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

Article Types

Countries / Regions

remove_circle_outline
remove_circle_outline
remove_circle_outline
remove_circle_outline
remove_circle_outline
remove_circle_outline
remove_circle_outline
remove_circle_outline
remove_circle_outline

Search Results (6,639)

Search Parameters:
Keywords = binding selectivity

Order results
Result details
Results per page
Select all
Export citation of selected articles as:
20 pages, 39430 KiB  
Article
Scutellarin Alleviates Neuronal Apoptosis in Ischemic Stroke via Activation of the PI3K/AKT Signaling Pathway
by Zhaoda Duan, Yingqi Peng, Dongyao Xu, Yujia Yang, Yuke Wu, Chunyun Wu, Shan Yan and Li Yang
Int. J. Mol. Sci. 2025, 26(5), 2175; https://doi.org/10.3390/ijms26052175 - 28 Feb 2025
Abstract
Among all stroke types, ischemic stroke (IS) occurs most frequently, resulting in neuronal death and tissue injury within both the central infarct region and surrounding areas. This study explored the neuroprotective mechanisms of scutellarin, a flavonoid compound, through an integrated strategy that merged [...] Read more.
Among all stroke types, ischemic stroke (IS) occurs most frequently, resulting in neuronal death and tissue injury within both the central infarct region and surrounding areas. This study explored the neuroprotective mechanisms of scutellarin, a flavonoid compound, through an integrated strategy that merged in silico analyses (including network pharmacology and molecular docking simulations) with both in vitro and in vivo experimental verification. We identified 1887 IS-related targets and 129 scutellarin targets, with 23 overlapping targets. PPI network analysis revealed five core targets, and molecular docking demonstrated strong binding affinities between scutellarin and these targets. Bioinformatic analyses, including GO functional annotation and KEGG pathway mapping, indicated that the PI3K/AKT cascade represents the primary signaling mechanism. An in vitro experimental system was developed using PC12 cells under oxygen-glucose deprivation conditions to investigate how scutellarin regulates neuronal cell death via the PI3K/AKT pathway. Western blot quantification demonstrated that treatment with scutellarin enhanced the expression of p-PI3K, p-AKT, and Bcl-2 proteins, while simultaneously reducing levels of apoptotic markers Bax and cleaved caspase-3. Furthermore, pharmacological intervention with the selective PI3K inhibitor LY294002 attenuated these molecular alterations, resulting in diminished expression of p-PI3K, p-AKT, and Bcl-2, accompanied by elevated levels of Bax and cleaved caspase-3. In a rat model of middle cerebral artery occlusion, scutellarin administration demonstrated comparable neuroprotective effects, maintaining neuronal survival and modulating apoptotic protein expression via PI3K/AKT pathway activation. Collectively, this study demonstrates the therapeutic potential of scutellarin in cerebral ischemia through PI3K/AKT pathway modulation, suggesting its possible application in treating ischemic disorders. Full article
(This article belongs to the Section Molecular Neurobiology)
Show Figures

Figure 1

Figure 1
<p>Differential expression analysis between MCAO and sham samples: (<b>A</b>) Distribution of gene expression in the samples. (<b>B</b>) Distance heatmap showing clustering relationships among the samples. (<b>C</b>,<b>D</b>) Volcano plot and heatmap illustrating differentially expressed genes.</p>
Full article ">Figure 2
<p>Screening of intersection gene targets: (<b>A</b>) Target intersection analysis between scutellarin and cerebral ischemia using Venn diagram. (<b>B</b>) STRING-based protein–protein interaction network construction. (<b>C</b>) Topological visualization of interaction networks. (<b>D</b>) Hierarchical representation of five key molecular targets based on degree centrality.</p>
Full article ">Figure 3
<p>Functional annotation analysis: (<b>A</b>) Enriched biological processes visualization. (<b>B</b>) Cellular component distribution analysis. (<b>C</b>) Molecular function enrichment patterns. (<b>D</b>) KEGG pathway enrichment landscape.</p>
Full article ">Figure 4
<p>Molecular docking visualization of scutellarin with core targets: (<b>A</b>) Visualization of scutellarin docking with the CASP1 molecule. (<b>B</b>) Visualization of scutellarin docking with the CASP3 molecule. (<b>C</b>) Visualization of scutellarin docking with the CCL2 molecule. (<b>D</b>) Visualization of scutellarin docking with the EGFR molecule. (<b>E</b>) Visualization of scutellarin docking with the PTGS2 molecule.</p>
Full article ">Figure 5
<p>Scutellarin alleviates neuronal apoptosis in the ischemic cortex of MCAO rats at three days: Bar = 50 μm, <span class="html-italic">n</span> = 3. (<b>A</b>) TUNEL immunofluorescence images; (<b>B</b>) Quantitative analysis. **** <span class="html-italic">p</span> &lt; 0.001 vs. Sham group; ## <span class="html-italic">p</span> &lt; 0.01 vs. MCAO group.</p>
Full article ">Figure 6
<p>Effects of scutellarin on Nissl staining in the ischemic cortex of MCAO rats (×100).</p>
Full article ">Figure 7
<p>PI3K and AKT phosphorylation analysis in MCAO rat ischemic cortex following scutellarin treatment at day 3: (<b>A</b>,<b>C</b>) Representative Western blots and immunofluorescence micrographs; (<b>B</b>,<b>D</b>) Statistical analyses. Green fluorescence indicates NeuN-positive neurons, blue represents DAPI-labeled nuclei, and red shows Cy3-tagged proteins. Statistical significance: ** <span class="html-italic">p</span> &lt; 0.01 vs. Sham; # <span class="html-italic">p</span> &lt; 0.05, ## <span class="html-italic">p</span> &lt; 0.01 vs. MCAO; Scale bar: 50 μm, <span class="html-italic">n</span> = 5.</p>
Full article ">Figure 8
<p>Apoptotic protein expression analysis in MCAO rat ischemic cortex after 3-day scutellarin treatment: (<b>A</b>,<b>C</b>) Western blot bands and immunofluorescence microscopy; (<b>B</b>,<b>D</b>) Quantitative data analysis. Neurons (NeuN, green), nuclei (DAPI, blue), target proteins (Cy3, red). Statistical significance: * <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. Sham; # <span class="html-italic">p</span> &lt; 0.05, ## <span class="html-italic">p</span> &lt; 0.01 vs. MCAO; Scale = 50 μm, <span class="html-italic">n</span> = 5.</p>
Full article ">Figure 9
<p>Effect of scutellarin on the mRNA expression of pathway-enriched genes in PC12 cells. Statistical significance: * <span class="html-italic">p</span> &lt; 0.05, ** <span class="html-italic">p</span> &lt; 0.01 vs. Control group; # <span class="html-italic">p</span> &lt; 0.05, ## <span class="html-italic">p</span> &lt; 0.01 vs. OGD group; <span class="html-italic">n</span> = 5.</p>
Full article ">Figure 10
<p>Impact of PI3K inhibition on scutellarin-mediated protein regulation in PC12 cells: (<b>A</b>,<b>C</b>) Western blot bands; (<b>B</b>,<b>D</b>) Quantitative data analysis. Statistical significance: * <span class="html-italic">p</span> &lt; 0.05, ** <span class="html-italic">p</span> &lt; 0.01 vs. Control; # <span class="html-italic">p</span> &lt; 0.05, ## <span class="html-italic">p</span> &lt; 0.01 vs. OGD; <sup>▲</sup> <span class="html-italic">p</span> &lt; 0.05, <sup>▲▲</sup> <span class="html-italic">p</span> &lt; 0.01 vs. OGD+S; <span class="html-italic">n</span> = 5.</p>
Full article ">
26 pages, 4067 KiB  
Article
Ru(II) Complexes with 3,4-Dimethylphenylhydrazine: Exploring In Vitro Anticancer Activity and Protein Affinities
by Jasmina Dimitrić Marković, Dušan Dimić, Thomas Eichhorn, Dejan Milenković, Aleksandra Pavićević, Dragoslava Đikić, Emilija Živković, Vladan Čokić, Tobias Rüffer and Goran N. Kaluđerović
Biomolecules 2025, 15(3), 350; https://doi.org/10.3390/biom15030350 - 28 Feb 2025
Abstract
Two new Ru(II) complexes, mononuclear [RuCl26-p-cymene)(3,4-dmph-κN)] (1) and the binuclear complex [{RuCl(η6-p-cymene)}2(μ-Cl)(μ-3,4-dmph-κ2N,N′)]Cl (2; 3,4-dmph = 3,4-dimethylphenylhydrazine), are synthesized and experimentally [...] Read more.
Two new Ru(II) complexes, mononuclear [RuCl26-p-cymene)(3,4-dmph-κN)] (1) and the binuclear complex [{RuCl(η6-p-cymene)}2(μ-Cl)(μ-3,4-dmph-κ2N,N′)]Cl (2; 3,4-dmph = 3,4-dimethylphenylhydrazine), are synthesized and experimentally and theoretically structurally characterized utilizing 1H and 13C NMR and FTIR spectroscopy, as well as DFT calculations. Degradation product of 2, thus ([{RuCl(η6-p-cymene)}2(μ-Cl)(μ-3,4-dmph-κ2N,N′)][RuCl36-p-cymene)] (2b) was characterized with SC-XRD. In the crystals of 2b, the cationic and anionic parts interact through N-H...Cl hydrogen bridges. The spectrofluorimetric measurements proved the spontaneity of the binding processes of both complexes and HSA. Spin probing EPR measurements implied that 1 and 2 decreased the amount of bound 16-doxylstearate and implicated their potential to bind to HSA more strongly than the spin probe. The cytotoxicity assessment of both complexes against the MDA-MB-231 and MIA PaCa-2 cancer cell lines demonstrated a clear dose-dependent decrease in cell viability and no effect on healthy HS-5 cells. Determination of the malondialdehyde and protein carbonyl concentrations indicated that new complexes could offer protective antioxidant benefits in specific cancer contexts. Gel electrophoresis measurements showed the reduction in MMP9 activity and indicated the potential of 1 in limiting the cancer cells’ invasion. The annexin V/PI apoptotic assay results showed that 1 and 2 exhibit different selectivity towards MIA PaCa-2 and MDA-MB-231 cancer cells. A comparative molecular docking analysis of protein binding, specifically targeting acetylcholinesterase (ACHE), matrix metalloproteinase-9 (MMP-9), and human serum albumin (HSA), demonstrated distinct binding interactions for each complex. Full article
Show Figures

Figure 1

Figure 1
<p>ORTEP (30% probability ellipsoids) of the hydrogen bridged ion pair of [{RuCl(η<sup>6</sup>-<span class="html-italic">p</span>-cymene)}<sub>2</sub>(μ-Cl)(μ-3,4-dmph-κ<sup>2</sup><span class="html-italic">N</span>,<span class="html-italic">N</span>′)]<sup>+</sup> (left) and of [RuCl<sub>3</sub>(<span class="html-italic">p</span>-cymene)]<sup>−</sup> (right). All carbon-bonded hydrogens and disordered atomic positions are omitted for clarity. Dashed lines represent aryl Ruthenium interactions as well as hydrogen bonds.</p>
Full article ">Figure 2
<p>Optimized structures of complexes <b>1</b> (<b>a</b>) and (<b>2</b>–Cl)<b><sup>+</sup></b> (<b>b</b>) at B3LYP/6-311++G(d,p)(H,C,N,Cl)/def2-TZVP level of theory (The hydrogen atoms are omitted for clarity, carbon-gray, nitrogen-blue, chlorine-gray, ruthenium-teal).</p>
Full article ">Figure 3
<p>The fluorescence emission spectra of HSA for the titration with various concentrations of complex <b>1</b> at (<b>a</b>) 27°, (<b>b</b>) 32°, and (<b>c</b>) 37 °C, and (<b>d</b>) the van ’t Hoff plot for the binding process.</p>
Full article ">Figure 4
<p>EPR spectra of 250 µM 16-DS: (<b>a</b>) in the phosphate buffer (100 mM, pH 7.4), and (<b>b</b>) incubated with HSA at [HSA]:[16-DS] molar ratio of 1:5. The spectra are not presented on the same intensity scale. The markers show the amplitudes of low-field (I<sub>lf</sub>) and high-field peak (I<sub>hf</sub>) that correspond to the bound and unbound 16-DS, respectively. (<b>c</b>) The amplitudes ratio of the high-field (I<sub>hf</sub>) and low-field (I<sub>lf</sub>) peaks calculated for the different molar ratios of [HSA]:[Ru-complexes]. All samples were prepared in the phosphate buffer (100 mM, pH 7.4) and contained 50 µM HSA, 250 µM 16-DS. The reference sample HSA+16-DS+DMSO contained 1.2% <span class="html-italic">v</span>/<span class="html-italic">v</span> DMSO, the same amount as the sample at the [HSA]:[Ru-complex] molar ratio 1:4.</p>
Full article ">Figure 5
<p>Viability of (<b>a</b>) HS-5, (<b>b</b>) MDA-MB-231, and (<b>c</b>) MIA PaCa-2 cancer cells under the effect of <b>1</b> and <b>2</b> for 24 h of treatment (MTT assay).</p>
Full article ">Figure 6
<p>The concentration of (<b>a</b>) MDA and (<b>b</b>) PC in the medium of treated MDA-MB-231 and MIA PaCa-2 cells. * <span class="html-italic">p</span> &lt; 0.05 (comparison to <b>2</b>); ** <span class="html-italic">p</span> &lt; 0.01 (comparison to <b>2</b>); <sup>#</sup> <span class="html-italic">p</span> &lt; 0.01 (comparison to control); <sup>##</sup> <span class="html-italic">p</span> &lt; 0.01 (comparison to control).</p>
Full article ">Figure 7
<p>Matrix metalloproteinase 2 and 9 activity in (<b>a</b>) MDA-MB-231 and (<b>b</b>) MIA PaCa-2 cancer cells under the effect of the complexes <b>1</b> and <b>2</b>. * <span class="html-italic">p</span> &lt; 0.05 (comparison to <b>1</b>); ** <span class="html-italic">p</span> &lt; 0.01 (comparison to <b>1</b>).</p>
Full article ">Figure 8
<p>The cell cycle of (<b>a</b>) MDA-MB-231 and (<b>b</b>) MIA PaCa-2 cells under the effect of complexes <b>1</b> and <b>2</b>.</p>
Full article ">Figure 9
<p>The hydrogen bond (green dotted lines) and hydrophobic (rose pink dotted lines) docking interactions of the most stable conformations of <b>1</b> with ACHE, MMP-9, and HSA.</p>
Full article ">Scheme 1
<p>Synthesis of Ru(II) complexes <b>1</b> and <b>2</b>.</p>
Full article ">
11 pages, 1807 KiB  
Communication
Rapid and Ultrasensitive Sensor for Point-of-Use Detection of Perfluorooctanoic Acid Based on Molecular Imprinted Polymer and AC Electrothermal Effect
by Niloufar Amin, Jiangang Chen, Ngoc Susie Nguyen, Qiang He, John Schwartz and Jie Jayne Wu
Micromachines 2025, 16(3), 283; https://doi.org/10.3390/mi16030283 - 28 Feb 2025
Abstract
Perfluorooctanoic acid (PFOA) is one of the most persistent and bioaccumulative water contaminants. Sensitive, rapid, and in-field analysis is needed to ensure safe water supplies. Here, we present a single step (one shot) and rapid sensor capable of measuring PFOA at the sub-quadrillion [...] Read more.
Perfluorooctanoic acid (PFOA) is one of the most persistent and bioaccumulative water contaminants. Sensitive, rapid, and in-field analysis is needed to ensure safe water supplies. Here, we present a single step (one shot) and rapid sensor capable of measuring PFOA at the sub-quadrillion (ppq) level, 4.5 × 10−4 ppq, within 10 s. This innovative sensor employs a synergistic combination of a molecularly imprinted polymer (MIP)-modified gold interdigitated microelectrode chip and AC electrothermal effects (ACETs), which enhance detection sensitivity by facilitating the accelerated movement of PFOA molecules towards specific recognition sites on the sensing surface. The application of a predetermined AC signal induces microfluidic enrichment and results in concentration-dependent changes in interfacial capacitance during the binding process. This enables real-time, rapid quantification with exceptional sensitivity. We achieved a linear dynamic range spanning from 0.4 to 40 fg/L (4 × 10−7–4 × 10−5 ppt) and demonstrated good selectivity (~1:100) against other PFAS compounds, including perfluorooctanoic acid (PFOS), in PBS buffer. The sensor’s straightforward operation, cost-effectiveness, elimination of the need for external redox probes, compact design, and functionality in relatively resistant environmental matrices position it as an outstanding candidate for deployment in practical applications. Full article
(This article belongs to the Special Issue Innovations in Biosensors, Gas Sensors and Supercapacitors)
Show Figures

Figure 1

Figure 1
<p>(<b>a</b>) A schematic of the PFOA-MIP film electropolymerization on the Au-IDME and a sensing mechanism based on the change in the electrical double layer (EDL). The top transparent layer shows the dielectric layer on the electrode’s surface. (<b>b</b>) A photo of a Au-IDME (<b>A</b>); the measurement setup for the ACEK-capacitive sensing (<b>B</b>); a simplified circuit model employed to represent the electrode–electrolyte interface (<b>C</b>). d<sub>EDL</sub>, d<sub>MIP</sub>, and d<sub>PFOA</sub>: the thickness of the EDL, MIP, and PFOA molecules.</p>
Full article ">Figure 2
<p>(<b>a</b>) Cyclic voltammograms during the electrodeposition of MIP layer on Au-IDME in acetate buffer (pH 5.8) with 10 mM o-PD and 1 mM PFOA (pink line), or 10 mM o-PD only (gray one); scan rate 50 mV/s; number of scans 25. Electrical spectrums for Au-IDME surface characterization. Impedance (<b>b</b>) and capacitance (<b>c</b>) spectrums before and after MIP electrodeposition and after template.</p>
Full article ">Figure 3
<p>(<b>a</b>) Normalized capacitances change as a function of time within 30 s for different levels of PFOA. (<b>b</b>) Capacitance change rates in response to different concentrations of PFOA, at 3 kHz and 100 mV.</p>
Full article ">Figure 4
<p>(<b>a</b>). MIP-AuIDME in the presence of the other PFAS compounds with similar head and alkyl groups. (<b>b</b>) The structure of the PFAS compounds studied.</p>
Full article ">
11 pages, 2147 KiB  
Technical Note
GPCRVS-AI-Driven Decision Support System for GPCR Virtual Screening
by Dorota Latek, Khushil Prajapati, Paulina Dragan, Matthew Merski and Przemysław Osial
Int. J. Mol. Sci. 2025, 26(5), 2160; https://doi.org/10.3390/ijms26052160 - 27 Feb 2025
Abstract
G protein-coupled receptors (GPCRs) constitute the largest and most frequently used family of molecular drug targets. The simplicity of GPCR drug design results from their common seven-transmembrane-helix topology and well-understood signaling pathways. GPCRs are extremely sensitive to slight changes in the chemical structure [...] Read more.
G protein-coupled receptors (GPCRs) constitute the largest and most frequently used family of molecular drug targets. The simplicity of GPCR drug design results from their common seven-transmembrane-helix topology and well-understood signaling pathways. GPCRs are extremely sensitive to slight changes in the chemical structure of compounds, which allows for the reliable design of highly selective and specific drugs. Only recently has the number of GPCR structures, both in their active and inactive conformations, together with their active ligands, become sufficient to comprehensively apply machine learning in decision support systems to predict compound activity in drug design. Here, we describe GPCRVS, an efficient machine learning system for the online assessment of the compound activity against several GPCR targets, including peptide- and protein-binding GPCRs, which are the most difficult for virtual screening tasks. As a decision support system, GPCRVS evaluates compounds in terms of their activity range, the pharmacological effect they exert on the receptor, and the binding mode they could demonstrate for different types and subtypes of GPCRs. GPCRVS allows for the evaluation of compounds ranging from small molecules to short peptides provided in common chemical file formats. The results of the activity class assignment and the binding affinity prediction are provided in comparison with predictions for known active ligands of each included GPCR. Multiclass classification in GPCRVS, handling incomplete and fuzzy biological data, was validated on ChEMBL and Google Patents-retrieved data sets for class B GPCRs and chemokine CC and CXC receptors. Full article
(This article belongs to the Special Issue G Protein-Coupled Receptors)
Show Figures

Figure 1

Figure 1
<p>The scheme of the input data, the implemented algorithms, and the output results of GPCRVS. Bottom left—example results for maraviroc showing GPCRVS performance in terms of the receptor subtype selectivity prediction.</p>
Full article ">Figure 2
<p>Examples of the feature importance obtained for the CCR6, CRF1R, and GLP1R orthosteric and allosteric ligand training sets. Here, the 1st and 2nd most important structural features for each training set were presented. The most important structural features (fingerprint bits) were selected using the feature importance gain method in LightGBM and drawn with RDKit. Yellow atoms are aromatic atoms, blue dots represent the centers of the fragment, and other atoms and bonds are marked with grey with the continuation of the fragment marked with asterisks.</p>
Full article ">Figure 3
<p>Example results for two patent compounds for which GPCRVS correctly predicted both the drug target and drug target class. Asterisks - the probability score provided by SwissTargetPrediction mean: ‘probability for the query molecule—assumed as bioactive—to have this protein as a target’.</p>
Full article ">
13 pages, 4594 KiB  
Article
Quantitative Criteria for Solvent Selection in Liquid-Phase Exfoliation: Balancing Exfoliation and Stabilization Efficiency
by Shunnian Wu, W. P. Cathie Lee, Hashan N. Thenuwara and Ping Wu
Nanomaterials 2025, 15(5), 370; https://doi.org/10.3390/nano15050370 (registering DOI) - 27 Feb 2025
Abstract
The selection of solvent is pivotal in liquid-phase exfoliation (LPE), as an ideal solvent minimizes the energy required to disrupt the interlayer attractive interactions while stabilizing exfoliated layers to prevent re-agglomeration. This study theoretically analyzed the LPE of Mg(OH)2 in different solvents, [...] Read more.
The selection of solvent is pivotal in liquid-phase exfoliation (LPE), as an ideal solvent minimizes the energy required to disrupt the interlayer attractive interactions while stabilizing exfoliated layers to prevent re-agglomeration. This study theoretically analyzed the LPE of Mg(OH)2 in different solvents, including water, dimethyl sulfoxide (DMSO), dimethylformamide, N-methyl-2-pyrrolidone (NMP), isopropyl alcohol, and 2-butanone, through first-principles calculations combined with experimental validation. DMSO was identified as the most effective solvent for reducing the interlayer attraction, based on exfoliation energy calculations, while NMP was the most efficient for stabilizing exfoliated layers, based on binding energy assessments. Principal component analysis of the solvents’ physicochemical properties reduced the original dataset of seven variables to two dominant factors. The binding energy correlates with planarity and polarity, whereas the exfoliation energy is governed by dipole moment and polarity. The biaxial straining theory successfully clarified the underlying mechanisms behind the established criteria for selecting the optimal solvent. Experimental results confirmed that DMSO outperforms water in the LPE of Mg(OH)2. These results establish a quantitative framework for solvent selection, enhancing the efficiency of the LPE processes. Full article
(This article belongs to the Special Issue Modeling, Simulation and Optimization of Nanomaterials)
Show Figures

Figure 1

Figure 1
<p>Binding of water (<b>a</b>), MEK (<b>b</b>), DMF (<b>c</b>), DMSO (<b>d</b>), IPA (<b>e</b>), and NMP to Mg(OH)<sub>2</sub> surface (<b>f</b>). Black, dark green, red, blue, yellow, and white balls represent C, Mg, O, N, S, and H atoms, respectively.</p>
Full article ">Figure 2
<p>Intercalation of water (<b>a</b>), MEK (<b>b</b>), DMF (<b>c</b>), DMSO (<b>d</b>), IPA (<b>e</b>), and NMP in Mg(OH)<sub>2</sub> interlayers (<b>f</b>). Black, dark green, red, blue, yellow, and white balls represent C, Mg, O, N, S, and H atoms, respectively.</p>
Full article ">Figure 3
<p>Percent variance explained by principal components.</p>
Full article ">Figure 4
<p>Variation in binding energy (<b>a</b>) and exfoliation energy (<b>b</b>) with the selected physical properties of solvent.</p>
Full article ">Figure 5
<p>Calculated strain indices of the studied solvents.</p>
Full article ">Figure 6
<p>(<b>a</b>) SEM image of the bulk Mg(OH)<sub>2</sub> particles; (<b>b</b>) TEM image of Sample A; (<b>c</b>) TEM image of Sample B; (<b>d</b>) Thickness of the nanosheets of sample B.</p>
Full article ">
22 pages, 5195 KiB  
Article
Therapeutic Mechanisms of Medicine Food Homology Plants in Alzheimer’s Disease: Insights from Network Pharmacology, Machine Learning, and Molecular Docking
by Shuran Wen, Ye Han, You Li and Dongling Zhan
Int. J. Mol. Sci. 2025, 26(5), 2121; https://doi.org/10.3390/ijms26052121 - 27 Feb 2025
Viewed by 75
Abstract
Alzheimer’s disease (AD) is a progressive neurodegenerative disorder characterized by a gradual decline in cognitive function. Currently, there are no effective treatments for this condition. Medicine food homology plants have gained increasing attention as potential natural treatments for AD because of their nutritional [...] Read more.
Alzheimer’s disease (AD) is a progressive neurodegenerative disorder characterized by a gradual decline in cognitive function. Currently, there are no effective treatments for this condition. Medicine food homology plants have gained increasing attention as potential natural treatments for AD because of their nutritional value and therapeutic benefits. In this work, we aimed to provide a deeper understanding of how medicine food homology plants may help alleviate or potentially treat AD by identifying key targets, pathways, and small molecule compounds from 10 medicine food homology plants that play an important role in this process. Using network pharmacology, we identified 623 common targets between AD and the compounds from the selected 10 plants, including crucial proteins such as STAT3, IL6, TNF, and IL1B. Additionally, the small molecules from the selected plants were grouped into four clusters using hierarchical clustering. The ConPlex algorithm was then applied to predict the binding capabilities of these small molecules to the key protein targets. Cluster 3 showed superior predicted binding capabilities to STAT3, TNF, and IL1B, which was further validated by molecular docking. Scaffold analysis of small molecules in Cluster 3 revealed that those with a steroid-like core—comprising three fused six-membered rings and one five-membered ring with a carbon–carbon double bond—exhibited better predicted binding affinities and were potential triple-target inhibitors. Among them, MOL005439, MOL000953, and MOL005438 were identified as the top-performing compounds. This study highlights the potential of medicine food homology plants as a source of active compounds that could be developed into new drugs for AD treatment. However, further pharmacokinetic studies are essential to assess their efficacy and minimize side effects. Full article
(This article belongs to the Special Issue Network Pharmacology: An Emerging Field in Drug Discovery)
Show Figures

Figure 1

Figure 1
<p>Ranking and interactions of the top 30 targets.</p>
Full article ">Figure 2
<p>Results of (<b>A</b>) gene ontology (GO) and (<b>B</b>) Kyoto Encyclopedia of Genes and Genomes (KEGG) pathway enrichment analysis based on the common genes between Alzheimer’s disease (AD) and medicine food homology plants.</p>
Full article ">Figure 3
<p>Hierarchical clustering analysis of small molecules from medicine food homology plants. (<b>A</b>) Silhouette coefficient plot used to determine the optimal number of clusters. A higher silhouette coefficient indicates better-defined and more-cohesive clustering groups. (<b>B</b>) Clustering results of 1142 small molecules based on their structural scaffolds.</p>
Full article ">Figure 4
<p>Predicted binding scores of the four clusters with (<b>A</b>) STAT3, (<b>B</b>) TNF, and (<b>C</b>) IL1B.</p>
Full article ">Figure 5
<p>Heatmap of the average predicted binding score.</p>
Full article ">Figure 6
<p>Molecular docking interactions of STAT3 with (<b>A</b>) MOL009678 and (<b>B</b>) MOL005438.</p>
Full article ">Figure 7
<p>Molecular docking interactions of TNF with (<b>A</b>) MOL000291, (<b>B</b>) MOL002442, (<b>C</b>) MOL002046, and (<b>D</b>) MOL008253.</p>
Full article ">Figure 8
<p>Molecular docking interactions of IL1B with (<b>A</b>) MOL011210, (<b>B</b>) MOL011442, (<b>C</b>) MOL005529, and (<b>D</b>) MOL000232.</p>
Full article ">Figure 9
<p>(<b>A</b>) Venn diagram showing the overlap of small molecules with predicted binding scores above the mean value for IL1B, TNF, and STAT3. (<b>B</b>) Statistical analysis of the scaffolds of the 46 small molecules, with the corresponding number of molecules for each scaffold indicated in brackets. The oxygen atom is represented in red.</p>
Full article ">Figure 10
<p>Molecular docking interactions of IL1B with (<b>A</b>) MOL005439, (<b>B</b>) MOL000953, and (<b>C</b>) MOL005438.</p>
Full article ">Figure 11
<p>Molecular docking interactions of TNF with (<b>A</b>) MOL005439, (<b>B</b>) MOL000953, and (<b>C</b>) MOL005438.</p>
Full article ">Figure 12
<p>Molecular docking interactions of STAT3 with (<b>A</b>) MOL005439 and (<b>B</b>) MOL000953. The docking results of MOL005438 with STAT3 are shown in <a href="#ijms-26-02121-f006" class="html-fig">Figure 6</a>B.</p>
Full article ">
18 pages, 4115 KiB  
Article
Development of an Anti-Zearalenone Nanobody Phage Display Library and Preparation of Specific Nanobodies
by Ying Zeng, Yiying Hu, Ganying Chen, Qingqing Feng, Ruiting Wang, Zhilin Zhang, Jinxian Chen, Junbin Liao, Danrong Lin and Wei Zhu
Curr. Issues Mol. Biol. 2025, 47(3), 157; https://doi.org/10.3390/cimb47030157 - 27 Feb 2025
Viewed by 54
Abstract
Zearalenone (ZEN), a toxic estrogenic mycotoxin in cereals, threatens human and animal health through reproductive, immune, and cytotoxic effects, necessitating sensitive detection methods. While nanobodies offer advantages over conventional antibodies for on-site ZEN detection, their application remains unexplored. This study aimed to develop [...] Read more.
Zearalenone (ZEN), a toxic estrogenic mycotoxin in cereals, threatens human and animal health through reproductive, immune, and cytotoxic effects, necessitating sensitive detection methods. While nanobodies offer advantages over conventional antibodies for on-site ZEN detection, their application remains unexplored. This study aimed to develop an anti-ZEN nanobody derived from an anti-ZEN phage display nanobody library. An alpaca was immunized with a ZEN-bovine serum albumin (ZEN-BSA) antigen, achieving peak serum antibody titers (1:25,600) following four immunizations. A high-capacity phage display nanobody library (1.0 × 1011 plaque-forming units/mL) was constructed. Following four rounds of biopanning, an enrichment factor of 479 was achieved. Phage ELISA screening identified six phage display nanobodies with specific ZEN-binding activity, and multiple sequence alignment revealed four unique nanobody sequences. The selected phage display nanobody, designated phage-V44, was expressed and purified, and its presence was validated by SDS-PAGE and western blotting, which detected a single approximately 17 kDa band consistent with the expected nanobody size. We established a working curve for an indirect competitive enzyme-linked immunoassay (ELISA) for ZEN, which showed an IC50 value of 7.55 ng/mL. The specificity and affinity of the V44 were also verified. Collectively, the study successfully constructed an anti-ZEN phage display nanobody library, screened four specific ZEN-binding phage display nanobodies, and prepared the anti-ZEN nanobody V44. Thereby establishing a foundation for the nanobody’s future integration into rapid on-site detection methods for ZEN in both animal feed and human food products. Full article
(This article belongs to the Section Molecular Microbiology)
Show Figures

Figure 1

Figure 1
<p>Library construction. (<b>A</b>) Total RNA extracted; (<b>B</b>) lanes 1 to 5: 0, 0.5, 1, 2, and 4 µL cDNA templates; (<b>C</b>) lane 6: gel recovery of PCR-1 products; (<b>D</b>) lane 1: VHH-F/VHH-R1 amplification product, and lane 2: VHH-F/VHH-R2 amplification product; (<b>E</b>) lane 3: gel recovery of PCR-2 products; and (<b>F</b>) colony PCR identification of the primary nanobody library.</p>
Full article ">Figure 2
<p>Identification library. (<b>A</b>) Amino acid sequence alignment analysis of the primary nanobody library. DNA sequences were translated into amino acid sequences using ExPASy, and sequence conservation and variability were analyzed using WebLogo to generate frequency distribution profiles. (<b>B</b>) Phylogenetic tree of 20 nanobody sequences. A neighbor-joining phylogenetic tree was constructed using MEGA 11, with branch support assessed by the bootstrap method (1000 replications).</p>
Full article ">Figure 3
<p>Indirect competition phage ELISA screening of phage clones: the (<b>A</b>) second round of panning; (<b>B</b>) third round of panning; and (<b>C</b>) fourth round of panning. Data are presented as means ± S.D. (<span class="html-italic">n</span> = 3). Clones (V8, V22, V44, V55, V59, and V62) with ELISA signal greater than 1.0 in the absence of ZEN and with &gt;40% signal inhibition in the presence of ZEN were selected for further analysis and marked with “*” in the image.</p>
Full article ">Figure 4
<p>Amino acid sequence alignment analysis of anti-ZEN VHH. Significant variability was observed in the CDR3 region, which is a critical determinant of epitope recognition and functional activity within nanobodies. Specifically, the CDR3 region length differed across variants: V8 exhibited 22 amino acids, while V44, V55, and V62 displayed shorter sequences of 15, 12, and 15 amino acids, respectively.</p>
Full article ">Figure 5
<p>Three-dimensional structure simulation and docking analysis of VHH. (<b>A</b>) Homology-modeled 3D structure of V8, V44, V55, and V62. FRs are shown as anti-parallel β-sheets (gray), with CDRs colored as follows: CDR1 (pale green), CDR2 (pale yellow), and CDR3 (light blue). (<b>B</b>) Ramachandran plot validation of model quality, with glycine and proline residues excluded. Regions are defined as the most favored [A, B, L], additional allowed [a, b, l, p], generously allowed [~a, ~b, ~l, ~p], and disallowed. All models met high-quality criteria, with 90% or more of the residues in the most favored regions. (<b>C</b>) ZEN binding mode predicted by semi-flexible docking (Autodock, Lamarckian genetic algorithm/LGA). Optimal conformations (50 cycles) revealed stable ligand–receptor complexes with binding energies that were all below 5 kJ/mol. The green part represents ZEN and the yellow part represents the amino acid residues of the nanobody.</p>
Full article ">Figure 6
<p>Expression and purification results of nanobody V44. (<b>A</b>) For extraction and purification of the nanobody, lane 1 represents the total protein in the original sample, while lane 2 shows the flow-through eluate. Lanes 3 to 10 manifest the eluate from 10 mM, 20 mM, 30 mM, 40 mM, 50 mM, 100 mM, 100 mM, and 200 mM imidazole solutions, respectively. The flow-through eluate was ascertained to be predominantly devoid of target proteins, signifying that the preponderance of target proteins were successfully bound to the Ni column. This also intimates that high concentrations of imidazole (100 mM and 200 mM) were requisite to competitively elute the nanobody. (<b>B</b>) SDS-PAGE images of a nanobody; lanes 1–3 represent different loading amounts of nanobody V44. (<b>C</b>) Western blot images of a nanobody; lanes 1–3 represent different loading amounts of nanobody V44.</p>
Full article ">Figure 7
<p>Specificity and Affinity of anti-ZEN VHH. (<b>A</b>) The binding specificity of purified nanobodies to ZEN was detected by indirect ELISA. The results are presented as mean absorbance values at OD450 nm ± SD (<span class="html-italic">n</span> = 3). Statistical analysis was performed with one-way ANOVA (*** <span class="html-italic">p</span>&lt; 0.001). (<b>B</b>) The standard inhibition curve of the icELISA for zearalenone, an IC<sub>50</sub> value of 7.55 ng/mL, a detection range of 4.52–12.62 ng/mL (IC<sub>20</sub>–IC<sub>80</sub> inhibition concentration), and a correlation coefficient of R<sup>2</sup> = 0.99. The analysis of the absorbance of blank wells and the inhibition rates of corresponding inhibitory wells helped determine the optimal conditions for the working curve development.</p>
Full article ">
19 pages, 7047 KiB  
Article
Overexpression of NKG2D and IL24 in NK Cell-Derived Exosomes for Cancer Therapy
by Chujun Huang, Qian Hu, Peiyun Wang, Mi Xie, Ying Zhang, Zhixing Li, Shuqing Tang, Yuxuan Zhang, Zhixin Tian, Xionghao Liu, Zhiqing Hu and Desheng Liang
Int. J. Mol. Sci. 2025, 26(5), 2098; https://doi.org/10.3390/ijms26052098 - 27 Feb 2025
Viewed by 21
Abstract
Natural killer (NK) cell-derived exosomes (NK-Exos) are emerging as a promising avenue in cancer immunotherapy due to their inherent tumor-targeting properties and their capacity to deliver therapeutic agents directly to malignant cells. This research delves into the boosted anti-tumor potency of NK-Exos that [...] Read more.
Natural killer (NK) cell-derived exosomes (NK-Exos) are emerging as a promising avenue in cancer immunotherapy due to their inherent tumor-targeting properties and their capacity to deliver therapeutic agents directly to malignant cells. This research delves into the boosted anti-tumor potency of NK-Exos that has been genetically enhanced to overexpress NKG2D, a vital activating receptor, along with interleukin-24 (IL24), a cytokine renowned for its selective suppressive impact on tumor cells. NKG2D facilitates the recognition of tumor cells by binding to stress-induced ligands, while IL24 induces apoptosis and modulates immune responses to enhance tumor destruction. The NK-Exos engineered to express both NKG2D and IL24 significantly enhanced tumor targeting and increased the apoptosis rate of tumor cells by 30% in A549 and by 20% in HELA at 48 h compared with non-modified NK-Exos, respectively. Furthermore, this enhancement also impacted cell proliferation, with inhibition rates increasing by 30%, 15%, and 15% in A549, HELA, and MCF-7 cells, respectively, and it reduced A549 cell migration by 10%. The integration of NKG2D and IL24 within NK-Exos confers a dual therapeutic mechanism, synergistically amplifying their efficacy in cancer treatment. The utility of NK-Exos co-expressing NKG2D and IL24 offers a novel approach to overcome the limitations of current therapies, providing prolonged tumor suppression and precise targeting of malignant cells and holding great promise for clinical application. Full article
(This article belongs to the Section Molecular Oncology)
Show Figures

Figure 1

Figure 1
<p>Lentiviral vector construction and identification of NKG2D-NK and IL24-NKG2D-NK cell lines. (<b>A</b>) NKG2D plasmid, (<b>B</b>) IL24-NKG2D plasmid, (<b>C</b>) flow cytometry detection of NKG2D overexpression, (<b>D</b>,<b>E</b>) Western blot (WB) analysis of NKG2D and IL24 expression, <span class="html-italic">n</span> = 5, (<b>F</b>) NK cell proliferation rate, and (<b>G</b>) NK cell viability, <span class="html-italic">n</span> = 3 (significance was determined using a one-way ANOVA followed by Tukey’s test; ns: not significant, **** <span class="html-italic">p</span> &lt; 0.0001).</p>
Full article ">Figure 2
<p>Characterization of NK-derived exosomes post-purification. (<b>A</b>) Transmission electron microscopy (TEM) for exosome morphology (scale bar = 100 nm), (<b>B</b>) nanoparticle tracking analysis (NTA) for exosome size distribution, (<b>C</b>) Western blot for exosome markers (CD9, CD81, TSG101, calnexin), (<b>D</b>,<b>E</b>) Western blot for IL24 and NKG2D in exosomes, <span class="html-italic">n</span> = 3, and (<b>F</b>,<b>G</b>) quantitative PCR (qPCR) for IL24 and NKG2D mRNA expression in exosomes, <span class="html-italic">n</span> = 3 (significance was determined using a one-way ANOVA followed by Tukey’s test; ns: not significant, * <span class="html-italic">p</span> &lt; 0.05, *** <span class="html-italic">p</span> &lt; 0.001, **** <span class="html-italic">p</span> &lt; 0.0001).</p>
Full article ">Figure 3
<p>Exosome uptake, localization, and targeting advantage of NKG2D overexpression. (<b>A</b>–<b>F</b>) exosome uptake by tumor cells as assessed by fluorescence intensity (exosomes—red fluorescence (CM-Dil); nucleus—blue fluorescence(DAPI)): (<b>A</b>,<b>B</b>) A549, <span class="html-italic">n</span> = 3 (scale bar = 50 µm); (<b>C</b>,<b>D</b>) HELA, <span class="html-italic">n</span> = 3 (scale bar = 25 µm); (<b>E</b>,<b>F</b>) MCF-7, <span class="html-italic">n</span> = 3 (scale bar = 50 µm); (<b>G</b>–<b>I</b>) flow cytometric analysis of exosome uptake over time, <span class="html-italic">n</span> = 3; (<b>J</b>–<b>L</b>) GFP-A549 and 293T co-culture with exosomes (exosomes—red fluorescence (CM-Dil); GFP-A549—green fluorescence (GFP-A549); nucleus of GFP-A549 and 293T cells—blue fluorescence (DAPI)) (scale bar = 50 µm), Colocalization analysis: (<b>K</b>) Pearson’s correlation coefficient (PCC) scatterplots; (<b>L</b>) Manders’ colocalization coefficient, <span class="html-italic">n</span> = 3 (significance was determined using a one-way ANOVA followed by Tukey’s test; ns: not significant, * <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 4
<p>Cytotoxic effect of NKG2D and IL24 overexpressing exosomes on tumor cells. (<b>A</b>) Immunofluorescence verification of IL24 in IL24-NKG2D-Exo entering tumor cells (exosomes—red fluorescence (CM-Dil); IL24—green fluorescence (IL24), nucleus of A549 or HELA—blue fluorescence (DAPI)) (scale bar = 25 µm); (<b>B</b>–<b>E</b>) annexin V-FITC/PI flow cytometry for assessing apoptosis in tumor cells incubated with exosomes at 24 h and 48 h: (<b>B</b>,<b>C</b>) A549, <span class="html-italic">n</span> = 3; (<b>D</b>,<b>E</b>) HELA, <span class="html-italic">n</span> = 3; (<b>F</b>) CCK-8 assay for the growth inhibition of tumor cells incubated with exosomes at 12 h, 24 h, and 48 h: A549, HELA, MCF-7, <span class="html-italic">n</span> = 5. (<b>G</b>,<b>H</b>) Scratch assay to assess the impact of exosomes on tumor cell migration at 12 h, 36 h, and 48 h, <span class="html-italic">n</span> = 3 (scale bar = 200 µm); (<b>I</b>–<b>L</b>) Western blot analysis of apoptotic pathways: (<b>I</b>,<b>J</b>) cleaved caspase-3, <span class="html-italic">n</span> = 3; (<b>K</b>,<b>L</b>) Bcl-2, <span class="html-italic">n</span> = 3; (<b>M</b>) The mechanism schematic diagram (significance was determined using a one-way ANOVA followed by Tukey’s test; ns: not significant, * <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 ">
14 pages, 1742 KiB  
Article
Characterization of Glycoprotein 5-Specific Response in Pigs Vaccinated with Modified Live Porcine Reproductive and Respiratory Syndrome Virus Vaccine Derived from Two Different Lineages
by Jing Huang, Venkatramana D. Krishna, Igor A. D. Paploski, Kimberly VanderWaal, Declan C. Schroeder and Maxim C.-J. Cheeran
Vaccines 2025, 13(3), 247; https://doi.org/10.3390/vaccines13030247 - 27 Feb 2025
Viewed by 139
Abstract
Background/Objectives: Porcine reproductive and respiratory syndrome virus (PRRSV) is classified into various lineages based on the phylogenetic variation of orf5, which encodes a major surface glycoprotein GP5 containing both neutralizing and non-neutralizing linear epitopes. Several positively selected sites have been identified on [...] Read more.
Background/Objectives: Porcine reproductive and respiratory syndrome virus (PRRSV) is classified into various lineages based on the phylogenetic variation of orf5, which encodes a major surface glycoprotein GP5 containing both neutralizing and non-neutralizing linear epitopes. Several positively selected sites have been identified on the GP5 ectodomain, indicating host immune pressure on these sites. This present study aimed to investigate the kinetics of antibody responses to GP5 and to map the epitope-specific response to the GP5 ectodomain from different PRRSV lineages after vaccination with commercially available modified live virus (MLV) vaccines. Methods: Post-weaning pigs were vaccinated with MLV vaccines derived from either lineage 1D (Prevacent PRRS®) or lineage 5 (Ingelvac PRRS®). Animals were challenged with a heterologous (lineage 1A) strain at 64 days post-vaccination (dpv). Blood samples were collected at various times post-vaccination and challenge. Kinetics of antibody response to different PRRSV antigens were monitored and virus neutralization against archetypal and contemporary strains belonging to lineage 5 and 1A were evaluated. In addition, antibody responses to peptides derived from the GP5 ectodomain of different viral lineages were assessed. Results: Our results showed that the GP5-specific antibody response observed between 18 and 35 dpv was delayed compared to responses to the viral nucleocapsid protein. The polyclonal antibody response in both vaccinated groups showed similar levels of binding to variant GP5 peptides from different sub-lineages. Notably, in both vaccinated groups, the antibody directed to a peptide representing the GP5 ectodomain of a lineage 1C strain (variant 1C.5) displayed a rise in titer at 64 dpv, which was further increased by the challenge with the lineage 1A strain. Less than 50% of animals developed heterologous neutralizing antibodies post-vaccination with both MLV vaccines. However, higher neutralization titers were observed in all vaccinated animal post-challenge. Conclusions: Together, these data provide insights into the antibody responses to the GP5 ectodomain in MLV-vaccinated swine herds. Full article
(This article belongs to the Special Issue Vaccines for Porcine Viruses)
Show Figures

Figure 1

Figure 1
<p>Timeline and experimental design. Pigs were either vaccinated with the respective modified live virus (MLV) vaccines or left unvaccinated, followed by a virus challenge (IA/2014) at 64 days post-vaccination (dpv). Thick ticks represent sample collections for both peripheral blood mononuclear cells and serum while the thin ticks represent sample collections for serum only. This figure was created with <a href="http://BioRender.com" target="_blank">BioRender.com</a>.</p>
Full article ">Figure 2
<p>Relative levels of antibody responses in animals vaccinated with L1D or L5 vaccines, and controls, against PRRSV isolates from different lineages. Serum samples were collected at 64 days post-vaccination. (<b>A</b>) D11-052871 is a viral isolate belonging to lineage 5, (<b>B</b>) IA/2014 belongs to lineage 1A and (<b>C</b>) 46/2020 belongs to lineage 1C.5. Data were presented as mean ± SEM with individual values denoted by closed symbols. All optical density (OD450) values were normalized relative to a reference serum recognizing the corresponding antigen. ns = not significant; ** <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, one-way ANOVA.</p>
Full article ">Figure 3
<p>Kinetics of antibody responses to PRRSV antigens post-vaccination with L1D and L5 MLV. Serum samples were collected at 0, 6, 13, 18, 35, 53, and 64 days post-vaccination. (<b>A</b>) Kinetics of the L5 whole virus-specific antibody response post-vaccination. (<b>B</b>) Kinetics of the nucleocapsid (N)-specific antibody response post-vaccination. (<b>C</b>) Kinetics of recombinant GP5-specific antibody responses post-vaccination. Data are shown as means ± SEMs. Values were normalized as relative levels compared to a reference serum recognizing the corresponding antigen.</p>
Full article ">Figure 4
<p>Frequencies of PRRSV antigen-specific antibody-secreting cells (ASCs) post-vaccination. PBMCs were isolated from unvaccinated animals or animals vaccinated with L5-derived MLV vaccine at different time points post-vaccination and then cultured to determine IgG-ASCs for different antigens including PRRSV D11-052871 (lineage 5) (<b>A</b>), N (<b>B</b>), and GP5 (<b>C</b>) using the ELISPOT assay. Data are shown as means ± SEMs.</p>
Full article ">Figure 5
<p>Antibody responses to peptides (aa 32–aa 61) designed from the GP5 ectodomain of different lineages. Relative levels of antibody response to the GP5 ectodomain are shown using serum samples collected at (<b>A</b>) 64 dpv and (<b>B</b>) 14 days post-challenge (dpc) from vaccinated and challenged (L1D MLV and L5 MLV), unvaccinated and challenged (control), or unvaccinated (naïve) groups. Peptides corresponding to L5 (D11-052871), L1C (46/2020), L1A (IA/2014), L1C* (KP283416) and L1D (KY348849) were used to assess vaccination response. Two-way ANOVA. (<b>C</b>) Kinetics of antibody response to the GP5 ectodomain peptide derived from 46/2020 (lineage 1C.5). The arrow indicates challenge with an L1A isolate (IA/2014) at 64 dpv. Data are shown as means ± SEMs. Values are normalized as relative levels compared to a reference serum.</p>
Full article ">
18 pages, 2871 KiB  
Article
Unveiling the Mechanism of Action of Palmitic Acid, a Human Topoisomerase 1B Inhibitor from the Antarctic Sponge Artemisina plumosa
by Alessio Ottaviani, Davide Pietrafesa, Bini Chhetri Soren, Jagadish Babu Dasari, Stine S. H. Olsen, Beatrice Messina, Francesco Demofonti, Giulia Chicarella, Keli Agama, Yves Pommier, Blasco Morozzo della Rocca, Federico Iacovelli, Alice Romeo, Mattia Falconi, Bill J. Baker and Paola Fiorani
Int. J. Mol. Sci. 2025, 26(5), 2018; https://doi.org/10.3390/ijms26052018 - 26 Feb 2025
Viewed by 61
Abstract
Cancer remains a leading cause of death worldwide, highlighting the urgent need for novel and more effective treatments. Natural products, with their structural diversity, represent a valuable source for the discovery of anticancer compounds. In this study, we screened 750 Antarctic extracts to [...] Read more.
Cancer remains a leading cause of death worldwide, highlighting the urgent need for novel and more effective treatments. Natural products, with their structural diversity, represent a valuable source for the discovery of anticancer compounds. In this study, we screened 750 Antarctic extracts to identify potential inhibitors of human topoisomerase 1 (hTOP1), a key enzyme in DNA replication and repair, and a target of cancer therapies. Bioassay-guided fractionation led to the identification of palmitic acid (PA) as the active compound from the Antarctic sponge Artemisina plumosa, selectively inhibiting hTOP1. Our results demonstrate that PA irreversibly blocks hTOP1-mediated DNA relaxation and specifically inhibits the DNA religation step of the enzyme’s catalytic cycle. Unlike other fatty acids, PA exhibited unique specificity, which we confirmed through comparisons with linoleic acid. Molecular dynamics simulations and binding assays further suggest that PA interacts with hTOP1-DNA complexes, enhancing the inhibitory effect in the presence of camptothecin (CPT). These findings identify PA as a hTOP1 inhibitor with potential therapeutic implications, offering a distinct mechanism of action that could complement existing cancer therapies. Full article
(This article belongs to the Special Issue Discovering Novel Bioactive Compounds Against Cancers)
Show Figures

Figure 1

Figure 1
<p>Relaxation of supercoiled DNA in presence of PA. (<b>A</b>) 2D structure representation of PA. (<b>B</b>) Relaxation of negative supercoiled DNA plasmid by hTOP1 at increasing PA concentrations (lanes 2–8), lane 1 DMSO and lane 9 with 200 μM PA and no protein added. (<b>C</b>) Relaxation of negative supercoiled DNA plasmid in a time course experiment with DMSO (lanes 1–6), with 100 µM PA (lanes 7–12), and 100 μM CPT (lanes 13–18); lanes 19 and 20 correspond to samples with 100 µM PA and 100 µM CPT, respectively, with no protein added. Reaction products were resolved on agarose gel and visualized with ethidium bromide (EtBr). DSC—dimer supercoiled DNA plasmid; MSC—monomer super-coiled DNA plasmid; C—negative control (corresponding to samples with 100 µM PA and 100 µM CPT, respectively, with no protein added).</p>
Full article ">Figure 2
<p>Analysis of religation of hTOP1 catalytic mechanism using FITC (fluorescein isothiocyanate) oligonucleotide labeled SS. (<b>A</b>) Top panel displays sequences of fluorescently FITC labeled SS used in religation assay, asterisk indicates that FITC was conjugated to guanine. (<b>B</b>) Representation of a denaturing polyacrylamide gel of the religation assay. Samples were incubated for 1 h at 25 °C followed by 30 min at 37 °C. Reaction was initiated by adding a 200-fold excess of R11 oligonucleotide, either with or without 100 µM PA, then stopped at various time points with 0.5% SDS. CL1 represents cleaved strand (TOP1cc), C is negative control (no protein added), and 0 denotes TOP1cc starting condition before addition of R11. (<b>C</b>) Plot illustrates percentage of religated bands over time from religation assay. Figure presents cumulative data with mean ± SD from three independent experiments. Statistical significance is indicated with asterisks: *** <span class="html-italic">p</span> ≤ 0.001.</p>
Full article ">Figure 3
<p>hTOP1–DNA cleavage complex reversal assay. (<b>A</b>) Top panel displays sequences of fluorescently labeled SS used in the assay. (<b>B</b>) Polyacrylamide gel reporting kinetics of formation of PA and CPT induced hTOP1-mediated DNA cleavage complexes. 3′-6-FAM end labeled 48 bp oligonucleotide was reacted with hTOP1 in presence or absence of 1 µM CPT, 10 µM PA, or both at 25 °C for 20 min. DNA cleavage was reversed by adding 0.35 M NaCl and monitored over time. (<b>C</b>) Graph reporting 35 bp band quantification as function of time for PA and CPT (blue line), CPT (red line), PA (black line), and hTOP1 (green line) as control. Samples are represented as mean value ± SD. * <span class="html-italic">p</span> &lt; 0.05 *** <span class="html-italic">p</span> &lt; 0.001.</p>
Full article ">Figure 4
<p>Pre-incubation dose-dependent relaxation assay. Relaxation of negative supercoiled plasmid DNA in a dose-dependent experiment with DMSO (lane 1), 150 μM and 200 μM PA in pre-incubation condition, indicated as PRE (lanes 2–3), and 150 μM and 200 μM PA in simultaneous condition, indicated as SIM (lane 4–5), with no protein added in lane 6. Reaction products are resolved on agarose gel and visualized with EtBr. C indicates negative control.</p>
Full article ">Figure 5
<p>Essential motions of MD simulations. (<b>A</b>) Representation of two extreme projections of motions described by first eigenvector (PC1), interpolated onto 3D structures of hTOP1-DNA (left), hTOP1-DNA-CPT (center) and hTOP1-DNA-CPT-PAs (right) systems. Direction and amplitude of the internal motions are shown as color shift from blue to red and width of ribbons, respectively. (<b>B</b>) 2D projections of first (PC1) and second (PC2) eigenvectors of hTOP1-DNA (left), hTOP1DNA-CPT (center) and hTOP1-DNA-CPT-PAs (right) systems. Color coding shows progression from starting (violet) to final (yellow) stages of the simulations.</p>
Full article ">Figure 6
<p>MD of PA-DNA systems. Representative snapshots for first replica are shown at 0 ns, 75 ns, 125 ns, and 250 ns. For each figure, DNA is shown using a cyan surface representation, while PA molecules are shown in Van der Waals representation.</p>
Full article ">
12 pages, 4034 KiB  
Article
Study of Intracellular Peptides of the Central Nervous System of Zebrafish (Danio rerio) in a Parkinson’s Disease Model
by Louise O. Fiametti, Camilla A. Franco, Leticia O. C. Nunes, Leandro M. de Castro and Norival A. Santos-Filho
Int. J. Mol. Sci. 2025, 26(5), 2017; https://doi.org/10.3390/ijms26052017 - 26 Feb 2025
Viewed by 41
Abstract
Although peptides have been shown to have biological functions in neurodegenerative diseases, their role in Parkinson’s disease has been understudied. A previous study by our group, which used a 6-hydroxydopamine zebrafish model, suggested that nine intracellular peptides may play a part in this [...] Read more.
Although peptides have been shown to have biological functions in neurodegenerative diseases, their role in Parkinson’s disease has been understudied. A previous study by our group, which used a 6-hydroxydopamine zebrafish model, suggested that nine intracellular peptides may play a part in this condition. In this context, our aim is to better understand the role of five of these nine peptides. The selection of peptides was made based on their precursor proteins, which are fatty acid binding protein 7, mitochondrial ribosomal protein S36, MARCKS-related protein 1-B, excitatory amino acid transporter 2 and thymosin beta-4. The peptides were chemically synthesized in solid phase and characterized by high-performance liquid chromatography and mass spectrometry. Circular dichroism was performed to determine the secondary structure of each peptide, which showed that all five peptides maintain a random structure in the aqueous solutions that were studied. Two molecules show a helical profile in trifluoroethanol, a known structuring agent. Cell viability by the MTT assay indicates that all five peptides are not cytotoxic in all concentrations tested in both mouse and human cell lines. Behavioral assay using a 6-OHDA zebrafish larvae model suggest that all peptides help in the recovery of motor function with 24 h treatment at two concentrations. Three peptides showed a complete recovery from the 6-OHDA-induced motor impairment. Further studies are needed to better understand the mechanism of action of these peptides and whether they are truly a potential ally against Parkinson’s disease. Full article
(This article belongs to the Special Issue New Challenges of Parkinson’s Disease)
Show Figures

Figure 1

Figure 1
<p>Chromatograms of pure peptides p-FABP7 (<b>A</b>), p-MRPS36 (<b>B</b>), p-MARCKS1B (<b>C</b>), p-EEAT2 (<b>D</b>) and p-Tbeta4 (<b>E</b>). The degree of purity of the fractions was determined on a Shimadzu chromatograph, using a 0.46 × 25 cm C18 reversed-phase analytical column (Kromasil). Solutions used for purification and verification of the purity degree of the peptides were 0.045% TFA in ultrapure water (solution A) and 0.036% TFA in acetonitrile (solution B) in a 30 min run at a flow rate of 1 mL/min in a 5–95% of solution B gradient.</p>
Full article ">Figure 2
<p>Mass spectra of p-FABP7 (<b>A</b>), p-MRPS36 (<b>B</b>), p-MARCKS1B (<b>C</b>), p-EEAT (<b>D</b>) and p-Tbeta4 (<b>E</b>). Arrows indicate peaks that confirm correct synthesis according to mass/charge ratio. Mass spectra were obtained by direct injection electrospray ionization in positive detection mode.</p>
Full article ">Figure 3
<p>Circular dichroism spectra of p-FABP7 (<b>A</b>), p_MRSP36 (<b>B</b>), p-MARCKS1B (<b>C</b>), p-EEAT2 (<b>D</b>) and p-Tbeta4 (<b>E</b>) at pHs 5 (black), 7.4 (red), 10 (blue) and in TFE (violet).</p>
Full article ">Figure 4
<p>Cell viability (%) of Neuro 2A cells treated in different concentrations of p-FABP7 (<b>A</b>), p-MRPS36 (<b>B</b>), p-MARCKS1B (<b>C</b>), p-EAAT2 (<b>D</b>) and p-Tbeta4 (<b>E</b>). Experiments were performed in duplicate on three different occasions. ****, <span class="html-italic">p</span> &lt; 0.0001.</p>
Full article ">Figure 5
<p>Cell viability (%) of SHSY5Y cells treated in different concentrations of p-FABP7 (<b>A</b>), p-MRPS36 (<b>B</b>), p-MARCKS1B (<b>C</b>), p-EAAT2 (<b>D</b>) and p-Tbeta4 (<b>E</b>). Experiments were performed in duplicate on three different occasions. ****, <span class="html-italic">p</span> &lt; 0.0001.</p>
Full article ">Figure 6
<p>Mean speed of zebrafish treated with 750 µM for 24 h, then by p-FABP7 (<b>A</b>), p-MRPS36 (<b>B</b>), p-MARCKS1B (<b>C</b>), p-EEAT2 (<b>D</b>) and p-Tbeta4 (<b>E</b>) at concentrations of 10 µM, 20 µM, 30 µM, 40 µM and 50 µM for 24 h. Experiments were performed with 4 embryos per group on three different occasions, adding up to 12 embryos per group. #, <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; 0001, ####, <span class="html-italic">p</span> &lt; 0.00001 compared to the control group. *, <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; ****, <span class="html-italic">p</span> &lt; 0.00001 compared to the 6-OHDA group.</p>
Full article ">Figure 7
<p>Distance traveled by zebrafish treated with 750 µM for 24 h, then by p-FABP7 (<b>A</b>), p-MRPS36 (<b>B</b>), p-MARCKS1B (<b>C</b>), p-EEAT2 (<b>D</b>) and p-Tbeta4 (<b>E</b>) at concentrations of 10 µM, 20 µM, 30 µM, 40 µM and 50 µM for 24 h. Experiments were performed with 4 embryos per group on three different occasions, adding up to 12 embryos per group. #, <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; 0001; ####, <span class="html-italic">p</span> &lt; 0.00001 compared to the control group. *, <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; ****, <span class="html-italic">p</span> &lt; 0.00001 compared to the 6-OHDA group.</p>
Full article ">Figure 8
<p>Zebrafish larva model. Zebrafish larvae were immersed in 6-OHDA 750 µM for 24 h, then in each peptide at 10 µM, 20 µM, 30 µM, 40 µM or 50 µM for 24 h. Videos were recorded for 5 min and then analyzed by software (ToxTrac Version 2.96). Parameters were mean speed and distance traveled.</p>
Full article ">
19 pages, 3777 KiB  
Article
Sustained Epigenetic Reactivation in Fragile X Neurons with an RNA-Binding Small Molecule
by Christina W. Kam, Jason G. Dumelie, Gabriele Ciceri, Wang-Yong Yang, Matthew D. Disney, Lorenz Studer and Samie R. Jaffrey
Genes 2025, 16(3), 278; https://doi.org/10.3390/genes16030278 - 25 Feb 2025
Viewed by 348
Abstract
Background/Objectives: Fragile X syndrome (FXS) is a disease of pathologic epigenetic silencing induced by RNA. In FXS, an expanded CGG repeat tract in the FMR1 gene induces epigenetic silencing during embryogenesis. FMR1 silencing can be reversed with 5-aza-deoxyctidine (5-aza-dC), a nonspecific epigenetic reactivator; [...] Read more.
Background/Objectives: Fragile X syndrome (FXS) is a disease of pathologic epigenetic silencing induced by RNA. In FXS, an expanded CGG repeat tract in the FMR1 gene induces epigenetic silencing during embryogenesis. FMR1 silencing can be reversed with 5-aza-deoxyctidine (5-aza-dC), a nonspecific epigenetic reactivator; however, continuous administration of 5-aza-dC is problematic due to its toxicity. We describe an approach to restore FMR1 expression in FXS neurons by transient treatment with 5-aza-dC, followed by treatment with 2HE-5NMe, which binds the CGG repeat expansion in the FMR1 mRNA and could block the resilencing of the FMR1 gene after withdrawal of 5-aza-dC. Methods: This study uses immunofluorescence and fluorescent in situ hybridization (FISH) to measure whether FMR1 expression is maintained in FXS post-mitotic neurons treated with 2HE-5NMe. Genome-wide profiling of histone marks was used to monitor epigenetic changes and drug selectivity in response to 5-aza-dC followed by 2HE-5NMe treatment. Changes to dendritic morphology were visualized using confocal microscopy. Results: In this study, we find that 2HE-5Nme maintains FMR1 in a reactivated state after reactivation using 5-aza-dC in post-mitotic neurons. FMR1 reactivation in neurons results in the re-expression of FMRP and reversal of FXS-associated dendritic spine defects. Conclusions: These results demonstrate that an RNA-binding small molecule can achieve gene-specific epigenetic control and provide an approach for the restoration of FMRP in FXS neurons. Full article
(This article belongs to the Section Epigenomics)
Show Figures

Figure 1

Figure 1
<p>5-aza-dC reactivates <span class="html-italic">FMR1</span> mRNA expression and FMRP protein levels in FXS neurons. (<b>A</b>) FXS neurons derived from WCMC37 hESCs lack <span class="html-italic">FMR1</span> mRNA expression and FMRP, and 5-aza-dC reactivates expression. Representative images (<span class="html-italic">n</span> = 3 experiments) of anti-MAP2 and anti-FMRP immunofluorescence staining and <span class="html-italic">FMR1</span> FISH imaging of untreated wild-type and FXS neurons and FXS neurons treated with 1 µM 5-aza-dC for 7 days (scale bar, 10 µm). Insets show images enlarged by 150% with FISH puncta in white. (<b>B</b>) 5-aza-dC reactivates <span class="html-italic">FMR1</span> mRNA expression in FXS neurons. <span class="html-italic">FMR1</span> FISH was performed in neurons derived from a wild-type hESC line (WCMC7) and an FXS hESC line (WCMC37). 5-aza-dC (7 days, 1 µM) increases the number of detected <span class="html-italic">FMR1</span> FISH puncta in FXS neurons, but it has minimal effect on wild-type neurons. <span class="html-italic">FMR1</span> FISH foci were quantified from <span class="html-italic">n</span> = 3 experiments. Shown are the mean and s.d.; univariate two-sided <span class="html-italic">t</span>-test. (<b>C</b>) 5-aza-dC treatment leads to the appearance of FMRP in FXS neurons. Shown is the average FMRP immunofluorescence intensity per cell body in wild-type and FXS neurons. FMRP is nearly undetectable in FXS neurons but achieves levels similar to wild-type FMRP levels after 7 d of 1 µM 5-aza-dC treatment. The quantification of FMRP mean intensity was performed in <span class="html-italic">n</span> = 3 experiments. Shown are the mean and s.d.; univariate two-sided <span class="html-italic">t</span>-test.</p>
Full article ">Figure 2
<p>2HE-5NMe maintains <span class="html-italic">FMR1</span> reactivation in FXS neurons following 5-aza-dC withdrawal. (<b>A</b>) 2HE-5NMe maintains <span class="html-italic">FMR1</span> mRNA expression and FMRP protein levels in FXS neurons following 5-aza-dC withdrawal. Representative images (<span class="html-italic">n</span> = 3 experiments) of anti-MAP2 immunofluoresence (green), anti-FMRP immunofluorescence (violed), <span class="html-italic">FMR1</span> FISH (red) and DAPI staining (blue) of FXS neurons derived from WCMC37 hESCs treated with 1 µM 5-aza-dC for 7 days followed by 14 days of either DMSO or 500 nM 2HE-5NMe (scale bar, 10 µm). Insets show images enlarged by 150% with FISH puncta in white. (<b>B</b>) 2HE-5NMe maintains the appearance of FMRP in FXS neurons. Shown is the average FMRP immunofluorescence intensity per cell body in FXS neurons treated with 7 d of 1 µM 5-aza-dC treatment followed by 14 d of 500 nM 2HE-5NMe or DMSO. FMRP levels drop when 5-aza-dC is withdrawn and replaced with vehicle (DMSO) for 14 d. The quantification of FMRP mean intensity was performed in <span class="html-italic">n</span> = 3 experiments. Shown are the mean and s.d.; univariate two-sided <span class="html-italic">t</span>-test. (<b>C</b>) The quantification of <span class="html-italic">FMR1</span> FISH foci in wild-type (WCMC7, left) and FXS (WCMC37, right) neurons following 7 days of treatment with 1 µM 5-aza-dC followed by either 14 days of DMSO or 14 days of 500 nM 2HE-5NMe (<span class="html-italic">n</span> = 3 experiments, mean and s.d.; univariate two-sided <span class="html-italic">t</span>-test).</p>
Full article ">Figure 3
<p>5-aza-dC induces H3K4me3 at the <span class="html-italic">FMR1</span> locus, which is selectively maintained by 2HE-5Nme. (<b>A</b>) A schematic depiction of 5-aza-dC treatments followed by 2HE-5Nme-mediated maintenance of <span class="html-italic">FMR1</span> gene activation. FXS neurons were treated with 5-aza-dC for 7 d to reactivate the <span class="html-italic">FMR1</span> allele, followed by replacement in media lacking 5-aza-dC and containing either DMSO (vehicle) or 2HE-5Nme for 14 d to prevent the resilencing of the <span class="html-italic">FMR1</span> allele. (<b>B</b>) The effect of 5-aza-dC and 2HE-5Nme on activating and repressive histone marks at the <span class="html-italic">FMR1</span> promoter. Neurons were treated with 5-aza-dC (1 µM, 7 d), which was then replaced with either DMSO or 2HE-5Nme (14 d). Histone marks were measured using CUT&amp;RUN. The activating mark H3K4me3 is largely absent on the 5′ region of the <span class="html-italic">FMR1</span> gene in untreated FXS neurons (DMSO), but it is markedly induced in 5-aza-dC-treated neurons. This signal is nearly completely depleted in FXS neurons switched to DMSO for 14 d, but it remains detected in FXS neurons switched to 2HE-5Nme (500 nM). The repressive mark H3K27me3 is readily detected on the <span class="html-italic">FMR1</span> locus in all conditions. (<b>C</b>) Shown is a scatter plot indicating the change in H3K4me3 and H3K27me3 as measured by CUT&amp;RUN for gene promoters in FXS neurons treated with 5-aza-dC relative to untreated FXS neurons. <span class="html-italic">FMR1</span> is the most prominently affected gene after 5-aza-dC treatment, and it shows a marked increase in H3K4me3 levels. (<b>D</b>) The withdrawal of 5-aza-dC and culturing in DMSO leads to the loss of H3K4me3 marks at the <span class="html-italic">FMR1</span> promoter. Here, we compared histone marks in neurons cultured in 5-aza-dC and then DMSO to untreated FXS neurons. The scatter plot is prepared as in <span class="html-italic">C</span>. (<b>E</b>) The withdrawal of 5-aza-dC and culturing in 2HE-5Nme maintains H3K4me3 marks at the <span class="html-italic">FMR1</span> promoter. Here, we compared histone marks in FXS neurons cultured in 5-aza-dC and then 2HE-5NMe (500 nM) to untreated FXS neurons. The scatter plot is prepared as in <span class="html-italic">C</span>.</p>
Full article ">Figure 4
<p>Transient 5-aza-dC treatment induces dendritic spine maturation, which can be maintained by 2HE-5NMe. (<b>A</b>) A schematic depicting different stages of dendritic spine maturation. Immature spines are common in FXS and are recognized by their long and thin morphology. (<b>B</b>) FXS neurons have a higher percentage of immature spines compared to wild-type neurons. FXS neurons are derived from WCMC37 hESCs. Representative images of 3D reconstructed dendritic spines from wild-type and FXS neurons. Long and thin spines are readily detected in FXS neurons. (<b>C</b>) The high prevalence of immature dendritic spine maturation in FXS neurons can be reversed by treatment with 1 µM 5-aza-dC for 7 d and maintained replacement with media containing 500 nM 2HE-5NMe for 14 d. 5-aza-dC treatment results in the appearance of mature spines and the loss of immature spines and thus resembles wild-type neurons. Switching to DMSO-containing media causes the number of immature spines to increase and a loss of mature spines. In contrast, 2HE-5NMe (500 nM) improves the maintenance of mature spines. The quantification of the percentage of immature versus mature dendritic spines in wild-type neurons or FXS neurons following treatments was performed in <span class="html-italic">n</span> = 3 experiments. Shown are the mean and s.d.; univariate two-sided <span class="html-italic">t</span>-test. (<b>D</b>) The spine length of dendritic spines in FXS neurons shortens when the neurons are treated with 1 µM 5-aza-dC for 7 d, which is indicative of dendritic spine maturation. 2HE-5Nme maintains spine maturation induced by 5-aza-dC. The quantification of dendritic spine length (μm) in FXS neurons following the indicated treatments was performed in <span class="html-italic">n</span> = 3 experiments (box limits, interquartile range; whiskers, minimum to maximum; center line, median; univariate two-sided <span class="html-italic">t</span>-test).</p>
Full article ">
16 pages, 3579 KiB  
Article
A Quantitative Approach to Potency Testing for Chimeric Antigen Receptor-Encoding Lentiviral Vectors and Autologous CAR-T Cell Products, Using Flow Cytometry
by Juan José Mata-Molanes, Leticia Alserawan, Carolina España, Carla Guijarro, Ana López-Pecino, Hugo Calderón, Ane Altuna, Lorena Pérez-Amill, Nela Klein-González, Carlos Fernández de Larrea, Europa Azucena González-Navarro, Julio Delgado, Manel Juan and Maria Castella
Pharmaceutics 2025, 17(3), 303; https://doi.org/10.3390/pharmaceutics17030303 - 25 Feb 2025
Viewed by 238
Abstract
Background/Objectives: Potency testing of clinical-grade lentiviral vectors (LVVs) is critical to support a drug’s commercial approval. Careful consideration should be paid to the development of a suitable potency test during the drug’s clinical development. We aimed to develop an affordable, quantitative test [...] Read more.
Background/Objectives: Potency testing of clinical-grade lentiviral vectors (LVVs) is critical to support a drug’s commercial approval. Careful consideration should be paid to the development of a suitable potency test during the drug’s clinical development. We aimed to develop an affordable, quantitative test for our CAR19-LVV, based on a measure of transgene’s functional activity. Methods: Several indicators of functional activity of CAR19-LVV were explored in a co-culture setting of CAR-transduced Jurkat cells and CD19-expressing target cells. The selected assay was further developed and subjected to validation. Assay’s adaptability to other CAR-encoding LVV and autologous CAR-T cell products was also investigated. Results: Measure of CD69 expression on the membrane of Jurkat-CAR-expressing cells is a specific indicator of CAR functionality. Quantification of CD69 in terms of mean fluorescence intensity (MFI), coupled with an intra-assay standard curve calibration, allows for a quantitative assay with high precision, specificity, robustness, linearity and accuracy. The assay has also shown optimal performance for a CARBCMA-LVV product. Importantly, we show that in primary T cells, CD69 expression reflects CAR-T cell cytotoxicity. After adaptation, we have applied a CD69-based potency test, with simultaneous measurement of CAR-T cell cytotoxicity, to autologous CAR-T cell products, demonstrating the assay’s specificity also in this context. Conclusions: We developed a validated, in vitro cell-based potency test, using a quantitative flow-cytometry method, for our CAR19-LVV. The assay is based on the detection of T-cell activation upon CAR binding to antigen, which is a measure of transgene functionality. The assay was easily adapted to another CAR-encoding LVV, targeting a different molecule. Furthermore, the same assay principle can be applied in the context of autologous CAR-T cell products. The quantitative CD69 potency assay shows reduced variability among autologous products compared to the IFNγ assay and allows for simultaneous evaluation of traditional semi-quantitative cytotoxicity, thereby directly evaluating the drug’s mechanism of action (MoA) in the same assay. Full article
Show Figures

Figure 1

Figure 1
<p>Readout selection for potency assay using CAR19-LVV-transduced Jurkat cells. (<b>A</b>) Jurkat cells expressing CAR19 co-cultured with or without NALM6 cells at the indicated E:T ratios. Histograms show MFI of CD25 staining in CD3+ cells at 24 h time point. (<b>B</b>) Presence of IL-2 in the supernatants of co-cultures used in (<b>A</b>). The mean of triplicates ±SD is shown. (<b>C</b>) Histogram plots of CD69 expression in CD3+ cells after 24 h of co-culture. (<b>D</b>) CD69 MFI quantification in co-cultures of untransduced and CAR19-expressing Jurkat cells and NALM6. The mean of triplicates ±SD is shown. “ns” indicates no statistical significance. “*” indicates <span class="html-italic">p</span> ≤ 0.05.</p>
Full article ">Figure 2
<p>CD69-based potency assay optimization. (<b>A</b>) Correlation between the number of LVV particles used per cell (also known as multiplicity of infection (MOI)) and percentage of CAR-expressing cells at 72 h. The red line indicates a linear correlation. (<b>B</b>) Representative flow-cytometry images of CAR19-expressing Jurkat cells transduced at different MOIs. (<b>C</b>) CD69 activation index test performed using Jurkat cells displaying various percentages of CAR-expressing cells. Mean ± SD of triplicates is shown. (<b>D</b>) Same as (<b>C</b>) but using small intervals of CAR-expressing Jurkat cells. Mean ± SD is shown. (<b>E</b>) CD69 activation index test performed at the indicated time points after co-culture initiation. Mean ± SD is shown. “ns” indicates no statistical significance. “*” indicates <span class="html-italic">p</span> ≤ 0.05.</p>
Full article ">Figure 3
<p>CD69-based potency assay test in routine LVV-batch analysis. (<b>A</b>) Diagram depicting the different steps of CD69-based potency test. (<b>B</b>) Results of potency test of 26 CAR19-LVV batches tested, 4 CARBCMA-LVV batches and 26 untransduced Jurkat cells. The dashed red line indicates the limit of the specification set for ARI-0001-LVV samples. Mean ± SD is shown. (<b>C</b>) Results of potency test of 12 CARBCMA-LVV batches and 10 untransduced Jurkat cells tested using U266 cells as target cells. The dashed red line indicates the limit of the specification set for CARBCMA-LVV samples. Mean ± SD is shown. “*” indicates <span class="html-italic">p</span> ≤ 0.05.</p>
Full article ">Figure 4
<p>CD69 activation assay as a surrogate potency assay for autologous CAR-T cell products. (<b>A</b>) Diagram of CD69-based potency test applied to autologous CAR-T cell products. T-cell activation (CD69) and target cell killing or cytotoxicity are evaluated simultaneously. (<b>B</b>) Results of T-cell activation (CD69 fold activation) and target cell killing, using NALM6 clones with a variable number of CD19-molecules per cell as target cells. Results of triplicate experiments are shown. (<b>C</b>) Non-linear fit analysis between CD69 fold activation and %Target cell killing using data generated in (<b>B</b>). (<b>D</b>) Results of CD69-fold activation and cytotoxicity values obtained for 10 ARI-0001 batches analyzed. (<b>E</b>) Representative flow cytometry images of CD69 T-cell activation of patient batches analyzed in (<b>D</b>).</p>
Full article ">
29 pages, 3039 KiB  
Review
Ligand-Induced Biased Activation of GPCRs: Recent Advances and New Directions from In Silico Approaches
by Shaima Hashem, Alexis Dougha and Pierre Tufféry
Molecules 2025, 30(5), 1047; https://doi.org/10.3390/molecules30051047 - 25 Feb 2025
Viewed by 116
Abstract
G-protein coupled receptors (GPCRs) are the largest family of membrane proteins engaged in transducing signals from the extracellular environment into the cell. GPCR-biased signaling occurs when two different ligands, sharing the same binding site, induce distinct signaling pathways. This selective signaling offers significant [...] Read more.
G-protein coupled receptors (GPCRs) are the largest family of membrane proteins engaged in transducing signals from the extracellular environment into the cell. GPCR-biased signaling occurs when two different ligands, sharing the same binding site, induce distinct signaling pathways. This selective signaling offers significant potential for the design of safer and more effective drugs. Although its molecular mechanism remains elusive, big efforts are made to try to explain this mechanism using a wide range of methods. Recent advances in computational techniques and AI technology have introduced a variety of simulations and machine learning tools that facilitate the modeling of GPCR signal transmission and the analysis of ligand-induced biased signaling. In this review, we present the current state of in silico approaches to elucidate the structural mechanism of GPCR-biased signaling. This includes molecular dynamics simulations that capture the main interactions causing the bias. We also highlight the major contributions and impacts of transmembrane domains, loops, and mutations in mediating biased signaling. Moreover, we discuss the impact of machine learning models on bias prediction and diffusion-based generative AI to design biased ligands. Ultimately, this review addresses the future directions for studying the biased signaling problem through AI approaches. Full article
(This article belongs to the Special Issue Protein-Ligand Interactions)
Show Figures

Graphical abstract

Graphical abstract
Full article ">Figure 1
<p>Structural organization of GPCRs. Example of the glucagon-like peptide-1 receptor (PDB: 7lcj).</p>
Full article ">Figure 2
<p>Conformational diversity of GPCRs and biased signaling. (<b>A</b>): Different conformations of GPCRs are represented, inactive (green) and active conformations (wheat), but also alternative conformations able to trigger different pathways by recruiting different transducers (yellow, indigo). (<b>B</b>): Ligand bias: different ligands of the same receptor are able to select different conformations recruiting different transducers, hence activating different pathways. (<b>C</b>): System bias: different pathways are activated upon the binding of the same ligand to different versions of the same receptor (mutants, isoforms, localization). Here, a mutation affects the selection of the conformation selection upon ligand binding.</p>
Full article ">Figure 3
<p>Illustration of modified protocols to generate diverse GPCR conformations with AlphaFold. Templates are retrieved from a state-annotated database, replacing the default database. Depending on the protocol, different depths of MSA can be used. The different colored triangles in the MSA table represent different amino acids. Modeling the structure of the GPCR-G<math display="inline"><semantics> <mi>α</mi> </semantics></math> subunit complex with AlphaFold-multimer can lead to a higher-quality GPCR structure. In this example, the Succinate receptor 1 structures are retrieved from GPCRdb and GproteinDb, including the active and inactive models in blue and yellow, respectively. The GPCR structure modeled with a G<math display="inline"><semantics> <mrow> <mi>α</mi> <mi>s</mi> </mrow> </semantics></math> subunit using AlphaFold-multimer is represented in red (G<math display="inline"><semantics> <mrow> <mi>α</mi> <mi>s</mi> </mrow> </semantics></math> subunit in salmon).</p>
Full article ">Figure 4
<p>Conformational changes of TMs upon G-protein or <math display="inline"><semantics> <mi>β</mi> </semantics></math>-arrestin signaling. G-protein signaling in most of the GPCRs showed an outward movement of TM6. On the other hand, <math display="inline"><semantics> <mi>β</mi> </semantics></math>-arrestin signaling was reported to be accompanied by structural alterations in TM7 and H8, reducing the distance between TM7 and TM2.</p>
Full article ">Figure 5
<p>Depiction of the challenges related to ligand bias prediction.</p>
Full article ">Figure 6
<p>(<b>a</b>) Design strategy incorporating AI methods to identify biased ligands for the <math display="inline"><semantics> <mrow> <mi>D</mi> <mn>2</mn> </mrow> </semantics></math> dopamine receptor, inspired by the structure-based approach by [<a href="#B202-molecules-30-01047" class="html-bibr">202</a>]. (<b>b</b>) Schematic representation of the workflow to generate a scaffold to interact with an allosteric site of D2R, leveraging the encoder–decoder architecture conditioned with a pharmacophore developed by [<a href="#B203-molecules-30-01047" class="html-bibr">203</a>]. The pharmacophore can be derived from the structure of the allosteric pocket. (<b>c</b>) Generative diffusion model to design a linker. A linker can be generated with a diffusion-based model [<a href="#B204-molecules-30-01047" class="html-bibr">204</a>] to link two scaffolds designed to interact with specific D2R residues involved in biased signaling.</p>
Full article ">
17 pages, 9227 KiB  
Article
Nanoparticle-Enhanced Acoustic Wave Biosensor Detection of Pseudomonas aeruginosa in Food
by Sandro Spagnolo, Katharina Davoudian, Brian De La Franier, Robert Kocsis, Tibor Hianik and Michael Thompson
Biosensors 2025, 15(3), 146; https://doi.org/10.3390/bios15030146 - 25 Feb 2025
Viewed by 144
Abstract
A biosensor was designed for detecting Pseudomonas aeruginosa (P. aeruginosa) bacteria in whole milk samples. The sensing layer involved the antifouling linking molecule 3-(2-mercaptoethanoxy)propanoic acid (HS-MEG-COOH), which was covalently linked to an aptamer for binding P. aeruginosa. The aptasensor uses [...] Read more.
A biosensor was designed for detecting Pseudomonas aeruginosa (P. aeruginosa) bacteria in whole milk samples. The sensing layer involved the antifouling linking molecule 3-(2-mercaptoethanoxy)propanoic acid (HS-MEG-COOH), which was covalently linked to an aptamer for binding P. aeruginosa. The aptasensor uses the thickness shear mode (TSM) system for mass-sensitive acoustic sensing of the bacterium. High concentrations (105 CFU mL−1) of nonspecific bacteria, E. coli, S. aureus, and L. acidophilus, were tested with the aptasensor and caused negligible frequency shifts compared to P. aeruginosa. The aptasensor has high selectivity for P. aeruginosa, with an extrapolated limit of detection (LOD) of 86 CFU mL−1 in phosphate-buffered saline (PBS) and 157 CFU mL−1 in milk. To improve the sensitivity of the sensor, gold nanoparticles (AuNPs) were functionalized with the same aptamer for P. aeruginosa and flowed through the sensor following bacteria, reducing the extrapolated LOD to 68 CFU mL−1 in PBS and 46 CFU mL−1 in milk. The frequency variations in the aptasensor are proportional to various concentrations of P. aeruginosa (102–105 CFU mL−1) with and without AuNPs, respectively. The low and rapid mass-sensitive detection demonstrates the ability of the aptasensor to quantitatively identify bacterial contamination in buffer and milk. Full article
(This article belongs to the Special Issue Advancements in Biosensors for Foodborne Pathogens Detection)
Show Figures

Figure 1

Figure 1
<p>The functionalization of the TSM crystal to build the aptasensor, where (<b>A</b>) aptamers are covalently bonded to the HS-MEG-COOH linker, (<b>B</b>) HS-MEG-OH antifouling molecules act as spacers, and (<b>C</b>) HS-MEG-COOH molecules that did not bind to aptamers are, instead, elongated with ethanolamine. The coloured dots of the aptamers represent cytosine (orange), guanine (blue), and thymine (green) nucleic acids.</p>
Full article ">Figure 2
<p>The scheme of the functionalization of AuNPs with DNA aptamers. MUA-functionalized AuNPs were modified with NHS/EDC, aptamer, and then ethanolamine. The number of bound aptamer molecules as well as the size of the AuNPs are only intended to describe the functionalization process and do not reflect the actual size. The coloured dots of the aptamers represent cytosine (orange), guanine (blue), and thymine (green) nucleic acids.</p>
Full article ">Figure 3
<p>A comparison of the contact angles on layers on TSM crystals. Results from our previous work are shown in grey, compared to the aptasensor for <span class="html-italic">P. aeruginosa</span> designed in this work. The measurements were conducted in triplicate and the error bars represent the standard deviation.</p>
Full article ">Figure 4
<p>The decrease in frequency upon exposing milk to various functionalized surfaces on the TSM crystal. The arrow indicates the addition of whole milk on the sensing surface.</p>
Full article ">Figure 5
<p>A comparison of the fouling of whole milk on the bare gold electrode and various surface modifications. The results of this study are shown in colour, while the results of the previous work are in grey [<a href="#B21-biosensors-15-00146" class="html-bibr">21</a>].</p>
Full article ">Figure 6
<p>The sensing area, functionalized with HS-MEG-Mix, was further functionalized with HS-MEG-OH thiol in water, NHS/EDC, aptamer, and ethanolamine in flow. Afterwards, different bacterial concentrations were tested, and the frequency varied proportionally with bacterial concentration. The variation in these parameters following exposure to nanoparticles was also proportional with the bacterial concentration.</p>
Full article ">Figure 7
<p>Changes in the resonance frequency of TSM crystals following exposure to increasing concentrations of <span class="html-italic">P. aeruginosa</span> (on a logarithmic scale), before the subsequent exposure to specific aptamer-coated AuNPs (blue) and after (orange).</p>
Full article ">Figure 8
<p>(<b>A</b>) The response of the aptasensor to different bacterial concentrations and AuNPs, in which the frequency was proportional to the bacterial concentration. (<b>B</b>) The variation in the frequency following exposure to nanoparticles also varied proportionally with the bacterial concentration.</p>
Full article ">Figure 9
<p>The response of the aptasensor to <span class="html-italic">P. aeruginosa</span> compared to <span class="html-italic">E. coli</span>, <span class="html-italic">S. aureus</span>, and <span class="html-italic">L. acidophilus</span> in whole milk (10<sup>5</sup> CFU mL<sup>−</sup><sup>1</sup>): (<b>A</b>) frequency variations and (<b>B</b>) average changes in frequency.</p>
Full article ">Figure 10
<p>The average number of counted cells as a function of the concentration of bacteria after the experiments of detection. The number of cells is an average of the bacterial count from twenty SEM images. The error bars were determined from the standard deviation of the twenty SEM measurements.</p>
Full article ">Figure 11
<p>SEM images after exposing (<b>A</b>) antifouling and (<b>B</b>) bare electrode surfaces to milk. The antifouling surface in (<b>A</b>) is the aptasensor, comprising HS-MEG-Mix thiols, aptamer, and HS-MEG-COOH elongated with ethanolamine.</p>
Full article ">
Back to TopTop