunofluorescence staining illustrating the expression of IL-1β, IL-18, IL-6, and TNF-α in the synovium of the knee joint of mice, along with quantification of expression. DAPI stains the nuclei blue, and white arrowheads indicate positively stained cells. Data are presented as mean ± SD. ns indicates no statistical significance. * <span class="html-italic">p</span> &lt; 0.05, ** <span class="html-italic">p</span> &lt; 0.01 (vs. 0 mg/kg group); ## <span class="html-italic">p</span> &lt; 0.01 (vs. C57BJ/6 group), with <span class="html-italic">n</span> = 6 per group.</p>
Full article ">Figure 4
<p>MPs aggravate synovial pyroptosis in SLE mice. (<b>A</b>–<b>C</b>) Immunofluorescence staining showing the expression of NLRP3, CASPASE-1, and GSDMD in the synovium of the knee joint of mice, along with (<b>D</b>–<b>F</b>) quantification of expression. DAPI stains nuclei blue, and white arrowheads indicate positively stained cells. Data are expressed as mean ± SD. ns indicates no statistical significance. * <span class="html-italic">p</span> &lt; 0.05, ** <span class="html-italic">p</span> &lt; 0.01 (vs. 0 mg/kg group); ## <span class="html-italic">p</span> &lt; 0.01 (vs. C57BL/6J group), with <span class="html-italic">n</span> = 6 per group.</p>
Full article ">Figure 5
<p>Effects of MPs exposure on oxidative stress and the NF-κB signaling pathway in SLE mice. (<b>A</b>–<b>C</b>) Immunofluorescence staining was conducted to assess the expression of NRF2, KEAP1, and HO-1 in the synovium of the knee joints of mice, while (<b>H</b>–<b>J</b>) illustrates the quantification of these expressions. (<b>D</b>–<b>G</b>) Immunofluorescence staining for P65, P-P65, IκBα, and p-IκBα in the knee joint synovium of mice is presented, with (<b>K</b>–<b>N</b>) showing the corresponding quantification. DAPI was used to stain nuclei in blue, and white arrowheads indicate positively stained cells. Data are expressed as mean ± SD. ns indicates no statistical significance. * <span class="html-italic">p</span> &lt; 0.05, ** <span class="html-italic">p</span> &lt; 0.01 (vs. 0 mg/kg group); ## <span class="html-italic">p</span> &lt; 0.01 (vs. C57BL/6J group), with <span class="html-italic">n</span> = 6 per group.</p>
Full article ">
Back to TopTop