Nothing Special   »   [go: up one dir, main page]

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

Years

Between: -

Subjects

remove_circle_outline
remove_circle_outline
remove_circle_outline
remove_circle_outline
remove_circle_outline
remove_circle_outline
remove_circle_outline
remove_circle_outline
remove_circle_outline

Journals

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

Article Types

Countries / Regions

remove_circle_outline
remove_circle_outline
remove_circle_outline
remove_circle_outline
remove_circle_outline
remove_circle_outline
remove_circle_outline

Search Results (6,194)

Search Parameters:
Keywords = flow cytometry

Order results
Result details
Results per page
Select all
Export citation of selected articles as:
15 pages, 20467 KiB  
Article
Identification and Functional Analysis of Ras-Related Associated with Diabetes Gene (rrad) in Edwardsiella piscicida-Resistant Individuals of Japanese Flounder (Paralichthys olivaceus)
by Ying Zhu, Xinsheng Yang, Yingming Yang, Xu Yan, Chao Li and Songlin Chen
Int. J. Mol. Sci. 2024, 25(19), 10532; https://doi.org/10.3390/ijms251910532 - 1 Oct 2024
Viewed by 99
Abstract
Ras-related associated with diabetes (RRAD) is a member of the Ras GTPase superfamily that plays a role in several cellular functions, such as cell proliferation and differentiation. In particular, the superfamily acts as an NF-κB signaling pathway inhibitor and calcium regulator to participate [...] Read more.
Ras-related associated with diabetes (RRAD) is a member of the Ras GTPase superfamily that plays a role in several cellular functions, such as cell proliferation and differentiation. In particular, the superfamily acts as an NF-κB signaling pathway inhibitor and calcium regulator to participate in the immune response pathway. A recent transcriptome study revealed that rrad was expressed in the spleen of disease-resistant Japanese flounder (Paralichthys olivaceus) individuals compared with disease-susceptible individuals, and the results were also verified by qPCR. Thus, the present study aimed to explore how rrad regulates antimicrobial immunity via the NF-κB pathway. First, the coding sequence of P. olivaceus rrad was identified. The sequence was 1092 bp in length, encoding 364 amino acids. Based on phylogenetic and structural relationship analyses, P. olivaceus rrad appeared to be more closely related to teleosts. Next, rrad expression differences between disease-resistant and disease-susceptible individuals in immune-related tissues were evaluated, and the results revealed that rrad was expressed preferentially in the spleen of disease-resistant individuals. In response to Edwardsiella piscicida infection, rrad expression in the spleen changed. In vitro, co-culture was carried out to assess the hypo-methylated levels of the rrad promoter in the disease-resistant spleen, which was consistent with the high mRNA expression. The siRNA-mediated knockdown of rrad performed with the gill cell line of P. olivaceus affected many rrad-network-related genes, i.e., dcp1b, amagt, rus1, rapgef1, ralbp1, plce1, rasal1, nckipsd, prkab2, cytbc-1, sh3, and others, as well as some inflammation-related genes, such as bal2 and Il-1β. In addition, flow cytometry analysis showed that rrad overexpression was more likely to induce cell apoptosis, with establishing a link between rrad‘s function and its potential roles in regulating the NF-κB pathway. Thus,. the current study provided some clarity in terms of understanding the immune response about rrad gene differences between disease-resistant and disease-susceptible P. olivaceus individuals. This study provides a molecular basis for fish rrad gene functional analysis and may serve as a reference for in-depth of bacterial disease resistance of teleost. Full article
(This article belongs to the Section Molecular Biology)
Show Figures

Figure 1

Figure 1
<p>cDNA and predicted amino acid sequences of <span class="html-italic">Po-rrad</span>. The small letters show nucleotides, and capital letters denote predicted amino acid sequences. The letters in the red box indicate the start codon (ATG). The double underline marks the poly-A sequence. The black star represents the stop codon (TGA). The conserved domains are shown based on prediction. The green, blue, red, and black line represent G1, G2, G3, and G4 box, respectively.</p>
Full article ">Figure 2
<p>Structural domains and rrad amino acid sequences of <span class="html-italic">P. olivaceus</span> and other vertebrates. All sequences were aligned using DNAMAN. The red box represent RGK domain.</p>
Full article ">Figure 3
<p>A phylogenetic tree was constructed with the neighbor-joining algorithm in MEGA 4.0. The relative genetic distances are indicated by the scale bar and the branch lengths. The protein sequences of the different species used to build the tree were as follows: <span class="html-italic">Oreochromis niloticus</span> rrad (XP_003445756.1), <span class="html-italic">Labrus bergylta</span>, rrad (XP_020508528.1), <span class="html-italic">Larimichthys crocea</span> rrad (XP_019116206.1), <span class="html-italic">Amphiprion ocellaris</span>, rrad (XP_003445756.1), <span class="html-italic">Acanthochromis polyacanthus</span>, rrad (XP_022057232.1), <span class="html-italic">Seriola dumerili</span> rrad (XP_022623546.1), <span class="html-italic">Paralichthys olivaceus</span> rrad (XP_019935401.1), <span class="html-italic">Cynoglossus semilaevis</span> rrad (XP_008308414.1), <span class="html-italic">Melopsittacus undulatus</span>, RRAD (XP_005152231.1), <span class="html-italic">Gallus gallus</span> RRAD (NP_001264535.3), <span class="html-italic">Canis lupus familiaris</span> RRAD (XP_038520230.1), <span class="html-italic">Trichechus manatus latirostris</span>, RRAD (XP_004371570.1), <span class="html-italic">Homo sapiens</span> RRAD (AAB17064.1), and <span class="html-italic">Pan troglodytes</span> RRAD (XP_001143391.3). The black triangle represents the target species.</p>
Full article ">Figure 4
<p><span class="html-italic">rrad</span> expression level in <span class="html-italic">P. olivaceus</span> evaluated using qPCR. (<b>A</b>) Relative <span class="html-italic">rrad</span> mRNA expression in the various tissues of normal fish. (<b>B</b>) Relative <span class="html-italic">rrad</span> mRNA expression in the immune-related tissues of disease-resistant and disease-susceptible individuals. The mean ± SEM values from three separate individuals (<span class="html-italic">n</span> = 3) are shown. The different letters “a” and “b” indicate significant differences (<span class="html-italic">p</span> &lt; 0.05).</p>
Full article ">Figure 5
<p>qPCR analysis of <span class="html-italic">rrad</span> expression profile in different immune-related tissues (spleen, liver, kidney, and intestines) after <span class="html-italic">E. piscicida</span> infection. The results were determined at different time points (0, 6, 12, 24, and 48 h), and PBS was used as the control. The transcription levels were normalized using <span class="html-italic">β-actin</span> levels. Data were analyzed using IBM SPSS Statistics 19 with the independent samples t-test. A asterisk stand for a significant difference in comparison with the 0 h group (<span class="html-italic">p</span> &lt; 0.05), two asterisks represented significantly different comparison with the 0 h group (<span class="html-italic">p</span> &lt; 0.01).</p>
Full article ">Figure 6
<p>In vitro stimulation of <span class="html-italic">rrad</span> in response to LPS, PGN, and poly I:C in the gill cell line of <span class="html-italic">P. olivaceus</span>. The data were measured using quantitative RT-PCR and normalized using <span class="html-italic">β</span>-actin gene as an internal control. The data are presented as the mean ± standard deviation of three biological replicates. The expression levels with different letters were significantly different, including a asterisk represented significantly different comprared to control (<span class="html-italic">p</span> &lt; 0.05), two asterisks represented significantly different comprared to control (<span class="html-italic">p</span> &lt; 0.01).</p>
Full article ">Figure 7
<p>DNA methylation of <span class="html-italic">rrad</span> in the promoter and gene body. Methylation differences in the <span class="html-italic">rrad</span> promoter and gene body in the spleen tissue of disease-resistant (DR_<span class="html-italic">Po</span>_Sp) and disease-susceptible (DS_<span class="html-italic">Po</span>_Sp) individuals. Red and blue vertical lines illustrate the methylation level of cytosines, whereas solid rims denote methylation and unmethylation positions, respectively, in disease-resistant individuals, and blue indicates disease-susceptible individuals. The red box is the difference in promoter region methylation between disease-resistant and disease-susceptible individuals.</p>
Full article ">Figure 8
<p>Methylation analysis. Luciferase assays for methylated and unmethylated recombinant plasmids. The <span class="html-italic">x</span>-axis shows different recombinant plasmids, and the <span class="html-italic">y</span>-axis shows the relative luciferase activity. The letters “a” and “b” indicate significant differences (<span class="html-italic">p</span> &lt; 0.05).</p>
Full article ">Figure 9
<p>Effect of <span class="html-italic">rrad</span> on the transcriptional activity of NF-κB. (<b>A</b>) Effect of overexpression and siRNA treatment of <span class="html-italic">rrad</span> on the transcriptional activity of NF-κB luciferase reporter gene for 48 h, after which the luciferase activity was measured. (<b>B</b>) After co-transfection of <span class="html-italic">Po-rrad</span>-pEGFP-N3, <span class="html-italic">Po-rrad</span>-siRNA, and LPS with the NF-κB luciferase reporter gene, cells were stimulated for 24 h. The letters “a” “b” and “c” indicate significant differences (<span class="html-italic">p</span> &lt; 0.05).</p>
Full article ">Figure 10
<p>Analysis of PPI interaction and siRNA effects after <span class="html-italic">rrad</span> RNAi in <span class="html-italic">P. olivaceus</span> gill cells. (<b>A</b>) The expression of <span class="html-italic">rrad</span>, <span class="html-italic">dcp1b</span>, <span class="html-italic">amagt</span>, <span class="html-italic">rsu1</span>, <span class="html-italic">rapgef1</span>, <span class="html-italic">ralbp1</span>, <span class="html-italic">plce1</span>, <span class="html-italic">rasal1</span>, <span class="html-italic">nckipsd</span>, <span class="html-italic">prkab2</span>, <span class="html-italic">cytbc-1</span>, <span class="html-italic">sh3</span>, <span class="html-italic">bcl2</span>, and <span class="html-italic">Il-1β</span> was analyzed in cultured gill cells after RNAi. (<b>B</b>) Three-dimensional protein prediction and PPI analysis of rrad. The correlation of proteins was predicted using the STRING 11.0 online database. (<b>C</b>) Compared with the control, the mRNA levels of the interaction predictor of <span class="html-italic">rrad</span> and other immune response genes, i.e., <span class="html-italic">dcp1b</span>, <span class="html-italic">amagt</span>, <span class="html-italic">rsu1</span>, <span class="html-italic">rapgef1</span>, <span class="html-italic">ralbp1</span>, <span class="html-italic">plce1</span>, <span class="html-italic">rasal1</span>, <span class="html-italic">nckipsd</span>, <span class="html-italic">prkab2</span>, <span class="html-italic">cytbc-1</span>, <span class="html-italic">sh3</span>, <span class="html-italic">bcl2</span>, and <span class="html-italic">Il-1β</span>, were detected after RNAi. The stars represented a significant differentce copared to the control group (<span class="html-italic">p</span> &lt; 0.05).</p>
Full article ">
22 pages, 6256 KiB  
Article
Endothelial Myosin IIA Is Required for the Maintenance of Blood–Brain Barrier Integrity
by Yanan Deng, Ziqi Qiao, Changping Zhou, Yujun Pei, Han Xu, Xuya Kang and Jincai Luo
Cells 2024, 13(19), 1635; https://doi.org/10.3390/cells13191635 - 1 Oct 2024
Viewed by 148
Abstract
Brain endothelial cells (ECs) are essential elements of the blood–brain barrier (BBB), maintaining its integrity through both paracellular junctions and transcellular transport systems. Myosin IIA, a multifunctional protein, plays a significant role in various cellular processes, including cytoskeletal maintenance, cell division, and signal [...] Read more.
Brain endothelial cells (ECs) are essential elements of the blood–brain barrier (BBB), maintaining its integrity through both paracellular junctions and transcellular transport systems. Myosin IIA, a multifunctional protein, plays a significant role in various cellular processes, including cytoskeletal maintenance, cell division, and signal transduction. While Myosin IIA has been implicated in bleeding and ischemic stroke, its role in regulating BBB integrity under physiological conditions remains unclear. In this study, we investigated the impact of Myosin IIA deficiency on BBB integrity using intravenous tracer injections and models of epilepsy. Flow cytometry, Western blot, and real-time PCR were employed to isolate brain cells and assess changes in protein and mRNA levels. Additionally, immunofluorescence staining and electron microscopy were used to explore alterations in protein expression and the structure of BBB. Our results demonstrate that endothelial Myosin IIA deficiency increased BBB permeability and exacerbated symptoms in BBB-related diseases. Mechanistically, we found that Myosin IIA modulates β-catenin transcription and protein interactions. The overexpression of β-catenin in brain endothelial Myosin IIA deficiency mice improved BBB integrity and reduced disease severity. This study establishes Myosin IIA as a critical regulator of BBB integrity and suggests new therapeutic targets for vascular diseases. Full article
Show Figures

Figure 1

Figure 1
<p>Brain endothelial deletion of Myosin IIA impairs the integrity of the BBB in mice. (<b>A</b>) Representative PCR gel image for genotyping mice using tail genomic DNA, showing floxed allele band at 480 bp, WT allele band at 349 bp, and Cre-positive band at 421 bp. (<b>B</b>) Representative Western blot image, showing Myosin IIA protein levels in primary brain ECs from <span class="html-italic">Myh9</span><sup>fl/fl</sup> and <span class="html-italic">Myh9</span><sup>ECKO</sup> mice (<span class="html-italic">n</span> = 3). (<b>C</b>) Statistical analysis of Myosin IIA protein levels in primary brain ECs from <span class="html-italic">Myh9</span><sup>fl/fl</sup> and <span class="html-italic">Myh9</span><sup>ECKO</sup> mice (<span class="html-italic">n</span> = 3). *** <span class="html-italic">p</span> &lt; 0.001; Student’s <span class="html-italic">t</span>-test. (<b>D</b>) Representative immunofluorescence staining of brain tissue sections from <span class="html-italic">Myh9</span><sup>fl/fl</sup> and <span class="html-italic">Myh9</span><sup>ECKO</sup> mice, showing CD31 (green), Myosin IIA (red), and DAPI (blue). Scale bar: 50 μm. (<b>E</b>) Representative image of brain tissues from <span class="html-italic">Myh9</span><sup>fl/fl</sup> and <span class="html-italic">Myh9</span><sup>ECKO</sup> mice following intravenous injection of Evans blue. Scale bar: 5 mm. (<b>F</b>) Statistical analysis of total Evans blue tracer leakage into the brain parenchyma of <span class="html-italic">Myh9</span><sup>fl/fl</sup> and <span class="html-italic">Myh9</span><sup>ECKO</sup> mice (<span class="html-italic">Myh9</span><sup>fl/fl</sup>, <span class="html-italic">n</span> = 4; <span class="html-italic">Myh9</span><sup>ECKO</sup>, <span class="html-italic">n</span> = 5). * <span class="html-italic">p</span> &lt; 0.05; Student’s <span class="html-italic">t</span>-test. (<b>G</b>) Representative immunofluorescence staining of brain sections from <span class="html-italic">Myh9</span><sup>fl/fl</sup> and <span class="html-italic">Myh9</span><sup>ECKO</sup> mice following intravenous injection of Sulfo-NHS-Biotin, showing Sulfo-NHS-Biotin (green) and DAPI (blue). Scale bar: 1 mm.</p>
Full article ">Figure 2
<p>Brain endothelial deletion of Myosin IIA increases seizure susceptibility and seizure-induced mortality. (<b>A</b>) Experimental design for a mouse epilepsy model, involving intraperitoneal injection of scopolamine followed by pilocarpine 30 min later, with mouse behavior monitored over the next 120 min. (<b>B</b>) Representative images of Evans blue tracer leakage in brain tissues of <span class="html-italic">Myh9</span><sup>fl/fl</sup> and <span class="html-italic">Myh9</span><sup>ECKO</sup> mice after pilocarpine-induced epilepsy. Scale bar: 5 mm. (<b>C</b>) Statistical analysis of total Evans blue tracer leakage into the brain tissues of <span class="html-italic">Myh9</span><sup>fl/fl</sup> and <span class="html-italic">Myh9</span><sup>ECKO</sup> mice post-pilocarpine administration (<span class="html-italic">n</span> = 6). * <span class="html-italic">p</span> &lt; 0.05, *** <span class="html-italic">p</span> &lt; 0.001; one-way ANOVA test. (<b>D</b>) Representative immunofluorescence co-staining images of brain sections from <span class="html-italic">Myh9</span><sup>fl/fl</sup> and <span class="html-italic">Myh9</span><sup>ECKO</sup> mice post-pilocarpine administration, showing CD31 (red) and Sulfo-NHS-Biotin (green). Scale bar: 100 μm. (<b>E</b>) Statistical analysis of Sulfo-NHS-Biotin leakage index in brain tissues of <span class="html-italic">Myh9</span><sup>fl/fl</sup> and <span class="html-italic">Myh9</span><sup>ECKO</sup> mice after pilocarpine-induced epilepsy (<span class="html-italic">n</span> = 3). * <span class="html-italic">p</span> &lt; 0.05; Student’s <span class="html-italic">t</span>-test. (<b>F</b>) Statistical analysis of the incidence of seizure occurrence in <span class="html-italic">Myh9</span><sup>fl/fl</sup> and <span class="html-italic">Myh9</span><sup>ECKO</sup> mice after pilocarpine administration (NS, no seizure group; Seizure, seizure group; <span class="html-italic">n</span> = 14). *** <span class="html-italic">p</span> &lt;0.001; Fisher’s exact test. (<b>G</b>) Time to first seizure onset in <span class="html-italic">Myh9</span><sup>fl/fl</sup> and <span class="html-italic">Myh9</span><sup>ECKO</sup> mice following pilocarpine administration (<span class="html-italic">n</span> = 6). * <span class="html-italic">p</span> &lt; 0.05; Student’s <span class="html-italic">t</span>-test. (<b>H</b>) Survival time of <span class="html-italic">Myh9</span><sup>fl/fl</sup> and <span class="html-italic">Myh9</span><sup>ECKO</sup> mice after pilocarpine administration, with a total monitoring duration of 120 min (<span class="html-italic">n</span> = 14). * <span class="html-italic">p</span> &lt; 0.05; Student’s <span class="html-italic">t</span>-test.</p>
Full article ">Figure 3
<p>Deletion of brain endothelial Myosin IIA downregulates junctional proteins of the BBB. (<b>A</b>) Representative electron microscopy image of vascular endothelial tight junction structures in brain sections from <span class="html-italic">Myh9</span><sup>fl/fl</sup> and <span class="html-italic">Myh9</span><sup>ECKO</sup> mice, with yellow arrows indicating disrupted tight junctions. Scale bar: 200 nm. (<b>B</b>) Statistical analysis of the proportion of abnormal tight junction structures in brain sections from <span class="html-italic">Myh9</span><sup>fl/fl</sup> and <span class="html-italic">Myh9</span><sup>ECKO</sup> mice (<span class="html-italic">n</span> = 3). *** <span class="html-italic">p</span> &lt; 0.001; Student’s <span class="html-italic">t</span>-test. (<b>C</b>) Western blot analysis of ZO-1, Ve-cadherin, Occludin, and Claudin-5 protein levels in primary brain ECs from <span class="html-italic">Myh9</span><sup>fl/fl</sup> and <span class="html-italic">Myh9</span><sup>ECKO</sup> mice. (<b>D</b>) Statistical analysis of protein expression levels of ZO-1, Ve-cadherin, Occludin, and Claudin-5 in primary brain ECs from <span class="html-italic">Myh9</span><sup>fl/fl</sup> and <span class="html-italic">Myh9</span><sup>ECKO</sup> mice (<span class="html-italic">n</span> = 3). * <span class="html-italic">p</span> &lt; 0.05, ** <span class="html-italic">p</span> &lt; 0.01; Student’s <span class="html-italic">t</span>-test. (<b>E</b>) Representative immunofluorescence co-staining image of brain sections from <span class="html-italic">Myh9</span><sup>fl/fl</sup> and <span class="html-italic">Myh9</span><sup>ECKO</sup> mice, showing CD31 (green) and ZO-1 (red). Scale bar: 50 μm. (<b>F</b>) Representative immunofluorescence co-staining image of brain sections from <span class="html-italic">Myh9</span><sup>fl/fl</sup> and <span class="html-italic">Myh9</span><sup>ECKO</sup> mice, showing CD31 (green) and Occludin (red). Scale bar: 50 μm. (<b>G</b>) Representative immunofluorescence co-staining image of brain sections from <span class="html-italic">Myh9</span><sup>fl/fl</sup> and <span class="html-italic">Myh9</span><sup>ECKO</sup> mice, showing CD31 (green) and Claudin-5 (red). Scale bar: 50 μm. (<b>H</b>) Statistical analysis of colocalization of ZO-1, Occludin, Claudin-5, and CD31 in brain sections from <span class="html-italic">Myh9</span><sup>fl/fl</sup> and <span class="html-italic">Myh9</span><sup>ECKO</sup> mice (<span class="html-italic">n</span> = 3). *** <span class="html-italic">p</span> &lt; 0.001; Student’s <span class="html-italic">t</span>-test.</p>
Full article ">Figure 4
<p>Myosin IIA mediates the transcription of <span class="html-italic">Ctnnb1</span> gene and interacts with its protein β-catenin. (<b>A</b>) RNA was extracted from primary brain ECs of <span class="html-italic">Myh9</span><sup>fl/fl</sup> and <span class="html-italic">Myh9</span><sup>ECKO</sup> mice, and RT-PCR was conducted to assess mRNA expression levels of key molecules regulating the blood–brain barrier (<span class="html-italic">n</span> = 3–4). ** <span class="html-italic">p</span> &lt; 0.01; Student’s <span class="html-italic">t</span>-test. (<b>B</b>) Western blot analysis of β-catenin protein levels in primary brain ECs from <span class="html-italic">Myh9</span><sup>fl/fl</sup> and <span class="html-italic">Myh9</span><sup>ECKO</sup> mice. (<b>C</b>) Statistical analysis of β-catenin protein levels in primary brain ECs from <span class="html-italic">Myh9</span><sup>fl/fl</sup> and <span class="html-italic">Myh9</span><sup>ECKO</sup> mice (<span class="html-italic">n</span> = 3). * <span class="html-italic">p</span> &lt; 0.05; Student’s <span class="html-italic">t</span>-test. (<b>D</b>) RT-PCR analysis of <span class="html-italic">Ctnnb1</span> mRNA levels in SCR-shRNA and <span class="html-italic">Myh9</span>-shRNA group cells (<span class="html-italic">n</span> = 3). ** <span class="html-italic">p</span> &lt; 0.01; Student’s <span class="html-italic">t</span>-test. (<b>E</b>) Representative immunofluorescence staining of Myosin IIA in HCMEC/D3 brain ECs, with Myosin IIA (red) and DAPI (blue). Scale bar: 20 μm. (<b>F</b>) Construction strategy for a luciferase reporter plasmid containing the Ctnnb1 gene promoter. (<b>G</b>) Brain ECs overexpressing either control or Myosin IIA were co-transfected with the <span class="html-italic">Ctnnb1</span> gene promoter luciferase reporter plasmid and Renilla luciferase reporter plasmid. Cells were collected 24 h post-transfection to measure their activity (<span class="html-italic">n</span> = 3). ** <span class="html-italic">p</span> &lt; 0.01; Student’s <span class="html-italic">t</span>-test. (<b>H</b>) Interaction between Myosin IIA and β-catenin. Once bEnd.3 brain ECs reached a confluent monolayer, cells were lysed, proteins were harvested, and co-immunoprecipitation was performed to detect interactions between Myosin IIA and β-catenin proteins.</p>
Full article ">Figure 5
<p>Overexpression of β-catenin ameliorates BBB leakage in <span class="html-italic">Myh9</span><sup>ECKO</sup> mice. (<b>A</b>) Construction strategy for AAV-GFP and AAV-HA-β-catenin overexpression plasmids. (<b>B</b>) Transfection of AAV-GFP and AAV-HA-β-catenin plasmids into 293T cells; 48 h later, cell proteins were harvested and analyzed by Western blot to assess β-catenin protein levels. (<b>C</b>) Representative immunofluorescence co-staining images of brain sections from <span class="html-italic">Myh9</span><sup>fl/fl</sup> and <span class="html-italic">Myh9</span><sup>ECKO</sup> mice 30 days post-injection with AAV-GFP virus showing CD31 (red) and GFP (green); and 30 days post-injection with AAV-HA-β-catenin virus showing CD31 (red) and HA-tag (green). Scale bar: 100 μm. (<b>D</b>) Intravenous injection of AAV-GFP or AAV-HA-β-catenin virus in <span class="html-italic">Myh9</span><sup>fl/fl</sup> and <span class="html-italic">Myh9</span><sup>ECKO</sup> mice; 30 days later, Evans blue was injected intravenously, and the total amount of Evans blue in the brain tissues of different treatment groups was quantitatively analyzed <span class="html-italic">(n</span> = 5). *** <span class="html-italic">p</span> &lt; 0.001; one-way ANOVA test. (<b>E</b>) Intravenous injection of AAV-GFP or AAV-HA-β-catenin virus in <span class="html-italic">Myh9</span><sup>fl/fl</sup> and <span class="html-italic">Myh9</span><sup>ECKO</sup> mice; 30 days later, Sulfo-NHS-Biotin was injected, and brain sections were co-stained for CD31 (red) and Sulfo-NHS-Biotin (green). Scale bar: 100 μm. (<b>F</b>) Statistical analysis of Sulfo-NHS-Biotin leakage index in brain tissues of <span class="html-italic">Myh9</span><sup>fl/fl</sup> and <span class="html-italic">Myh9</span><sup>ECKO</sup> mice 30 days post-injection with AAV-GFP or AAV-HA-β-catenin virus (<span class="html-italic">n</span> = 3). Scale bar: 200 μm. * <span class="html-italic">p</span> &lt; 0.05; one-way ANOVA test.</p>
Full article ">Figure 6
<p>Overexpression of β-catenin in cerebral endothelium of <span class="html-italic">Myh9</span><sup>ECKO</sup> mice ameliorates epilepsy-induced injury. (<b>A</b>) Statistical analysis of the total amount of Evans blue tracer in the brain tissues of <span class="html-italic">Myh9</span><sup>fl/fl</sup> and <span class="html-italic">Myh9</span><sup>ECKO</sup> mice injected intravenously with AAV-GFP or AAV-HA-β-catenin viruses, followed by pilocarpine-induced epilepsy 30 days later (<span class="html-italic">n</span> = 5). *** <span class="html-italic">p</span> &lt; 0.001; one-way ANOVA test. (<b>B</b>) Statistical analysis of the incidence of epilepsy in <span class="html-italic">Myh9</span><sup>fl/fl</sup> and <span class="html-italic">Myh9</span><sup>ECKO</sup> mice after intravenous injection of AAV-GFP or AAV-HA-β-catenin viruses and subsequent pilocarpine administration 30 days later (NS, no seizure group; Seizure, seizure group; <span class="html-italic">n</span> = 10). ** <span class="html-italic">p</span> &lt; 0.01, *** <span class="html-italic">p</span> &lt; 0.001; Fisher’s exact test. (<b>C</b>) Statistical analysis of the time to first seizure in <span class="html-italic">Myh9</span><sup>fl/fl</sup> and <span class="html-italic">Myh9</span><sup>ECKO</sup> mice treated with AAV-GFP or AAV-HA-β-catenin viruses and subsequently induced with pilocarpine (<span class="html-italic">Myh9</span><sup>fl/fl</sup>+Vector and <span class="html-italic">Myh9</span><sup>fl/fl</sup>+HA-β-catenin, <span class="html-italic">n</span> = 5; <span class="html-italic">Myh9</span><sup>ECKO</sup>, <span class="html-italic">n</span> = 8; <span class="html-italic">Myh9</span><sup>ECKO</sup> +HA-β-catenin, <span class="html-italic">n</span> = 6). ** <span class="html-italic">p</span> &lt; 0.01; one-way ANOVA test. (<b>D</b>) Statistical analysis of survival time for <span class="html-italic">Myh9</span><sup>fl/fl</sup> and <span class="html-italic">Myh9</span><sup>ECKO</sup> mice injected with AAV-GFP or AAV-HA-β-catenin viruses and then induced with pilocarpine 30 days later, with a total monitoring duration of 120 min (<span class="html-italic">n</span> = 10). * <span class="html-italic">p</span> &lt; 0.05; one-way ANOVA test.</p>
Full article ">
18 pages, 1785 KiB  
Article
Multi-Omics Analysis Unravels the Impact of Stool Sample Logistics on Metabolites and Microbial Composition
by Jannike L. Krause, Beatrice Engelmann, David J. D. Lallinger, Ulrike Rolle-Kampczyk, Martin von Bergen and Hyun-Dong Chang
Microorganisms 2024, 12(10), 1998; https://doi.org/10.3390/microorganisms12101998 - 30 Sep 2024
Viewed by 228
Abstract
Human health and the human microbiome are inevitably intertwined, increasing their relevance in clinical research. However, the collection, transportation and storage of faecal samples may introduce bias due to methodological differences, especially since postal shipping is a common practise in large-scale clinical cohort [...] Read more.
Human health and the human microbiome are inevitably intertwined, increasing their relevance in clinical research. However, the collection, transportation and storage of faecal samples may introduce bias due to methodological differences, especially since postal shipping is a common practise in large-scale clinical cohort studies. Using four different Omics layer, we determined the structural (16S rRNA sequencing, cytometric microbiota profiling) and functional integrity (SCFAs, global metabolome) of the microbiota in relation to different easy-to-handle conditions. These conditions were storage at −20 °C, −20 °C as glycerol stock, 4 °C and room temperature with and without oxygen exposure for a maximum of one week. Storage time affected the microbiota on all Omics levels. However, the magnitude was donor-dependent, highlighting the need for purpose-optimized sample collection in clinical multi-donor studies. The effects of oxygen exposure were negligible for all analyses. At ambient temperature, SCFA and compositional profiles were stable for 24 h and 48 h, respectively, while at 4 °C, SCFA profiles were maintained for 48 h. The global metabolome was highly susceptible, already changing at 24 h in non-frozen conditions. Thus, faecal microbiota was best preserved on all levels when transported as a native sample frozen within 24 h, leading to the least biased outcomes in the analysis. We conclude that the immediate freezing of native stool samples for transportation to the lab is best suited for planned multi-Omics analyses that include metabolomics to extend standard sequencing approaches. Full article
(This article belongs to the Special Issue Effects of Gut Microbiota on Human Health and Disease)
33 pages, 4414 KiB  
Article
Effects of Pressure, Hypoxia, and Hyperoxia on Neutrophil Granulocytes
by Richard F. Kraus, Daniel Panter, Michael A. Gruber, Stephanie Arndt, Petra Unger, Michael T. Pawlik and Diane Bitzinger
Biomolecules 2024, 14(10), 1242; https://doi.org/10.3390/biom14101242 - 30 Sep 2024
Viewed by 252
Abstract
Background: The application of normo- and hyperbaric O2 is a common therapy option in various disease patterns. Thereby, the applied O2 affects the whole body, including the blood and its components. This study investigates influences of pressure and oxygen fraction on [...] Read more.
Background: The application of normo- and hyperbaric O2 is a common therapy option in various disease patterns. Thereby, the applied O2 affects the whole body, including the blood and its components. This study investigates influences of pressure and oxygen fraction on human blood plasma, nutrient media, and the functions of neutrophil granulocytes (PMNs). Methods: Neutrophil migration, reactive oxygen species (ROS) production, and NETosis were examined by live cell imaging. The treatment of various matrices (Roswell Park Memorial Institute 1640 medium, Dulbecco’s Modified Eagle’s Medium, H2O, human plasma, and isolated PMNs) with hyperbaric oxygen (HBO) was performed. In addition, the expression of different neutrophil surface epitopes (CD11b, CD62L, CD66b) and the oxidative burst were investigated by flow cytometry (FACS). The application of cold atmospheric plasma (CAP) to normoxic and normobaric culture media served as a positive control. Soluble reaction products such as H2O2, reactive nitrogen species (RNS: NO2 and NO3), and ROS-dependent dihydrorhodamine oxidation were quantified by fluoro- and colorimetric assay kits. Results: Under normobaric normoxia, PMNs migrate slower and shorter in comparison with normobaric hyper- or hypoxic conditions and hyperbaric hyperoxia. The pressure component has less effect on the migration behavior of PMNs than the O2 concentration. Individual PMN cells produce prolonged ROS at normoxic conditions. PMNs showed increased expression of CD11b in normobaric normoxia, lower expression of CD62L in normobaric normoxia, and lower expression of CD66b after HBO and CAP treatment. Treatment with CAP increased the amount of ROS and RNS in common culture media. Conclusions: Hyperbaric and normobaric O2 influences neutrophil functionality and surface epitopes in a measurable way, which may have an impact on disorders with neutrophil involvement. In the context of hyperbaric experiments, especially high amounts of H2O2 in RPMI after hyperbaric oxygen should be taken into account. Therefore, our data support a critical indication for the use of normobaric and hyperbaric oxygen and conscientious and careful handling of oxygen in everyday clinical practice. Full article
20 pages, 5950 KiB  
Article
The Synergistic Combination of Curcumin and Polydatin Improves Temozolomide Efficacy on Glioblastoma Cells
by Annalucia Serafino, Ewa Krystyna Krasnowska, Sabrina Romanò, Alex De Gregorio, Marisa Colone, Maria Luisa Dupuis, Massimo Bonucci, Giampietro Ravagnan, Annarita Stringaro and Maria Pia Fuggetta
Int. J. Mol. Sci. 2024, 25(19), 10572; https://doi.org/10.3390/ijms251910572 - 30 Sep 2024
Viewed by 311
Abstract
Glioblastoma (GBL) is one of the more malignant primary brain tumors; it is currently treated by a multimodality strategy including surgery, and radio- and chemotherapy, mainly consisting of temozolomide (TMZ)-based chemotherapy. Tumor relapse often occurs due to the establishment of TMZ resistance, with [...] Read more.
Glioblastoma (GBL) is one of the more malignant primary brain tumors; it is currently treated by a multimodality strategy including surgery, and radio- and chemotherapy, mainly consisting of temozolomide (TMZ)-based chemotherapy. Tumor relapse often occurs due to the establishment of TMZ resistance, with a patient median survival time of <2 years. The identification of natural molecules with strong anti-tumor activity led to the combination of these compounds with conventional chemotherapeutic agents, developing protocols for integrated anticancer therapies. Curcumin (CUR), resveratrol (RES), and its glucoside polydatin (PLD) are widely employed in the pharmaceutical and nutraceutical fields, and several studies have demonstrated that the combination of these natural products was more cytotoxic than the individual compounds alone against different cancers. Some of us recently demonstrated the synergistic efficacy of the sublingual administration of a new nutraceutical formulation of CUR+PLD in reducing tumor size and improving GBL patient survival. To provide some experimental evidence to reinforce these clinical results, we investigated if pretreatment with a combination of CUR+PLD can improve TMZ cytotoxicity on GBL cells by analyzing the effects on cell cycle, viability, morphology, expression of proteins related to cell proliferation, differentiation, apoptosis or autophagy, and the actin network. Cell viability was assessed using the MTT assay or a CytoSmart cell counter. CalcuSyn software was used to study the CUR+PLD synergism. The morphology was evaluated by optical and scanning electron microscopy, and protein expression was analyzed by Western blot. Flow cytometry was used for the cell cycle, autophagic flux, and apoptosis analyses. The results provide evidence that CUR and PLD, acting in synergy with each other, strongly improve the efficacy of alkylating anti-tumor agents such as TMZ on drug-resistant GBL cells through their ability to affect survival, differentiation, and tumor invasiveness. Full article
(This article belongs to the Special Issue New Agents and Novel Drugs Use for the Oncological Diseases Treatment)
Show Figures

Figure 1

Figure 1
<p>Features of cell lines used (<b>a</b>,<b>b</b>) and dose- and time-dependent responses to TMZ of U87 and LN18 cells (<b>c</b>). (<b>a</b>) Phase contrast microscopy (<b>top</b> panels) and scanning electron microscopy (<b>bottom</b> panels) images illustrating the morphological features of U87 and LN18 glioblastoma cell lines. The scale bars represent 100 μm. (<b>b</b>) Baseline expression, measured by WB, of the DNA repair protein O6-methylguanine-DNA-methyltransferase (MGMT) in U87 and LN18 cells. GAPDH was used as the loading control. (<b>c</b>) Time- and dose-dependent response experiments performed using the MTT assay to define the optimal condition for TMZ treatment of U87 and LN18 cells. Results are the mean ± SD from three independent experiments (<span class="html-italic">n</span> = 3). Significance vs. untreated control (two-tailed Student’s <span class="html-italic">t</span>-test): * <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; ns: not significant.</p>
Full article ">Figure 2
<p>Experimental scheme (<b>a</b>) and MTT cell viability assay results (<b>b</b>,<b>c</b>) from the preliminary assessment of the efficacy of the selected synergic combination of CUR+PLD. (<b>a</b>) Scheme of the protocol used for evaluating the effects of treatments on morphology, viability, and protein expression in both cell lines, as detailed in the Methods section. (<b>b</b>,<b>c</b>) Viability assay (MTT assay) results for U87 and LN18 cells subjected to TMZ treatments in the presence or absence of CUR and PLD as single or combined pretreatment. Values are the mean ± SD from three independent experiments (<span class="html-italic">n</span> = 3). Significance vs untreated control (two-tailed Student’s <span class="html-italic">t</span>-test): *** <span class="html-italic">p</span> &lt; 0.001.</p>
Full article ">Figure 3
<p>Effects of the synergic combination of CUR+PLD before TMZ treatment on U87 and LN18 cells. (<b>a</b>,<b>c</b>) Phase contrast microscopy images showing the morphological changes induced by 24 h CUR+PLD treatment on U87 (<b>a</b>) and LN18 (<b>c</b>) cells. The scale bars represent 100 μm. (<b>b</b>,<b>d</b>,<b>e</b>) WB and analysis of modifications induced by 24 h CUR+PLD treatment on expression levels of the proliferation marker c-Myc, the astroglial differentiation marker GFAP, the autophagosomal marker LC3B and the marker of caspase-dependent apoptosis PARP. Histograms in (<b>b</b>,<b>d</b>) report the densitometric analysis results for GFAP, c-Myc, and LC3B expression; values were normalized to GAPDH. Results are the mean ± SD from three independent experiments (<span class="html-italic">n</span> = 3). Significance vs. control (two-tailed Student’s <span class="html-italic">t</span>-test): * <span class="html-italic">p</span> &lt; 0.05, ** <span class="html-italic">p</span> &lt; 0.01; ns: not significant.</p>
Full article ">Figure 4
<p>Effects of pretreatment with the synergic combination of CUR+PLD on TMZ-induced effects on U87 cells. (<b>a</b>) Phase contrast microscopy images showing the morphological modifications induced in untreated controls and in samples pretreated for 24 h with CUR+PLD with or without TMZ post-treatment. Arrows point to rare cells displaying morphological features suggestive of apoptosis/autophagy. The scale bars represent 100 μm. (<b>b</b>–<b>e</b>) Analysis results from the automated CytoSmart cell counter: the number of live cells (<b>b</b>), cell growth between T0 and T 96 h (<b>c</b>), percent viable cells (<b>d</b>), and cell size (<b>e</b>). Results are the mean ± SD from three independent experiments (<span class="html-italic">n</span> = 3). Significance (one-way ANOVA + Tukey multiple comparison test): ** <span class="html-italic">p</span> &lt; 0.01, *** <span class="html-italic">p</span> &lt; 0.001 vs. untreated control (CTR); ## <span class="html-italic">p</span> &lt; 0.01, ### <span class="html-italic">p</span> &lt; 0.001 vs. TMZ alone or pretreated with CUR+PLD; ns: not significant.</p>
Full article ">Figure 5
<p>Effects of pretreatment with the synergic combination of CUR+PLD on TMZ-induced effects on LN18 cells. (<b>a</b>) Phase contrast microscopy images showing the morphological modification induced in untreated controls and in samples pretreated for 24 h with CUR+PLD with or without TMZ post-treatment. Arrows point to numerous cells displaying morphological features suggestive of apoptosis/autophagy. The scale bars represent 100 μm. (<b>b</b>–<b>e</b>) Analysis results from the automated CytoSmart cell counter: the number of live cells (<b>b</b>), cell growth between T0 and T 96 h (<b>c</b>), percent viable cells (<b>d</b>), and cell size (<b>e</b>). Results are the mean ± SD from three independent experiments (<span class="html-italic">n</span> = 3). Significance (one-way ANOVA + Tukey multiple comparison test): * <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 vs. untreated control (CTR); # <span class="html-italic">p</span> &lt; 0.05, ## <span class="html-italic">p</span> &lt; 0.01 vs. TMZ alone or pretreated with CUR+PLD; ns: not significant.</p>
Full article ">Figure 6
<p>Effects of TMZ treatment with or without pretreatment with CUR+PLD on cell cycle, autophagic flux, and apoptosis. (<b>a</b>,<b>b</b>) The bar graphs illustrate the results of the cytofluorimetric analysis of the cell cycle in U87 (<b>a</b>) and LN18 (<b>b</b>) cells between 48 and 72 h of culture. (<b>c</b>,<b>d</b>) Cytofluorimetric analysis of autophagy performed using the Cyto-ID autophagy detection kit after 72 h of TMZ treatment in U87 (<b>c</b>) and LN18 (<b>d</b>) cells. (<b>e</b>,<b>f</b>) Cytofluorimetric evaluation of apoptosis using Annexin V (AV) staining in U87 (<b>e</b>) and LN18 (<b>f</b>) cells. PI−/AV+, PI+/AV+, and PI+/AV− counts indicate early apoptotic, late apoptotic, and necrotic cells, respectively. Significance vs control (two-tailed Student’s <span class="html-italic">t</span>-test): * <span class="html-italic">p</span> &lt; 0.05, ** <span class="html-italic">p</span> &lt; 0.01; ns: not significant.</p>
Full article ">Figure 7
<p>Effects on expression of MGMT and proteins related to cell proliferation, apoptosis, and autophagy. Western blot (<b>a</b>,<b>b</b>) and densitometric analysis (<b>c</b>,<b>d</b>) of the expression levels of MGMT, the proliferation marker c-Myc, the astroglial differentiation marker GFAP, the autophagosomal marker LC3B, and the marker of caspase-dependent apoptosis PARP, evaluated after 72 h of TMZ treatment with or without pretreatment with CUR+PLD in U87 (<b>a</b>,<b>c</b>) and LN18 (<b>b</b>,<b>d</b>) cells; values were normalized to GAPDH. Results are the mean ± SD from three independent experiments (<span class="html-italic">n</span> = 3). Significance (one-way ANOVA + Tukey multiple comparison test): * <span class="html-italic">p</span> &lt; 0.05, ** <span class="html-italic">p</span> &lt; 0.01 vs. untreated control (CTR), # <span class="html-italic">p</span> &lt; 0.05, ## <span class="html-italic">p</span> &lt; 0.01 vs. TMZ alone or pretreated with CUR+PLD; ns: not significant.</p>
Full article ">Figure 8
<p>Effect of pretreatment with CUR+PLD on the actin network and ultrastructural morphology of U87 cells. Confocal microscopic analysis of the actin network (<b>left</b> panels) and SEM analysis of morphology (<b>right</b> panels), evaluated after 72 h of TMZ treatment with or without pretreatment with CUR+PLD. The scale bars represent 25 μm.</p>
Full article ">Figure 9
<p>Effect of pretreatment with CUR+PLD on the actin network and ultrastructural morphology of LN18 cells. Confocal microscopic analysis of the actin network (<b>left</b> panels) and SEM analysis of morphology (<b>right</b> panels), evaluated after 72 h of TMZ treatment with or without pretreatment with CUR+PLD. The scale bars represent 25 μm.</p>
Full article ">
20 pages, 4246 KiB  
Article
TongGuanWan Alleviates Doxorubicin- and Isoproterenol-Induced Cardiac Hypertrophy and Fibrosis by Modulating Apoptotic and Fibrotic Pathways
by Jung-Joo Yoon, Ai-Lin Tai, Hye-Yoom Kim, Byung-Hyuk Han, Sarah Shin, Ho-Sub Lee and Dae-Gill Kang
Int. J. Mol. Sci. 2024, 25(19), 10573; https://doi.org/10.3390/ijms251910573 - 30 Sep 2024
Viewed by 217
Abstract
Heart failure, a major public health issue, often stems from prolonged stress or damage to the heart muscle, leading to cardiac hypertrophy. This can progress to heart failure and other cardiovascular problems. Doxorubicin (DOX), a common chemotherapy drug, and isoproterenol (ISO), a β-adrenergic [...] Read more.
Heart failure, a major public health issue, often stems from prolonged stress or damage to the heart muscle, leading to cardiac hypertrophy. This can progress to heart failure and other cardiovascular problems. Doxorubicin (DOX), a common chemotherapy drug, and isoproterenol (ISO), a β-adrenergic agonist, both induce cardiac hypertrophy through different mechanisms. This study investigates TongGuanWan (TGW,), a traditional herbal remedy, for its effects on cardiac hypertrophy and fibrosis in DOX-induced H9c2 cells and ISO-induced mouse models. TGW was found to counteract DOX-induced increases in H9c2 cell surface area (n = 8, p < 0.01) and improve biomarkers like ANP (n = 3, p < 0.01)) and BNP (n = 3, p < 0.01). It inhibited the MAPK pathway (n = 4, p < 0.01) and GATA-4/calcineurin/NFAT-3 signaling, reduced inflammation by decreasing NF-κB p65 translocation, and enhanced apoptosis-related factors such as caspase-3 (n = 3, p < 0.01), caspase-9 (n = 3, p < 0.01), Bax (n = 3, p < 0.01), and Bcl-2 (n = 3, p < 0.01). Flow cytometry showed TGW reduced apoptotic cell populations. In vivo, TGW reduced heart (n = 8~10, p < 0.01), and left ventricle weights (n = 6~7), cardiac hypertrophy markers (n = 3, p < 0.01), and perivascular fibrosis in ISO-induced mice, with Western blot analysis confirming decreased levels of fibrosis-related factors like fibronectin, α-SMA (n = 3, p < 0.05), and collagen type I (n = 3, p < 0.05). These findings suggest TGW has potential as a therapeutic option for cardiac hypertrophy and fibrosis. Full article
(This article belongs to the Section Bioactives and Nutraceuticals)
15 pages, 954 KiB  
Article
Differential Diagnosis of Tuberculosis and Sarcoidosis by Immunological Features Using Machine Learning
by Nikolay Osipov, Igor Kudryavtsev, Dmitry Spelnikov, Artem Rubinstein, Ekaterina Belyaeva, Anastasia Kulpina, Dmitry Kudlay and Anna Starshinova
Diagnostics 2024, 14(19), 2188; https://doi.org/10.3390/diagnostics14192188 - 30 Sep 2024
Viewed by 177
Abstract
Despite the achievements of modern medicine, tuberculosis remains one of the leading causes of mortality globally. The difficulties in differential diagnosis have particular relevance in the case of suspicion of tuberculosis with other granulomatous diseases. The most similar clinical and radiologic changes are [...] Read more.
Despite the achievements of modern medicine, tuberculosis remains one of the leading causes of mortality globally. The difficulties in differential diagnosis have particular relevance in the case of suspicion of tuberculosis with other granulomatous diseases. The most similar clinical and radiologic changes are sarcoidosis. The aim of this study is to apply mathematical modeling to determine diagnostically significant immunological parameters and an algorithm for the differential diagnosis of tuberculosis and sarcoidosis. Materials and methods: The serum samples of patients with sarcoidosis (SD) (n = 29), patients with pulmonary tuberculosis (TB) (n = 32) and the control group (n = 31) (healthy subjects) collected from 2017 to 2022 (the average age 43.4 ± 5.3 years) were examined. Circulating ‘polarized’ T-helper cell subsets were analyzed by multicolor flow cytometry. A symbolic regression method was used to find general mathematical relations between cell concentrations and diagnosis. The parameters of the selected model were finally fitted through multi-objective optimization applied to two conflicting indices: sensitivity to sarcoidosis and sensitivity to tuberculosis. Results: The difference in Bm2 and CD5−CD27− concentrations was found to be more significant for the differential diagnosis of sarcoidosis and tuberculosis than any individual concentrations: the combined feature Bm2 − [CD5−CD27−] differentiates sarcoidosis and tuberculosis with p < 0.00001 and AUC = 0.823. An algorithm for differential diagnosis was developed. It is based on the linear model with two variables: the first variable is the difference Bm2 − [CD5−CD27−] mentioned above, and the second is the naïve-Tregs concentration. The algorithm uses the model twice and returns “dubious” in 26.7% of cases for patients with sarcoidosis and in 16.1% of cases for patients with tuberculosis. For the remaining patients with one of these two diagnoses, its sensitivity to sarcoidosis is 90.5%, and its sensitivity to tuberculosis is 88.5%. Conclusions: A simple algorithm was developed that can distinguish, by certain immunological features, the cases in which sarcoidosis is likely to be present instead of tuberculosis. Such cases may be further investigated to rule out tuberculosis conclusively. The mathematical model underlying the algorithm is based on the analysis of “naive” T-regulatory cells and “naive” B-cells. This may be a promising approach for differential diagnosis between pulmonary sarcoidosis and pulmonary tuberculosis. The findings may be useful in the absence of clear differential diagnostic criteria between pulmonary tuberculosis and sarcoidosis. Full article
Show Figures

Figure 1

Figure 1
<p>The distributions of the difference Bm2 − [CD5−CD27−] depending on the group.</p>
Full article ">Figure 2
<p>Visualization of the differential diagnosis algorithm on training and test samples. The high-risk zone for sarcoidosis is below the red graph, the high-risk zone for tuberculosis is above the blue graph, and the zone of uncertainty is between these two graphs.</p>
Full article ">Figure 3
<p>The final algorithm for differential diagnosis between sarcoidosis and tuberculosis derived from the entire dataset.</p>
Full article ">
23 pages, 6688 KiB  
Article
1,2,4-Oxadiazole Derivatives: Physicochemical Properties, Antileishmanial Potential, Docking and Molecular Dynamic Simulations of Leishmania infantum Target Proteins
by Deyzi C. S. Barbosa, Vanderlan N. Holanda, Elton M. A. Lima, Marton K. A. Cavalcante, Maria Carolina A. Brelaz-de-Castro, Elton J. F. Chaves, Gerd B. Rocha, Carla J. O. Silva, Ronaldo N. Oliveira and Regina C. B. Q. Figueiredo
Molecules 2024, 29(19), 4654; https://doi.org/10.3390/molecules29194654 - 30 Sep 2024
Viewed by 324
Abstract
Visceral leishmaniasis (VL), caused by protozoa of the genus Leishmania, remains a significant public health concern due to its potentially lethal nature if untreated. Current chemotherapy options are limited by severe toxicity and drug resistance. Derivatives of 1,2,4-oxadiazole have emerged as promising [...] Read more.
Visceral leishmaniasis (VL), caused by protozoa of the genus Leishmania, remains a significant public health concern due to its potentially lethal nature if untreated. Current chemotherapy options are limited by severe toxicity and drug resistance. Derivatives of 1,2,4-oxadiazole have emerged as promising drug candidates due to their broad biological activity. This study investigated the effects of novel 1,2,4-oxadiazole derivatives (Ox1Ox7) on Leishmania infantum, the etiological agent of VL. In silico predictions using SwissADME suggest that these compounds have high oral absorption and good bioavailability. Among them, Ox1 showed the most promise, with higher selectivity against promastigotes and lower cytotoxicity towards L929 fibroblasts and J774.G8 macrophages. Ox1 exhibited selectivity indices of 18.7 and 61.7 against L. infantum promastigotes and amastigotes, respectively, compared to peritoneal macrophages. Ultrastructural analyses revealed severe morphological damage in both parasite forms, leading to cell death. Additionally, Ox1 decreased the mitochondrial membrane potential in promastigotes, as shown by flow cytometry. Molecular docking and dynamic simulations indicated a strong affinity of Ox1 for the L. infantum CYP51 enzyme. Overall, Ox1 is a promising and effective compound against L. infantum. Full article
(This article belongs to the Special Issue The Design, Synthesis, and Biological Activity of New Drug Candidates)
Show Figures

Graphical abstract

Graphical abstract
Full article ">Figure 1
<p>Representative graphs of the effects of <b>Ox1</b> on mammalian cells and promastigotes. (<b>A</b>) Activity of <b>Ox1</b> on the viability of fibroblast (L929) and (<b>B</b>) macrophages (J774.G8) after 48 h of drug treatment. (<b>C</b>) Effects of <b>Ox1</b> on promastigotes after 24 (black bars) and 48 (gray bars) hours of treatment. Values represent the mean ± standard deviation of three independent experiments in triplicate. * Significant differences at <span class="html-italic">p</span> &lt; 0.05 compared to untreated control.</p>
Full article ">Figure 2
<p>Effects of <b>Ox1</b> on the viability of PeMs and intracellular amastigotes of forms of <span class="html-italic">L. infantum</span>. (<b>A</b>) Percentage of viable treated PeMs compared to untreated ones. (<b>B</b>) Total number of amastigotes in 300 infected cells. (<b>C</b>) Survival index (SuI) of amastigotes in PeMs. Glu—glucantime; AmB—Amphotericin B. Each bar represents the mean ± SD of three independent experiments performed in duplicate. * Significant differences at <span class="html-italic">p</span> &lt; 0.05 compared to the untreated control. (<b>D</b>) Representative images of untreated infected-PeMs negative control (NC) or treated with Glu, AmB, or <b>Ox1</b> at ¼ (8.24 µM), ½ IC<sub>50</sub> (16.42 µM), and IC<sub>50</sub> (32.98 µM) for promastigote forms. Intracellular amastigotes are indicated by arrows. Note in the culture treated with AmB the presence of cellular debris (thick arrow). Cells treated with IC<sub>50</sub> and ½ IC<sub>50</sub> of <b>Ox1</b> or AmB, presenting empty parasitophorous vacuoles (*) can be also observed. Bars 25 µm.</p>
Full article ">Figure 3
<p>Scanning electron microscopy of control and treated promastigotes of <span class="html-italic">L. infantum</span>. (<b>A</b>) Detail of untreated promastigote showing elongated cell body, smooth plasma membrane, and long flagellum. (<b>B</b>) Untreated dividing promastigotes showing a preserved morphology. (<b>C</b>) Low magnification of untreated culture showing the predominance of elongated and dividing cells. (<b>D</b>) treated culture showing rounded cells with twisted flagellum around the parasite cell body next to an elongated cell. (<b>E</b>) High magnification of treated cell showing rounded cell body, septation, and altered and short flagellum (arrow). (<b>F</b>) Low magnification of culture treated with IC<sub>50</sub> <b>Ox1</b> showing numerous promastigote rosettes (R). (<b>G</b>) Severe injured cells treated with 2× IC<sub>50</sub> <b>Ox1</b>. Note the presence of cell membrane perforations (arrow), the loss of characteristic morphology, and the absence of visible flagellum. (<b>H</b>) At low magnification, it is possible to observe the decrease in cell number and the increase in cells presenting altered rounded morphology. (<b>I</b>) Representative graph of the percentage of rounded promastigotes treated with IC<sub>50</sub> (32.98 µM) and 2× IC<sub>50</sub> (65.96 µM). A total of 500 cells were randomly counted per sample. * Significant differences at <span class="html-italic">p</span> &lt; 0.05 compared to the untreated control. Bars: (<b>A</b>,<b>B</b>,<b>D</b>,<b>E</b>) =2 µm; (<b>C</b>,<b>F</b>,<b>H</b>) =20 µm; (<b>G</b>) =1 µm.</p>
Full article ">Figure 4
<p>Transmission electron microscopy of control and treated promastigotes. (<b>A</b>) Untreated promastigote showing dispersed endoplasmic reticulum (arrow), preserved nucleus (N), mitochondrion (m), and lipid droplets (L). (<b>B</b>) Promastigote treated with IC<sub>50</sub> <b>Ox1</b> showing altered mitochondrion (asterisk) and increased endoplasmic reticulum (arrow). (<b>C</b>) Promastigote treated with 2× IC<sub>50</sub> showing multiple flagella within the flagellar pocket (arrow). (<b>D</b>) Note the presence of varying levels of cellular damage (1–3), increase in endoplasmic reticulum (1), and lipid droplets (2 and 3). Bars: (<b>A</b>,<b>B</b>)—2 µm. (<b>C</b>,<b>D</b>)—1 µm.</p>
Full article ">Figure 5
<p>Representative histogram of untreated cells (control) and cells treated with half, once, and twice the IC<sub>50</sub> of <b>Ox1</b>.</p>
Full article ">Figure 6
<p>Transmission electron microscopy of the effects of <b>Ox1</b> on (PeM) infected with amastigote. (<b>A</b>,<b>B</b>) Untreated infected cells displaying several amastigotes (star) inside the parasitophorous vacuole (PV). (<b>B</b>) Detail of dividing amastigote (star) with preserved cellular structure. (<b>C</b>,<b>D</b>) Infected macrophages treated with IC<sub>50</sub> <b>Ox1</b>. (<b>C</b>) Amastigotes (stars) presenting cytoplasmic vacuoles surrounded by membrane (arrowhead). Stretched kinetoplast showing partial rupture can be observed (arrow). (<b>D</b>) Amastigote with loss of cytoplasmic content, asterisk. (<b>E</b>,<b>F</b>) Infected PeM treated with 2× IC<sub>50</sub> <b>Ox1</b>. (<b>E</b>) Amastigotes (star) presenting numerous membrane-bound vacuoles with or without internal content. Note the presence of high amounts of cellular debris inside the PV. (<b>F</b>) Detail of partially degraded amastigote (Am) inside the PV. (<b>G</b>) Infected macrophages treated with Amphotericin B. Amastigotes were indicated by stars. (<b>H</b>) Non-infected and non-treated macrophage. PV, parasitophorous vacuole; Am, amastigote; PeM, peritoneal macrophages infected with amastigote; FP, flagellar pocket; N, nucleus; K, kinetoplast. Bars: (<b>A</b>,<b>C</b>,<b>E</b>,<b>F</b>) =1µm; (<b>G</b>,<b>H</b>) =2 µm; (<b>B</b>,<b>D</b>) =500 nm.</p>
Full article ">Figure 7
<p>Summary of the in-silico results. (<b>A</b>) Docking scores obtained with GOLD software; label (1) corresponds to the scores obtained with the Chemscore scoring function, (2) corresponds to the modified version of this scoring function, with iron parameters. (<b>B</b>) RMSD profile during MD simulation. (<b>C</b>) Analysis of the center of mass distance during MD simulation. (<b>D</b>) The CYP51/<b>Ox1</b> complex is shown on the left and a detailed view of the binding site is shown on the right, showing the hydrophobic contacts and two hydrogen bonds (hb) between CYP51 and <b>Ox1</b> compound: (i) the hydroxyl group of Tyr74 and the nitrogen atom (N1) of <b>Ox1</b>, and (ii) the oxygen atom of Met329 and hydrogen atom attached to the nitrogen (N) to the <b>Ox1</b>. The interaction energy profile between CYP51 (per residue) and <b>Ox1</b> is shown below. It is important to note that the <b>Ox1</b> binding pose does not reflect the interaction analysis entirely since this analysis was carried out considering more than one binding pose. The structure of the CYP51 receptor in the cartoon model is shown in gray, and the carbon atoms of the binding site are in sticks model and colored in gray, whereas the carbon atoms of <b>Ox1</b> are shown in green. The carbon atoms of the heme propionate are shown in pink. The hydrogen atoms were omitted to improve visualization.</p>
Full article ">Scheme 1
<p>Representative scheme of the synthesis of 1,2,4-oxadiazole derivatives <b>Ox1</b>–<b>Ox7</b>.</p>
Full article ">
19 pages, 2205 KiB  
Article
The PreS-Based Recombinant Vaccine VVX001 Induces Hepatitis B Virus Neutralizing Antibodies in a Low-Responder to HBsAg-Based HBV Vaccines
by Inna Tulaeva, Felix Lehmann, Nora Goldmann, Alexandra Dubovets, Daria Trifonova, Mikhail Tulaev, Carolin Cornelius, Milena Weber, Margarete Focke-Tejkl, Alexander Karaulov, Rainer Henning, David Niklas Springer, Ursula Wiedermann, Dieter Glebe and Rudolf Valenta
Vaccines 2024, 12(10), 1123; https://doi.org/10.3390/vaccines12101123 - 30 Sep 2024
Viewed by 422
Abstract
Background: Approximately 10–20% of subjects vaccinated with HBsAg-based hepatitis B virus (HBV) vaccines are non-responders. BM32 is a recombinant grass pollen allergy vaccine containing the HBV-derived preS surface antigen as an immunological carrier protein. PreS includes the binding site of HBV to its [...] Read more.
Background: Approximately 10–20% of subjects vaccinated with HBsAg-based hepatitis B virus (HBV) vaccines are non-responders. BM32 is a recombinant grass pollen allergy vaccine containing the HBV-derived preS surface antigen as an immunological carrier protein. PreS includes the binding site of HBV to its receptor on hepatocytes. We investigated whether immunological non-responsiveness to HBV after repeated HBsAg-based vaccinations could be overcome by immunization with VVX001 (i.e., alum-adsorbed BM325, a component of BM32). Methods: A subject failing to develop protective HBV-specific immunity after HBsAg-based vaccination received five monthly injections of 20 µg VVX001. PreS-specific antibody responses were measured by enzyme-linked immunosorbent assay (ELISA) and micro-array technology. Serum reactivity to subviral particles of different HBV genotypes was determined by sandwich ELISA. PreS-specific T cell responses were monitored by carboxyfluorescein diacetate succinimidyl ester (CFSE) staining and subsequent flow cytometry. HBV neutralization was assessed using cultured HBV-infected HepG2 cells. Results: Vaccination with VVX001 induced a strong and sustained preS-specific antibody response composed mainly of the IgG1 subclass. PreS-specific IgG antibodies were primarily directed to the N-terminal part of preS containing the sodium taurocholate co-transporting polypeptide (NTCP) attachment site. IgG reactivity to subviral particles as well as to the N-terminal preS-derived peptides was comparable for HBV genotypes A–H. A pronounced reactivity of CD3+CD4+ lymphocytes specific for preS after the complete injection course remaining up to one year after the last injection was found. Maximal HBV neutralization (98.4%) in vitro was achieved 1 month after the last injection, which correlated with the maximal IgG reactivity to the N-terminal part of preS. Conclusions: Our data suggest that VVX001 may be used as a preventive vaccination against HBV even in non-responders to HBsAg-based HBV vaccines. Full article
(This article belongs to the Special Issue 2nd Edition of Antibody Response to Infection and Vaccination)
Show Figures

Figure 1

Figure 1
<p>Vaccinations with conventional HBsAg-based vaccines in the study subject. Time points and administered vaccines are indicated.</p>
Full article ">Figure 2
<p>Development of antibodies to HBV surface proteins and their virus neutralization capacity. (<b>a</b>) Anti-HBs antibody levels (IU/L) measured after the last booster immunization with Engerix-B and after the preS-based vaccination. (<b>b</b>) PreS-specific IgG levels (OD) measured by ELISA after VVX001 immunization. (<b>c</b>) Percentages reduction in HBeAg secretion of infected NTCP-expressing HepG2 cells after pre-incubation of HBV inocula with sera obtained in the course of immunization compared to the baseline. Neutralization: ≥90% (strong neutralization), ≥50–90% (partial neutralization), ≥10–50% (low neutralization).</p>
Full article ">Figure 3
<p>VVX001-induced IgG antibodies react mainly with the N-terminus of preS1. (<b>a</b>) Localization scheme of preS-derived peptide within the preS sequence (amino acid numbering indicated for genotype A) and the levels of preS/peptide-specific IgG measured by (<b>b</b>) ELISA (OD) and (<b>c</b>) micro-array technology (fluorescence intensity, FI).</p>
Full article ">Figure 4
<p>VVX001-induced antibodies cross-react to the NTCP binding site of preS of HBV genotypes A–H. Shown are the IgG levels to the synthetic peptides representing the NTCP attachment site of HBV genotypes A–H measured by (<b>a</b>) ELISA (OD) and (<b>b</b>) micro-array technology (fluorescence intensity, FI).</p>
Full article ">Figure 5
<p>VVX001-induced antibodies react to SVPs of different HBV genotypes. Shown are IgG levels (OD) to SVPs of HBV genotypes A2, B2, C2, D1, E, F4, H (LHBs, MHBs, SHBs), and D3 (SHBs only) in the serum samples of the study subject before and after immunization as well as in two subjects successfully vaccinated with conventional HBsAg-based vaccines.</p>
Full article ">Figure 6
<p>IC<sub>50</sub> determination. IC<sub>50</sub> values for serum dilutions determined based on the HBeAg results are presented with a 95% confidence interval for (<b>a</b>) the anti-preS-positive human serum obtained in April 2019 and (<b>b</b>) for the anti-S-positive human serum (conventional vaccine, 2600 IU/L anti-HBs).</p>
Full article ">Figure 6 Cont.
<p>IC<sub>50</sub> determination. IC<sub>50</sub> values for serum dilutions determined based on the HBeAg results are presented with a 95% confidence interval for (<b>a</b>) the anti-preS-positive human serum obtained in April 2019 and (<b>b</b>) for the anti-S-positive human serum (conventional vaccine, 2600 IU/L anti-HBs).</p>
Full article ">Figure 7
<p>PreS-specific T cell responses after vaccination with VVX001. Shown are the percentages of proliferated CD3<sup>+</sup>CD4<sup>+</sup> and CD3<sup>+</sup>CD8<sup>+</sup> cells upon stimulation with (<b>a</b>) preS over time, (<b>b</b>) preS, P1-P8, equimolar mix of P1–P8, and (<b>c</b>) peptides representing the NTCP attachment site of genotypes A–H at the time point 4 months after the last injection (26 June 2019).</p>
Full article ">Figure 7 Cont.
<p>PreS-specific T cell responses after vaccination with VVX001. Shown are the percentages of proliferated CD3<sup>+</sup>CD4<sup>+</sup> and CD3<sup>+</sup>CD8<sup>+</sup> cells upon stimulation with (<b>a</b>) preS over time, (<b>b</b>) preS, P1-P8, equimolar mix of P1–P8, and (<b>c</b>) peptides representing the NTCP attachment site of genotypes A–H at the time point 4 months after the last injection (26 June 2019).</p>
Full article ">
24 pages, 1403 KiB  
Review
Advancement and Challenges in Monitoring of CAR-T Cell Therapy: A Comprehensive Review of Parameters and Markers in Hematological Malignancies
by Weronika Ploch, Karol Sadowski, Wioletta Olejarz and Grzegorz W. Basak
Cancers 2024, 16(19), 3339; https://doi.org/10.3390/cancers16193339 - 29 Sep 2024
Viewed by 630
Abstract
Chimeric antigen receptor T-cell (CAR-T) therapy has revolutionized the treatment for relapsed/refractory B-cell lymphomas. Despite its success, this therapy is accompanied by a significant frequency of adverse events, including cytokine release syndrome (CRS), immune-effector-cell-associated neurotoxicity syndrome (ICANS), or cytopenias, reaching even up to [...] Read more.
Chimeric antigen receptor T-cell (CAR-T) therapy has revolutionized the treatment for relapsed/refractory B-cell lymphomas. Despite its success, this therapy is accompanied by a significant frequency of adverse events, including cytokine release syndrome (CRS), immune-effector-cell-associated neurotoxicity syndrome (ICANS), or cytopenias, reaching even up to 80% of patients following CAR-T cell therapy. CRS results from the uncontrolled overproduction of proinflammatory cytokines, which leads to symptoms such as fever, headache, hypoxia, or neurological complications. CAR-T cell detection is possible by the use of flow cytometry (FC) or quantitative polymerase chain reaction (qPCR) assays, the two primary techniques used for CAR-T evaluation in peripheral blood, bone marrow (BM), and cerebrospinal fluid (CSF). State-of-the-art imaging technologies play a crucial role in monitoring the distribution and persistence of CAR-T cells in clinical trials. Still, they can also be extended with the use of FC and digital PCR (dPCR). Monitoring the changes in cell populations during disease progression and treatment gives an important insight into how the response to CAR-T cell therapy develops on a cellular level. It can help improve the therapeutic design and optimize CAR-T cell therapy to make it more precise and personalized, which is crucial to overcoming the problem of tumor relapse. Full article
(This article belongs to the Special Issue Innovative Immunotherapies: CAR-T Cell Therapy for Cancers)
Show Figures

Figure 1

Figure 1
<p>The immune checkpoint molecules on the surface of CAR-T cell. Abbreviations: Glucocorticoid-induced tumor-necrosis-factor-receptor-related protein (GITR), cytotoxic T-lymphocyte-associated antigen 4 (CTLA-4), programmed death-1 (PD-1), T-cell immunoglobulin and mucin domain 3 (TIM-3), B and T lymphocyte attenuator (BTLA), V-domain immunoglobulin suppressor of T cell activation (VISTA), lymphocyte activation gene 3 (LAG-3), and chimeric antigen receptor T (CAR-T) [<a href="#B122-cancers-16-03339" class="html-bibr">122</a>].</p>
Full article ">Figure 2
<p>Mechanism leading to CRS and ICANS. The red dots indicate TNF-<math display="inline"><semantics> <mi mathvariant="sans-serif">α</mi> </semantics></math> and IFN-<math display="inline"><semantics> <mi mathvariant="sans-serif">γ</mi> </semantics></math>, yellow dots indicate TNF-<math display="inline"><semantics> <mi mathvariant="sans-serif">α</mi> </semantics></math>, and purple dots indicate IL-1, IL-6, IFN-<math display="inline"><semantics> <mi mathvariant="sans-serif">γ</mi> </semantics></math>, MIP, MCP-1, and NOS. The interactions between them are bidirectional. Abbreviations: tumor necrosis factor α (TNF-<math display="inline"><semantics> <mi mathvariant="sans-serif">α</mi> </semantics></math>), interferon gamma (IFN-<math display="inline"><semantics> <mrow> <mi mathvariant="sans-serif">γ</mi> <mo stretchy="false">)</mo> </mrow> </semantics></math>, macrophage inflammatory proteins (MIP), monocyte chemotactic protein 1 (MCP-1), nitric oxide synthases (NOS), chimeric antigen receptor (CAR), cytokine release syndrome (CRS), immune-effector-cell-associated neurotoxicity syndrome (ICANS), and blood–brain barrier (BBB) [<a href="#B143-cancers-16-03339" class="html-bibr">143</a>,<a href="#B144-cancers-16-03339" class="html-bibr">144</a>,<a href="#B145-cancers-16-03339" class="html-bibr">145</a>].</p>
Full article ">
13 pages, 2486 KiB  
Article
Immune Profile in COVID-19: Unveiling TR3-56 Cells in SARS-CoV-2 Infection
by Flavia Carriero, Valentina Rubino, Monica Gelzo, Giulia Scalia, Maddalena Raia, Massimo Ciccozzi, Ivan Gentile, Biagio Pinchera, Giuseppe Castaldo, Giuseppina Ruggiero and Giuseppe Terrazzano
Int. J. Mol. Sci. 2024, 25(19), 10465; https://doi.org/10.3390/ijms251910465 - 28 Sep 2024
Viewed by 404
Abstract
The emergence of COronaVIrus Disease 2019 (COVID-19), caused by severe acute respiratory syndrome coronavirus 2 (SARS-CoV-2), presented a global health challenge since its identification in December 2019. With clinical manifestations ranging from mild respiratory symptoms to severe multi-organ dysfunction, COVID-19 continues to affect [...] Read more.
The emergence of COronaVIrus Disease 2019 (COVID-19), caused by severe acute respiratory syndrome coronavirus 2 (SARS-CoV-2), presented a global health challenge since its identification in December 2019. With clinical manifestations ranging from mild respiratory symptoms to severe multi-organ dysfunction, COVID-19 continues to affect populations worldwide. The complex interactions between SARS-CoV-2 variants and the human immune system are crucial for developing effective therapies, vaccines, and preventive measures. Understanding these immune responses highlights the intricate nature of COVID-19 pathogenesis. This retrospective study analyzed, by flow cytometry approach, a cohort of patients infected with SARS-CoV-2 during the initial pandemic waves from 2020 to 2021. It focused on untreated individuals at the time of hospital admission and examined the presence of TR3-56 cells in their immune profiles during the anti-viral immune response. Our findings provide additional insights into the complex immunological dynamics of SARS-CoV-2 infection and highlight the potential role of TR3-56 cells as crucial components of the immune response. We suggest that TR3-56 cells could serve as valuable biomarkers for identifying more severe cases of COVID-19, aiding in the assessment and management of the disease. Full article
(This article belongs to the Special Issue COVID-19: Advances in Pathophysiology and Therapeutics)
Show Figures

Figure 1

Figure 1
<p>Analysis of white blood cells (WBCs) in groups of COronaVIrus Disease 2019 (COVID-19) patients based on increasing severity. (<b>a</b>) The percentage of whole lymphocytes, (<b>b</b>) monocytes, and (<b>c</b>) neutrophils in Groups 1 (white columns), 2 (grey columns), and 3 (black columns) of patients. Standard error (SE) bars are reported at the top of the columns. Statistical analysis (<span class="html-italic">Mann–Whitney test</span>) is reported: <span class="html-italic">p</span> ≤ 0.05 (*); <span class="html-italic">p</span> ≤ 0.005 (**); not significant (NS).</p>
Full article ">Figure 2
<p>Analysis of lymphocyte subtypes in groups of COVID-19 patients based on increasing severity. (<b>a</b>) The percentage of T, (<b>b</b>) cytotoxic T cells (CTLs), (<b>c</b>) T helper (Th), (<b>d</b>) B, and (<b>e</b>) Natural Killer (NK) lymphocytes in Groups 1 (white columns), 2 (grey columns), and 3 (black columns) of patients. Standard error (SE) bars are reported at the top of the columns. Statistical analysis (<span class="html-italic">Mann–Whitney test</span>) is reported: <span class="html-italic">p</span> ≤ 0.05 (*); <span class="html-italic">p</span> ≤ 0.005 (**); <span class="html-italic">p</span> ≤0.0005 (***); <span class="html-italic">p</span> &lt; 0.0001 (****); and not significant (NS).</p>
Full article ">Figure 3
<p>Analysis of activated T, Th1 and Th17 cells in groups of COVID-19 patients based on increasing severity. (<b>a</b>) The percentage of activated T, (<b>b</b>) Th1, and (<b>c</b>) Th17 lymphocytes in Groups 1 (white columns), 2 (grey columns), and 3 (black columns) of patients. Standard error (SE) bars are reported at the top of the columns. Statistical analysis (<span class="html-italic">Mann–Whitney test</span>) is reported: <span class="html-italic">p</span> ≤ 0.005 (**); <span class="html-italic">p</span> &lt; 0.0001 (****); and not significant (NS).</p>
Full article ">Figure 4
<p>Analysis of T regulatory (Treg) and T<sub>R3-56</sub> cells in groups of COVID-19 patients based on increasing severity. (<b>a</b>) The percentage of Treg and (<b>b</b>) T<sub>R3-56</sub> lymphocytes in Groups 1 (white columns), 2 (grey columns), and 3 (black columns) of patients. Standard error (SE) bars are reported at the top of the columns. Statistical analysis (<span class="html-italic">Mann–Whitney test</span>) is reported: <span class="html-italic">p</span> ≤ 0.05 (*); <span class="html-italic">p</span> ≤ 0.005 (**); <span class="html-italic">p</span> &lt; 0.0001 (****) and not significant (NS).</p>
Full article ">
10 pages, 1440 KiB  
Article
Photothermal Effect of 970 nm Diode Laser Irradiation on Enterococcus faecalis Biofilms in Single-Rooted Teeth Ex Vivo
by Soraya Tanner, Anna Thibault, Julian Grégoire Leprince and Serge Bouillaguet
Dent. J. 2024, 12(10), 308; https://doi.org/10.3390/dj12100308 - 27 Sep 2024
Viewed by 337
Abstract
Objective: The aim of this study was to evaluate the photothermal effect of a 970 nm diode laser on Enterococcus faecalis biofilms. Methods: 72 extracted human single-rooted teeth were prepared, sterilized, and inoculated with Enterococcus faecalis to establish a two-week-old biofilm. [...] Read more.
Objective: The aim of this study was to evaluate the photothermal effect of a 970 nm diode laser on Enterococcus faecalis biofilms. Methods: 72 extracted human single-rooted teeth were prepared, sterilized, and inoculated with Enterococcus faecalis to establish a two-week-old biofilm. The specimens were divided into six groups (n = 12): Group 1 (G1)—negative control (PBS—no laser), Group 2 (G2)—positive control (1% NaOCl rinse—no laser), Group 3 (G3)—a 970 nm laser in 1.5 W pulse mode, Group 4 (G4)—a 970 nm laser in 2 W pulse mode, Group 5 (G5)—a 970 nm laser in 1.5 W continuous mode, Group 6 (G6)—a 970 nm laser in 2 W continuous mode. Bacterial viability was evaluated using the LIVE/DEAD BacLight kit and analyzed by flow cytometry (FCM). Temperature changes on the root surface during irradiation were analyzed using a K-type thermocouple. Data were statistically analyzed using one-way ANOVA and Tukey’s multiple comparison test (α = 0.05). Results: Bacterial viability was significantly reduced after laser irradiation in continuous mode using 1.5 W (21% of live bacteria) and 2 W (14% of live bacteria). When the pulsed mode was applied, the reduction in bacterial viability was less, with a mean survival of 53% (1.5 PF, whereas 29% of bacteria survived after 2 W irradiation). Conclusions: The 970 nm diode laser at 2 W continuous mode effectively reduced the viability of E. faecalis biofilms in root canals without causing unacceptable temperature rises at the root surface. Full article
(This article belongs to the Special Issue Endodontics and Restorative Sciences: 2nd Edition)
Show Figures

Figure 1

Figure 1
<p>Bacterial biofilms were grown for 2 weeks. In order to eliminate dead bacteria and debris, the bacterial culture was passed through a peristaltic pump and refreshed at regular intervals. Scanning electron microphotography of the bacterial biofilm grown on dentin root canals (mag. ×1200).</p>
Full article ">Figure 2
<p>Illustration of the experimental configuration for heat control in the root canal, showing thermocouple placement in the apical third. The thermocouple was connected to a thermometer that displayed the temperature directly on the screen. The temperature was recorded every 5 s, until returning to 37 °C.</p>
Full article ">Figure 3
<p>Bacterial survival after laser application. Letters represent statistical differences between groups (ANOVA, Tukey, <span class="html-italic">p</span> = 0.05). CW indicates continuous wave and PF indicates the pulse frequency in pulse mode.</p>
Full article ">Figure 4
<p>A morphological observation of specimens exposed to 2 W in continuous (<b>a</b>) or pulsed mode (<b>b</b>). A slight carbonization was observed inside the root canals exposed to 2 W continuous mode (<b>a</b>).</p>
Full article ">
13 pages, 6038 KiB  
Article
The VEGFA-Induced MAPK-AKT/PTEN/TGFβ Signal Pathway Enhances Progression and MDR in Gastric Cancer
by Hongming Fang, Yujuan Zhou, Xue Bai, Wanlin Che, Wenxuan Zhang, Danying Zhang, Qingmei Chen, Wei Duan, Guochao Nie and Yingchun Hou
Genes 2024, 15(10), 1266; https://doi.org/10.3390/genes15101266 - 27 Sep 2024
Viewed by 220
Abstract
Background/Objectives: Gastric cancer (GC) is a globally frequent cancer, in particular leading in mortality caused by digestive tract cancers in China. Vascular endothelial growth factor A (VEGFA) is excessively expressed in cancers including GC; its involvement in GC development, particularly in multidrug resistance [...] Read more.
Background/Objectives: Gastric cancer (GC) is a globally frequent cancer, in particular leading in mortality caused by digestive tract cancers in China. Vascular endothelial growth factor A (VEGFA) is excessively expressed in cancers including GC; its involvement in GC development, particularly in multidrug resistance (MDR), and the signal route it affects in GC remain unknown. To explore the roles VEGFA plays during progression and MDR formation in GC, we studied its function in a VEGFA-deleted GC cell platform. Methods: We initially assessed the importance of VEGFA in GC and MDR using database analysis. Then, using CCK8, wound healing, transwell, scanning electron microscopy, immunofluorescence, flow cytometry, and other techniques, the alterations in tumor malignancy-connected cell behaviors and microstructures were photographed and evaluated in a VEGFA-gene-deleted GC cell line (VEGFA−/−SGC7901). Finally, the mechanism of VEGFA in GC progression and MDR was examined by Western blot. Results: Database analysis revealed a strong correlation between high VEGFA expression and a poor prognosis for GC. The results showed that VEGFA deletion reduced GC cell proliferation and motility and altered microstructures important for motility, such as the depolymerized cytoskeleton. VEGFA deletion inhibited the growth of pseudopodia/filopodia and suppressed the epithelial–mesenchymal transition (EMT). The occurrence of MDR is induced by overactivation of the MAPK-AKT and TGFβ signaling pathways, while PTEN inhibits these pathways. Conclusions: All findings suggested that VEGFA acts as a cancer enhancer and MDR inducer in GC via the MAPK-AKT/PTEN/TGFβ signal pathway. Full article
(This article belongs to the Section Human Genomics and Genetic Diseases)
Show Figures

Figure 1

Figure 1
<p>Different databases were evaluated to determine the effect of VEGFA on GC progression. (<b>A</b>,<b>B</b>) VEGFA expression in various cancers (<b>A</b>) and in GC (<b>B</b>). (<b>C</b>) The HPA database immunohistochemical samples (pictured above is normal tissue; below is GC tissue). (<b>D</b>) The relationship between VEGFA mutation and prognosis in GC patients. (<b>E</b>,<b>F</b>) VEGFA in different differentiation degree and pathological staging in GC. (<b>G</b>) Survival curve analysis of VEGFA in GC patients (TCGA). (<b>H</b>) VEGF family expression in various cancers. (<b>I</b>) The expression of VEGFA in SGC7901 cells was detected by RT-PCR. * <span class="html-italic">p</span> &lt; 0.05, ** <span class="html-italic">p</span> &lt; 0.01, *** <span class="html-italic">p</span> &lt; 0.001, **** <span class="html-italic">p</span> &lt;0.0001.</p>
Full article ">Figure 2
<p>VEGFA-deleted SGC7901 cell line was generated with CRISPR/Cas9n. (<b>A</b>) The designation for VEGFA deletion. (<b>B</b>) The efficiency (&gt;80%) of the transfection of CRISPR/Cas9n vector at 60 h post-transfection (4×). (<b>C</b>,<b>D</b>) Sequencing assay for positive clone selection. (<b>E</b>–<b>G</b>) The validation of the VEGFA-deleted clone by Western blot ((<b>E</b>,<b>F</b>), **** <span class="html-italic">p</span> &lt; 0.0001) and immunocytochemistry assay ((<b>G</b>), 20×).</p>
Full article ">Figure 3
<p>VEGFA deletion reduced the multiplication and viability of SGC7901 cells. (<b>A</b>) Cell morphology of each group under light microscope (10×). (<b>B</b>,<b>C</b>) Results of CCK8 (<b>B</b>) and colony formation (<b>C</b>). (<b>G</b>,<b>H</b>) Digital assays for colony formatting efficiency (<b>G</b>) and clone counting (<b>H</b>). (<b>D</b>,<b>I</b>) Cell cycle analysis by FCM (<b>D</b>) and data assay (<b>I</b>). (<b>E</b>,<b>J</b>) Results of transwell. (<b>F</b>,<b>K</b>) Results of wound healing. (** <span class="html-italic">p</span> &lt; 0.01, *** <span class="html-italic">p</span> &lt; 0.001, **** <span class="html-italic">p</span> &lt; 0.0001).</p>
Full article ">Figure 4
<p>VEGFA deletion caused cyto/nucleoskeletal remodeling. (<b>A</b>,<b>B</b>) Results of Coomassie Brilliant Blue staining (40×) under optical microscope (<b>A</b>) and SEM (<b>B</b>) (scale bar: 20 μm). (<b>C</b>–<b>E</b>) The cyto/nucleoskeletal microstructures displayed by immunocytochemistry (<b>C</b>) and immunofluorescence assays ((<b>D</b>), 40×, scale bar: 20 μm), and data assay for F-actin ((<b>E</b>), ** <span class="html-italic">p</span> &lt; 0.01, **** <span class="html-italic">p</span> &lt;0.0001).</p>
Full article ">Figure 5
<p>The tendency of apoptosis in SGC7901 cells induced by VEGFA deletion. (<b>A</b>) MitoScene 633 staining and (<b>B</b>) DAPI staining (scale bar: 10 μm). (<b>C</b>–<b>E</b>) Apoptosis detection by annexin V FITC antibody and PI under a fluorescence microscope ((<b>C</b>), scale bar: 10 μm). (*** <span class="html-italic">p</span> &lt; 0.001, **** <span class="html-italic">p</span> &lt; 0.0001).</p>
Full article ">Figure 6
<p>The mechanism VEGFA utilizes to enhance oncogenesis, progression, and MDR in GC. (<b>A</b>) Simulated spatial configuration of VEGFA (SWISS-MODEL database). (<b>B</b>) Prediction of interactions between VEGFA and other genes (STRING database). (<b>C</b>–<b>H</b>) Detection of the expression of the key signal genes (<b>C</b>,D), EMT-relevant genes (<b>E</b>,<b>F</b>), and MDR-relevant genes (<b>G</b>,<b>H</b>) in SGC7901 cells. (* <span class="html-italic">p</span> &lt; 0.05, ** <span class="html-italic">p</span> &lt; 0.01, *** <span class="html-italic">p</span> &lt; 0.001, **** <span class="html-italic">p</span> &lt; 0.0001).</p>
Full article ">Figure 7
<p>An illustration of the VEGFA regulatory mechanism in GC.</p>
Full article ">
19 pages, 4491 KiB  
Article
Exploring Liraglutide in Lithium–Pilocarpine-Induced Temporal Lobe Epilepsy Model in Rats: Impact on Inflammation, Mitochondrial Function, and Behavior
by Fatma Merve Antmen, Zeynep Fedaioglu, Dilan Acar, Ahmed Kerem Sayar, Ilayda Esma Yavuz, Ece Ada, Bengisu Karakose, Lale Rzayeva, Sevcan Demircan, Farah Kardouh, Simge Senay, Meltem Kolgazi, Guldal Suyen and Devrim Oz-Arslan
Biomedicines 2024, 12(10), 2205; https://doi.org/10.3390/biomedicines12102205 - 27 Sep 2024
Viewed by 353
Abstract
Background/Objectives: Glucagon-like peptide-1 receptor agonists such as liraglutide are known for their neuroprotective effects in neurodegenerative disorders, but their role in temporal lobe epilepsy (TLE) remains unclear. We aimed to investigate the effects of liraglutide on several biological processes, including inflammation, antioxidant [...] Read more.
Background/Objectives: Glucagon-like peptide-1 receptor agonists such as liraglutide are known for their neuroprotective effects in neurodegenerative disorders, but their role in temporal lobe epilepsy (TLE) remains unclear. We aimed to investigate the effects of liraglutide on several biological processes, including inflammation, antioxidant defense mechanisms, mitochondrial dynamics, and function, as well as cognitive and behavioral changes in the TLE model. Methods: Low-dose, repeated intraperitoneal injections of lithium chloride–pilocarpine hydrochloride were used to induce status epilepticus (SE) in order to develop TLE in rats. Fifty-six male Sprague Dawley rats were subjected and allocated to the groups. The effects of liraglutide on inflammatory markers (NLRP3, Caspase-1, and IL-1β), antioxidant pathways (Nrf-2 and p-Nrf-2), and mitochondrial dynamics proteins (Pink1, Mfn2, and Drp1) were evaluated in hippocampal tissues via a Western blot. Mitochondrial function in peripheral blood mononuclear cells (PBMCs) was examined using flow cytometry. Cognitive-behavioral outcomes were assessed using the open-field, elevated plus maze, and Morris water maze tests. Results: Our results showed that liraglutide modulates NLRP3-mediated inflammation, reduces oxidative stress, and triggers antioxidative pathways through Nrf2 in SE-induced rats. Moreover, liraglutide treatment restored Pink1, Mfn2, and Drp1 levels in SE-induced rats. Liraglutide treatment also altered the mitochondrial function of PBMCs in both healthy and epileptic rats. This suggests that treatment can modulate mitochondrial dynamics and functions in the brain and periphery. Furthermore, in the behavioral aspect, liraglutide reversed the movement-enhancing effect of epilepsy. Conclusions: This research underscores the potential of GLP-1RAs as a possibly promising therapeutic strategy for TLE. Full article
(This article belongs to the Section Endocrinology and Metabolism Research)
Show Figures

Graphical abstract

Graphical abstract
Full article ">Figure 1
<p>Timeline and experimental design (created using BioRender.com).</p>
Full article ">Figure 2
<p>Western blotting of neuroinflammation-related proteins in hippocampal tissues of rats. (<b>a</b>–<b>f</b>) Relative expression of NLRP3, Caspase-1, and IL-1β protein bands were quantified using Image Lab. Actin was used for internal loading control. The plots show “median ± IQR” with each point corresponding to the measurement from a single animal. Comparisons of sample groups with either the control (C) group (* <span class="html-italic">p</span> &lt; 0.05) or the E+SAL group (# <span class="html-italic">p</span> &lt; 0.05) were performed via a Mann–Whitney U test. NLRP3, nucleotide-binding oligomerization domain-like receptor pyrin domain containing 3; IL-1β, interleukin-1β; C, control group (circles); LIR, liraglutide control group (squares); E+SAL, SE-induced group (triangles up); E+LIR, SE-induced and liraglutide-treated group (triangles down).</p>
Full article ">Figure 2 Cont.
<p>Western blotting of neuroinflammation-related proteins in hippocampal tissues of rats. (<b>a</b>–<b>f</b>) Relative expression of NLRP3, Caspase-1, and IL-1β protein bands were quantified using Image Lab. Actin was used for internal loading control. The plots show “median ± IQR” with each point corresponding to the measurement from a single animal. Comparisons of sample groups with either the control (C) group (* <span class="html-italic">p</span> &lt; 0.05) or the E+SAL group (# <span class="html-italic">p</span> &lt; 0.05) were performed via a Mann–Whitney U test. NLRP3, nucleotide-binding oligomerization domain-like receptor pyrin domain containing 3; IL-1β, interleukin-1β; C, control group (circles); LIR, liraglutide control group (squares); E+SAL, SE-induced group (triangles up); E+LIR, SE-induced and liraglutide-treated group (triangles down).</p>
Full article ">Figure 3
<p>Western blotting of antioxidant pathway-related proteins in hippocampal tissues of rats. (<b>a</b>–<b>c</b>) The relative expression of Nrf2 and p-Nrf2 protein bands was quantified using Image Lab. Actin was used as an internal loading control. (<b>d</b>) The ratio of p-Nrf2 to total Nrf2. The plots show “median ± IQR” with each point corresponding to the measurement from a single animal. Comparisons of sample groups with either the control (C) group (* <span class="html-italic">p</span> &lt; 0.05) or the E+SAL group (# <span class="html-italic">p</span> &lt; 0.05) were performed via a Mann–Whitney U test. Nrf2, nuclear factor E2-related factor 2; p-Nrf2, phosphor-Nrf2; C, control group (circles); LIR, liraglutide control group (squares); E+SAL, SE-induced group (triangles up); E+LIR, SE-induced and liraglutide-treated group (triangles down).</p>
Full article ">Figure 4
<p>Variations in (<b>a</b>) MDA and (<b>b</b>) GSH levels in hippocampal tissues of rats. The plots show “median ± IQR” with each point corresponding to the measurement from a single animal. Comparisons of sample groups with either the control (C) group (* <span class="html-italic">p</span> &lt; 0.05 and ** <span class="html-italic">p</span> &lt; 0.001) or the E+SAL group (## <span class="html-italic">p</span> &lt; 0.001) were performed via a Mann–Whitney U test. MDA, malondialdehyde; GSH, glutathione; C, control group (circles); LIR, liraglutide control group (squares); E+SAL, SE-induced group (triangles up); E+LIR, SE-induced and liraglutide-treated group (triangles down).</p>
Full article ">Figure 5
<p>Western blotting mitochondrial dynamics-related proteins in hippocampal tissues of rats. (<b>a</b>–<b>d</b>) Relative expression of Pink1, Mfn2, and Drp1 protein bands was quantified using Image Lab. Actin was used as an internal loading control. The plots show “median ± IQR” with each point corresponding to the measurement from a single animal. Comparisons of sample groups with either the control (C) group (* <span class="html-italic">p</span> &lt; 0.05) or the E+SAL group (# <span class="html-italic">p</span> &lt; 0.05) were performed via a Mann–Whitney U test. Pink1, PTEN-induced kinase 1; Mfn2, mitofusin 2; Drp1, dynamin-related protein-1; C, control group (circles); LIR, liraglutide control group (squares); E+SAL, SE-induced group (triangles up); E+LIR, SE-induced and liraglutide-treated group (triangles down).</p>
Full article ">Figure 5 Cont.
<p>Western blotting mitochondrial dynamics-related proteins in hippocampal tissues of rats. (<b>a</b>–<b>d</b>) Relative expression of Pink1, Mfn2, and Drp1 protein bands was quantified using Image Lab. Actin was used as an internal loading control. The plots show “median ± IQR” with each point corresponding to the measurement from a single animal. Comparisons of sample groups with either the control (C) group (* <span class="html-italic">p</span> &lt; 0.05) or the E+SAL group (# <span class="html-italic">p</span> &lt; 0.05) were performed via a Mann–Whitney U test. Pink1, PTEN-induced kinase 1; Mfn2, mitofusin 2; Drp1, dynamin-related protein-1; C, control group (circles); LIR, liraglutide control group (squares); E+SAL, SE-induced group (triangles up); E+LIR, SE-induced and liraglutide-treated group (triangles down).</p>
Full article ">Figure 6
<p>Measurement of mitochondrial function in PMBCs. Rat PBMCs were isolated using the density-gradient centrifugation method. The fluorescence intensities of MitoTracker Green, MitoTracker Red CMXRos, and MitoSox were measured in (<b>a</b>) monocyte and (<b>b</b>) lymphocyte populations. Furthermore, to estimate normalized MMP and MitoSOX, the MitoTracker Red CMXRos/ MitoTracker Green and MitoSox/MitoTracker Green ratios were calculated, respectively. The plots show “median ± IQR” with each point corresponding to the measurement from a single animal. Comparisons of sample groups with the control (C) group (* <span class="html-italic">p</span> &lt; 0.05, ** <span class="html-italic">p</span> &lt; 0.001, and *** <span class="html-italic">p</span> &lt; 0.0001) were performed using a Mann–Whitney U test. C, control group; LIR, liraglutide control group; E+SAL, SE-induced group; E+LIR, SE-induced and liraglutide-treated group.</p>
Full article ">Figure 7
<p>The effects of liraglutide treatment (300 µg/kg; 300 µg/1 mL/day) on locomotor activity in rats. The plots show “mean ± SD” with each point corresponding to the measurement from a single animal. A post hoc analysis for multiple comparisons was performed using Tukey’s HSD test after one-way ANOVA (*** <span class="html-italic">p</span> &lt; 0.0001, # <span class="html-italic">p</span> &lt; 0.05). C, control group (circles); LIR, liraglutide control group (squares); E+SAL, SE-induced group (triangles up); E+LIR, SE-induced and liraglutide-treated group (triangles down).</p>
Full article ">Figure 8
<p>The effects of liraglutide treatment (300 µg/kg; 300 µg/1 mL/day) on anxiety-related behavior in rats. The percentage of total time spent in (<b>a</b>) closed and (<b>b</b>) open arms of the EPM. The plots show “mean ± SD” with each point corresponding to the measurement from a single animal. A post hoc analysis for multiple comparisons was performed using Tukey’s HSD test after a one-way ANOVA (* <span class="html-italic">p</span> &lt; 0.05 and ** <span class="html-italic">p</span> &lt; 0.001). C, control group (circles); LIR, liraglutide control group (squares); E+SAL, SE-induced group (triangles up); E+LIR, SE-induced and liraglutide-treated group (triangles down).</p>
Full article ">Figure 9
<p>The effects of liraglutide treatment (300 µg/kg; 300 µg/1 mL/day) on learning and memory abilities in rats. (<b>a</b>) The escape latency and (<b>b</b>) time spent in the target quadrant during the MWM test. A repeated-measures ANOVA was used to analyze the escape latency data. The plots show “mean ± SD” with each point corresponding to the measurement from a single animal. A post hoc analysis for multiple comparisons was performed using Tukey’s HSD test (* <span class="html-italic">p</span> &lt; 0.05, **** <span class="html-italic">p</span> &lt; 0.00001 compared with the C group; ## <span class="html-italic">p</span> &lt; 0.001 compared with the E+SAL group; &amp;&amp; <span class="html-italic">p</span> &lt; 0.001, &amp;&amp;&amp; <span class="html-italic">p</span> &lt; 0.0001, &amp;&amp;&amp;&amp; <span class="html-italic">p</span> &lt; 0.00001 compared with day 1 in their group). C, control group (circles); LIR, liraglutide control group (squares); E+SAL, SE-induced group (triangles up); E+LIR, SE-induced and liraglutide-treated group (triangles down).</p>
Full article ">
19 pages, 3196 KiB  
Article
The Effect of Different Surfactants and Polyelectrolytes on Nano-Vesiculation of Artificial and Cellular Membranes
by Urška Zagorc, Darja Božič, Vesna Arrigler, Žiga Medoš, Matej Hočevar, Anna Romolo, Veronika Kralj-Iglič and Ksenija Kogej
Molecules 2024, 29(19), 4590; https://doi.org/10.3390/molecules29194590 - 27 Sep 2024
Viewed by 298
Abstract
Nano- and micro-sized vesicular and colloidal structures mediate cell–cell communication. They are important players in the physiology of plants, animals, and humans, and are a subject of increasing interest. We investigated the effect of three surfactants, N-cetylpyridinium chloride (CPC), sodium dodecyl sulfate (SDS), [...] Read more.
Nano- and micro-sized vesicular and colloidal structures mediate cell–cell communication. They are important players in the physiology of plants, animals, and humans, and are a subject of increasing interest. We investigated the effect of three surfactants, N-cetylpyridinium chloride (CPC), sodium dodecyl sulfate (SDS), and Triton X-100 (TX100), and two anionic polyelectrolytes, sodium polystyrene sulfonate (NaPSS) and sodium polymethacrylate (NaPMA), on nanoliposomes. In addition, the effect of SDS and TX100 on selected biological membranes (erythrocytes and microalgae) was investigated. The liposomes were produced by extrusion and evaluated by microcalorimetry and light scattering, based on the total intensity of the scattered light (Itot), hydrodynamic radius (Rh), radius of gyration (Rg), shape parameter p (=Rh/Rg,0), and polydispersity index. The EPs shed from erythrocytes and microalgae Dunaliella tertiolecta and Phaeodactylum tricornutum were visualized by scanning electron microscopy (SEM) and analyzed by flow cytometry (FCM). The Rh and Itot values in POPC liposome suspensions with added CPC, SDS, and TX100 were roughly constant up to the respective critical micelle concentrations (CMCs) of the surfactants. At higher compound concentrations, Itot dropped towards zero, whereas Rh increased to values higher than in pure POPC suspensions (Rh ≈ 60–70 nm), indicating the disintegration of liposomes and formation of larger particles, i.e., various POPC–S aggregates. Nanoliposomes were stable upon the addition of NaPSS and NaPMA, as indicated by the constant Rh and Itot values. The interaction of CPC, SDS, or TX100 with liposomes was exothermic, while there were no measurable heat effects with NaPSS or NaPMA. The SDS and TX100 increased the number density of EPs several-fold in erythrocyte suspensions and up to 30-fold in the conditioned media of Dunaliella tertiolecta at the expense of the number density of cells, which decreased to less than 5% in erythrocytes and several-fold in Dunaliella tertiolecta. The SDS and TX100 did not affect the number density of the microalgae Phaeodactylum tricornutum, while the number density of EPs was lower in the conditioned media than in the control, but increased several-fold in a concentration-dependent manner. Our results indicate that amphiphilic molecules need to be organized in nanosized particles to match the local curvature of the membrane for facilitated uptake. To pursue this hypothesis, other surfactants and biological membranes should be studied in the future for more general conclusions. Full article
(This article belongs to the Section Physical Chemistry)
Show Figures

Graphical abstract

Graphical abstract
Full article ">Figure 1
<p>Scheme of the study. (<b>Left</b>) study of POPC liposome membrane stability; (<b>right</b>) study of cellular membrane vesiculation.</p>
Full article ">Figure 2
<p>The average <span class="html-italic">R</span><sub>h</sub><span class="html-italic">/R</span><sub>h,0</sub> values of the POPC liposome population and the total intensity of light scattered from POPC suspensions at an angle of 90°, presented as the ratio <span class="html-italic">I</span><sub>tot</sub>/<span class="html-italic">I</span><sub>tot,0</sub>, as functions of the added compound concentration; (<b>A</b>,<b>B</b>) SDS; (<b>C</b>,<b>D</b>) CPC; (<b>E</b>,<b>F</b>) TX100, (<b>G</b>) NaPSS; and (<b>H</b>) NaPMA. The <span class="html-italic">R</span><sub>h,0</sub> is the hydrodynamic radius, and <span class="html-italic">I</span><sub>tot,0</sub> is the total intensity of scattered light in pure POPC liposome suspensions. Time-dependence data (obtained after 24 h) are depicted by the dashed red lines in Panels (<b>A</b>–<b>F</b>). The nominal concentration of POPC in suspensions was 132 µM. The black vertical dashed lines in Panels (<b>A</b>–<b>F</b>) indicate the CMC values of surfactants in 5 mM TRIS buffer at 25 °C.</p>
Full article ">Figure 3
<p>Distribution of the hydrodynamic radii <span class="html-italic">R</span><sub>h</sub> of POPC liposomes in 5 mM TRIS buffer suspensions (pH = 8) at different temperatures (indicated in the plots); (<b>A</b>) pure POPC suspensions; (<b>B</b>) POPC suspensions with added SDS; (<b>C</b>) POPC suspensions with added CPC; (<b>D</b>) POPC suspensions with added TX100; (<b>E</b>) POPC suspensions with added NaPSS; and (<b>F</b>) POPC suspensions with added NaPMA. The nominal POPC concentration was <span class="html-italic">c</span><sub>POPC</sub> = 132 µM and the nominal compound:POPC molar ratio was 1:1. Temperature legend is the same for all Figures.</p>
Full article ">Figure 4
<p>Enthalpograms obtained by titrating (<b>A</b>) SDS; (<b>B</b>) CPC; (<b>C</b>) TX100; (<b>D</b>) NaPSS; and (<b>E</b>) NaPMA solution (empty circles) into 0.66 mM POPC suspension (filled circles) in 5 mM TRIS buffer. The vertical dashed lines indicate the surfactant’s CMC and the red arrows indicate the concentration of POPC (<span class="html-italic">c</span><sub>POPC</sub>). All data were collected on the VP–ITC calorimeter at 15 °C.</p>
Full article ">Figure 5
<p>(<b>A</b>) EVs isolated from aged erythrocyte suspension stimulated by 50 mM SDS; (<b>B</b>) EVs isolated from aged erythrocyte suspension stimulated by 50 mM SDS and incubated at 80 °C; (<b>C</b>) EVs in the <span class="html-italic">Phaeodactylum tricornutum</span> culture; (<b>D</b>) <span class="html-italic">Dunaliella tertiolecta</span> culture stimulated by TX100. Insets in (<b>A</b>–<b>C</b>) show histograms of particle sizes.</p>
Full article ">Figure 6
<p>The FCM analysis of the effect of surfactants SDS and TX100 on (<b>A</b>) erythrocyte suspension; (<b>B</b>) culture of microalgae <span class="html-italic">Phaeodactylum tricornutum</span>; and (<b>C</b>) culture of microalgae <span class="html-italic">Dunaliella tertiolecta</span>. The number density of cells is shown by gray bars and the number density of EPs by white bars.</p>
Full article ">
Back to TopTop