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

Next Issue
Volume 26, February-2
Previous Issue
Volume 26, January-2
You seem to have javascript disabled. Please note that many of the page functionalities won't work as expected without javascript enabled.
 
 
ijms-logo

Journal Browser

Journal Browser

Int. J. Mol. Sci., Volume 26, Issue 3 (February-1 2025) – 530 articles

Cover Story (view full-size image): Transient Receptor Potential (TRP) ion channels play a key role in cancer-induced bone pain (CIBP). They are involved in the pain pathway at different sites both in the peripheral and central nervous system. The acidic tumour environment, caused by the release of protons by osteoclast activity, is a potent activator of TRP channels on primary afferent fibres. They are also expressed in astrocytes and glial cells, where they promote neuroinflammatory processes by interacting with the physiological cross-talk between neurons and immune cells. They also act directly on bone cells by modulating the differentiation of osteoclasts and osteoblasts. Preclinical data suggest a potential role in bone cancer pain management for the modulators of some TRP channels, namely TRPV1, TRPA1, TRPM7, and TRPM8. View this paper
  • Issues are regarded as officially published after their release is announced to the table of contents alert mailing list.
  • You may sign up for e-mail alerts to receive table of contents of newly released issues.
  • PDF is the official format for papers published in both, html and pdf forms. To view the papers in pdf format, click on the "PDF Full-text" link, and use the free Adobe Reader to open them.
Order results
Result details
Section
Select all
Export citation of selected articles as:
18 pages, 465 KiB  
Article
Optimising Aripiprazole Long-Acting Injectable: A Comparative Study of One- and Two-Injection Start Regimens in Schizophrenia with and Without Substance Use Disorders and Relationship to Early Serum Levels
by Giada Trovini, Ginevra Lombardozzi, Georgios D. Kotzalidis, Luana Lionetto, Felicia Russo, Angela Sabatino, Elio Serra, Simone Castorina, Giorgia Civita, Sara Frezza, Donatella De Bernardini, Giuseppe Costanzi, Marika Alborghetti, Maurizio Simmaco, Ferdinando Nicoletti and Sergio De Filippis
Int. J. Mol. Sci. 2025, 26(3), 1394; https://doi.org/10.3390/ijms26031394 - 6 Feb 2025
Viewed by 831
Abstract
Aripiprazole as a long-acting injectable (LAI) is initiated in oral aripiprazole-stabilised patients and needs, after first injection, 14 days supplementation of oral aripiprazole (one-injection start, OIS). Recently, an alternative two-injection start (TIS) was advanced, involving two 400 mg injections with a single 20 [...] Read more.
Aripiprazole as a long-acting injectable (LAI) is initiated in oral aripiprazole-stabilised patients and needs, after first injection, 14 days supplementation of oral aripiprazole (one-injection start, OIS). Recently, an alternative two-injection start (TIS) was advanced, involving two 400 mg injections with a single 20 mg oral supplementation of aripiprazole. We tested the two regimens in patients with schizophrenia (SCZ, n = 152, 90 men and 62 women) with (SUD+; n = 93) or without (SUD; n = 59) substance use disorders (SUDs), comparing OIS (n = 66) with TIS (n = 86) and SUD+ vs. SUD. For 26 patients, we measured weekly for one month, aripiprazole + dehydroaripiprazole (active moiety) levels. Patients were followed for three months after LAI with psychopathology and quality-of-life scales (BPRS, CGI-S, ACES, BIS-11, and WHOQOL). All groups improved in psychopathology with no differences between OSI and TIS and between SCZ–SUD+ and SCZ–SUD. The TIS group was associated with serum blood levels of the active moiety within the therapeutic window, while the OIS group showed peaks above the window, possibly exposing patients to toxicity. Treatments were well-tolerated. Here we showed no disadvantages for TIS vs. OIS and possibly increased safety. Shifting the initiation of aripiprazole LAIs to the TIS modality may be safe and pharmacokinetically advantageous. Full article
Show Figures

Figure 1

Figure 1
<p>Course of aripiprazole plus dehydroaripiprazole serum levels during the first month post-injection in patients assigned to OIS (blue line) or TIS (orange line). OIS, one-injection start with aripiprazole LAI 400 mg plus oral supplementation of 10 mg/day aripiprazole for 14 days; TIS, two-injection start with two 400 mg aripiprazole injections plus one dose of 20 mg oral aripiprazole on the first day. The thick grey dashed lines represent the therapeutic window of the active moiety (between 150 and 500 ng/mL).</p>
Full article ">
14 pages, 2834 KiB  
Article
Complement Factor B Deficiency Is Dispensable for Female Fertility but Affects Microbiome Diversity and Complement Activity
by Manato Sunamoto, Kazunori Morohoshi, Ban Sato, Ryo Mihashi, Masafumi Inui, Mitsutoshi Yamada, Kenji Miyado and Natsuko Kawano
Int. J. Mol. Sci. 2025, 26(3), 1393; https://doi.org/10.3390/ijms26031393 - 6 Feb 2025
Viewed by 496
Abstract
Complement factor B (CFB) is a crucial component for the activation of the alternative pathway due to the formation of the C3 convertase with C3b, which further produces C3b to enhance the overall complement activity. Although Cfb is expressed not only in the [...] Read more.
Complement factor B (CFB) is a crucial component for the activation of the alternative pathway due to the formation of the C3 convertase with C3b, which further produces C3b to enhance the overall complement activity. Although Cfb is expressed not only in the immune tissues, but also in the reproductive tract, the physiological role of the alternative complement pathway in reproduction remains unclear. In this study, we addressed this issue by producing Cfb-knockout (KO) mice and analyzing their phenotypes. Sperm function, number of ovulated oocytes, and litter size were normal in KO mice. In contrast, the diversity of microbiomes in the gut and vaginal tract significantly increased in KO mice. Some serine protease activity in the serum from KO mice was lower than that of wild-type mice. Since the serum from KO mice showed significantly lower activity of the alternative complement pathway, CFB was found to be essential for this pathway. Our results indicate that although the alternative pathway is dispensable for normal fertility and development, it maintains the gut and vaginal microbiomes by suppressing their diversity and activating the alternative complement pathway. Full article
Show Figures

Figure 1

Figure 1
<p><span class="html-italic">Complement factor b (Cfb)</span> expression analysis and generation of <span class="html-italic">Cfb</span>-KO mice. (<b>a</b>) <span class="html-italic">Cfb</span> mRNA expression in several tissues was analyzed by qPCR. <span class="html-italic">n</span> = 3. Data are expressed as means ± SE. (<b>b</b>) Immunoblotting of several tissue lysates with anti-CFB Ab and anti-GAPDH. GAPDH is an endogenous control. The densitometry analyses of band intensity normalized with GAPDH. (<b>c</b>) Schematic illustration of the wild-type (WT) and <span class="html-italic">Cfb</span>-KO (KO) allele of the <span class="html-italic">Cfb</span> gene. Black boxes mean exons. We designed gRNAs at exon 2 and exon 12, and KO allele deleted a 4 kbp long sequence in chromosome 17. (<b>d</b>) Sequences of PCR products with using Primer F/R in an WT and F1 mouse. gRNA sequences were found in WT allele, but not in F1 allele. Green characters indicated intron sequences between exon 1 and 2, and pink characters means exon 12 sequence. (<b>e</b>) Genomic DNA sequence at breakpoint junction in KO mice. Green and pink shading characters indicate the intron between exon 1 and 2 and the exon 12, respectively. (<b>f</b>) Efficiencies of <span class="html-italic">i</span>-GONAD treatment. (<b>g</b>) Representative image of WT and KO mice. (<b>h</b>) Genotyping of the F2 by using primer F and R. (<b>i</b>) Immunoblotting the serum from KO mice with anti-CFB Ab.</p>
Full article ">Figure 2
<p>Fertility of KO mice. (<b>a</b>) A method for measuring mouse fertility. For the analysis of male fertility, cauda epididymal sperm were collected from male KO mice. After the incubation in TYH medium, sperm were subjected to sperm motility and concentration analysis. For female fertility, Cumulus-oocyte complexes were isolated from super-ovulated KO mice. The number of pups produced after mating was counted to determine the reproductive potential of both males and females. (<b>b</b>) Sperm motility during 3 h of incubation between WT mice (triangles) and KO mice (solid circles): * <span class="html-italic">p</span> &lt; 0.05. <span class="html-italic">n</span> = 3. Data are expressed as means ± SE. (<b>c</b>) Rates of acrosome reacted sperm during 3 h of incubation between WT mice (triangles) and KO mice (solid circles): ** <span class="html-italic">p</span> &lt; 0.01. <span class="html-italic">n</span> = 3. Data are expressed as means ± SE. (<b>d</b>) Epididymal sperm concentration: ns, not significant, <span class="html-italic">n</span> = 3. Data are expressed as means ± SE. (<b>e</b>) The number of ovulated eggs: ns, not significant, <span class="html-italic">n</span> = 3. Data are expressed as means ± SE. (<b>f</b>) Representative image showing an entire reproductive tract of estrous female. (<b>g</b>) Fecundity of KO mice: ns, not significant, <span class="html-italic">n</span> = 3. Data are expressed as means ± SE.</p>
Full article ">Figure 3
<p>Characterization of gut and vaginal microbiota. (<b>a</b>) Stool samples and vaginal lavages were collected from identical mice at the estrous stage. (<b>b</b>) Relative abundance of <span class="html-italic">Bacteroides</span> in the gut microbiota: ns, not significant, <span class="html-italic">n</span> = 3. Data are expressed as means ± SE. (<b>c</b>) Relative abundance of <span class="html-italic">Firmicutes</span> in the gut microbiota: * <span class="html-italic">p</span> &lt; 0.05, <span class="html-italic">n</span> = 3. Data are expressed as means ± SE. (<b>d</b>) Total alpha diversity (Chao 1) of the gut microbiota: * <span class="html-italic">p</span> &lt; 0.05, <span class="html-italic">n</span> = 3. Boxes represent the interquartile range, lines indicate medians, and whiskers indicate the range. (<b>e</b>) Total alpha diversity (Shannon) of the gut microbiota: * <span class="html-italic">p</span> &lt; 0.05, <span class="html-italic">n</span> = 3. Boxes represent the interquartile range, lines indicate medians, and whiskers indicate the range. (<b>f</b>) Relative abundance of <span class="html-italic">Firmicutes</span> in the vaginal microbiota: ns, not significant, <span class="html-italic">n</span> = 3. Data are expressed as means ± SE. (<b>g</b>) Relative abundance of <span class="html-italic">Proteobacteria</span> in the vaginal microbiota: ns, not significant, <span class="html-italic">n</span> = 3. Data are expressed as means ± SE. (<b>h</b>) Total alpha diversity (Chao 1) of the vaginal microbiota: * <span class="html-italic">p</span> &lt; 0.05, <span class="html-italic">n</span> = 3. Boxes represent the interquartile range, lines indicate medians, and whiskers indicate the range. (<b>i</b>) Total alpha diversity (Shannon) of the vaginal microbiota: ns, not significant, <span class="html-italic">n</span> = 3. Boxes represent the interquartile range, lines indicate medians, and whiskers indicate the range.</p>
Full article ">Figure 4
<p>Enzyme assay of serum from KO mice. (<b>a</b>) Experimental flow. Serum was collected from wild-type and KO mice and then incubated with various MCA substrates at 37 °C for 1 h. Released fluorescence was measured every 10 min for 1 h. (<b>b</b>) Substrates with strong enzyme activity (<b>i</b>), with moderate enzyme activity (<b>ii</b>), with weak enzyme activity (<b>iii</b>). * <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. <span class="html-italic">n</span> = 3. Data are expressed as means ± SE.</p>
Full article ">Figure 5
<p>Hemolysis assay of serum from KO mice. (<b>a</b>) Experimental flow. For measuring hemolytic activity of complement classical pathway, sheep erythrocyte was sensitized with rabbit IgG and then incubated with the serum from WT and KO mice. To measure that of complement alternative pathway, rabbit erythrocyte was incubated with the serum in the presence of EGTA. After incubation, the rates of hemolysis were calculated. (<b>b</b>) Classical pathway hemolytic activity: ns, not significant, <span class="html-italic">n</span> = 3. Data are expressed as means ± SE. (<b>c</b>) Alternative pathway hemolytic activity: * <span class="html-italic">p</span> &lt; 0.05, <span class="html-italic">n</span> = 3. Data are expressed as means ± SE. (<b>d</b>) Schematic model of the relationship between innate immunity and the microbiota in female mice. The relationship between microbiota diversity and female fertility remains unclear.</p>
Full article ">
19 pages, 5241 KiB  
Article
Quantitative Proteomic Analysis of Lysine Malonylation in Response to Salicylic Acid in the Roots of Platycodon grandiflorus
by Wanyue Ding, Yingying Duan, Yuqing Wang, Jizhou Fan, Weiyi Rao and Shihai Xing
Int. J. Mol. Sci. 2025, 26(3), 1392; https://doi.org/10.3390/ijms26031392 - 6 Feb 2025
Viewed by 478
Abstract
Salicylic acid, as a plant hormone, significantly affects the physiological and biochemical indexes of soluble sugar, malondialdehyde content, peroxidase, and superoxide dismutase enzyme activity in Platycodon grandiflorus. Lysine malonylation is a post-translational modification that involves various cellular functions in plants, though it [...] Read more.
Salicylic acid, as a plant hormone, significantly affects the physiological and biochemical indexes of soluble sugar, malondialdehyde content, peroxidase, and superoxide dismutase enzyme activity in Platycodon grandiflorus. Lysine malonylation is a post-translational modification that involves various cellular functions in plants, though it is rarely studied, especially in medicinal plants. In this study, the aim was to perform a comparative quantitative proteomic study of malonylation modification on P. grandiflorus root proteins after salicylic acid treatment using Western blot with specific antibodies, affinity enrichment and LC-MS/MS analysis methods. The analysis identified 1907 malonyl sites for 809 proteins, with 414 proteins and 798 modification sites quantified with high confidence. Post-treatment, 361 proteins were upregulated, and 310 were downregulated. Bioinformatics analysis revealed that malonylation in P. grandiflorus is primarily involved in photosynthesis and carbon metabolism. Physiological and biochemical analysis showed that salicylic acid treatment increased the malondialdehyde levels, soluble protein, superoxide dismutase, and peroxidase activity but did not significantly affect the total saponins content in P. grandiflorus. These findings provide an important basis for exploring the molecular mechanisms of P. grandiflorus following salicylic acid treatment and enhance understanding of the biological function of protein lysine malonylation in plants. Full article
(This article belongs to the Special Issue Advanced Plant Molecular Responses to Abiotic Stresses)
Show Figures

Figure 1

Figure 1
<p><span class="html-italic">P. grandiflorus</span> and malonyl-lysine formation. (<b>A</b>) The plant of <span class="html-italic">P. grandiflorus</span>. (<b>B</b>) The process of protein lysine malonylation.</p>
Full article ">Figure 2
<p>The variation in physiological and biochemical indexes, total saponin content, and gene transcription level by salicylic acid treatment. (<b>A</b>) Chlorophyll a content; (<b>B</b>) chlorophyll b content; (<b>C</b>) the carotenoid content; (<b>D</b>) relative expression of <span class="html-italic">CAO</span>; (<b>E</b>) relative expression of <span class="html-italic">CHLG</span> gene; (<b>F</b>) the SP content; (<b>G</b>) the MDA content; (<b>H</b>) the POD activity; (<b>I</b>) the SOD activity; (<b>J</b>) the total saponin content. The data represent the average value and standard error of three replicates. The asterisks of *, **, ***, and **** above each column show significant differences in <span class="html-italic">p</span> ≤ 0.05, <span class="html-italic">p</span> ≤ 0.01, <span class="html-italic">p</span> ≤ 0.001, and <span class="html-italic">p</span> ≤ 0.0001, respectively.</p>
Full article ">Figure 3
<p>Malonylated protein analysis in <span class="html-italic">P. grandiflorus</span>. (<b>A</b>) SDS-PAGE of proteins of <span class="html-italic">P. grandiflorus</span>, dyed with Coomassie bright blue; (<b>B</b>) Western blot image of proteins’ malonylation in <span class="html-italic">P. grandiflorus</span>. C1 and C2 are samples of control, SA1 and SA2 are samples treated with SA; the molecular weight of protein is marked on the right; 20 mg of protein was loaded in each sample; (<b>C</b>) the overall number of malonyl acyl sites and proteins identified; (<b>D</b>) a PCA analysis between controls and treatments. The horizontal and vertical axes show the interpretability of PC1 and PC2. C1 and C2 are samples of control, SA1 and SA2 are samples treated with SA.</p>
Full article ">Figure 4
<p>Classification of identified proteins in <span class="html-italic">P. grandiflorus</span>. (<b>A</b>) GO Classification of malonylated proteins in <span class="html-italic">P. grandiflorus</span> based on biological process. (<b>B</b>) GO classification of malonylated proteins based on cellular compartment. (<b>C</b>) GO Classification of malonylated proteins based on molecular function. (<b>D</b>) Subcellular localization of identified proteins corresponding to modified sites.</p>
Full article ">Figure 5
<p>Classification of <span class="html-italic">P. grandiflorus</span> proteins at identification sites. (<b>A</b>) The GO classification of the malonylated proteins in <span class="html-italic">P. grandiflorus</span> based on biological processes, cellular components, and molecular functions. The enrichment bar graph shows the most significantly enriched functions, with the vertical axis representing the description information of the corresponding GO function, and the horizontal axis representing the enrichment significance <span class="html-italic">p</span> value of the log10 transformation. The larger the value, the stronger the enrichment significance. (<b>B</b>) KEGG pathway diagram of malonylation modification site. The vertical axis is the description information of the KEGG pathway, and the horizontal axis is the degree of enrichment of the differentially modified proteins after log2 conversion in this function (fold enrichment). The larger the value, the higher the degree; a blue dot color signifies stronger enrichment significance; the larger the dot, the more types of the different modified proteins. (<b>C</b>) COG/KOG pathway diagram of malonylation modification site.</p>
Full article ">Figure 6
<p>PPI relationship of malobylation proteins in <span class="html-italic">P. grandiflorus</span>. The blue color represents the downregulated modified proteins, and the red color represents the upregulated modified ones. The size indicates the number of proteins interacting with it.</p>
Full article ">Figure 7
<p>Motif analysis of the detected malonylation sites. (<b>A</b>) Sequence probability markers of the first eight enriched malonylated motifs around the malonylation sites. (<b>B</b>) A heat map of the amino acid composition around the malonylation site shows the frequencies of different types of amino acids around the residue. Red indicates enrichment and green indicates depletion.</p>
Full article ">Figure 8
<p>Heat map of differential modification sites. Each row is a differentially modified site, and each column is a sample. Red represents high expression, blue represents low expression, and gray represents unquantifiable in the corresponding sample.</p>
Full article ">Figure 9
<p>Heat map of differential proteins. (<b>A</b>) GO classification of differentially expressed proteins in different groups of different Q groups by cluster analysis. (<b>B</b>) KEGG pathways of differentially expressed proteins after cluster analysis. (<b>C</b>) Protein domains of differentially expressed proteins after cluster analysis. The color blocks corresponding to the functional descriptions of different Q groups and differentially expressed modified protein enrichment indicate the significance of enrichment. Blue indicates high significance, blue and white indicate low significance. *, **, and *** indicate <span class="html-italic">p</span> &lt; 0.05, <span class="html-italic">p</span> &lt; 0.01, and <span class="html-italic">p</span> &lt; 0.0001, respectively.</p>
Full article ">
20 pages, 4383 KiB  
Article
Adsorption of Serum Fetuin onto Octacalcium Phosphate and Its Relation to Osteogenic Property
by Yuki Tsuboi, Ryo Hamai, Kyosuke Okuyama, Kaori Tsuchiya, Yukari Shiwaku, Kensuke Yamauchi and Osamu Suzuki
Int. J. Mol. Sci. 2025, 26(3), 1391; https://doi.org/10.3390/ijms26031391 - 6 Feb 2025
Viewed by 436
Abstract
This study aimed to investigate how the chemical elements in relation to octacalcium phosphate (OCP) hydrolysis affect the osteoblastic differentiation in the presence of serum fetuin. The adsorption of fetuin onto OCP was examined in buffers having different degrees of supersaturation (DS) with [...] Read more.
This study aimed to investigate how the chemical elements in relation to octacalcium phosphate (OCP) hydrolysis affect the osteoblastic differentiation in the presence of serum fetuin. The adsorption of fetuin onto OCP was examined in buffers having different degrees of supersaturation (DS) with respect to OCP and hydroxyapatite (HA) at pH 7.4 and 37 °C. The osteoblastic differentiation of mesenchymal stem cells (MSCs) was evaluated in cultures with OCP and 0 to 0.8 mg/mL of fetuin. The amount of fetuin adsorbed increased with increasing DS in the buffer. In the MSC culture, the coexistence of OCP and 0.2–0.4 mg/mL of fetuin close to serum level increased alkaline phosphatase activity; however, the activity was suppressed by 0.2–0.8 mg/mL of fetuin. Transmission electron microscopy revealed de novo crystal formation on OCP in supersaturated buffer and culture media with respect to OCP and HA at lower fetuin concentrations. Infrared spectroscopy and DS estimation indicate that the hydrolysis of OCP with de novo apatite formation was promoted in the culture media at 0.2–0.4 mg/mL of fetuin. These results suggest that OCP may promote osteoblastic differentiation if the suitable conditions are attained regarding the chemical elements and fetuin adsorption around OCP. Full article
Show Figures

Figure 1

Figure 1
<p>(<b>A</b>) Adsorption isotherms of fetuin onto OCP plotted as adsorption amount per unit surface area of OCP as a function of fetuin equilibrium concentration in the Tris-HCl buffers––0.5 mM Ca<sup>2+</sup> with 0.5 mM Pi ions (Ca0.5Pi0.5) and 3.0 mM Ca<sup>2+</sup> with 1.0 mM Pi ions (Ca3.0Pi1.0) at 37 °C. (<b>B</b>) CD spectra. (<b>C</b>) Secondary structural elements of fetuin in the Ca0.5Pi0.5 and Ca3.0Pi1.0 buffers in the presence of OCP with 0.25 mg/mL of fetuin.</p>
Full article ">Figure 2
<p>(<b>A</b>) FTIR spectra of OCP over the range of 400–1800 cm<sup>−1</sup> and (<b>B</b>) 1300–1800 cm<sup>−1</sup> before and after the incubations in the Ca0.5Pi0.5 and Ca3.0Pi1.0 buffers containing 0.25 and 0.75 mg/mL of fetuin. Arrows indicate amide II derived from fetuin.</p>
Full article ">Figure 3
<p>(<b>A</b>) Lower magnified TEM images and (<b>B</b>) SAED patterns of OCP before (original) and after the incubations in the Ca3.0Pi1.0 and Ca0.5Pi0.5 Tris-HCl buffers with 0, 0.25, and 1.0 mg/mL of fetuin. The SAED patterns indicate the reflections along the [100] zone axis of OCP. Bars in the TEM images and SAED patterns represent 200 nm and 10 nm<sup>–1</sup>, respectively.</p>
Full article ">Figure 4
<p>Higher magnified TEM images of OCP before (original) and after the incubations in the Ca3.0Pi1.0 and Ca0.5Pi0.5 Tris-HCl buffers with 0, 0.25, and 1.0 mg/mL of fetuin. Bars in the TEM images represent 50 nm. Open arrows indicate the de novo depositions formed on the edge of plate-like OCP particles.</p>
Full article ">Figure 5
<p>(<b>A</b>) DNA concentrations and (<b>B</b>) ALP activities of D1 cells in the presence and absence of OCP granules at 0, 0.2, 0.4, and 0.8 mg/mL of fetuin after the cultivations for 4, 7, and 14 days. ** <span class="html-italic">p</span> &lt; 0.01, * <span class="html-italic">p</span> &lt; 0.05. <sup>a</sup> <span class="html-italic">p</span> &lt; 0.05 indicates a significant difference between day 4 and day 7. <sup>b</sup> <span class="html-italic">p</span> &lt; 0.05 indicates a significant difference between day 4 and day 14. <sup>c</sup> <span class="html-italic">p</span> &lt; 0.05 indicates a significant difference between day 7 and day 14.</p>
Full article ">Figure 6
<p>(<b>A</b>) TEM images and (<b>B</b>) FTIR spectra of OCP before and after the cultivations of D1 cells in the presence of fetuin at day 7 (<b>A</b>,<b>B</b>) and day 14 (<b>A</b>). Bars in the TEM images represent 50 nm. Open arrows indicate the de novo depositions formed on the <span class="html-italic">a</span>-plane of OCP crystals.</p>
Full article ">Figure 7
<p>Flowchart of adsorption experiments. The experiments were designed to examine how the adsorption behavior of OCP is regulated by the surrounding chemical environment under physiological conditions. Adsorption isotherms were measured by the incubations of OCP granules (&lt;53 μm) in the fetuin-containing Tris-HCl buffer saturated (0.5 mM Ca<sup>2+</sup> and 0.5 mM Pi) and supersaturated (3.0 mM Ca<sup>2+</sup> and 1.0 mM Pi) with respect to OCP. The adsorption isotherms were approximated using the Langmuir model to estimate the adsorption affinity of fetuin for OCP. The conformation changes in fetuin were examined in the physiological conditions in the presence of OCP by measurements of CD spectra. The structural and morphological changes in OCP during the fetuin adsorption were analyzed by FTIR and SAED. The morphological changes in OCP after the incubations were observed using TEM.</p>
Full article ">Figure 8
<p>Flowchart of cell culture experiments. The experiments were designed to examine the relationship between OCP hydrolysis in the presence of fetuin and the osteoblastic differentiation of MSCs. D1 cells were cultured in the osteogenic media containing fetuin (0, 0.2, 0.4, and 0.8 mg/mL) in the absence of and presence of OCP granules (300–500 μm). At days 4, 7, and 14 of incubations, the DNA concentration and ALP activity of cells were determined to analyze the proliferation and osteoblastic differentiation of cells. The incubated OCP granules were characterized by FTIR to evaluate the progress of OCP hydrolysis in the presence of fetuin and cells. The formation of de novo crystals during the hydrolysis was observed using TEM. The concentrations of ions and pH in the supernatants of culture media were measured. The solubility of calcium phosphates in the media was estimated by calculating DS with respect to HA, OCP, and DCPD using the analytical values of ion concentrations and pH.</p>
Full article ">
17 pages, 3816 KiB  
Article
SMURF1-Induced Ubiquitination of FTH1 Disrupts Iron Homeostasis and Suppresses Myogenesis
by Xia Xiong, Wen Li, Chunlin Yu, Mohan Qiu, Zengrong Zhang, Chenming Hu, Shiliang Zhu, Li Yang, Han Pen, Xiaoyan Song, Jialei Chen, Bo Xia, Shunshun Han and Chaowu Yang
Int. J. Mol. Sci. 2025, 26(3), 1390; https://doi.org/10.3390/ijms26031390 - 6 Feb 2025
Viewed by 413
Abstract
Ferritin heavy chain 1 (FTH1) is pivotal in the storage, release, and utilization of iron, plays a crucial role in the ferroptosis pathway, and exerts significant impacts on various diseases. Iron influences skeletal muscle development and health by promoting cell growth, ensuring energy [...] Read more.
Ferritin heavy chain 1 (FTH1) is pivotal in the storage, release, and utilization of iron, plays a crucial role in the ferroptosis pathway, and exerts significant impacts on various diseases. Iron influences skeletal muscle development and health by promoting cell growth, ensuring energy metabolism and ATP synthesis, maintaining oxygen supply, and facilitating protein synthesis. However, the precise molecular mechanisms underlying iron’s regulation of skeletal muscle growth and development remain elusive. In this study, we demonstrated that the conditional knockout (cKO) of FTH1 in skeletal muscle results in muscle atrophy and impaired exercise endurance. In vitro studies using FTH1 cKO myoblasts revealed notable decreases in GSH concentrations, elevated levels of lipid peroxidation, and the substantial accumulation of Fe2+, collectively implying the induction of ferroptosis. Mechanistically, E3 ubiquitin-protein ligase SMURF1 (SMURF1) acts as an E3 ubiquitin ligase for FTH1, thereby facilitating the ubiquitination and subsequent degradation of FTH1. Consequently, this activation of the ferroptosis pathway by SMURF1 impedes myoblast differentiation into myotubes. This study identifies FTH1 as a novel regulator of muscle cell differentiation and skeletal muscle development, implicating its potential significance in maintaining skeletal muscle health through the regulation of iron homeostasis. Full article
(This article belongs to the Special Issue Molecular Mechanisms of Skeletal Muscle Metabolism)
Show Figures

Figure 1

Figure 1
<p>Morphological and functional alterations in FTH1 knockout mice: (<b>A</b>) Representative images of gastrocnemius (GAS), soleus (SOL), extensor digitorum longus (EDL), and tibialis anterior (TA) muscles from control and <span class="html-italic">FTH1</span> conditional knockout (cKO) mice. (<b>B</b>) Histological examination of GAS muscle sections stained with hematoxylin and eosin. The mean cross-sectional area of muscle fibers is presented on the right. (<b>C</b>) Immunohistochemical staining of GAS muscle fibers from mice using an anti-dystrophin antibody. The relative mean area is displayed on the right. The green is the dystrophin protein, and the blue is the nucleus. (<b>D</b>) Size distribution and the average diameter of muscle fibers in the gastrocnemius muscle of <span class="html-italic">FTH1</span> cKO mice compared to control mice. (<b>E</b>) Force measurements obtained during the tetanic contraction of GAS muscles from control and <span class="html-italic">FTH1</span> cKO mice. (<b>F</b>) Assessment of muscle endurance through forced treadmill running to exhaustion. Data are reported as means ± standard error of the mean (s.e.m.). * indicates <span class="html-italic">p</span> &lt; 0.05; ** indicates <span class="html-italic">p</span> &lt; 0.01.</p>
Full article ">Figure 2
<p>Impact of FTH1 on ferroptosis in mouse myoblast cells: (<b>A</b>) Gene Ontology (GO) enrichment analysis conducted to identify significantly differentially expressed genes between <span class="html-italic">FTH1</span> cKO myoblasts and their control counterparts. (<b>B</b>) Pathway enrichment analysis performed to elucidate the pathways associated with the significantly differentially expressed genes observed between <span class="html-italic">FTH1</span> cKO myoblasts and controls. (<b>C</b>) Western blot analysis employed to assess the expression levels of ferroptosis-related proteins ACSL4 and COX2 in FTH1 cKO myoblasts compared to controls. (<b>D</b>) qPCR utilized to measure the mRNA expression levels of ACSL4 and COX2 in <span class="html-italic">FTH1</span> cKO myoblasts relative to control cells. (<b>E</b>–<b>H</b>) The concentrations of GSH, MDA, and Fe<sup>2+</sup> quantified and compared between <span class="html-italic">FTH1</span> cKO myoblasts and control myoblasts to determine their relative levels. Data are reported as means ± standard error of the mean (s.e.m.). * indicates <span class="html-italic">p</span> &lt; 0.05; ** indicates <span class="html-italic">p</span> &lt; 0.01.</p>
Full article ">Figure 3
<p>FTH1 regulates myoblast differentiation into myotubes via the ferroptosis pathway: (<b>A</b>) qPCR analysis of MyoG and MyoD mRNA expression in FTH1 cKO myoblasts relative to control cells. (<b>B</b>) Western blot assessment of MyoD protein levels in FTH1 cKO myoblasts compared to controls. (<b>C</b>) MyHC immunofluorescence staining to determine the myotube fusion index in FTH1 cKO myoblasts versus controls. Red: MyHC, Green, nucleus. (<b>D</b>–<b>G</b>) Quantification of GSH, MDA, and Fe<sup>2+</sup> concentrations in FTH1 cKO myoblasts and control myoblasts, with and without DFO treatment. (<b>H</b>) qPCR analysis of MyoG and MyoD mRNA expression in FTH1 cKO myoblasts relative to controls, with and without DFO treatment. (<b>I</b>) MyHC immunofluorescence staining to assess the myotube fusion index in FTH1 cKO myoblasts compared to controls, with and without DFO treatment. Red: MyHC, Green, nucleus. Data are reported as means ± standard error of the mean (s.e.m.). * indicates <span class="html-italic">p</span> &lt; 0.05; ** indicates <span class="html-italic">p</span> &lt; 0.01, ns: not significant.</p>
Full article ">Figure 4
<p>SMURF1 identified as the E3 ligase for FTH1: (<b>A</b>,<b>B</b>) Prediction of SMURF1 as the specific E3 ligase for FTH1 using the UbiBrowser database. (<b>C</b>) Confocal microscopy images showing the colocalization of SMURF1 (red) and FTH1 (green). (<b>D</b>) Reciprocal co-immunoprecipitation analysis revealing the interaction between SMURF1 and FTH1 in primary myoblasts cultured for 4 days in a differentiation medium. Atrogin1 used as native control, GAPDH used as loading control. (<b>E</b>) Validation of the interaction by reciprocal co-immunoprecipitation between GFP-tagged SMURF1 and Flag-tagged FTH1 in HEK293T cells. Data are reported as means ± standard error of the mean (s.e.m.).</p>
Full article ">Figure 5
<p>SMURF1 promotes the degradation of FTH1: (<b>A</b>,<b>B</b>) The wild type SMURF1 (<b>A</b>) decreases the FTH1 protein level, but the C276S mutant (<b>C</b>) cannot downregulate FTH1. Cells expressing Flag-SMURF1 or Flag-SMURF1C276S treated with cycloheximide (CHX, 200 μg/mL). (<b>C</b>) SMURF1 is critical to the stability of FTH1 before or after erastin treatment. Cells expressing indicated shRNA constructs were treated with cycloheximide (CHX, 200 μg/mL) and erastin (0–12 h, 5 μM). Data are reported as means ± standard error of the mean (s.e.m.). ** indicates <span class="html-italic">p</span> &lt; 0.01. ns: not significant.</p>
Full article ">Figure 6
<p>SMURF1 promotes FTH1 ubiquitination and degradation: (<b>A</b>) Representative Western blot analysis and quantification of FTH1 protein levels in cells treated with Fer-1 (5 µM) for 12 h. (<b>B</b>) qPCR analysis of FTH1 mRNA expression in cells treated with Fer-1 (5 µM) for 12 h. (<b>C</b>) The immunoprecipitation (IP) of lysates from HEK 293T cells transfected with either control (Ctrl) or SMURF1-shRNA and pretreated with MG132 prior to collection, followed by detection with the indicated antibodies. (<b>D</b>) The IP of lysates from HEK 293T cells transfected with either Vector or SMURF1 overexpression constructs and pretreated with MG132 prior to collection, followed by detection with the indicated antibodies. Data are reported as means ± standard error of the mean (s.e.m.). ** indicates <span class="html-italic">p</span> &lt; 0.01.</p>
Full article ">Figure 7
<p>SMURF1 targets FTH1 to facilitate ferroptosis and impede skeletal muscle development: (<b>A</b>) Overview of the experimental groups and treatments employed in this study phase. (<b>B</b>) Western blot analysis depicting the protein levels of ACSL4 and COX2 in the specified cell treatments. (<b>C</b>–<b>F</b>) Quantification of GSH, MDA, GSSG, and Fe<sup>2+</sup> concentrations in the indicated treated cells. (<b>G</b>) Immunofluorescence staining for MyHC in the specified cell treatments. Red: MyHC, Green, nucleus. (<b>H</b>,<b>I</b>) Western blot analysis of MyoG and MyoD protein expression in the indicated treated cells. (<b>I</b>) qRT-PCR analysis of MyoG and MyoD mRNA expression in the indicated treated cells. * indicates <span class="html-italic">p</span> &lt; 0.05; ** indicates <span class="html-italic">p</span> &lt; 0.01.</p>
Full article ">
18 pages, 3512 KiB  
Review
Cancer Stem Cells and the Renin–Angiotensin System in the Tumor Microenvironment of Melanoma: Implications on Current Therapies
by Ethan J. Kilmister and Swee T. Tan
Int. J. Mol. Sci. 2025, 26(3), 1389; https://doi.org/10.3390/ijms26031389 - 6 Feb 2025
Viewed by 527
Abstract
Multiple signaling pathways are dysregulated in melanoma, notably the Ras/RAF/MAPK/ERK and PI3K/AKT/mTOR pathways, which can be targeted therapeutically. The high immunogenicity of melanoma has been exploited using checkpoint inhibitors. Whilst targeted therapies and immune checkpoint inhibitors have improved the survival of patients with [...] Read more.
Multiple signaling pathways are dysregulated in melanoma, notably the Ras/RAF/MAPK/ERK and PI3K/AKT/mTOR pathways, which can be targeted therapeutically. The high immunogenicity of melanoma has been exploited using checkpoint inhibitors. Whilst targeted therapies and immune checkpoint inhibitors have improved the survival of patients with advanced melanoma, treatment resistance, their side effect profiles, and the prohibitive cost remain a challenge, and the survival outcomes remain suboptimal. Treatment resistance has been attributed to the presence of cancer stem cells (CSCs), a small subpopulation of pluripotent, highly tumorigenic cells proposed to drive cancer progression, recurrence, metastasis, and treatment resistance. CSCs reside within the tumor microenvironment (TME) regulated by the immune system, and the paracrine renin–angiotensin system, which is expressed in many cancer types, including melanoma. This narrative review discusses the role of CSCs and the paracrine renin–angiotensin system in the melanoma TME, and its implications on the current treatment of advanced melanoma with targeted therapy and immune checkpoint blockers. It also highlights the regulation of the Ras/RAF/MAPK/ERK and PI3K/AKT/mTOR pathways by the renin–angiotensin system via pro-renin receptors, and how this may relate to CSCs and treatment resistance, underscoring the potential for improving the efficacy of targeted therapy and immunotherapy by concurrently modulating the renin–angiotensin system. Full article
(This article belongs to the Special Issue Melanoma: Molecular Mechanisms and Therapy)
Show Figures

Figure 1

Figure 1
<p>(<b>A</b>) The hierarchical model of cancer proposes the presence of a highly tumorigenic cancer stem cell (CSC) sitting atop the tumor cellular hierarchy, giving rise to identical CSCs and differentiated cancer cells through asymmetric cell division, resulting in heterogenous tumor cell populations. (<b>B</b>) The stochastic model of cancer proposes that a normal somatic cell accumulates oncogenic mutations in a stepwise manner and becomes a cancer cell that undergoes clonal expansion to form a tumor. Figure modified and reproduced with permission from <span class="html-italic">Biomedicines</span> [<a href="#B13-ijms-26-01389" class="html-bibr">13</a>].</p>
Full article ">Figure 2
<p>A schema showing the paracrine renin–angiotensin system (see text). PRR, pro-renin receptor; Cath G, cathepsin G; Cath B, cathepsin B; Cath D, cathepsin D; ACE1, angiotensin-converting enzyme 1; ACE2, angiotensin-converting enzyme 2; AGT, angiotensinogen; Ang(1–7), angiotensin (1–7); Ang(1–9), angiotensin (1–9); AP-A, aminopeptidase-A; AP-N, aminopeptidase-N; ATI, angiotensin I; ATII, angiotensin II; ATIII, angiotensin III; ATIV, angiotensin IV; AT<sub>1</sub>R, angiotensin II receptor 1; AT<sub>2</sub>R, angiotensin II receptor 2; AT<sub>4</sub>R, angiotensin II receptor 4; MrgD, Mas-related-G protein coupled receptor; MasR, Mas receptor. Reproduced and adapted with permission from <span class="html-italic">Biomedicines</span> [<a href="#B13-ijms-26-01389" class="html-bibr">13</a>].</p>
Full article ">Figure 3
<p>A proposed model demonstrating the role of the paracrine renin–angiotensin system (RAS) in the cancer stem cell (CSC) niche. A CSC (with the cytoplasm in light blue and the nucleus in purple) residing within the tumor microenvironment (TME). Angiotensin II (ATII) activates ATII receptor 1 (AT<sub>1</sub>R), promoting the hallmarks of cancer, including creating an inflammatory state through multiple mechanisms. AT<sub>1</sub>R activates phosphatidylinositol signaling, which increases cytosolic Ca<sup>2+</sup> to promote mitogenesis. Hypoxia increases paracrine RAS activity by upregulating angiotensin-converting enzyme (ACE) and the expression of hypoxia-inducible factor 1α (HIF-1α) and HIF-2α, which increase tumor progression and treatment resistance. HIF-1α, HIF-2α, and hypoxia increase the expression of vascular endothelial growth factor (VEGF) which increases angiogenesis. The binding of AT<sub>1</sub>R to C-X-C chemokine receptor type 4 (CXCR4) promotes tumor cell migration and invasion, leading to metastatic progression. AT<sub>1</sub>R, via MAPK-STAT3 signaling, contributes to a cytokine release that leads to CSC renewal. AT<sub>1</sub>R signaling also contributes to the migration of fibroblasts in an epidermal growth factor receptor (EGFR)-dependent fashion. AT<sub>1</sub>R signaling and the pro-renin receptor, which act in a feedback loop with Wnt/β-catenin, increase Wnt signaling, which promotes CSC stemness by upregulating stemness-associated markers. Myeloid-derived suppressor cells (MDSCs) promote CSC characteristics by increasing microRNA-101 expression that induces expression of stemness-related genes in CSCs. Under the influence of the TME, the polarization of tumor-associated macrophages (TAMs)—immune cells that are located within the TME—changes from the M1 to the M2 phenotype. M2 TAMs induce the proliferation of CSCs via interleukin 6 (IL-6)-induced activation of STAT3, leading to cytokine release and positive feedback contributing to CSC renewal. Abbreviations: ATI, angiotensin I; MAPK, mitogen-activated protein kinase. Reproduced from <span class="html-italic">Journal of Histochemistry and Cytochemistry</span> [<a href="#B31-ijms-26-01389" class="html-bibr">31</a>].</p>
Full article ">Figure 4
<p>A schema showing the effect of the paracrine renin–angiotensin system and its convergent signaling pathways on the tumor microenvironment, and its influence on cellular proliferation, invasiveness, and cell survival in cancer development. The renin–angiotensin system interacts with downstream pathways, such as the Ras/RAF/MEK/ERK pathway (light purple) and the PI3K/AKT/mTOR pathway (light green), and the upstream Wnt/β-catenin pathway that influences cellular proliferation, migration, inhibition of apoptosis, migration, and invasion. It also influences gene expression of components of the renin–angiotensin system (see text). PRR, pro-renin receptor; LRP6, low-density lipoprotein receptor-related protein; Fzd, frizzled receptor; Cath G, cathepsin G; Cath B, cathepsin B; Cath D, cathepsin D; ACE1, angiotensin-converting enzyme 1; ACE2, angiotensin-converting enzyme 2; ADP, adenosine diphosphate; AGT, angiotensinogen; ATP, adenosine triphosphate; Ang(1–7), angiotensin (1–7); Ang(1–9), angiotensin (1–9); AP-A, aminopeptidase-A; NEP, neutral endopeptidase; AP-N, aminopeptidase-N; ATI, angiotensin I; ATII, angiotensin II; ATIII, angiotensin III; ATIV, angiotensin IV; AT<sub>1</sub>R, angiotensin II receptor 1; AT<sub>2</sub>R, angiotensin II receptor 2; AT<sub>4</sub>R, angiotensin II receptor 4; MrgD, Mas-related-G protein coupled receptor; MasR, Mas receptor; mTOR, mammalian target of rapamycin; NF-κB, nuclear factor kappa B; TGF-β1, transforming growth factor-β1; V-ATPase, vacuolar H+-adenosine triphosphate. Reproduced and adapted with permission from <span class="html-italic">Biomedicines</span> [<a href="#B13-ijms-26-01389" class="html-bibr">13</a>].</p>
Full article ">Figure 5
<p>A schema showing how the Ras/RAF/MAPK/ERK and PI3K/AKT/mTOR pathways can be inhibited in melanoma. BRAF can be targeted by BRAF inhibitors, and MEK by MEK inhibitors. Pro-renin receptor (PRR) and angiotensin II receptor 1 (AT<sub>2</sub>R), which activate Ras, can be inhibited through inhibition of the paracrine renin–angiotensin system (RAS). PRR also increases the production of reactive oxygen species (ROS), which contributes to the overactivation of the Ras/RAF/MAPK/ERK and PI3K/AKT/mTOR pathways. PRR also acts on vacuolar ATPase and Wnt/β-catenin signaling to facilitate cancer progression. Angiotensin II (ATII) induces phosphorylation of PI3K/AKT, driving tumor progression. mTORC1 signaling can be inhibited by rapamycin (sirolimus) and metformin. Figure created using Biorender.</p>
Full article ">Figure 6
<p>The renin–angiotensin system and its bypass loops and converging signaling pathways can be targeted at different points. The renin–angiotensin system (black) regulates blood pressure, stem cell differentiation, and tumor development. Bypass loops in the system involving cathepsins B, D, and G, and chymase (green) provide redundancy. Multiple points of the pathway can be targeted by specific inhibitors (red). ACE, angiotensin-converting enzyme; ARBs, AT<sub>1</sub>R blockers. Adapted with permission from <span class="html-italic">Frontiers in Oncology</span> [<a href="#B93-ijms-26-01389" class="html-bibr">93</a>]. Diagram recreated with BioRender.com, accessed on 13 June 2024.</p>
Full article ">
10 pages, 1174 KiB  
Article
Bisabolane Sesquiterpenes with Anti-Chlamydial Activity Isolated from Ligularia narynensis
by Na Gao, Yi-Lin He, Hui-Ming Qi, Hong-Ying Yang, Guo-Li Li, Zhao-Cai Li and Tong Shen
Int. J. Mol. Sci. 2025, 26(3), 1388; https://doi.org/10.3390/ijms26031388 - 6 Feb 2025
Viewed by 357
Abstract
Chlamydia are obligate intracellular bacterial pathogens affecting humans and animals, causing miscarriage, stillbirth, or weak fetuses in the late stages of pregnancy of goats and sheep. Because there is no commercial vaccine for chlamydia in animals, drug treatment has become the most effective [...] Read more.
Chlamydia are obligate intracellular bacterial pathogens affecting humans and animals, causing miscarriage, stillbirth, or weak fetuses in the late stages of pregnancy of goats and sheep. Because there is no commercial vaccine for chlamydia in animals, drug treatment has become the most effective curative method. Natural products, also known as secondary metabolites, are becoming one of the main sources used in new drug development because of their structural diversity and biodiversity. In natural products, plant sources play a major role in the development process of new drugs. In this study, five undescribed highly oxygenated bisabolane sesquiterpenes (Pararubin W, Pararubin X, Pararubin Y., Pararubin Z, and Pararubin AA) were isolated from whole plants of Ligularia narynensis. Their chemical structures were determined via analyses of HRESIMS, IR, 1D, and 2D NMR data, along with the assignment of their relative configurations. These compounds were tested for their anti-chlamydial activity. The results show that compounds 1 and 5 inhibited the growth of Chlamydia abortus in host cells in a dose-dependent manner. Full article
Show Figures

Figure 1

Figure 1
<p>Structures of compounds <b>1</b>–<b>5</b>.</p>
Full article ">Figure 2
<p><sup>1</sup>H−<sup>1</sup>H COSY (blue bold) and key HMBC (red arrows) correlations of compounds <b>1</b>–<b>5</b>.</p>
Full article ">Figure 3
<p>NOESY (red double arrow) correlations of compound <b>1</b>.</p>
Full article ">Figure 4
<p>Anti-chlamydial effects of compounds <b>1</b>–<b>5</b>. Various concentrations of the compounds were applied to the <span class="html-italic">C. abortus</span> strain GN6 cultured in McCoy cells. Anti-chlamydial activity was represented by the inclusion formation ratio, based on immunofluorescent staining of <span class="html-italic">C. abortus</span> inclusions. (<b>A</b>) Observation of <span class="html-italic">C. abortus</span> inclusions in cell cultures under treatment with tetracycline (5 μM) as a positive control or the tested compounds at the concentration of 100 μg/mL. (<b>B</b>) <span class="html-italic">C. abortus</span> inclusion formation ratio in cell cultures treated with the tested compounds at 100 μg/mL final concentrations. (<b>C</b>) Treatment with compound <b>1</b> reduces the inclusion formation ratio of <span class="html-italic">C. abortus</span> in cell cultures. (<b>D</b>) Treatment with compound <b>5</b> reduces the inclusion formation ratio of <span class="html-italic">C. abortus</span> in the cell cultures. These data are the mean ± SD and representative of three independent experiments.</p>
Full article ">
17 pages, 9073 KiB  
Article
Genetic Diversity and Environmental Adaptation Signatures of the Great Seahorse (Hippocampus kelloggi) in the Coastal Regions of the Indo-Pacific as Revealed by Whole-Genome Re-Sequencing
by Wen-Xin Hao, Ying-Yi Zhang, Xin Wang, Meng Qu, Shi-Ming Wan and Qiang Lin
Int. J. Mol. Sci. 2025, 26(3), 1387; https://doi.org/10.3390/ijms26031387 - 6 Feb 2025
Viewed by 403
Abstract
The great seahorse (Hippocampus kelloggi) is one of the larger species within the seahorse group and is widely distributed in coastal areas of the Indo-Pacific. However, the natural resources of this species continue to decrease, rendering it a vulnerable species that [...] Read more.
The great seahorse (Hippocampus kelloggi) is one of the larger species within the seahorse group and is widely distributed in coastal areas of the Indo-Pacific. However, the natural resources of this species continue to decrease, rendering it a vulnerable species that faces a high risk of extinction. Therefore, there is an urgent need to conduct research on the genetic diversity of this species to protect its genetic resources. In this study, we conducted whole-genome re-sequencing (WGRS) on three H. kelloggi populations from the Red Sea (RS, n = 30), the Andaman Sea (AS, n = 13), and the South China Sea (SCS, n = 13), and a total of 1,398,936 high-quality single-nucleotide polymorphisms (SNPs) were identified. The results indicate that the average observed heterozygosity (Ho) and the average expected heterozygosity (He) for the RS, AS, and SCS populations are 0.2031 and 0.1987, 0.1914 and 0.1822, and 0.2083 and 0.2001, respectively. The three geographic populations exhibit a high degree of genetic differentiation with only a minimal gene flow between them. Consistently, in a population structure analysis, the three groups are also clearly distinguished, which is consistent with the results of the population differentiation coefficient. Demographic analyses revealed that the effective population size (Ne) of the SCS population underwent a dramatic bottleneck during the Last Glacial Maximum (LGM), followed by a substantial recovery, whereas the RS and AS populations maintained stable Ne values throughout this period. To investigate adaptive responses to climate change in the SCS population, we employed selective elimination analysis, which identified 21 candidate genes potentially involved in environmental adaptation. Of particular significance were myo5a, hps4, znf385a, msh3, and pfkfb4, which likely play crucial roles in the adaptive mechanisms of H. kelloggi. This comprehensive study not only illuminates the genetic diversity patterns of H. kelloggi but also provides a valuable foundation for future investigations into the species’ evolutionary adaptations. Full article
(This article belongs to the Section Molecular Genetics and Genomics)
Show Figures

Figure 1

Figure 1
<p>The genetic information of <span class="html-italic">Hippocampus kelloggi</span>. The identification of high-quality SNPs and their distribution across 21 chromosomes of <span class="html-italic">H. kelloggi</span>. The gradient colors from green to red indicate an increase in SNP density within 0.1 Mb interval.</p>
Full article ">Figure 2
<p>Population structure analyses of <span class="html-italic">H. kelloggi</span>. (<b>A</b>) A phylogenetic tree of the three analyzed populations based on genotype data. The blue, red and yellow backgrounds represent individuals in the SCS, AS, RS population, respectively. (<b>B</b>) A population structure map for K = 2~5. (<b>C</b>) Population structure revealed by PCA. The blue, red and yellow dots represent individuals in the SCS, AS and RS populations, respectively. (<b>D</b>) The decay of linkage disequilibrium in the three experimental populations. The <span class="html-italic">X</span>-axis represents physical location. The <span class="html-italic">Y</span>-axis represents the <span class="html-italic">LD</span> value (<span class="html-italic">r</span><sup>2</sup>). The blue, red and yellow line represent the SCS, AS and RS populations, respectively. (<b>E</b>) The error rate of the cross validation (CV) for K = 1~10 (K value represents the number of subgroups of the population).</p>
Full article ">Figure 3
<p>Demographic history of the three <span class="html-italic">H. kelloggi</span> populations in this study. Pink, green and blue dots represent individuals in the SCS, AS, RS populationd, respectively. The orange background represents the period of the Last Glacial Maximum (LGM).</p>
Full article ">Figure 4
<p>Candidate gene exploration and enrichment analysis of <span class="html-italic">H. kelloggi</span> from the SCS population (SCS vs. RS). (<b>A</b>) A plot of the moving average <span class="html-italic">F<sub>ST</sub></span> (SCS vs. RS) values of the SNPs per chromosome. The blue line indicates the significant threshold for identifying putative selection regions (top 5 <span class="html-italic">F<sub>ST</sub></span> = 0.479, <span class="html-italic">p</span>-value &lt; 0.05). (<b>B</b>) Distribution of the <span class="html-italic">π</span>-<span class="html-italic">ratio</span> (SCS/RS) on 21 chromosomes. The blue line indicates the significant threshold for identifying putative selection regions (top 5 log2(<span class="html-italic">π</span>-<span class="html-italic">ratio</span> SCS/RS) = 2.265, <span class="html-italic">p</span>-value &lt; 0.05). (<b>C</b>) The distribution of the log2 (<span class="html-italic">π</span>-<span class="html-italic">ratio</span>) and <span class="html-italic">F<sub>ST</sub></span>. The RS population is the control group and the SCS population is the selection group. (<b>D</b>,<b>E</b>) Results of the GO and KEGG enrichment analysis of selected genes in the SCS population.</p>
Full article ">Figure 5
<p>Candidate gene exploration and enrichment analysis of <span class="html-italic">H. kelloggi</span> from the SCS population (SCS vs. AS). (<b>A</b>) A plot of the moving average <span class="html-italic">F<sub>ST</sub></span> (SCS vs. AS) values of SNPs per chromosome. The blue line indicates the significant threshold for identifying putative selection regions (top 5 <span class="html-italic">F<sub>ST</sub></span> = 0.525, <span class="html-italic">p</span>-value &lt; 0.05). (<b>B</b>) Distribution of the <span class="html-italic">π</span>-<span class="html-italic">ratio</span> (SCS/AS) on 21 chromosomes. The blue line indicates the significant threshold for identifying putative selection regions (top 5 log2(<span class="html-italic">π</span>-<span class="html-italic">ratio</span> SCS/AS) = 1.937, <span class="html-italic">p</span>-value &lt; 0.05). (<b>C</b>) Distribution of the log2 (<span class="html-italic">π</span>-<span class="html-italic">ratio</span>) and <span class="html-italic">F<sub>ST</sub></span>. The AS population is the control group and the SCS population is the selection group. (<b>D</b>,<b>E</b>) Results of the GO and KEGG enrichment analysis of selected genes in the SCS population.</p>
Full article ">Figure 6
<p>Information on the natural distribution of <span class="html-italic">H. kelloggi</span> and sampling sites. The <span class="html-italic">H. kelloggi</span> pattern map indicates their natural distribution density, and the blue, red, and yellow triangles represent the population sampling points for the SCS, AS, RS.</p>
Full article ">
14 pages, 3027 KiB  
Article
Antibiofilm Activities of Halogenated Pyrimidines Against Enterohemorrhagic Escherichia coli O157:H7
by Hyejin Jeon, Yong-Guy Kim, Jin-Hyung Lee and Jintae Lee
Int. J. Mol. Sci. 2025, 26(3), 1386; https://doi.org/10.3390/ijms26031386 - 6 Feb 2025
Viewed by 313
Abstract
Enterohemorrhagic Escherichia coli (EHEC) is a significant public health concern due to its ability to form biofilms, enhancing its resistance to antimicrobials and contributing to its persistence in food processing environments. Traditional antibiotics often fail to target these biofilms effectively, leading to increased [...] Read more.
Enterohemorrhagic Escherichia coli (EHEC) is a significant public health concern due to its ability to form biofilms, enhancing its resistance to antimicrobials and contributing to its persistence in food processing environments. Traditional antibiotics often fail to target these biofilms effectively, leading to increased bacterial resistance. This study aims to explore the efficacy of novel antibiofilm agents, specifically halogenated pyrimidine derivatives, against EHEC. We screened pyrimidine and 31 halogenated pyrimidine derivatives for their antimicrobial and antibiofilm activities against EHEC using biofilm quantification assays, SEM analysis, motility, and curli production assessments. Our findings reveal that certain halogenated pyrimidine derivatives, notably 2-amino-5-bromopyrimidine (2A5BP), 2-amino-4-chloropyrrolo[2,3-d]pyrimidine (2A4CPP), and 2,4-dichloro-5-iodo-7H-pyrrolo[2,3-d]pyrimidine (2,4DC5IPP) at 50 µg/mL, exhibited significant inhibitory effects on EHEC biofilm formation without affecting bacterial growth, suggesting a targeted antibiofilm action. These compounds effectively reduced curli production and EHEC motility, essential factors for biofilm integrity and development. qRT-PCR analysis revealed that two active compounds downregulated the expression of key curli genes (csgA and csgB), leading to reduced bacterial adhesion and biofilm formation. Additionally, in silico ADME–Tox profiles indicated that these compounds exhibit favorable drug-like properties and lower toxicity compared with traditional pyrimidine. This study highlights the potential of halogenated pyrimidine derivatives as effective antibiofilm agents against EHEC, offering a promising strategy for enhancing food safety and controlling EHEC infections. The distinct mechanisms of action of these compounds, particularly in inhibiting biofilm formation and virulence factors without promoting bacterial resistance, underscore their therapeutic potential. Full article
(This article belongs to the Special Issue Mechanisms in Biofilm Formation, Tolerance and Control: 2nd Edition)
Show Figures

Figure 1

Figure 1
<p>Antibiofilm screening of pyrimidine and 31 halogenated pyrimidine derivatives and pyrimidine against EHEC. Biofilm formation (indicated as bars) was quantified after 24 h culture in 96-well plates without shaking in the presence of each compound at 0, 20, or 100 μg/mL. Planktonic cell growth was also measured and indicated as dots. The red-fonted numbers 5, 23, and 31 in the x-axis of the upper panel represent the most active compounds. * <span class="html-italic">p</span> &lt; 0.05 versus non-treated controls. “N” indicates “none”, referring to the absence of any treatment. Full chemical names and structures are shown in <a href="#app1-ijms-26-01386" class="html-app">Table S1</a>.</p>
Full article ">Figure 2
<p>Effects of halogenated pyrimidines on biofilm formation and cell growth: EHEC biofilm formation was quantified with or without the presence of 2A5BP (<b>A</b>), 2A4CPP (<b>B</b>), or (2,4DC5IPP) (<b>C</b>) after 24 h culture in 96-well polystyrene plates. Planktonic cell growths in the presence of 2A5BP (<b>D</b>) and 2A4CPP (<b>E</b>). Bacteriostatic activity of 2A4CPP (<b>F</b>). * <span class="html-italic">p</span> &lt; 0.05 versus non-treated controls.</p>
Full article ">Figure 3
<p>Effects of halogenated pyrimidines on curli production and cell motility: Colony morphology in the presence of 2A5BP and 2A4CPP (<b>A</b>). Curli formation in the presence of 2A5BP and 2A4CPP with Congo red staining (<b>B</b>). Swimming motility in the presence of 2A5BP and 2A4CPP (<b>C</b>,<b>D</b>). Swarming motility (<b>E</b>,<b>F</b>). * <span class="html-italic">p</span> &lt; 0.05 versus non-treated controls. Yellow scale bars represent 0.5 cm (<b>A</b>).</p>
Full article ">Figure 4
<p>SEM analysis of EHEC biofilms and qRT-PCR analysis. SEM images of cells treated with 2A5BP (<b>A</b>) or 2A4CPP (<b>B</b>) and gene expression by 2A5BP and 2A4CPP (<b>C</b>). Red and yellow scale bars represent 5 µm and 2 µm, respectively. * <span class="html-italic">p</span> &lt; 0.05 versus non-treated controls.</p>
Full article ">Figure 5
<p>Toxicity of halogenated pyrimidines in the plant and nematode models: Radish seed germination rate (<b>A</b>), total length of radish seedlings (<b>B</b>), and representative plant images (<b>C</b>). cultured with or without different concentrations of 2A5BP, 2A4CPP, and PP at 25 °C. <span class="html-italic">C. elegans</span> survival was assessed in the presence or absence of 2A5BP (<b>D</b>), 2A4CPP (<b>E</b>), and PP (<b>F</b>) for 10 days. The red scale bar in (<b>C</b>) represents 2 cm.</p>
Full article ">
14 pages, 11016 KiB  
Article
Rubiadin Mediates the Upregulation of Hepatic Hepcidin and Alleviates Iron Overload via BMP6/SMAD1/5/9-Signaling Pathway
by Xueting Xie, Linyue Chang, Xinyue Zhu, Fengbei Gong, Linlin Che, Rujun Zhang, Lixin Wang, Chenyuan Gong, Cheng Fang, Chao Yao, Dan Hu, Weimin Zhao, Yufu Zhou and Shiguo Zhu
Int. J. Mol. Sci. 2025, 26(3), 1385; https://doi.org/10.3390/ijms26031385 - 6 Feb 2025
Viewed by 395
Abstract
Iron overload disease is characterized by the excessive accumulation of iron in the body. To better alleviate iron overload, there is an urgent need for safe and effective small molecule compounds. Rubiadin, the active ingredient derived from the Chinese herb Prismatomeris tetrandra, possesses [...] Read more.
Iron overload disease is characterized by the excessive accumulation of iron in the body. To better alleviate iron overload, there is an urgent need for safe and effective small molecule compounds. Rubiadin, the active ingredient derived from the Chinese herb Prismatomeris tetrandra, possesses notable anti-inflammatory and hepatoprotective properties. Nevertheless, its impact on iron metabolism remains largely unexplored. To determine the role of rubiadin on iron metabolism, Western blot analysis, real-time PCR analysis, and the measurement of serum iron were performed. Herein, we discovered that rubiadin significantly downregulated the expression of transferrin receptor 1, ferroportin 1, and ferritin light chain in ferric-ammonium-citrate-treated or -untreated HepG2 cells. Moreover, intraperitoneal administration of rubiadin remarkably decreased serum iron and duodenal iron content and upregulated expression of hepcidin mRNA in the livers of high-iron-fed mice. Mechanistically, bone morphogenetic protein 6 (BMP6) inhibitor LDN-193189 completely reversed the hepcidin upregulation and suppressor of mother against decapentaplegic 1/5/9 (SMAD1/5/9) phosphorylation induced by rubiadin. These results suggested that rubiadin increased hepcidin expression through the BMP6/SMAD1/5/9-signaling pathway. Collectively, our findings uncover a crucial mechanism through which rubiadin modulates iron metabolism and highlight it as a potential natural compound for alleviating iron-overload-related diseases. Full article
Show Figures

Graphical abstract

Graphical abstract
Full article ">Figure 1
<p>Rubiadin significantly upregulates hepcidin expression in HepG2 cells. (<b>A</b>) Effect of varying concentrations (0 to 40 μM) of rubiadin on the proliferation of HepG2 cells for 24 h. (<b>B</b>) Hepcidin expression was quantified in HepG2 cells that had been treated with the specified concentrations of rubiadin for 24 h. (<b>C</b>) The expression of hepcidin was quantified after treatment of HepG2 cells with 20 μM rubiadin at specific time points. (<b>D</b>) Hepcidin protein expression in HepG2 cells treated with the indicated concentrations of rubiadin for 24 h was evaluated using an immunofluorescence assay. (<b>E</b>) Hepcidin protein expression in HepG2 cells treated for 20 μM rubiadin for the indicated time points was evaluated via the immunofluorescence assay. Scale bars are 50 μm. Significance was calculated using a one-way ANOVA. ns, no significance, * <span class="html-italic">p</span> &lt; 0.05, *** <span class="html-italic">p</span> &lt; 0.001.</p>
Full article ">Figure 2
<p>Rubiadin decreases the protein expression of TfR1, Fpn1 and FtL in FAC treated or untreated HepG2 cells. (<b>A</b>) Western blot analysis of proteins related to iron metabolism in HepG2 cells after rubiadin treatment for 24 h. Quantification of transferrin receptor 1 (TfR1) (<b>B</b>), ferroportin 1 (Fpn1) (<b>C</b>), ferritin light chain (FtL) (<b>D</b>), divalent metal transporter 1 (DMT1) (<b>E</b>), and ferritin heavy chain (FtH) (<b>F</b>) were shown. (<b>G</b>) HepG2 cells were loaded with iron by incubation with ferric ammonium citrate (FAC) (100 μM) for 24 h. After removing the medium, cells were washed and then treated with 20 μM rubiadin for 24 h. Western blot analysis of TfR1, DMT1, Fpn1, FtH, and FtL expression. Quantification of TfR1 (<b>H</b>), Fpn1 (<b>I</b>), FtL (<b>J</b>), DMT1 (<b>K</b>), and FtH (<b>L</b>) were shown. Significance was calculated using a one-way ANOVA. ns, no 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.</p>
Full article ">Figure 3
<p>Rubiadin increased phosphorylation of signal transducer and activator of transcription 3 (STAT3) and suppressor mothers against decapentaplegic 1/5/9 (SMAD1/5/9) in HepG2 cells (<b>A</b>,<b>B</b>) Interleukin 6 (IL-6) and bone morphogenetic protein 6 (BMP6) mRNA expression were measured in HepG2 cells that were treated with the specified concentrations of rubiadin for 24 h. (<b>C</b>–<b>E</b>) IL-6 and BMP6 protein expression were measured in HepG2 cells that were treated with the specified concentrations of rubiadin for 24 h. (<b>F</b>) Phosphorylation of STAT3 and SMAD1/5/9 in HepG2 cells analyzed by Western blot after rubiadin treatment for 24 h. (<b>G</b>,<b>H</b>) Quantification of p-STAT3 (<b>G</b>) and p-SMAD1/5/9 (<b>H</b>) were shown. Significance was calculated using a one-way ANOVA. ns, no 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.</p>
Full article ">Figure 4
<p>Rubiadin remarkably reverses the abnormal elevation of serum iron (SI) and alleviates splenomegaly caused by iron overload. Six-week-old mice with a successful high-iron-diet-induced iron overload model were continuously fed the high-iron diet for 4 weeks and intraperitoneally injected with different doses of rubiadin (5 mg/kg, 20 mg/kg). (<b>A</b>) Daily weight changes were recorded. (<b>B</b>) The size of the spleen in wild-type mice, model group mice, and rubiadin-treated mice was compared. (<b>C</b>) Spleen weight index (mg/g) was calculated. (<b>D</b>) SI and (<b>E</b>) unsaturated iron-binding capacity (UIBC) were measured. (<b>F</b>) Total iron-binding capacity (TIBC) and (<b>G</b>) transferrin saturation (TF%) were calculated. Significance was calculated using the two-way ANOVA. ns, no 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.</p>
Full article ">Figure 5
<p>Rubiadin remarkably reverses the elevation of duodenal iron content and further enhances the hepcidin mRNA and the phosphorylation of STAT3 and SMAD1/5/9 caused by iron overload. The tissue samples were isolated and tested for iron in each organ. (<b>A</b>) Duodenal iron content. (<b>B</b>) Liver iron content. (<b>C</b>) Heart iron content. (<b>D</b>) Spleen iron content. (<b>E</b>) Kidney iron content. (<b>F</b>) <span class="html-italic">Hepcidin</span> mRNA levels in the liver of high-iron-fed mice after rubiadin treatment were measured by RT-qPCR. (<b>G</b>) Liver sections from different groups were immunohistochemically stained for hepcidin. Scale bars are 100 μm. (<b>H</b>) Western blotting assessed the phosphorylation levels of STAT3 and SMAD1/5/9 in the liver of the treated mice. Quantification of p-STAT3 (<b>I</b>) and p-SMAD1/5/9 (<b>J</b>) were shown. Significance was calculated using a two-way ANOVA. ns, no 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.</p>
Full article ">Figure 6
<p>Rubiadin upregulates hepcidin via BMP6/SMAD1/5/9 signaling pathway. (<b>A</b>,<b>B</b>) Hepcidin mRNA was assessed in HepG2 cells treated with 20 μM rubiadin for 24 h, either with or without stattic or LDN-193189. (<b>C</b>) Hepcidin protein expression in HepG2 cells treated for 20 μM rubiadin in the presence or absence of LDN-193189 for 24 h was evaluated via an immunofluorescence assay. Scale bars are 100 μm. (<b>D</b>) Western blot analysis of phosphorylated SMAD1/5/9 in HepG2 cells treated with 20 μM rubiadin for 3 h in the presence or absence of LDN-193189 for 24 h. (<b>E</b>) Quantification of p-SMAD1/5/9 was shown. Significance was calculated using a two-way ANOVA. ns, no 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.</p>
Full article ">
17 pages, 5427 KiB  
Article
The NbCBP1-NbSAMS1 Module Promotes Ethylene Accumulation to Enhance Nicotiana benthamiana Resistance to Phytophthora parasitica Under High Potassium Status
by Sadia Noorin, Youwei Du, Yi Liu, Shuanghong Wang, Yan Wang, Hongchen Jia, Tom Hsiang, Rong Zhang and Guangyu Sun
Int. J. Mol. Sci. 2025, 26(3), 1384; https://doi.org/10.3390/ijms26031384 - 6 Feb 2025
Viewed by 419
Abstract
Potassium (K) fertilization is crucial for plant resistance to pathogens, but the underlying mechanisms remain unclear. Here, we investigate the molecular mechanism by which the addition of K promotes resistance in Nicotiana benthamiana to Phytophthora parasitica. We found that N. benthamiana with [...] Read more.
Potassium (K) fertilization is crucial for plant resistance to pathogens, but the underlying mechanisms remain unclear. Here, we investigate the molecular mechanism by which the addition of K promotes resistance in Nicotiana benthamiana to Phytophthora parasitica. We found that N. benthamiana with high K content (HK, 52.3 g/kg) produced more ethylene in response to P. parasitica infection, compared to N. benthamiana with low-K content (LK, 22.4 g/kg). An exogenous ethylene application effectively increased resistance in LK N. benthamiana to the level under HK status, demonstrating the involvement of ethylene in the HK-associated resistance in N. benthamiana. Further, transcriptome analysis showed that NbSAMS1, encoding ethylene biosynthesis, was induced to upregulate P. parasitica about five times higher in HK than in LK N. benthamiana. NbSAMS1 overexpression enhanced resistance in LK plants, whereas NbSAMS1 silencing reduced resistance in HK plants, confirming its importance in conferring resistance. Furthermore, we identified a calcium-binding protein, NbCBP1, which interacted with NbSAMS1, promoting its expression in HK N. benthamiana. Silencing NbCBP1 compromised resistance in HK N. benthamiana, whereas its overexpression improved resistance in LK N. benthamiana. Notably, NbCBP1 protected NbSAMS1 from degradation by the 26S proteasome, thereby sustaining ethylene accumulation in HK N. benthamiana in response to P. parasitica infection. Thus, our research elucidated some mechanisms of the NbCBP1-NbSAMS1 module associated with disease resistance in HK N. benthamiana. Full article
(This article belongs to the Section Molecular Plant Sciences)
Show Figures

Figure 1

Figure 1
<p>Ethylene is important in enhancing disease resistance. (<b>a</b>) Heatmap displays the expression variation in ethylene-synthesis and associated transcription factors in plants with high potassium content (HK), low potassium (LK), <span class="html-italic">P. parasitica</span>-infected HK (IHK), and <span class="html-italic">P. parasitica</span>-infected LK (ILK) shown in red and green. The scale bar represents the normalized FPKM of each gene. (<b>b</b>) Ethylene production in HK, LK, IHK, and ILK <span class="html-italic">N. benthamiana</span>. (<b>c</b>) Representative images and (<b>d</b>) comparative analysis of lesion diameter in HK plants treated with ethephon (HK-ETH), LK plants treated with ethephon (LK-ETH), HK plants without ethephon treatment (HK-MOCK), or LK plants without ethephon treatment (LK-MOCK). Data are shown as means ± SD and different letters represent significant differences among treatments in lesion diameter at <span class="html-italic">p</span> &lt; 0.05 using Student’s <span class="html-italic">t</span>-tests.</p>
Full article ">Figure 2
<p><span class="html-italic">NbSAMS1</span> plays a positive role in enhancing disease resistance in <span class="html-italic">Nicotiana benthamiana</span>. (<b>a</b>) <span class="html-italic">NbSAMS1</span> expression in LK, HK, ILK, and IHK <span class="html-italic">N. benthamiana.</span> (<b>b</b>) Upregulation of PR genes in overexpressed <span class="html-italic">NbSAMS1</span> in <span class="html-italic">N. benthamiana</span> inoculated with <span class="html-italic">P. parasitica</span>. (<b>c</b>) Lesion diameter analyses of <span class="html-italic">NbSAMS1</span> overexpressing (OENbSAMS1) HK and LK <span class="html-italic">N. benthamiana</span> leaves (<b>d</b>) Lesion diameter analyses of <span class="html-italic">NbSAMS1</span> silenced (VNbSAMS1) HK and LK <span class="html-italic">N. benthamiana</span> leaves. Data are shown as means ± SD using Student’s <span class="html-italic">t</span>-test analysis. *: <span class="html-italic">p</span> &lt; 0.05; different letters denote significant differences at <span class="html-italic">p</span> &lt; 0.05.</p>
Full article ">Figure 3
<p><span class="html-italic">NbCBP1</span> interacts with <span class="html-italic">NbSAMS1</span>. (<b>a</b>) Coexpression network of <span class="html-italic">NbCBP1</span> and <span class="html-italic">NbSAMS1</span> with genes involved in the hub module. (<b>b</b>) Molecular docking analysis of NbSAMS1 and NbCBP1 interaction. (<b>c</b>) Yeast two-hybrid assay displays the interaction between NbCBP1 and NbSAMS1. (<b>d</b>) Co-immunoprecipitation (Co-IP) assay confirms the NbCBP1-NbSAMS1 interaction in vivo.</p>
Full article ">Figure 4
<p><span class="html-italic">NbCBP1</span> enhances <span class="html-italic">Nicotiana benthamiana</span> resistance under high potassium conditions. (<b>a</b>) <span class="html-italic">NbCBP1</span> expression in HK, LK, IHK, and ILK <span class="html-italic">N. benthamiana.</span> (<b>b</b>) Upregulation of PR genes in overexpressed <span class="html-italic">NbCBP1</span> in <span class="html-italic">N. benthamiana</span> inoculated with <span class="html-italic">P. parasitica.</span> (<b>c</b>) Lesion diameter analyses of <span class="html-italic">N. benthamiana</span> leaves (HK or LK) with <span class="html-italic">NbCBP1</span> overexpressing. (<b>d</b>) Lesion diameter analyses of <span class="html-italic">NbCBP1</span> silenced HK <span class="html-italic">N. benthamiana</span> leaves. (<b>e</b>) Ethylene production in <span class="html-italic">NbCBP1</span> overexpression and silenced IHK and ILK <span class="html-italic">N. benthamiana</span>. (<b>f</b>) Plant height of <span class="html-italic">NbCBP1</span> silenced LK and HK <span class="html-italic">N. benthamiana.</span> Data are shown as means ± SD, *: <span class="html-italic">p</span> &lt; 0.05; different letters denote significant differences at <span class="html-italic">p</span> &lt; 0.05 from Student’s <span class="html-italic">t</span>-test.</p>
Full article ">Figure 5
<p>NbCBP1 stabilizes NbSAMS1 and promotes ethylene accumulation in response to K availability and pathogen stress. Expression of <span class="html-italic">NbSAMS1</span> in ILK and IHK <span class="html-italic">N. benthamiana</span> with (<b>a</b>) <span class="html-italic">NbCBP1</span> overexpression or (<b>b</b>) <span class="html-italic">NbCBP1</span> silencing. (<b>c</b>) Immunoblot analysis in <span class="html-italic">NbSAMS1</span> overexpressing HK and LK <span class="html-italic">N. benthamiana</span>. (<b>d</b>) NbSAMS1 degradation assays performed with or without MG132 treatment in LK <span class="html-italic">N. benthamiana.</span> Data are shown as means ± SD; different letters denote significant differences at <span class="html-italic">p</span> &lt; 0.05 from Student’s <span class="html-italic">t</span>-test.</p>
Full article ">Figure 6
<p>Proposed Mechanism of high resistance in <span class="html-italic">Nicotiana benthamiana</span> with high potassium status. In low potassium <span class="html-italic">N. benthamiana</span> (left), the <span class="html-italic">NbCBP1</span> shows low expression, leading to E3 ligase and SAMS1 interaction. SAMS1 is then degraded via 26S proteasome. This degradation results in lower ethylene production, weakening the plant’s defense response. In high potassium <span class="html-italic">N. benthamiana</span> (right), the <span class="html-italic">NbCBP1</span> shows high expression; CBP1 competes with E3 ligase to interact with NbSAMS1, thereby protecting SAMS1 from degradation, enhancing the stability of SAMS1, resulting in increased ethylene production. The elevated ethylene levels contribute to the plant’s enhanced plant disease resistance.</p>
Full article ">
19 pages, 3754 KiB  
Article
Differential Regulation of miRNA and Protein Profiles in Human Plasma-Derived Extracellular Vesicles via Continuous Aerobic and High-Intensity Interval Training
by Zhenghao Wang, Yiran Ou, Xinyue Zhu, Ye Zhou, Xiaowei Zheng, Meixia Zhang, Sheyu Li, Shao-Nian Yang, Lisa Juntti-Berggren, Per-Olof Berggren and Xiaofeng Zheng
Int. J. Mol. Sci. 2025, 26(3), 1383; https://doi.org/10.3390/ijms26031383 - 6 Feb 2025
Viewed by 490
Abstract
Both continuous aerobic training (CAT) and high-intensity interval training (HIIT) are recommended to promote health and prevent diseases. Exercise-induced circulating extracellular vesicles (EX-EVs) have been suggested to play essential roles in mediating organ crosstalk, but corresponding molecular mechanisms remain unclear. To assess and [...] Read more.
Both continuous aerobic training (CAT) and high-intensity interval training (HIIT) are recommended to promote health and prevent diseases. Exercise-induced circulating extracellular vesicles (EX-EVs) have been suggested to play essential roles in mediating organ crosstalk, but corresponding molecular mechanisms remain unclear. To assess and compare the systemic effects of CAT and HIIT, five healthy male volunteers were assigned to HIIT and CAT, with a 7-day interval between sessions. Plasma EVs were collected at rest or immediately after each training section, prior to proteomics and miRNA profile analysis. We found that the differentially expressed (DE) miRNAs in EX-EVs were largely involved in the regulation of transcriptional factors, while most of the DE proteins in EX-EVs were identified as non-secreted proteins. Both CAT and HIIT play common roles in neuronal signal transduction, autophagy, and cell fate regulation. Specifically, CAT showed distinct roles in cognitive function and substrate metabolism, while HIIT was more associated with organ growth, cardiac muscle function, and insulin signaling pathways. Interestingly, the miR-379 cluster within EX-EVs was specifically regulated by HIIT, involving several biological functions, including neuroactive ligand–receptor interaction. Furthermore, EX-EVs likely originate from various tissues, including metabolic tissues, the immune system, and the nervous system. Our study provides molecular insights into the effects of CAT and HIIT, shedding light on the roles of EX-EVs in mediating organ crosstalk and health promotion. Full article
(This article belongs to the Special Issue Molecular Insights into the Role of Exercise in Disease and Health)
Show Figures

Figure 1

Figure 1
<p>Obtention and characterization of human plasma-derived EVs before and after two types of exercise. (<b>A</b>) Schematic illustration of the study design. In brief, five healthy volunteers who underwent acute CAT and HIIT at an interval of 7 days were enrolled in the experiment. Blood samples were collected in heparin-coated blood collection tubes at rest or immediately after each training session, followed by centrifugation at 1600× <span class="html-italic">g</span> for 10 min at 4 °C to separate plasma. The plasma fraction was then subjected to ultracentrifugation for EV isolation, and aliquots from the same EV preparations were subsequently used for EV characterization, proteomics, and miRNA profile analysis, respectively. (<b>B</b>) Real-time heart rate monitoring of volunteers during CAT and HIIT (n = 5). (<b>C</b>) Western blot analysis of EV markers. (<b>D</b>) Evaluation of morphology and size distribution of EVs via TEM and NTA, respectively. Scale bar = 50 nm. (<b>E</b>) Quantification of the average diameter and concentration of plasma EVs obtained from the indicated groups (n = 5); ns = non-significant difference.</p>
Full article ">Figure 2
<p>miRNA profiles and functional enrichment analysis of plasma EVs. (<b>A</b>) Correlation matrix analysis of miRNA signatures of the indicated groups. (<b>B</b>) Two-way hierarchical clustering heatmaps of DE EV miRNAs in the CAT vs. REST comparison (left panel) and in the HIIT vs. REST comparison (right panel). (<b>C</b>) Volcano plots showing DE EV miRNAs in the CAT vs. REST comparison (left panel) and in the HIIT vs. REST comparison (right panel). (<b>D</b>) Bar plots illustrating the most upregulated and downregulated miRNAs in the CAT vs. REST comparison (left panel) and in the HIIT vs. REST comparison (right panel). (<b>E</b>) Commonly enriched GO biological processes of target genes of the DE miRNAs in both CAT vs. REST comparison and the HIIT vs. REST comparison. Same colors of the bubble contour represent similar biological functions that different GO biological processes involved. (<b>F</b>) Differentially enriched GO biological processes of target genes of the DE miRNAs in the CAT vs. REST comparison (left panel) and the HIIT vs. REST comparison (right panel). In each subgroup (CAT vs. REST or HIIT vs. REST), same colors of the bubble contour represent similar biological functions that different GO biological processes involved.</p>
Full article ">Figure 3
<p>Tissue-specific enrichment analysis of DE EV miRNAs. (<b>A</b>,<b>B</b>) Sankey diagrams visualizing the tissue origin of the upregulated (<b>A</b>) and downregulated (<b>B</b>) EV miRNAs in the CAT group compared to REST group. (<b>C</b>,<b>D</b>) Sankey diagrams visualizing the tissue origin of the upregulated (<b>C</b>) and downregulated (<b>D</b>) EV miRNAs in the HIIT group compared to REST group.</p>
Full article ">Figure 4
<p>Regulation of the miR-379 cluster and miR-154 family within plasma EVs by HIIT. (<b>A</b>) Genomic locations of the miR-379 cluster and miR-154 family. The promoter region is located approximately 19 kb upstream of Chr14MC, which contains the miR-379 cluster and the mir-154 family. The orange squares denote HIIT-regulated miR-379 cluster members, while the green frames denote HIIT-regulated miR-154 family members. (<b>B</b>) The bar graph showing fold changes in the expression of HIIT-regulated miR-379 cluster members. (<b>C</b>) Enriched GO biological processes of target genes of the DE miRNAs in the HIIT vs. REST comparison. (<b>D</b>) PPI analysis of target genes of the DE miRNAs in the HIIT vs. REST comparison. Line thickness indicates the strength of the PPI.</p>
Full article ">Figure 5
<p>Proteomic profiles and functional enrichment analysis of plasma EVs. (<b>A</b>) Correlation matrix analysis of protein signatures of the indicated groups. (<b>B</b>) Rank plot of total proteins based on their average log10 intensity. Some of the known EV marker proteins are highlighted in red. (<b>C</b>) Two-way hierarchical clustering heatmaps of DE EV proteins in the CAT vs. REST comparison (left panel) and in the HIIT vs. REST comparison (right panel). (<b>D</b>) Volcano plots of DE EV proteins in the CAT vs. REST comparison (left panel) and in the HIIT vs. REST comparison (right panel). (<b>E</b>) Bar plots of the 10 most upregulated and downregulated proteins in the CAT vs. REST comparison (left panel) and in the HIIT vs. REST comparison (right panel). (<b>F</b>) Percentage stacked bar charts showing the presence or absence of a signal peptide sequence of DE EV proteins in the indicated comparisons. (<b>G</b>) Prediction of the subcellular localization of DE EV proteins in the CAT vs. REST comparison (left panel) and in the HIIT vs. REST comparison (right panel) using Hum-mPLoc3. (<b>H</b>) Commonly enriched GO biological processes of DE EV proteins in the CAT vs. REST comparison and HIIT vs. REST comparison. Same colors of the bubble contour represent similar biological functions that different GO biological processes involved. (<b>I</b>) Differentially enriched GO biological processes of DE EV proteins in the CAT vs. REST comparison and the HIIT vs. REST comparison. In each subgroup (CAT vs. REST or HIIT vs. REST), same colors of the bubble contour represent similar biological functions that different GO biological processes involved.</p>
Full article ">Figure 6
<p>Tissue-specific enrichment analysis of DE EV proteins. (<b>A</b>,<b>B</b>) Sankey diagrams visualizing the tissue origin of the upregulated (<b>A</b>) and downregulated (<b>B</b>) EV proteins in the CAT group compared to REST group. (<b>C</b>,<b>D</b>) Sankey diagrams visualizing the tissue origin of the upregulated (<b>C</b>) and downregulated (<b>D</b>) EV proteins in the HIIT group compared to REST group.</p>
Full article ">Figure 7
<p>Multivariate Venn diagram of GO terms based on the DE EV proteins and target genes of the DE EV miRNAs in both CAT and HIIT groups.</p>
Full article ">
48 pages, 2940 KiB  
Review
Molecular Regulation of Palatogenesis and Clefting: An Integrative Analysis of Genetic, Epigenetic Networks, and Environmental Interactions
by Hyuna Im, Yujeong Song, Jae Kyeom Kim, Dae-Kyoon Park, Duk-Soo Kim, Hankyu Kim and Jeong-Oh Shin
Int. J. Mol. Sci. 2025, 26(3), 1382; https://doi.org/10.3390/ijms26031382 - 6 Feb 2025
Viewed by 520
Abstract
Palatogenesis is a complex developmental process requiring temporospatially coordinated cellular and molecular events. The following review focuses on genetic, epigenetic, and environmental aspects directing palatal formation and their implication in orofacial clefting genesis. Essential for palatal shelf development and elevation (TGF-β, BMP, FGF, [...] Read more.
Palatogenesis is a complex developmental process requiring temporospatially coordinated cellular and molecular events. The following review focuses on genetic, epigenetic, and environmental aspects directing palatal formation and their implication in orofacial clefting genesis. Essential for palatal shelf development and elevation (TGF-β, BMP, FGF, and WNT), the subsequent processes of fusion (SHH) and proliferation, migration, differentiation, and apoptosis of neural crest-derived cells are controlled through signaling pathways. Interruptions to these processes may result in the birth defect cleft lip and/or palate (CL/P), which happens in approximately 1 in every 700 live births worldwide. Recent progress has emphasized epigenetic regulations via the class of non-coding RNAs with microRNAs based on critically important biological processes, such as proliferation, apoptosis, and epithelial–mesenchymal transition. These environmental risks (maternal smoking, alcohol, retinoic acid, and folate deficiency) interact with genetic and epigenetic factors during palatogenesis, while teratogens like dexamethasone and TCDD inhibit palatal fusion. In orofacial cleft, genetic, epigenetic, and environmental impact on the complex epidemiology. This is an extensive review, offering current perspectives on gene-environment interactions, as well as non-coding RNAs, in palatogenesis and emphasizing open questions regarding these interactions in palatal development. Full article
(This article belongs to the Special Issue Gene Regulatory and Signaling Pathways in Palatogenesis)
Show Figures

Figure 1

Figure 1
<p>Developmental progression of secondary palate formation in mouse embryos from E11.5 to E15.5. (<b>A</b>–<b>E</b>) Frontal views show the developmental sequence of palatal shelf elevation and fusion. At E11.5 (<b>A</b>), the medial nasal process (MNP) and maxillary processes (MxP) are visible with the initial formation of the primary palate (PP). From E13.5 to E15.5 (<b>B</b>–<b>E</b>), the palatal shelves (PS) undergo vertical-to-horizontal elevation, with concurrent development of the nasal septum (NS). Progressive fusion of the PS occurs, resulting in the formation of the secondary palate (SP) by E15.5. (<b>F</b>–<b>T</b>) Frontal sections through the developing palate’s anterior, middle, and posterior regions at corresponding developmental stages. The sections demonstrate the progressive growth, elevation, and fusion of the PS around the tongue (T). The medial edge epithelium (MEE) is visible in anterior sections at early stages. The dynamic process of palatal shelf reorientation and fusion proceeds in an anterior-to-posterior sequence, with complete fusion achieved by E15.5. Color code: pink—medial nasal process (MNP); turquoise—maxillary process (MxP) (E11.5) and maxillary tissue (E13.5–E15.5) and lip; yellow—primary palate (PP); light blue—nasal septum (NS); light pink—palatal shelves (PS); black arrowheads indicate the gap between the primary and secondary palates, which will close following fusion between these tissues. Also, the black curved arrows indicate the direction of palatal growth.</p>
Full article ">Figure 2
<p>The anatomical spectrum of normal and cleft lip and/or palate malformations in humans. (<b>A</b>–<b>C</b>) Frontal facial views show variations in lip formation. (<b>A</b>) Normal lip morphology with complete fusion. (<b>B</b>) The unilateral cleft lip is showing incomplete fusion on one side. (<b>C</b>) Bilateral cleft lip presenting incomplete fusion on both sides of the upper lip. (<b>D</b>–<b>F</b>) Oral views of the palatal region depicting normal and cleft phenotypes. (<b>D</b>) Normal palate showing complete fusion of palatal shelves. (<b>E</b>) Unilateral cleft lip and palate with the incomplete fusion of the palatal shelf on one side. (<b>F</b>) Bilateral cleft lip and palate showing incomplete fusion of palatal shelves on both sides. Color code: turquoise—maxillary tissue; yellow—primary palate; pink—palatal and soft tissues.</p>
Full article ">Figure 3
<p>Complex molecular networks that coordinate palatal shelf growth, patterning, and morphogenesis along both the anterior–posterior and medial–lateral axes during palatogenesis. (<b>A</b>–<b>D</b>) Lateral and oral views of developing mouse embryos at E12.5 and E13.5 with corresponding schematic diagram The anterior and posterior orientation is indicated by dashed lines. (<b>E</b>–<b>G</b>) Schematic representations of molecular networks controlling palate development: (<b>E</b>) Key factors involved in palate growth and patterning along the anterior–posterior axis, showing interactions between epithelial and mesenchymal factors and their downstream targets. (<b>F</b>) Molecular regulation of anterior palatal shelf development. (<b>G</b>) Posterior palatal shelf patterning network showing interactions. Color code: pink—palatal epithelium; apricot—mesenchyme; blue—transcription factors; green—receptors; yellow—ligands; orange—other regulatory molecules.</p>
Full article ">Figure 4
<p>Development and molecular regulation of palatal shelf adhesion and fusion. (<b>A</b>–<b>C</b>) Schematic representation of mouse embryo development from E14.5 to E15.5. (<b>A</b>) Lateral view of E12.5 mouse embryo showing the anterior–posterior axis of palatal development. (<b>B</b>) Oral view at E14.5 showing elevated palatal shelves before fusion. (<b>C</b>) Oral view at E15.5 showing palatal fusing. (<b>D</b>–<b>F</b>) Molecular pathways controlling three key stages of palatal fusion. (<b>D</b>) Epithelial differentiation and periderm maintenance pathway showing genetic interactions. (<b>E</b>) Palatal adhesion and medial edge epithelium (MEE) formation pathway involving β-catenin, Tgf-β3, and downstream effectors. (<b>F</b>) Midline epithelial seam (MES) degeneration process leading to palatal fusion. Color code: pink—epithelium; apricot—mesenchyme; blue—transcription factors; green—receptors; yellow—ligands; orange—other regulatory molecules; black dotted line—remaining MES during palatal fusion.</p>
Full article ">Figure 5
<p>Epigenetic regulation during craniofacial development. (<b>A</b>,<b>B</b>) Schematic representation of early craniofacial development showing the morphological changes from initial facial prominences to their fusion. (<b>A</b>) The left panel shows the initial facial prominences, including midbrain, forebrain, lateral and medial nasal processes, nasal pit, maxillary and mandibular processes, and second arch. (<b>B</b>) The right panel shows the subsequent development of the frontonasal region, maxillary region, and the mandibular region. (<b>C</b>) Four major epigenetic mechanisms regulating craniofacial development: DNA methylation: addition of methyl groups to promoter regions controlling target gene expression. The red cross symbol means that the expression of the target gene is suppressed. Histone modification: post-translational modifications, including methylation (Me) and acetylation (Ac) of histone proteins. Non-coding RNAs: involvement of microRNAs (miRNA) and long non-coding RNAs (lncRNA) in gene regulation. Chromatin remodeling: ATP-dependent nucleosome ejection and sliding mediated by SWI/SNF complexes. Color code: dark blue—midbrain; light blue—forebrain; green—lateral nasal process; red—medial nasal process; navy blue—nasal pit; turquoise—maxillary process; purple—mandibular process and second arch; orange—frontonasal region; gray—other facial region behind maxillary/mandibular regions.</p>
Full article ">Figure 6
<p>Summary of the miRNAs and genes associated with cleft lip and/or palate (CL/P) in mice and human. The complex miRNA-mediated regulatory networks involved in cleft lip and/or palate (CL/P). Environmental factors also affect cleft palate development by modulating miRNA activity, and how their dysregulation contributes to cleft formation through cell proliferation defects, differentiation defects, and cell death. Red arrows; graph showing genes going up, blue arrows; graph showing genes going down. CL, cleft lip; CP, cleft palate;</p>
Full article ">
22 pages, 13927 KiB  
Article
Discovery of TRPV4-Targeting Small Molecules with Anti-Influenza Effects Through Machine Learning and Experimental Validation
by Yan Sun, Jiajing Wu, Beilei Shen, Hengzheng Yang, Huizi Cui, Weiwei Han, Rongbo Luo, Shijun Zhang, He Li, Bingshuo Qian, Lingjun Fan, Junkui Zhang, Tiecheng Wang, Xianzhu Xia, Fang Yan and Yuwei Gao
Int. J. Mol. Sci. 2025, 26(3), 1381; https://doi.org/10.3390/ijms26031381 - 6 Feb 2025
Viewed by 457
Abstract
Transient receptor potential vanilloid 4 (TRPV4) is a calcium-permeable cation channel critical for maintaining intracellular Ca2+ homeostasis and is essential in regulating immune responses, metabolic processes, and signal transduction. Recent studies have shown that TRPV4 activation enhances influenza A virus infection, promoting [...] Read more.
Transient receptor potential vanilloid 4 (TRPV4) is a calcium-permeable cation channel critical for maintaining intracellular Ca2+ homeostasis and is essential in regulating immune responses, metabolic processes, and signal transduction. Recent studies have shown that TRPV4 activation enhances influenza A virus infection, promoting viral replication and transmission. However, there has been limited exploration of antiviral drugs targeting the TRPV4 channel. In this study, we developed the first machine learning model specifically designed to predict TRPV4 inhibitory small molecules, providing a novel approach for rapidly identifying repurposed drugs with potential antiviral effects. Our approach integrated machine learning, virtual screening, data analysis, and experimental validation to efficiently screen and evaluate candidate molecules. For high-throughput virtual screening, we employed computational methods to screen open-source molecular databases targeting the TRPV4 receptor protein. The virtual screening results were ranked based on predicted scores from our optimized model and binding energy, allowing us to prioritize potential inhibitors. Fifteen small-molecule drugs were selected for further in vitro and in vivo antiviral testing against influenza. Notably, glecaprevir and everolimus demonstrated significant inhibitory effects on the influenza virus, markedly improving survival rates in influenza-infected mice (protection rates of 80% and 100%, respectively). We also validated the mechanisms by which these drugs interact with the TRPV4 channel. In summary, our study presents the first predictive model for identifying TRPV4 inhibitors, underscoring TRPV4 inhibition as a promising strategy for antiviral drug development against influenza. This pioneering approach lays the groundwork for future clinical research targeting the TRPV4 channel in antiviral therapies. Full article
Show Figures

Graphical abstract

Graphical abstract
Full article ">Figure 1
<p>(<b>A</b>) Distributions of various molecular descriptors (MolWt, TPSA, NumHDonors, NumHAcceptors, LogP, and MolVolume) for active and inactive molecules. (<b>B</b>) The heatmap presents the performance metrics of various machine learning models using different molecular fingerprints (MACCS keys, Morgan, RDKit, and atom pairs) and classifiers (XGBoost 2.0.0, SVM, random forest, and MLP).</p>
Full article ">Figure 2
<p>Performance evaluation of machine learning models for predicting TRPV4 inhibitor activity. (<b>A</b>) Confusion matrices of selected models show the classification performance in terms of true positives, true negatives, false positives, and false negatives. (<b>B</b>) AUC curves for each classifier (MLP, RandomForest, SVM, XGBoost 2.0.0) across different fingerprints (MACCS, Morgan, RDKit (Greg Landrum, Switzerland), AtomPairsFP), illustrating the classifiers’ ability to differentiate between active and inactive compounds.The black dashed line is random classifier.</p>
Full article ">Figure 3
<p>Viral inhibition of drug candidates in vitro. (<b>A</b>) In this figure, CCK−8 assay was used to detect the cytotoxic effect of selected drugs on MDCK cells. By staining drug-treated MDCK cells in response to CCK−8 reagents, the effects of the drugs on cell viability can be observed and quantified, and thus their potential cytotoxicity can be assessed. (<b>B</b>) In this figure, the expression of viral nucleoprotein (NP) in influenza virus-infected MDCK cells was examined using immunofluorescence after 48 h of drug incubation. By labeling the viral NP protein with a specific antibody and combining it with a fluorescently labeled secondary antibody, the distribution of the NP protein within the cell can be visualized under a microscope to understand the effect of the drug on viral replication. (<b>C</b>) Samples of the 15 selected drugs were collected 48 h after drug delivery and analyzed for H1N1-UI182 viral nucleoprotein (NP) expression and quantification using western blot images. (<b>D</b>) This figure shows the results of quantitative analysis of the immunofluorescence image obtained in (<b>B</b>). The results represented the mean of three independent experiments and were compared with a viral control group (±SD) (<span class="html-italic">n</span> = 3). *** <span class="html-italic">p</span> &lt; 0.0001 for significant difference; ** <span class="html-italic">p</span> &lt; 0.01 * <span class="html-italic">p</span> &lt; 0.05 indicates.</p>
Full article ">Figure 3 Cont.
<p>Viral inhibition of drug candidates in vitro. (<b>A</b>) In this figure, CCK−8 assay was used to detect the cytotoxic effect of selected drugs on MDCK cells. By staining drug-treated MDCK cells in response to CCK−8 reagents, the effects of the drugs on cell viability can be observed and quantified, and thus their potential cytotoxicity can be assessed. (<b>B</b>) In this figure, the expression of viral nucleoprotein (NP) in influenza virus-infected MDCK cells was examined using immunofluorescence after 48 h of drug incubation. By labeling the viral NP protein with a specific antibody and combining it with a fluorescently labeled secondary antibody, the distribution of the NP protein within the cell can be visualized under a microscope to understand the effect of the drug on viral replication. (<b>C</b>) Samples of the 15 selected drugs were collected 48 h after drug delivery and analyzed for H1N1-UI182 viral nucleoprotein (NP) expression and quantification using western blot images. (<b>D</b>) This figure shows the results of quantitative analysis of the immunofluorescence image obtained in (<b>B</b>). The results represented the mean of three independent experiments and were compared with a viral control group (±SD) (<span class="html-italic">n</span> = 3). *** <span class="html-italic">p</span> &lt; 0.0001 for significant difference; ** <span class="html-italic">p</span> &lt; 0.01 * <span class="html-italic">p</span> &lt; 0.05 indicates.</p>
Full article ">Figure 4
<p>Drug candidate in mice infected with a lethal H1N1 virus. (<b>A</b>,<b>B</b>) Weight changes in mice infected with and treated with H1N1-UI182. (<b>C</b>,<b>D</b>) Survival status of mice infected with H1N1-UI182 and mice treated with drugs. (<b>E</b>,<b>F</b>) Lung pathology in mice infected with H1N1-UI182 and in mice treated with drugs 5 days after infection. (<b>G</b>,<b>H</b>) Changes in lung index of mice in the drug treatment group after 5 days of infection. The coloured lines with dots is error bar. (<b>I</b>,<b>J</b>) Changes of viral load in the lungs of mice in the drug treatment group 5 days after infection. (<b>K</b>,<b>L</b>) Western blot images showed the expression of viral nucleoprotein (NP), hemagglutinin (HA), and structural protein (M1). The results represented the mean of three independent experiments and were compared with a viral control group (±SD). *** <span class="html-italic">p</span> &lt; 0.001 for significant difference; ** <span class="html-italic">p</span> &lt; 0.01 * <span class="html-italic">p</span> &lt; 0.05 indicates.</p>
Full article ">Figure 4 Cont.
<p>Drug candidate in mice infected with a lethal H1N1 virus. (<b>A</b>,<b>B</b>) Weight changes in mice infected with and treated with H1N1-UI182. (<b>C</b>,<b>D</b>) Survival status of mice infected with H1N1-UI182 and mice treated with drugs. (<b>E</b>,<b>F</b>) Lung pathology in mice infected with H1N1-UI182 and in mice treated with drugs 5 days after infection. (<b>G</b>,<b>H</b>) Changes in lung index of mice in the drug treatment group after 5 days of infection. The coloured lines with dots is error bar. (<b>I</b>,<b>J</b>) Changes of viral load in the lungs of mice in the drug treatment group 5 days after infection. (<b>K</b>,<b>L</b>) Western blot images showed the expression of viral nucleoprotein (NP), hemagglutinin (HA), and structural protein (M1). The results represented the mean of three independent experiments and were compared with a viral control group (±SD). *** <span class="html-italic">p</span> &lt; 0.001 for significant difference; ** <span class="html-italic">p</span> &lt; 0.01 * <span class="html-italic">p</span> &lt; 0.05 indicates.</p>
Full article ">Figure 5
<p>Glecaprevir and everolimus protect mice infected with H1N1 lethal virus. (<b>A</b>,<b>B</b>) Weight and survival trends in H1N1-UI182-infected mice and glecaprevir treated mice. (<b>C</b>,<b>D</b>) Weight and survival trends in H1N1-UI182-infected mice and everolimus-treated mice. (<b>E</b>) Changes of lung index in mice infected with H1N1-UI182 and treated with glecaprevir. (<b>F</b>) The effect of glecaprevir at 5 dpi on virus titer in the lung of mice was detected. (<b>G</b>) Changes of lung index in mice infected with H1N1-UI182 and treated with everolimus. (<b>H</b>) The effect of everolimus on virus titers in the lungs of mice at 5 dpi was detected. (<b>I</b>,<b>J</b>) Five mice in glecaprevir and everolimus groups were randomly selected 5 days after infection, and their lungs were dissected after euthanasia for examination by H&amp;E staining analysis (<span class="html-italic">n</span> = 5). The yellow arrow represents granulocytes, the gray arrow represents the alveolar wall, the purple arrow represents alveolar dilation, the red arrow represents bleeding around the alveoli and blood vessels, the brown arrow represents epithelial cells, and the blue arrow represents macrophages. (<b>K</b>,<b>L</b>) Western blot images showed in glecaprevir and everolimus groups the expression of viral nucleoprotein (NP), hemagglutinin (HA) and structural protein (M1) in mouse lungs. The results represented the mean of three independent experiments and were compared with a viral control group (±SD). *** <span class="html-italic">p</span> &lt; 0.001 for significant difference; * <span class="html-italic">p</span> &lt; 0.05 indicates.</p>
Full article ">Figure 5 Cont.
<p>Glecaprevir and everolimus protect mice infected with H1N1 lethal virus. (<b>A</b>,<b>B</b>) Weight and survival trends in H1N1-UI182-infected mice and glecaprevir treated mice. (<b>C</b>,<b>D</b>) Weight and survival trends in H1N1-UI182-infected mice and everolimus-treated mice. (<b>E</b>) Changes of lung index in mice infected with H1N1-UI182 and treated with glecaprevir. (<b>F</b>) The effect of glecaprevir at 5 dpi on virus titer in the lung of mice was detected. (<b>G</b>) Changes of lung index in mice infected with H1N1-UI182 and treated with everolimus. (<b>H</b>) The effect of everolimus on virus titers in the lungs of mice at 5 dpi was detected. (<b>I</b>,<b>J</b>) Five mice in glecaprevir and everolimus groups were randomly selected 5 days after infection, and their lungs were dissected after euthanasia for examination by H&amp;E staining analysis (<span class="html-italic">n</span> = 5). The yellow arrow represents granulocytes, the gray arrow represents the alveolar wall, the purple arrow represents alveolar dilation, the red arrow represents bleeding around the alveoli and blood vessels, the brown arrow represents epithelial cells, and the blue arrow represents macrophages. (<b>K</b>,<b>L</b>) Western blot images showed in glecaprevir and everolimus groups the expression of viral nucleoprotein (NP), hemagglutinin (HA) and structural protein (M1) in mouse lungs. The results represented the mean of three independent experiments and were compared with a viral control group (±SD). *** <span class="html-italic">p</span> &lt; 0.001 for significant difference; * <span class="html-italic">p</span> &lt; 0.05 indicates.</p>
Full article ">Figure 6
<p>Glecaprevir and everolimus improved the inflammatory response as well as the expression of viral proteins after viral infection. (<b>A</b>–<b>C</b>) Western blot was used to detect the expression of cytokines and protein in the two drug groups. (<b>D</b>–<b>H</b>) Real-time fluorescence quantitative PCR (qPCR) was applied to detect the expression of cytokines in the two drug groups. (<b>I</b>–<b>L</b>) Western blot analysis of TRPV4, NP, and M1 expression after virus infection. The results represented the mean of three independent experiments and were compared with a viral control group (±SD) (<span class="html-italic">n</span> = 3). *** <span class="html-italic">p</span> &lt; 0.001 for significant difference; ** <span class="html-italic">p</span> &lt; 0.01 * <span class="html-italic">p</span> &lt; 0.05 indicates.</p>
Full article ">Figure 7
<p>Binding mode of the compounds to 4DX2. (<b>A</b>,<b>B</b>) Binding mode of 4DX2 and glecaprevir. (<b>C</b>,<b>D</b>) Binding mode of 4DX2 and everolimus. Green cartoon represents the ligand compounds, purple sticks represent key residues in the covalent interaction, yellow dash lines represent hydrogen bonds. The corresponding 2D interaction maps are shown in (<b>B</b>,<b>D</b>) respectively, in which the hydrogen bond interactions are colored in dark green, with the van der Waals interactions in light green, the π−π stacking in dark pink, the π–alkyl interactions in light pink, and the π–anion and salt bridge interactions in orange.</p>
Full article ">
29 pages, 12007 KiB  
Article
Molecular Simulation of the Binding of Amyloid Beta to Apolipoprotein A-I in High-Density Lipoproteins
by Chris J. Malajczuk and Ricardo L. Mancera
Int. J. Mol. Sci. 2025, 26(3), 1380; https://doi.org/10.3390/ijms26031380 - 6 Feb 2025
Viewed by 337
Abstract
Disrupted clearance of amyloid beta (Aβ) from the brain enhances its aggregation and formation of amyloid plaques in Alzheimer’s disease. The most abundant protein constituent of circulating high-density lipoprotein (HDL) particles, apoA-I, readily crosses the blood–brain barrier from periphery circulation, exhibits low-micromolar binding [...] Read more.
Disrupted clearance of amyloid beta (Aβ) from the brain enhances its aggregation and formation of amyloid plaques in Alzheimer’s disease. The most abundant protein constituent of circulating high-density lipoprotein (HDL) particles, apoA-I, readily crosses the blood–brain barrier from periphery circulation, exhibits low-micromolar binding affinity for soluble, neurotoxic forms of Aβ, and modulates Aβ aggregation and toxicity in vitro. Its highly conserved N-terminal sequence, 42LNLKLLD48 (‘LN’), has been proposed as a binding region for Aβ. However, high-resolution structural characterisation of the mechanism of HDL–Aβ interaction is very difficult to attain. Molecular dynamics simulations were conducted to investigate for the first time the interaction of Aβ and the ‘LN’ segment of apoA-I. Favourable binding of Aβ by HDLs was found to be driven by hydrophobic and hydrogen-bonding interactions predominantly between the ‘LN’ segment of apoA-I and Aβ. Preferential binding of Aβ may proceed in small, protein-rich HDLs whereby solvent-exposed hydrophobic ‘LN’ segments of apoA-I interact specifically with Aβ, stabilising it on the HDL surface in a possibly non-amyloidogenic conformation, facilitating effective Aβ clearance. These findings rationalise the potentially therapeutic role of HDLs in reducing Aβ aggregation and toxicity, and of peptide mimics of the apoA-I interacting region in blocking Aβ aggregation. Full article
(This article belongs to the Special Issue Advances in Protein Dynamics)
Show Figures

Figure 1

Figure 1
<p>(<b>a</b>) Free energy landscape corresponding to the MD simulation of Aβ<sub>42CC</sub> and ‘LN’ as a function of intermolecular hydrophobic contacts (&lt;0.45 nm) and hydrogen bonds. (<b>b</b>) Average number of residue-wise intermolecular hydrophobic contacts (left) and intermolecular hydrogen bonds (right) between Aβ<sub>42CC</sub> (<b>top</b>) and ‘LN’ (<b>bottom</b>) throughout the final 150 ns of a conventional MD simulation. Select residues are labelled according to their single-letter amino acid code followed by their residue number. Red lettering corresponds to the residues within ‘LN’ region of apoA-I that are equivalent to the high-affinity Aβ-binding ‘GNLLTLD’ peptide. Horizontal bars are coloured according to general residue typing (basic—blue; acidic—red; non-polar—grey; polar—green). (<b>c</b>) Final snapshot of the Aβ<sub>42CC</sub>–‘LN’ complex corresponding to the free energy minimum for the simulation. Mainchain atoms drawn in a cartoon representation with ‘LN’ coloured green and the Aβ<sub>42CC</sub> peptide coloured according to secondary structure. ‘LN’ sidechains and interacting Aβ<sub>42CC</sub> residues are drawn in a licorice representation, with ‘LN’ residues coloured according to residue type whilst Aβ<sub>42CC</sub> residues are all coloured grey. Intermolecular hydrogen bonds are drawn as dashed magenta lines.</p>
Full article ">Figure 2
<p>(<b>a</b>) Free energy landscape as a function of the number of intermolecular hydrophobic contacts and hydrogen bonds formed between the ‘LN’ fragment and the Aβ<sub>42CC</sub> peptide at 310 K. (<b>b</b>) Secondary structures sampled by the Aβ<sub>42CC</sub> peptide (<b>top</b>) and the ‘LN’ fragment (<b>bottom</b>) at 310 K. Red lettering corresponds to the residues within ‘LN’ region of apoA-I that are equivalent to the high-affinity Aβ-binding ‘GNLLTLD’ peptide.</p>
Full article ">Figure 3
<p>Identification of binding poses between the ‘LN’ fragment of apoA-I and an Aβ<sub>42CC</sub> peptide at 310 K followed a two-step RMSD clustering regime considering low free energy states (ΔG ≤ 2 kcal/mol, 83.9% of the total replica time) taken from a T-REMD simulation. Initial clustering of ‘LN’ identified four high-occupancy clusters ranging from 3.9 to 68.4% of all low free energy states, followed by a secondary clustering of each ‘LN’ cluster based upon Aβ<sub>42CC</sub>. Bracketed percentages in black and red indicate the overall cluster occupancy as a percentage of the total considered low free energy states, with purple numbering denoting the relative proportion of the given ‘LN’ cluster and red numbering highlighting the top five highest-occupancy poses. Top five poses are denoted by an asterisk. Four of the top five poses were identified within the first ‘LN’ cluster, Cluster 1, wherein ‘LN’ exhibits a predominantly extended coil structure as shown in the corresponding secondary-structure plot. The fifth-highest-occupancy pose came from the mostly helical ‘LN’ Cluster 3. Note: Red lettering in ‘LN’ secondary structure plots correspond to the residues within ‘LN’ region of apoA-I that are equivalent to the high-affinity Aβ-binding ‘GNLLTLD’ peptide.</p>
Full article ">Figure 4
<p>Secondary structures and the average intermolecular properties for the top five binding poses between the ‘LN’ fragment of apoA-I and an Aβ<sub>42CC</sub> peptide from a T-REMD replica system at 310K. Each binding pose exhibits unique intermolecular interactions between ‘LN’ and Aβ<sub>42CC</sub> as highlighted by the distinctive minimum distance contact maps and Aβ<sub>42CC</sub>–‘LN’ intermolecular interaction residues. Note: Contact map colouring reflects a red to royal blue rainbow mapping scheme where prevailing contacts are represented by red and no contact is represented by blue, with residue numbering beginning from the bottom-left of the axis; horizontal bar plots are coloured according to general residue typing, with select residues labelled according to their single-letter amino acid code followed by their residue number.; red lettering in ‘LN’ secondary structure plots and ‘LN’ intermolecular contact bar plots correspond to the residues within ‘LN’ region of apoA-I that are equivalent to the high-affinity Aβ-binding ‘GNLLTLD’ peptide.</p>
Full article ">Figure 5
<p>Snapshots of the final MD simulation configuration for each Aβ42CC-HDL system showing that the two species are retained as a complex. Aβ42CC peptides and interacting ‘LN’ segments of the apoA-I chain are drawn in a deep blue and bright green surface, respectively. HDL surface lipids have been drawn as a white opaque surface, with HDL apoA-I chains drawn in an opaque cartoon representation surrounded by a transparent surface representation and each chain has been coloured separately as sky blue, red or green.</p>
Full article ">Figure 6
<p>(<b>a</b>) Summary of the 100 ns time-interval-wise binding free energies calculated throughout MD simulations of six HDL<sub>3c</sub>-Aβ<sub>42CC</sub> systems, each starting with a different initial binding pose between Aβ<sub>42CC</sub> and the full HDL<sub>3c</sub> particle via the exposed ‘LN’ region of apoA-I. Three initial binding poses (Cluster 1-1, Cluster 1-2, and Cluster 1-4) ultimately exhibited favourable and remarkably similar binding free energy values following 500 ns of MD simulation. The remaining three initial binding poses (Cluster 1-3, Cluster 3-1 and the cMD pose) ultimately exhibited unfavourable binding free energies following 500 ns of conventional MD simulation. (<b>b</b>) Decomposition of binding free energies for system components across the six binding poses throughout the final 100 ns of the MD simulations, with relative standard error shown as off-set error bars between the respective stacked bars.</p>
Full article ">Figure 7
<p>Average residue-wise decomposition of the free energies of binding across the interacting ‘LN’ region of apoA-I and the Aβ<sub>42CC</sub> peptide during the final 100 ns interval of a 500 ns MD simulation of HDL–Aβ<sub>42CC</sub> complexes for (<b>a</b>) the three favourable binding poses, and (<b>b</b>) the three unfavourable binding poses. Red lettering in the labels for ‘LN’ residues correspond to the residues within ‘LN’ region of apoA-I that are equivalent to the high-affinity Aβ-binding ‘GNLLTLD’ peptide.</p>
Full article ">Figure 8
<p>Snapshot of the final MD simulation configuration of the HDL–Aβ<sub>42CC</sub> complex for (<b>a</b>) C1-1, (<b>b</b>) C1-2 and (<b>c</b>) C1-4, with the surfaces of interacting ‘LN’ and Aβ<sub>42CC</sub> coloured green and blue, respectively. Insets show the binding interface with interacting residues drawn in a liquorice representation and the protein mainchain drawn in a cartoon representation. Aβ<sub>42CC</sub> is coloured according to secondary structure motif as defined in corresponding time-wise secondary structure profiles for Aβ<sub>42CC</sub> as well as the interacting ‘LN’ segment (far right). Note: Red lettering in ‘LN’ secondary structure plots correspond to the residues within ‘LN’ region of apoA-I that are equivalent to the high-affinity Aβ-binding ‘GNLLTLD’ peptide.</p>
Full article ">Figure 9
<p>Diagrammatic representation of the intermolecular interactions observed at the end of the HDL simulations in the binding interface of ‘LN’ (green) with Aβ<sub>42CC</sub> (blue) for (<b>a</b>) the (<b>a</b>) C1-1, (<b>b</b>) C1-2 and (<b>c</b>) C1-4 binding poses. Binding interface schematics generated according to data derived from Visual Molecular Dynamics (VMD) v. 1.9.3.</p>
Full article ">Figure 10
<p>Helical wheel diagrams of the H1 domain of lipidated apoA-I showing the position of Asp48 with respect to the prevailing non-polar face of the helix when residues within the ‘LN’ segment (circles with a green outline and green text) are (<b>a</b>) in a helical arrangement, and (<b>b</b>) when no longer contributing a part of the helix. Note: The succeeding <sup>61</sup>KEQLG<sup>65</sup> segment of H1 has been incorporated into the second helical wheel for context (circles with a blue outline and blue text).</p>
Full article ">
28 pages, 1219 KiB  
Review
Antioxidant and Anti-Inflammatory Effects of Bioactive Compounds in Atherosclerosis
by Ştefan Horia Roşian, Ioana Boarescu and Paul-Mihai Boarescu
Int. J. Mol. Sci. 2025, 26(3), 1379; https://doi.org/10.3390/ijms26031379 - 6 Feb 2025
Viewed by 549
Abstract
Atherosclerosis, a chronic inflammatory disease characterized by the accumulation of lipids and immune cells within arterial walls, remains a leading cause of cardiovascular morbidity and mortality worldwide. Oxidative stress and inflammation are central to its pathogenesis, driving endothelial dysfunction, foam cell formation, and [...] Read more.
Atherosclerosis, a chronic inflammatory disease characterized by the accumulation of lipids and immune cells within arterial walls, remains a leading cause of cardiovascular morbidity and mortality worldwide. Oxidative stress and inflammation are central to its pathogenesis, driving endothelial dysfunction, foam cell formation, and plaque instability. Emerging evidence highlights the potential of bioactive compounds with antioxidant and anti-inflammatory properties to mitigate these processes and promote vascular health. This review explores the mechanisms through which bioactive compounds—such as polyphenols, carotenoids, flavonoids, omega-3 fatty acids, coenzyme Q10, and other natural compounds—modulate oxidative stress and inflammation in atherosclerosis. It examines their effects on key molecular pathways, including the inhibition of reactive oxygen species (ROS) production, suppression of nuclear factor-κB (NF-κB), and modulation of inflammatory cytokines. By integrating current knowledge, this review underscores the therapeutic potential of dietary and supplemental bioactive compounds as complementary strategies for managing atherosclerosis, paving the way for future research and clinical applications. Full article
(This article belongs to the Special Issue Effects of Bioactive Compounds in Oxidative Stress and Inflammation)
Show Figures

Figure 1

Figure 1
<p>Pathogenesis of atherosclerosis.</p>
Full article ">Figure 2
<p>Effects of bioactive compounds on oxidative stress and inflammation in atherosclerosis. Bioactive compounds reduce oxidative stress preventing endothelial dysfunction and LDL oxi-dation. By reducing inflammation they also preserve endothelial function and prevent the recruitment of monocytes and their subsequent transformation into macrophages. The result is a reduction in formation of foam cells and inhibition of the development of atherosclerotic lesions.</p>
Full article ">Figure 3
<p>Chemical structure of (<b>a</b>) Polyphenol (Curcumin) and (<b>b</b>) Flavonoid.</p>
Full article ">
16 pages, 1891 KiB  
Article
Mitochondrial COX3 and tRNA Gene Variants Associated with Risk and Prognosis of Idiopathic Pulmonary Fibrosis
by Li-Na Lee, I-Shiow Jan, Wen-Ru Chou, Wei-Lun Liu, Yen-Liang Kuo, Chih-Yueh Chang, Hsiu-Ching Chang, Jia-Luen Liu, Chia-Lin Hsu, Chia-Nan Lin, Ke-Yun Chao, Chi-Wei Tseng, I-Hsien Lee, Jann-Tay Wang and Jann-Yuan Wang
Int. J. Mol. Sci. 2025, 26(3), 1378; https://doi.org/10.3390/ijms26031378 - 6 Feb 2025
Viewed by 423
Abstract
Idiopathic pulmonary fibrosis (IPF) has been associated with mitochondrial dysfunction. We investigated whether mitochondrial DNA variants in peripheral blood leukocytes (PBLs), which affect proteins of the respiratory chain and mitochondrial function, could be associated with an increased risk and poor prognosis of IPF. [...] Read more.
Idiopathic pulmonary fibrosis (IPF) has been associated with mitochondrial dysfunction. We investigated whether mitochondrial DNA variants in peripheral blood leukocytes (PBLs), which affect proteins of the respiratory chain and mitochondrial function, could be associated with an increased risk and poor prognosis of IPF. From 2020 to 2022, we recruited 36 patients (age: 75.3 ± 8.5; female: 19%) with IPF, and 80 control subjects (age: 72.3 ± 9.0; female: 27%). The mitochondrial genome of peripheral blood leukocytes was determined using next-generation sequencing. During a 45-month follow-up, 10 (28%) patients with IPF remained stable and the other 26 (72%) progressed, with 12 (33%) mortalities. IPF patients had more non-synonymous (NS) variants (substitution/deletion/insertion) in mitochondrial COX3 gene (coding for subunit 3 of complex IV of the respiratory chain), and more mitochondrial tRNA variants located in the anticodon (AC) stem, AC loop, variable loop, T-arm, and T-loop of the tRNA clover-leaf structure in PBLs than the control group. The succumbed IPF patients were older, had lower initial diffusion capacity, and higher initial fibrosis score on high-resolution computerized tomography (HRCT) than the alive group. NS variants in mitochondrial COX3 gene and tRNA variants in PBLs were associated with shorter survival. Our study shows that (1) leukocyte mitochondrial COX3 NS variants are associated with risk and prognosis of IPF; (2) leukocyte mitochondrial tRNA variants located in the AC stem, AC loop, variable loop, T-arm, and T-loop of the tRNA clover-leaf structure are associated with risk, and the presence of tRNA variants is associated with poor prognosis of IPF. Full article
(This article belongs to the Special Issue Advanced Molecular Research in Lung Diseases)
Show Figures

Figure 1

Figure 1
<p>The map of human mitochondrial DNA, a double-stranded circular DNA with H- and L-strands. It has 37 genes coding for 13 proteins of the respiratory chain (complex I, III, IV, and V), 2 rRNAs, and 22 tRNAs.</p>
Full article ">Figure 2
<p>Distribution of mitochondrial tRNA variants in regions of the secondary clover-leaf tRNA structure. Patients with IPF had more tRNA variants located in the right half of the clover-leaf structure (including anticodon loop, anticodon arm, variable loop, T-arm, and T loop) than the control subjects, with the majority of tRNA variants located in these regions (73% vs. 31% in the control group, <span class="html-italic">p</span> = 0.008). Numbering of nucleotides follows the standard system [<a href="#B19-ijms-26-01378" class="html-bibr">19</a>].</p>
Full article ">Figure 3
<p>Survival curves of patients with IPF, stratified by mitochondrial <span class="html-italic">COX3</span> gene NS variants (substitution/deletion/insertion) (<b>A</b>), mitochondrial tRNA variants (<b>B</b>), the presence of neither, either, or both <span class="html-italic">COX3</span> gene NS variants and mitochondrial tRNA variants (<b>C</b>), and anti-fibrotic treatment (<b>D</b>). For the 3 curves in (<b>C</b>): <span class="html-italic">p</span> = 0.127 when comparing between “Both mutated” and “Either one mutated”; <span class="html-italic">p</span> = 0.011 when comparing between “Either one mutated” and “Neither mutated”; <span class="html-italic">p</span> = 0.001 when comparing between “Both mutated” and “Neither mutated”. The four curves in (<b>D</b>) are: IPF patients treated with pirfenidone only, nintedanib only, pirfenidone and nintedanib sequentially (or vice versa), or not treated.</p>
Full article ">Figure 3 Cont.
<p>Survival curves of patients with IPF, stratified by mitochondrial <span class="html-italic">COX3</span> gene NS variants (substitution/deletion/insertion) (<b>A</b>), mitochondrial tRNA variants (<b>B</b>), the presence of neither, either, or both <span class="html-italic">COX3</span> gene NS variants and mitochondrial tRNA variants (<b>C</b>), and anti-fibrotic treatment (<b>D</b>). For the 3 curves in (<b>C</b>): <span class="html-italic">p</span> = 0.127 when comparing between “Both mutated” and “Either one mutated”; <span class="html-italic">p</span> = 0.011 when comparing between “Either one mutated” and “Neither mutated”; <span class="html-italic">p</span> = 0.001 when comparing between “Both mutated” and “Neither mutated”. The four curves in (<b>D</b>) are: IPF patients treated with pirfenidone only, nintedanib only, pirfenidone and nintedanib sequentially (or vice versa), or not treated.</p>
Full article ">
28 pages, 3097 KiB  
Review
Epigenetic Regulation by lncRNA GAS5/miRNA/mRNA Network in Human Diseases
by Lam Ngoc Thao Nguyen, Jaeden S. Pyburn, Nhat Lam Nguyen, Madison B. Schank, Juan Zhao, Ling Wang, Tabitha O. Leshaodo, Mohamed El Gazzar, Jonathan P. Moorman and Zhi Q. Yao
Int. J. Mol. Sci. 2025, 26(3), 1377; https://doi.org/10.3390/ijms26031377 - 6 Feb 2025
Viewed by 369
Abstract
The interplay between long noncoding RNAs (lncRNAs) and microRNAs (miRNAs) is crucial in the epigenetic regulation of mRNA and protein expression, impacting the development and progression of a plethora of human diseases, such as cancer, cardiovascular disease, inflammatory-associated diseases, and viral infection. Among [...] Read more.
The interplay between long noncoding RNAs (lncRNAs) and microRNAs (miRNAs) is crucial in the epigenetic regulation of mRNA and protein expression, impacting the development and progression of a plethora of human diseases, such as cancer, cardiovascular disease, inflammatory-associated diseases, and viral infection. Among the many lncRNAs, growth arrest-specific 5 (GAS5) has garnered substantial attention for its evident role in the regulation of significant biological processes such as proliferation, differentiation, senescence, and apoptosis. Through miRNA-mediated signaling pathways, GAS5 modulates disease progression in a cell-type-specific manner, typically by influencing proteins involved in inflammation and cell death. While GAS5 is recognized as a tumor suppressor in cancer, recent reports highlight its broader regulatory capacity in non-cancerous diseases. Its modulation of protein expression through the GAS5/miRNA network has been shown to both mitigate and exacerbate disease, depending on the specific context. Furthermore, the therapeutic potential of GAS5 manipulation, via knockdown or overexpression, offers promising avenues for targeted interventions across human diseases. This review explores the dualistic impacts of the GAS5/miRNA network in conditions such as cancer, cardiovascular disease, viral infections, and inflammatory disorders. Through the evaluation of current evidence, we aim to provide insight into GAS5’s biological functions and its implications for future research and therapeutic development. Full article
(This article belongs to the Special Issue Role of MicroRNAs in Human Diseases)
Show Figures

Figure 1

Figure 1
<p>The GAS5/miRNA-regulated pathways in cancers. This figure depicts the pathways modulated by the GAS5/miRNA-mediated mechanisms that regulate the progression of cancers. The upregulation of GAS5 results in the downregulation of its target miRNAs, thereby regulating downstream proteins involved in cellular activation, proliferation, differentiation or apoptosis, and cell cycle arrest. Thus, increasing GAS5 levels is considered a promising therapeutic approach for suppressing the progression of various cancers. PTEN, phosphatase and tensin homolog; AKT, protein kinase B; mTOR, mammalian target of rapamycin; SPRY1, Sprouty RTK signaling antagonist 1; RAF, rapidly accelerated fibrosarcoma; ERK, extracellular signal-regulated kinase; PDCD4, programmed cell death 4; P21, cyclin-dependent kinase inhibitor 1; SMAD7, SMAD family member 7; BIM, Bcl-2-like protein 11; BAX, BCL2-associated X; FOXO1, forkhead box protein O1; SOCS3, suppressor of cytokine signaling 3; ARHI, aplasia Ras homolog member I; DKK2, Dickkopf 2; ATG12, autophagy related 12; ATG3, autophagy related 3; ATG8, autophagy related 8; LC3II, LC3-phosphatidylethanolamine conjugate.</p>
Full article ">Figure 2
<p>The GAS5/miRNA-regulated pathways in cardiovascular diseases. This figure depicts the pathways modulated by GAS5/miRNA-mediated mechanisms that have either debilitating or protective effects on cardiovascular diseases. The pathways involved in modulating cell activity in a debilitating manner exert their effects, in general, through the suppression of cellular proliferation and differentiation while promoting cell death and apoptosis. The pathways involved in modulating cell activity in a protective manner exert their effects, in general, through the suppression of senescence, inflammation, or pyroptosis while promoting cellular proliferation. These pathways are largely cell-type- and context-dependent. PDCD4, programmed cell death protein 4; PTEN, phosphatase and tensin homolog; PI3K, phosphoinositide 3-kinase; AKT, protein kinase B; MEK, mitogen-activated protein kinase; ERK, extracellular signal-regulated kinase; ROCK1, rho-associated coiled-coil containing protein kinase 1; GSK-3β, glycogen synthase kinase 3 beta; TXNIP, thioredoxin-interacting protein; NAMPT, nicotinamide phosphoribosyltransferase; NLRP3, NOD-, LRR-, and pyrin domain-containing protein 3; TLR4, toll-like receptor 4; MI, myocardial infarction.</p>
Full article ">Figure 3
<p>The GAS5/miRNA-regulated pathways with debilitating roles in inflammation-associated diseases. This figure depicts the pathways modulated by GAS5/miRNA-mediated mechanisms that exert debilitating effects on inflammation-associated diseases. The proteins involved in these pathways are known for their roles in cell activation, proliferation, differentiation, apoptosis, and cell cycle arrest. By inhibiting specific miRNAs and their target proteins, elevated GAS5 levels promote apoptosis and cell death, thereby promoting inflammation and apoptosis, increasing the severity of inflammatory diseases. TSP-1, thrombospondin-1; GSK-3β, glycogen synthase kinase 3; PTEN, phosphatase and tensin homolog; NLRP3, NOD-like receptor family pyrin domain containing 3; BDNF, brain-derived neurotrophic factor; SMAD1, SMAD family member 1; FOXO3, forkhead box protein O3; PUMA, p53-upregulated modulator of apoptosis; JNK, c-Jun N-terminal kinase; γH2AX, phosphorylated histone H2AX; AKT, protein kinase B; Notch-1, neurogenic locus notch homolog protein 1; KCNQ3, potassium voltage-gated channel subfamily Q member 3.</p>
Full article ">Figure 4
<p>The GAS5/miRNA-regulated pathways with protective roles in inflammation-associated diseases. This figure depicts the pathways modulated by GAS5/miRNA-mediated mechanisms that exert protective effects and differential effects (sepsis) on inflammation-associated diseases. For sepsis, miRNAs appear to have a stage-dependent role in inflammation, where early sepsis relies on mitochondrial autophagy for antioxidant defense and intracellular stability, while GAS5 depletion in late-stage sepsis exacerbates inflammation, increasing tissue damage. Regarding the other disease models displayed, the overexpression of GAS5 leads to the repression of miRNAs displaying roles in inflammation, pyroptosis, proliferation, and apoptosis. This suppression of inflammation, apoptosis, and pyroptosis is significant in reducing the progression of these inflammation-associated diseases. MITF, microphthalmia-associated transcription factor; NRF2, nuclear factor erythroid 2–related factor 2; TLR4, toll-like receptor 4; SIRT1, Sirtuin 1; HMGB1, high mobility group box 1; MARCH7, membrane-associated ring-CH-type finger 7; NLRP3, NOD-like receptor family pyrin domain containing 3; PDK4, pyruvate dehydrogenase kinase 4; SMAD4, SMAD family member 4; GSDMD-N, gasdermin D N-terminal fragment; NAFLD, nonalcoholic fatty liver disease; OA, osteoarthritis; DN; diabetic nephropathy.</p>
Full article ">Figure 5
<p>The GAS5/miRNA-mediated regulation of viral infections. This figure depicts the mechanisms by which GAS5 regulates the progression of infectious diseases caused by viral infections. GAS5 plays a critical role in regulating infectious diseases associated with HCV and HIV infections by directly binding to viral proteins or miRNAs that promote viral replication. GAS5 also demonstrates therapeutic potential in PLWH, displaying an ability to improve the activity and longevity of CD4 T cells in ART-treated PLWH. Finally, GAS5 serves as an important biomarker of severe SARS-CoV-2 infection through its ability to function as a ceRNA for miR-200. Unbound miR-200 can inhibit ACE-2 expression, a phenomenon which is linked to increased inflammation and cytokine storms in SARS-CoV-2 infection. HCV, hepatitis C virus; HIV, human immunodeficiency virus; PLWH, patients living with HIV; ART, antiretroviral therapy; ACE-2, angiotensin-converting enzyme 2.</p>
Full article ">
32 pages, 1452 KiB  
Review
Modification in Structures of Active Compounds in Anticancer Mitochondria-Targeted Therapy
by Agnieszka Pyrczak-Felczykowska and Anna Herman-Antosiewicz
Int. J. Mol. Sci. 2025, 26(3), 1376; https://doi.org/10.3390/ijms26031376 - 6 Feb 2025
Viewed by 522
Abstract
Cancer is a multifaceted disease characterised by uncontrolled cellular proliferation and metastasis, resulting in significant global mortality. Current therapeutic strategies, including surgery, chemotherapy, and radiation therapy, face challenges such as systemic toxicity and tumour resistance. Recent advancements have shifted towards targeted therapies that [...] Read more.
Cancer is a multifaceted disease characterised by uncontrolled cellular proliferation and metastasis, resulting in significant global mortality. Current therapeutic strategies, including surgery, chemotherapy, and radiation therapy, face challenges such as systemic toxicity and tumour resistance. Recent advancements have shifted towards targeted therapies that act selectively on molecular structures within cancer cells, reducing off-target effects. Mitochondria have emerged as pivotal targets in this approach, given their roles in metabolic reprogramming, retrograde signalling, and oxidative stress, all of which drive the malignant phenotype. Targeting mitochondria offers a promising strategy to address these mechanisms at their origin. Synthetic derivatives of natural compounds hold particular promise in mitochondrial-targeted therapies. Innovations in drug design, including the use of conjugates and nanotechnology, focus on optimizing these compounds for mitochondrial specificity. Such advancements enhance therapeutic efficacy while minimizing systemic toxicity, presenting a significant step forward in modern anticancer strategies. Full article
Show Figures

Figure 1

Figure 1
<p>Development of novel anticancer drugs. Stages of drug creation are marked in green arrows. Descriptions of particular steps are provided below the arrows.</p>
Full article ">Figure 2
<p>Mitochondrial targets in anticancer drug design.</p>
Full article ">Figure 3
<p>Examples of structural modifications of natural compounds in mitochondrial-targeted anticancer therapy. Direct modifications in structures (<b>A</b>) and conjugates with a TPP moiety (<b>B</b>) are presented.</p>
Full article ">
21 pages, 5178 KiB  
Article
The Disruptions of Sphingolipid and Sterol Metabolism in the Short Fiber of Ligon-Lintless-1 Mutant Revealed Obesity Impeded Cotton Fiber Elongation and Secondary Cell Wall Deposition
by Huidan Tian, Qiaoling Wang, Xingying Yan, Hongju Zhang, Zheng Chen, Caixia Ma, Qian Meng, Fan Xu and Ming Luo
Int. J. Mol. Sci. 2025, 26(3), 1375; https://doi.org/10.3390/ijms26031375 - 6 Feb 2025
Viewed by 433
Abstract
Boosting evidence indicated lipids play important roles in plants. To explore lipid function in cotton fiber development, the lipid composition and content were detected by untargeted and targeted lipidomics. Compared with rapid elongation fibers, the lipid intensity of 16 sub-classes and 56 molecular [...] Read more.
Boosting evidence indicated lipids play important roles in plants. To explore lipid function in cotton fiber development, the lipid composition and content were detected by untargeted and targeted lipidomics. Compared with rapid elongation fibers, the lipid intensity of 16 sub-classes and 56 molecular species decreased, while only 7 sub-classes and 26 molecular species increased in the fibers at the stage of secondary cell wall deposition. Unexpectedly, at the rapid elongation stage, 20 sub-classes and 60 molecular species increased significantly, while only 5 sub-classes and 8 molecular species decreased in the ligon lintless-1 (li-1) mutant compared with its wild-type Texas Maker-1 (TM-1). Particularly, campesteryl, sitosteryl, and total steryl ester increased by 21.8-, 48.7-, and 45.5-fold in the li-1 fibers, respectively. All the molecular species of sphingosine-1-P, phytoceramide-OHFA, and glucosylceramide increased while all sphingosine, phytosphingosine, and glycosyl inositol phospho ceramides decreased in the li-1 fibers. Similarly, the different expression genes between the mutant and wild type were enriched in many pathways involved in the lipid metabolism. Furthermore, the number of lipid droplets also increased in the li-1 leaf and fiber cells when compared with the wild type. These results illuminated that fiber cell elongation being blocked in the li-1 mutant was not due to a lack of lipids, but rather lipid over-accumulation (obesity), which may result from the disruption of sphingolipid and sterol metabolism. This study provides a new perspective for further studying the regulatory mechanisms of fiber development. Full article
(This article belongs to the Section Molecular Plant Sciences)
Show Figures

Figure 1

Figure 1
<p>OPLS-DA score plot and the number of lipid species in each detected lipid class. (<b>A</b>): The OPLS-DA score plot between the 10-DPA fiber group and 20-DPA fiber group, * represents the multiplication sign. (<b>B</b>): The OPLS-DA score plot between the 10-DPA fiber group of wild type (TM-1) and the 10-DPA fiber group of li-1 mutant. (<b>C</b>): Thirty-three lipid classes detected in three samples and the number of lipid molecule species in each detected lipid class. The number on each column represents the number of molecular species in each lipid class. 10 D: 10-DPA fibers of TM-1 (wild type); 20 D: 20-DPA fibers of TM-1 (wild type); 10 D-li: 10-DPA fibers of <span class="html-italic">li-1</span> mutant. AGlcSiE, AcylGlcSitosterol ester; Cer, ceramides; CerG1, glucocerebroside; CerP, ceramides phosphate; CL, cardiolipin; Co, coenzyme; DG, diglyceride; DGMG, Digalactosylmonoacylglycerol; DGDG, digalactosyldiacylglycerol; FA, fatty acid; LPA, lysophosphatidic acid; LPC, lysophosphatidylcholine; LPE, lysophosphatidylethanolamine; LPI, lysophosphatidylinositol; MG, monoglyceride; MGDG, monogalactosyldiacylglycerol; MGMG, Monogalactosylmonoacylglycerol; OAHFA, O-Acetylated Hydroxy Fatty Acid. PA, phosphatidic acid; PC, phosphatidylserine; PE, phosphatidylcholine; PG, phosphatidylglycerol; phSM, phytosphingosine; PI, phosphatidylinositol; PS, phosphatidylserine; SiE, sitosterol ester; SM, sphingomyelin; So, sphingosine; MG, monoglyceride; SQDG, Sulfoquinovosyldiacylglycerol; StE, stigmasterol ester; TG, triglyceride; WE, wax esters.</p>
Full article ">Figure 2
<p>The lipid intensity of various lipid classes. The intensity of 7 classes of lipids including GP, SP, GL, SL, ST, FA, and PL and 33 kinds of lipid sub-classes in the fiber cell. GP, glycerophospholipid; SP, sphingolipid; GL, glycerolipid; ST, sterol lipids; PL, prenol lipid; FA, fatty acid; SL, saccharolipid; AGlcSiE, AcylGlcSitosterol ester; Cer, ceramides; CerG1, glucocerebroside; CerP, ceramides phosphate; DG, diglyceride; CL, cardiolipin; DGMG, Digalactosylmonoacylglycerol; DGDG, digalactosyldiacylglycerol; LPA, lysophosphatidic acid; LPC, lysophosphatidylcholine; LPE, lysophosphatidylethanolamine; LPI, lysophosphatidylinositol; MG, monoglyceride; MGDG, monogalactosyldiacylglycerol; MGMG, Monogalactosylmonoacylglycerol; OAHFA, O-Acetylated Hydroxy Fatty Acid; PA, phosphatidic acid; PC, phosphatidylserine; PE, phosphatidylcholine; PG, phosphatidylglycerol; phSM, phytosphingosine; PI, phosphatidylinositol; PS, phosphatidylserine; SiE, sitosterol ester; SM, sphingomyelin; So, sphingosine; SQDG, Sulfoquinovosyldiacylglycerol; StE, stigmasterol ester; TG, triglyceride; WE, wax esters; Co, coenzyme.</p>
Full article ">Figure 3
<p>The intensity difference in lipid sub-classes and lipid molecule species between 10-DPA fiber cells and 20-DPA fiber cells. (<b>A</b>): The fold changes in various lipid sub-classes between 10-DPA fiber cells and 20-DPA fiber cells; (<b>B</b>): the fold changes in various lipid molecule species between 10-DPA fiber cells and 20-DPA fiber cells; (<b>C</b>): the proportion of various lipid compounds in the total significantly changed lipids; (<b>D</b>): the proportion of various lipid molecule species in the total significantly changed lipids. 10 DPA: 10-DPA fibers of TM-1 (wild type); 20 DPA: 20-DPA fibers of TM-1 (wild type).</p>
Full article ">Figure 4
<p>The intensity difference in lipid sub-classes and lipid molecule species between 10-DPA fiber cells of TM-1 (wild type) and 10-DPA fiber cells of <span class="html-italic">li-1</span> mutant. (<b>A</b>): The fold changes in various lipid sub-classes between 10-DPA fiber cells of TM-1 and 10-DPA fiber cells of <span class="html-italic">li-1</span> mutant; (<b>B</b>): the fold changes in various lipid molecule species between 10-DPA fiber cells of TM-1 and 10-DPA fiber cells of li-1 mutant; (<b>C</b>): the proportion of various lipid compounds in the total significantly changed lipids; (<b>D</b>): the proportion of various lipid molecule species in the total significantly changed lipids. 10 DPA: 10-DPA fibers of TM-1 (wild type); 10 <span class="html-italic">li-1</span>: 10-DPA fibers of <span class="html-italic">li-1</span> mutant.</p>
Full article ">Figure 5
<p>Sphingolipid classes in cotton fiber cells and their alteration in <span class="html-italic">li-1</span> mutant fiber cells compared with its wild-type TM-1. The change ratio represents the percentage of increase and decrease in sphingolipid classes and molecular species in <span class="html-italic">li-1</span> fiber cells compared to wild-type fiber cells. (<b>A</b>): The number of classes and molecular species of sphingolipids detected in fiber cells of <span class="html-italic">li-1</span> and TM-1; (<b>B</b>): the change percentage of sphingolipid content in <span class="html-italic">li-1</span> fiber cells; (<b>C</b>): the change percentage of molecular species of t-S1P and Sph; (<b>D</b>): the change percentage of molecular species of Cer; (<b>E</b>): the change percentage of molecular species of Phyto-Cer; (<b>F</b>): the change percentage of molecular species of GIPC; (<b>G</b>): the change percentage of molecular species of PhytoCer-OHFA; (<b>H</b>): the change percentage of molecular species of GluCer. Cer, ceramides; PhytoCer, phytoceramides; PhytoCer-OHFA, phytoceramides with hydroxylated fatty acyls; S1P, sphingosine-1-phosphate; t-S1P, phytosphingosine-1-phosphate; Sph, sphingosines; PhytoSph, phytosphingosines; GluCer, glucosylceramides; Phyto-GluCer, phyto-glucosylceramides; GIPC, glycosyl inositol phospho ceramides.</p>
Full article ">Figure 6
<p>The content changes in sterol and steryl ester in <span class="html-italic">li-1</span> fiber cells. (<b>A</b>) The content changes in total sterol and various sterol classes in <span class="html-italic">li-1</span> fiber cells. (<b>B</b>) The content changes in total steryl ester and two steryl esters in <span class="html-italic">li-1</span> fiber cells. (<b>C</b>) The ratio of stigmasterol to sitosterol (St/Si) and campesterol to sitosterol (C/S) in 10-DPA fibers of mutant and wild type.</p>
Full article ">Figure 7
<p>GO annotations and KEGG enrichment analysis for the differentially expressed genes in 10-DPA fiber cells between <span class="html-italic">li-1</span> mutant and TM-1 wild type. Red arrows indicated the pathway involved in lipid metabolism.</p>
Full article ">Figure 8
<p>The expression changes in selected genes in <span class="html-italic">li-1</span> mutant fiber cells. Gh_D12G0217, LAG1 homologue 2; Gh_A07G0513, LAG1 longevity assurance homolog 3; Gh_D10G0211, Lactosylceramide 4-alpha-galactosyltransferase; Gh_A06G0144, phospholipase A 2A; Gh_A02G0884, GDSL-like Lipase/Acylhydrolase superfamily protein; Gh_D03G1074 and Gh_A05G3810, HXXXD-type acyl-transferase family protein; Gh_A08G1600, phospholipase D beta 1; Gh_A01G1605, alcohol dehydrogenase 1; Gh_D06G2376, 3-ketoacyl-CoA synthase 19. Error bars represent the SD for three independent experiments and asterisks indicate statistically significant differences between <span class="html-italic">li-1</span> and TM-1 fiber cells, as determined by Student’s <span class="html-italic">t</span>-test (** <span class="html-italic">p</span> &lt; 0.01).</p>
Full article ">Figure 9
<p>Oil bodies in the leaf and fiber cell of TM-1 and <span class="html-italic">li-1</span> mutant. (<b>A</b>) Oil bodies in the leaf of TM-1 and <span class="html-italic">li-1</span> mutant; (<b>B</b>) oil bodies in the fiber cell of TM-1 and <span class="html-italic">li-1</span> mutant. <span class="html-italic">Li-1</span>, <span class="html-italic">li-1</span> mutant; TM-1, wild type; UV, ultraviolet light; BF, bright light.</p>
Full article ">
11 pages, 4355 KiB  
Case Report
Peripheral Blood Mononuclear Cells Cytokine Profile in a Patient with Toxic Epidermal Necrolysis Triggered by Lamotrigine and COVID-19: A Case Study
by Margarita L. Martinez-Fierro, Idalia Garza-Veloz, Sidere Monserrath Zorrilla-Alfaro, Andrés Eduardo Campuzano-Garcia and Monica Rodriguez-Borroel
Int. J. Mol. Sci. 2025, 26(3), 1374; https://doi.org/10.3390/ijms26031374 - 6 Feb 2025
Viewed by 370
Abstract
Stevens–Johnson Syndrome (SJS)/toxic epidermal necrolysis (TEN) is a severe mucocutaneous reaction often induced by medications. The co-occurrence of SJS/TEN and COVID-19 presents a unique challenge due to overlapping inflammatory pathways. This case study examined the cytokine profile of a patient with both TEN [...] Read more.
Stevens–Johnson Syndrome (SJS)/toxic epidermal necrolysis (TEN) is a severe mucocutaneous reaction often induced by medications. The co-occurrence of SJS/TEN and COVID-19 presents a unique challenge due to overlapping inflammatory pathways. This case study examined the cytokine profile of a patient with both TEN (triggered by lamotrigine) and COVID-19. The clinical history of the patient, including lamotrigine exposure and COVID-19 diagnosis, was documented. A 13-cytokine profile assessment was performed in peripheral blood mononuclear cells from the patient and their parents by using quantitative Real Time-Polymerase Chain Reaction (qRT-PCR). A 6-year-old male patient developed lamotrigine-induced TEN with concomitant COVID-19 affecting 90% of the body surface area. Compared with their parents, who were positive for COVID-19, IL-6, IL-4, and IL-12 were modulated (downregulated) by TEN. The cytokine profile showed elevated levels of IL-1α, IL-1β, IL-5, IL-8, NF-κβ, and interferons (IFN; α, β, and γ), indicating a robust antiviral response. The immune profile suggested a hyperactivated immune state that contributed to the severity of the patient’s clinical manifestations, leading to death 18 days after hospitalization. Understanding the immune response is important for developing future targeted therapeutic strategies and improving patient outcomes. Further research is needed to explore the interaction between drug-induced SJS/TEN and infections. Full article
(This article belongs to the Special Issue Targeted Therapy for Immune Diseases)
Show Figures

Figure 1

Figure 1
<p>Toxic epidermal necrolysis clinical spectrum. (<b>A</b>–<b>D</b>): exanthematous rash. Lesions start on the face and thorax before spreading to other areas and are symmetrically distributed. Early lesions typically begin with ill-defined, coalescing, erythematous macules; (<b>E</b>–<b>H</b>): extensive, sheet-like detachment and erosions, and Nikolsky sign is present.</p>
Full article ">Figure 2
<p>Expression profile of evaluated cytokines. Figure shows the expression level of a 13-citokine panel for the patient with toxic epidermal necrolysis (TEN) and concurrent COVID-19, and for his parents, grouped by immune response in which they participate (<b>A</b>); and according with their status of overexpression and underexpression profile (<b>B</b>). Expression levels were calculated by quantitative real-time polymerase chain reaction by using GAPDH as endogenous control and RNA of peripheral blood mononuclear cell obtained from healthy controls (with negative qRT-PCR for SARS-CoV-2) as calibrator.</p>
Full article ">Figure 3
<p>Cell and immune pathophysiology of TEN. Infiltration of the epidermis by activated T lymphocytes (CD8+ epidermis; CD4+ dermis) and natural killer cells induce an immune response against the drug-reactive metabolites. TCRs recognize the molecules and produce interleukins (mainly TNFα) which cause epidermal detachment secondary to keratinocyte apoptosis induced by granzymes, perforins, and Fas/Fas ligand. MHC-II: major histocompatibility complex class II; TCR: T lymphocyte receptor; IFNγ: interferon gamma; TNFα: tumor necrosis factor alpha; IL: interleukin.</p>
Full article ">
26 pages, 2623 KiB  
Review
The Role of Gut Microbiota-Derived Trimethylamine N-Oxide in the Pathogenesis and Treatment of Mild Cognitive Impairment
by Haihua Xie, Jia Jiang, Sihui Cao, Xuan Xu, Jingyin Zhou, Ruhan Zhang, Bo Huang, Penghui Lu, Liang Peng and Mi Liu
Int. J. Mol. Sci. 2025, 26(3), 1373; https://doi.org/10.3390/ijms26031373 - 6 Feb 2025
Viewed by 490
Abstract
Mild cognitive impairment (MCI) represents a transitional stage between normal aging and dementia, often considered critical for dementia prevention. Despite its significance, no effective clinical treatment for MCI has yet been established. Emerging evidence has demonstrated a strong association between trimethylamine-N-oxide (TMAO), a [...] Read more.
Mild cognitive impairment (MCI) represents a transitional stage between normal aging and dementia, often considered critical for dementia prevention. Despite its significance, no effective clinical treatment for MCI has yet been established. Emerging evidence has demonstrated a strong association between trimethylamine-N-oxide (TMAO), a prominent metabolite derived from the gut microbiota, and MCI, highlighting its potential as a biomarker and therapeutic target. TMAO has been implicated in increasing MCI risk through its influence on factors such as hypertension, cardiovascular disease, depression, diabetes, and stroke. Moreover, it contributes to MCI by promoting oxidative stress, disrupting the blood–brain barrier, impairing synaptic plasticity, inducing inflammation, causing mitochondrial metabolic disturbances, and facilitating abnormal protein aggregation. This review further explores therapeutic strategies targeting TMAO to mitigate MCI progression. Full article
(This article belongs to the Section Molecular Microbiology)
Show Figures

Graphical abstract

Graphical abstract
Full article ">Figure 1
<p>Origins and Excretion of TMAO. TMAO is absorbed directly from dietary sources through the intestines. Exogenous TMAO is subsequently produced via oxidation by the gut microbiome and the liver. The primary excretion pathways for TMAO include urine, feces, and respiration. Abbreviations: TMA, trimethylamine; TMAO, trimethylamine N-oxide; FMO, flavin-containing monooxygenase; OCT, organic cation transporter.</p>
Full article ">Figure 2
<p>Contributions of TMAO to the pathogenesis of MCI. TMAO potentially contributes to the pathogenesis of MCI by promoting oxidative stress, neuroinflammation, and abnormal protein accumulation. TMAO induces oxidative stress by enhancing the production of reactive oxygen species (ROS) and reducing antioxidant activity. It also triggers neuroinflammation by activating NF-κβ and the NLRP3 inflammasome. Furthermore, TMAO exacerbates the formation of amyloid plaques and neurofibrillary tangles by impairing the intracellular ubiquitin–proteasome system. Abbreviations: TMAO, trimethylamine N-oxide; GSH, glutathione; GPX, glutathione peroxidase; SOD, superoxide dismutase; MsrA, methionine sulfoxide reductase A; NF-κB, nuclear factor kappa B; NLRP3, NOD-like receptor family pyrin domain containing 3; Sirt3, sirtuin 3; mtROS, mitochondrial reactive oxygen species; IL, interleukin; TXNIP, thioredoxin-interacting protein; NFTs, neurofibrillary tangles.</p>
Full article ">Figure 3
<p>Effects of TMAO on the blood–brain barrier and synaptic plasticity. TMAO impairs the structural integrity and function of the blood–brain barrier (BBB) and reduces synaptic plasticity, contributing to the pathogenesis of MCI. It reduces hippocampal synaptic plasticity by activating the PI3K/Akt/mTOR and PERK signaling pathways. Simultaneously, TMAO disrupts the BBB, facilitating the accumulation of neurotoxic molecules in the brain and inducing oxidative stress and neuroinflammation. Abbreviations: SYN, synaptophysin; NMDAR, N-methyl-D-aspartate receptor; GluA1, glutamate receptor ionotropic AMPA 1; GluN2A, glutamate receptor ionotropic NMDA 2A; PSD95, postsynaptic density protein 95; PERK, protein kinase R-like endoplasmic reticulum kinase; ATF4, activating transcription factor 4; CREB, cAMP response element-binding protein; p-PI3K, phosphorylated phosphoinositide 3-kinase; p-Akt, phosphorylated Akt protein; p-mTOR, phosphorylated mammalian target of rapamycin; ZO-1, zonula occludens-1; PDGFRβ, platelet-derived growth factor receptor beta.</p>
Full article ">Figure 4
<p>Effects of TMAO on mitochondrial metabolism. TMAO adversely affects mitochondrial metabolism, contributing to the pathogenesis of MCI. It significantly inhibits the oxidation of pyruvate and fatty acids in mitochondria, leading to energy metabolism disorders. Abbreviations: CAT, carnitine acylcarnitine translocase; CPT2, carnitine palmitoyl transferase II; TCA cycle, tricarboxylic acid cycle; IMM, inner mitochondrial membrane; OMM, outer mitochondrial membrane; ATP, adenosine triphosphate.</p>
Full article ">Figure 5
<p>Treatment Strategy for improving MCI by affecting TMAO.</p>
Full article ">
24 pages, 1083 KiB  
Review
Correlations Between Gut Microbiota Composition, Medical Nutrition Therapy, and Insulin Resistance in Pregnancy—A Narrative Review
by Robert-Mihai Enache, Oana Alexandra Roşu, Monica Profir, Luciana Alexandra Pavelescu, Sanda Maria Creţoiu and Bogdan Severus Gaspar
Int. J. Mol. Sci. 2025, 26(3), 1372; https://doi.org/10.3390/ijms26031372 - 6 Feb 2025
Viewed by 480
Abstract
Many physiological changes accompany pregnancy, most of them involving metabolic perturbations. Alterations in microbiota composition occur both before and during pregnancy and have recently been correlated with an important role in the development of metabolic complications, such as insulin resistance and gestational diabetes [...] Read more.
Many physiological changes accompany pregnancy, most of them involving metabolic perturbations. Alterations in microbiota composition occur both before and during pregnancy and have recently been correlated with an important role in the development of metabolic complications, such as insulin resistance and gestational diabetes mellitus (GDM). These changes may be influenced by physiological adaptations to pregnancy itself, as well as by dietary modifications during gestation. Medical nutritional therapy (MNT) applied to pregnant women at risk stands out as one of the most important factors in increasing the microbiota’s diversity at both the species and genus levels. In this review, we discuss the physiological changes during pregnancy and their impact on the composition of the intestinal microbiota, which may contribute to GDM. We also discuss findings from previous studies regarding the effectiveness of MNT in reducing insulin resistance. In the future, additional studies should aim to identify specific gut microbial profiles that serve as early indicators of insulin resistance during gestation. Early diagnosis, achievable through stool analysis or metabolite profiling, may facilitate the timely implementation of dietary or pharmaceutical modifications, thereby mitigating the development of insulin resistance and its associated sequelae. Full article
(This article belongs to the Special Issue Molecular Insight into Gestational Diabetes Mellitus)
Show Figures

Figure 1

Figure 1
<p>During pregnancy, various physiological changes occur in the endocrine, cardiovascular, respiratory, renal, and digestive systems. These changes lead to hormonal, immunological, and metabolic alterations that impact the gut microbiota composition. The intestinal microbiome, in turn, can influence these changes, affecting normal pregnancy progression. An inappropriate diet or lifestyle during pregnancy can disrupt the normal gut microbiota (dysbiosis), negatively influencing maternal and fetal health and increasing pregnancy risks. Conversely, appropriate manipulation of the microbiota through dietary and lifestyle adaptations promotes a eubiotic state, enhancing the health of both the mother and the fetus during pregnancy. Created with <a href="http://BioRender.com" target="_blank">BioRender.com</a> (accessed on 6 October 2024).</p>
Full article ">Figure 2
<p>During pregnancy, early increases in adiposity and decreased insulin sensitivity typically emerge by the third trimester. Physiological increases in insulin secretion help regulate glucose levels. However, changes in gut microbiota composition from the first to the third trimester can result in dysbiosis, leading to elevated levels of LPSs from Gram-negative bacteria and immune changes, such as higher levels of TNF-α and IL-6, which contribute to a pro-inflammatory state. When insulin secretion becomes insufficient, combined with this inflammatory environment, maternal hyperglycemia and GDM may develop. Created with <a href="http://BioRender.com" target="_blank">BioRender.com</a> (accessed on 9 October 2024).</p>
Full article ">
14 pages, 4642 KiB  
Article
Dynamics of Immune Cell Infiltration and Fibroblast-Derived IL-33/ST2 Axis Induction in a Mouse Model of Post-Surgical Lymphedema
by Kazuhisa Uemura, Kei-ichi Katayama, Toshihiko Nishioka, Hikaru Watanabe, Gen Yamada, Norimitsu Inoue and Shinichi Asamura
Int. J. Mol. Sci. 2025, 26(3), 1371; https://doi.org/10.3390/ijms26031371 - 6 Feb 2025
Viewed by 371
Abstract
Lymphedema is an intractable disease most commonly associated with lymph node dissection for cancer treatment and can lead to a decreased quality of life. Type 2 T helper (Th2) lymphocytes have been shown to be important in the progression of lymphedema. The activation [...] Read more.
Lymphedema is an intractable disease most commonly associated with lymph node dissection for cancer treatment and can lead to a decreased quality of life. Type 2 T helper (Th2) lymphocytes have been shown to be important in the progression of lymphedema. The activation of IL-33 and its receptor, the suppression of tumorigenicity 2 (ST2) signaling pathway, induces the differentiation of Th2 cells, but its involvement in lymphedema remains unclear. In the present study, we analyzed the dynamics of immune cell infiltration, including the IL-33/ST2 axis, in a mouse tail lymphedema model. Neutrophil infiltration was first detected in the lymphedema tissue on postoperative day (POD) 2. Macrophage infiltration increased from POD 2 to 5. The number of CD4+ T cells, including 50% Tregs, gradually increased from POD 14. The mRNA expression of ll13 and Ifng increased on POD 21. The expression of IL-33 was induced in fibroblast nuclei within dermal and subcutaneous tissues from POD 2, and the expression of the Il1rl1 gene encoding ST2 increased from POD 7. We demonstrated the infiltration process from innate to acquired immune cells through the development of a mouse tail lymphedema. The IL-33/ST2 axis was found to be induced during the transition from innate to acquired immunity. Full article
(This article belongs to the Special Issue Lymphedema: From Mechanism to Treatment)
Show Figures

Figure 1

Figure 1
<p>Tail dermal and subcutaneous tissue thickness in the tail lymphedema tissue in non-operated controls (Ctrl) and on postoperative days (PODs) 2, 5, 7, 14, 21, and 42. (<b>A</b>) Representative images of the tail lymphedema on different PODs. (<b>B</b>) Time course of tail thickness after surgery in C57BL/6 J mice (Ctrl: n = 3; PODs 2, 5, 7, 14, and 21: n = 8; POD 42: n = 6). Boxes represent 50% of the data, with medians (lines) and interquartile ranges (whiskers). The Steel–Dwass test was used to determine which time points were significantly different from POD 2. * <span class="html-italic">p</span> &lt; 0.05. (<b>C</b>) Representative histological images of lymphedema tissues stained with Masson’s trichrome, with collagen fibers stained blue and muscle fibers stained red. Dermal and subcutaneous tissue is indicated by the double-headed arrow.</p>
Full article ">Figure 2
<p>Infiltration of CD4<sup>+</sup> T cells and spatial relations of CD4<sup>+</sup> T cells and lymphatic vessels. (<b>A</b>) Representative images of CD4<sup>+</sup> T cell (green) infiltrating the lymphedema tissues (Ctrl and PODs 2, 5, 7, 14, 21, and 42). Nuclei were stained with 4′,6-diamidino-2-phenylindole (DAPI) (blue). Yellow arrowheads indicate CD4<sup>+</sup> T cells. Scale bar = 100 μm. (<b>B</b>) Variation in the numbers of CD4<sup>+</sup> T cells per field (8 fields/mouse) infiltrating the lymphedema tissue (Ctrl: n = 3; PODs 2, 5, 7, 14, and 21: n = 8; POD 42: n = 6). Boxes represent 50% of the data, with medians (lines) and interquartile ranges (whiskers). The Steel–Dwass test was used to determine which time points were significantly different from POD 2. * <span class="html-italic">p</span> &lt; 0.05. (<b>C</b>) The spatial relations of CD4<sup>+</sup> T cells and lymphatic vessels on postoperative days (PODs) 14, 21, and 42. Histograms indicate the number of CD4<sup>+</sup> T cells localized at various distances from lymphatic vessels. The distance between CD4<sup>+</sup> T cells present in eight fields/mouse, and lymphatic vessels were measured at each time point (443 cells on POD 14, n = 8; 764 cells on POD 21, n = 8; 1153 cells on POD 42, n = 8). (<b>D</b>) Averages of distances between CD4<sup>+</sup> T cells and lymphatic vessels at each time point. Boxes represent 50% of the data, with medians (lines) and interquartile ranges (whiskers). The Steel–Dwass test was used to determine the time points that were significantly different from POD 14. * <span class="html-italic">p</span> &lt; 0.05.</p>
Full article ">Figure 3
<p>Infiltration of CD11c<sup>+</sup> dendritic cells in the tail lymphedema tissue in non-operated controls (Ctrl) and on postoperative days (PODs) 2, 5, 7, 14, 21, and 42. (<b>A</b>) Representative images of CD11c<sup>+</sup> dendritic cells (green) infiltrating the lymphedema tissues (Ctrl and PODs 2, 5, 7, 14, 21, and 42). Nuclei were stained with 4′,6-diamidino-2-phenylindole (DAPI) (blue). Yellow arrowheads indicate CD11c<sup>+</sup> dendritic cells. Scale bar = 100 μm. (<b>B</b>) Variation in the numbers of CD11c<sup>+</sup> dendritic cells per field (8 fields/mouse) infiltrating the lymphedema tissue (Ctrl: n = 3; POD 2, 5, 7, 14, and 21: n = 8; POD 42: n = 6). Boxes represent 50% of the data, with medians (lines) and interquartile ranges (whiskers). The Steel–Dwass test was used to determine the time points that were significantly different from POD 2. * <span class="html-italic">p</span> &lt; 0.05.</p>
Full article ">Figure 4
<p>Expression of cytokines in the tail lymphedema tissue in non-operated controls (Ctrl) and on postoperative days (PODs) 2, 7, and 21. Relative expression levels of <span class="html-italic">ll-13</span> (<b>A</b>), <span class="html-italic">Ifng</span> (<b>B</b>), and <span class="html-italic">Csf2</span> (<b>C</b>) mRNA in tail lymphedema tissues (PODs 2,7, and 21: n = 4) were compared with those in non-operated controls (Ctrl) (n = 3). <span class="html-italic">Ifng</span> expression levels were below the detection sensitivity in the Ctrl group and, therefore, compared to POD 2. Relative expression levels of mRNA were normalized to 18s rRNA expression levels and were calculated with the ΔΔCT method as the ratio of the averages of expression levels of non-operated control or POD 2. Data are shown as mean ± standard error. The Games–Howell test was used for statistical analyses.</p>
Full article ">Figure 5
<p>Infiltration of Ly6G<sup>+</sup> neutrophils and F4/80<sup>+</sup> macrophages in the tail lymphedema tissue in non-operated controls (Ctrl) and on postoperative days (PODs) 2, 5, 7, 14, 21, and 42. (<b>A</b>) Representative images of Ly6G<sup>+</sup> neutrophils (green) infiltrating the lymphedema tissues (Ctrl and PODs 2, 5, 7, 14, 21, and 42). (<b>B</b>) Variation in the numbers of Ly6G<sup>+</sup> neutrophils per field (8 fields/mouse) infiltrating the lymphedema tissue (Ctrl: n = 3; PODs 2, 5, 7, 14, and 21: n = 8; POD 42: n = 6). Boxes represent 50% of the data, with medians (lines) and interquartile ranges (whiskers). (<b>C</b>) Representative images of F4/80+ macrophages (green) infiltrating the lymphedema tissues (Ctrl and PODs 2, 5, 7, 14, 21, and 42). (<b>D</b>) Variation in the numbers of F4/80+ macrophages per field (8 fields/mouse) infiltrating the lymphedema tissue (Ctrl: n = 3; PODs 2, 5, 7, 14, and 21: n = 8; POD 42: n = 6). Boxes represent 50% of the data, with medians (lines) and interquartile ranges (whiskers). The Steel–Dwass test was used to determine which time points were significantly different from POD 2. * <span class="html-italic">p</span> &lt; 0.05. Nuclei in (<b>A</b>,<b>C</b>) were stained with 4′,6-diamidino-2-phenylindole (DAPI) (blue). Scale bar = 100 μm.</p>
Full article ">Figure 6
<p>Infiltration of interleukin (IL)-33<sup>+</sup> cells and express of <span class="html-italic">Il-33/Il1rl1</span> in the tail lymphedema tissue. (<b>A</b>) Representative images of IL-33<sup>+</sup> cells (green) in the lymphedema tissues (non-operated controls [Ctrl] and postoperative days [PODs] 2, 5, 7, 14, 21, and 42). Nuclei are stained with 4′,6-diamidino-2-phenylindole (DAPI) (blue). Yellow arrowheads indicate IL-33<sup>+</sup> cells in dermal and subcutaneous tissues. Scale bar = 100 μm. (<b>B</b>) Variation in the numbers of IL-33<sup>+</sup> cells per field (8 fields/mouse) in the lymphedema tissue (Ctrl: n = 3; PODs 2, 5, 7, 14, and 21: n = 8; POD 42: n = 6). Data are shown as the means ± standard error. The Tukey test was used to determine which time points were significantly different from Ctrl. * <span class="html-italic">p</span> &lt; 0.05. Relative expression levels of <span class="html-italic">Il-33</span> (<b>C</b>) and <span class="html-italic">Il1rl1</span> (<b>D</b>) mRNA in tail lymphedema tissues (PODs 2, 7, and 21; n = 4) were compared with those in Ctrl tissues (n = 3). Relative mRNA expression levels were normalized to 18s rRNA expression levels and calculated as the ratio of the average levels in non-operated controls. Data are shown as mean ± standard error. Tukey’s test was used to determine the time points that were significantly different from those of the Ctrl group. * <span class="html-italic">p</span> &lt; 0.05.</p>
Full article ">Figure 7
<p>Expression of interleukin (IL)-33 in fibroblast within the lymphedema tissue on postoperative day (POD) 2. The tail lymphedema tissues on POD 2 were co-stained with IL-33 (magenta) and CD45 (<b>A</b>–<b>C</b>), α smooth muscle actin (SMA) (<b>D</b>–<b>F</b>), vimentin (<b>G</b>–<b>I</b>), heat shock protein (HSP) 47 (<b>J</b>–<b>L</b>), and S100A4 (<b>M</b>–<b>O</b>) (green). Nuclei were stained with 4′,6-diamidino-2-phenylindole (DAPI) (blue). Yellow arrowheads indicate IL-33 and fibroblast marker double-positive cells. Scale bar = 100 μm.</p>
Full article ">
20 pages, 10581 KiB  
Article
Phylogeny of Camphora and Cinnamomum (Lauraceae) Based on Plastome and Nuclear Ribosomal DNA Data
by Jian Xu, Haorong Zhang, Fan Yang, Wen Zhu, Qishao Li, Zhengying Cao, Yu Song and Peiyao Xin
Int. J. Mol. Sci. 2025, 26(3), 1370; https://doi.org/10.3390/ijms26031370 - 6 Feb 2025
Viewed by 380
Abstract
Camphora Fabr. is a genus in the family Lauraceae, comprising over 20 tropical and subtropical tree species. Since the genera Camphora and Cinnamomum Schaeff. were described, there has been a long-lasting controversy regarding the phylogenetic relationships among taxa in both genera. In particular, [...] Read more.
Camphora Fabr. is a genus in the family Lauraceae, comprising over 20 tropical and subtropical tree species. Since the genera Camphora and Cinnamomum Schaeff. were described, there has been a long-lasting controversy regarding the phylogenetic relationships among taxa in both genera. In particular, phylogenetic inferences derived from plastid data remain debated, with varying hypotheses proposed and occasional disputes concerning the monophyly of Camphora taxa. To further investigate the relationships, We analyzed plastomes and nuclear ribosomal cistron sequences (nrDNA) of 22 Camphora taxa, 15 Cinnamomum taxa, and 13 representative taxa of related genera. The Camphora plastomes range from 152,745 to 154,190 bp, with a GC content of 39.1% to 39.2%. A total of 128 genes were identified in the Camphora plastomes, including 84 protein-coding genes, 8 rRNA genes, and 36 tRNA genes. A total of 1130 SSR loci were detected from plastomes of Camphora, and A/T base repeats looked like the most common. Comparative analyses revealed that the plastomes of Camphora exhibit high similarity in overall structure. The loci ycf1, ycf2, trnK (UUU), psbJ-psbL, and ccsA-ndhD were identified as candidate DNA barcodes for these taxa. Plastome phylogenetic analysis revealed that Camphora is not monophyletic, whereas the nrDNA dataset supported the monophyly of Camphora. We propose that intergeneric hybridization may underlie the observed discordance between plastid and nuclear data in Camphora, and we recommend enhanced taxonomic sampling and precise species identification to improve phylogenetic resolution and accuracy. Full article
(This article belongs to the Section Molecular Genetics and Genomics)
Show Figures

Figure 1

Figure 1
<p>Circular gene map of <span class="html-italic">Camphora</span> taxa plastomes. Genes shown inside and outside the circle are transcribed in clockwise and counterclockwise directions, respectively. Genes belonging to different functional groups are color-coded. The GC and AT content are denoted by the dark gray and light gray colors in the inner circle, respectively. * indicates that the gene has introns.</p>
Full article ">Figure 2
<p>Repeats of the 22 newly sequenced plastomes. C represents complement repeats, F represents forward repeats, P represents palindromic repeats, and R represents reverse repeats.</p>
Full article ">Figure 3
<p>Statistics of SSR loci for plastomes of 22 species in the genus <span class="html-italic">Camphora.</span> The left vertical coordinate represents the number of SSR loci, the right vertical axis and different color columns represent different species in the genus <span class="html-italic">Camphora</span>, and the horizontal axis represents different single nucleotide repeat sequences.</p>
Full article ">Figure 4
<p>Nucleic acid variability in 22 species of the genus <span class="html-italic">Camphora</span>. The vertical axis indicates the nucleotide polymorphism value (Pi), and the horizontal axis indicates the sequence length. <span class="html-italic">Camphora</span> V represents the species of <span class="html-italic">Camphora</span> classified in the genus <span class="html-italic">Cinnamomum</span>.</p>
Full article ">Figure 5
<p>Visualization of the boundaries of the IR region of chloroplast genes in 22 species of the genus <span class="html-italic">Camphora</span>.</p>
Full article ">Figure 6
<p>Phylogenetic tree based on plastome sequences. (<b>Cp</b>) Phylogenetic tree inferred from maximum likelihood analysis based on concatenated plastome sequences. (<b>LSC</b>) LSC region. (<b>SSC</b>) SSC region. (<b>IR</b>) IR region. (<b>Coding</b>) Coding region. (<b>Noncoding</b>) Noncoding region. <span class="html-italic">Lindera obtusiloba</span>, <span class="html-italic">Litsea semecarpifolia, Neolitsea aurata and Actinodaphne koshepangii</span> were used as the outgroup. The number on each node is a bootstrap support value. The evolutionary branch length represents the degree of branch variation. Different tribal clades are highlighted with different colors. The number after the species name is the accession number. One-thousand bootstrap replicates were conducted to obtain node support.</p>
Full article ">Figure 7
<p>The plastome phylogenetic tree constructed via the Astral method. Pie charts for major clades representing the proportion of gene trees supporting the Astral species tree topology (blue), the main alternative bifurcation (red), and the remaining alternatives (yellow). The QS scores are shown in the branches.</p>
Full article ">Figure 8
<p>Phylogenetic tree based on nrDNA sequences. <span class="html-italic">Lindera obtusiloba</span>, <span class="html-italic">Litsea semecarpifolia, Neolitsea aurata and Actinodaphne koshepangii</span> were used as the outgroup. The number on each node is the bootstrap support value. The evolutionary branch length represents the degree of branch variation. Different tribal clades are highlighted with different colors. The number after the species name is the accession number. One-thousand bootstrap replicates were conducted to obtain node support.</p>
Full article ">Figure 9
<p>Phylogenetic tree based on plastome sequences. (<b>A</b>) Phylogenetic tree inferred by based on maximum likelihood analysis of plastome sequences. (<b>B</b>) Phylogenetic tree inferred by Xiao et al. based on maximum likelihood analysis of plastome sequences. (<b>C</b>) Phylogenetic tree inferred by Rhode et al. based on maximum likelihood analysis of plastome sequences. (<b>D</b>) Phylogenetic tree inferred by Yang et al. based on maximum likelihood analysis of plastome sequences.</p>
Full article ">Figure 10
<p>The phylogenetic trees based on the maximum likelihood method show plastome sequences result on the left and nrDNA result on the right. The purple lines indicate the emergence of conflicting topologies in phylogenetic analyses of nrDNA sequences and plastome sequences. The number on each node is bootstrap support value. The evolutionary branch length represents the degree of branch variation. Different tribal clades are highlighted with different colors. The number after the species name is the accession number. One-thousand bootstrap replicates were conducted to obtain node support.</p>
Full article ">Figure 11
<p>Distribution of the first nucleoplasmic conflict group.</p>
Full article ">Figure 12
<p>Distribution of the second nucleoplasmic conflict group.</p>
Full article ">
20 pages, 11711 KiB  
Article
CITE-Seq Analysis Reveals a Differential Natural Killer Cell SPON2 Expression in Cardiovascular Disease Patients Impacted by Human-Cytomegalovirus Serostatus and Diabetes
by Sujit Silas Armstrong, Daniel G. Chen, Sunil Kumar, James R. Heath, Matthew J. Feinstein, John R. Greenland, Daniel R. Calabrese, Lewis L. Lanier, Klaus Ley and Avishai Shemesh
Int. J. Mol. Sci. 2025, 26(3), 1369; https://doi.org/10.3390/ijms26031369 - 6 Feb 2025
Viewed by 562
Abstract
Coronary artery disease (CAD) is linked to atherosclerosis plaque formation. In pro-inflammatory conditions, human Natural Killer (NK) cell frequencies in blood or plaque decrease; however, NK cells are underexplored in CAD pathogenesis, inflammatory mechanisms, and CAD comorbidities, such as human cytomegalovirus (HCMV) infection [...] Read more.
Coronary artery disease (CAD) is linked to atherosclerosis plaque formation. In pro-inflammatory conditions, human Natural Killer (NK) cell frequencies in blood or plaque decrease; however, NK cells are underexplored in CAD pathogenesis, inflammatory mechanisms, and CAD comorbidities, such as human cytomegalovirus (HCMV) infection and diabetes. Analysis of PBMC CITE-seq data from sixty-one CAD patients revealed higher blood NK cell SPON2 expression in CAD patients with higher stenosis severity. Conversely, NK cell SPON2 expression was lower in pro-inflammatory atherosclerosis plaque tissue with an enriched adaptive NK cell gene signature. In CAD patients with higher stenosis severity, peripheral blood NK cell SPON2 expression was lower in patients with high HCMV-induced adaptive NK cell frequencies and corresponded to lower PBMC TGFβ transcript expression with dependency on diabetes status. These results suggest that high NK cell SPON2 expression is linked to atherosclerosis pro-homeostatic status and may have diagnostic and prognostic implications in cardiovascular disease. Full article
Show Figures

Figure 1

Figure 1
<p>NK cell <span class="html-italic">SPON2</span> expression significantly increases in CAD patients with high stenosis. (<b>A</b>) PBMC from 61 CAD patients were clustered by CITE-seq protein expression. UMAP of PBMC clusters based on CITE-seq protein expression. NK cells (purple cluster, black arrow). (<b>B</b>) Differential gene expression (DGE) analysis of NK cell gene expression between CAD<sup>low</sup> vs. CAD<sup>high</sup> at the single-cell level (red: upregulated in CAD<sup>high</sup>; blue: upregulated in CAD<sup>low</sup>). (<b>C</b>) Patients’ mean NK cell <span class="html-italic">SPON2</span> expression in CAD<sup>low</sup> vs. CAD<sup>high</sup> patients (dot = patient). (<b>D</b>) UMAP of <span class="html-italic">SPON2</span> RNA expression relative to PBMC clusters. (<b>E</b>) NK cell <span class="html-italic">SPON2</span> expression in patients grouped by CAD stenosis severity score. Patients were grouped the patients by stenosis severity score (combined percent stenosis of each artery segment score): I, 0–6 [<span class="html-italic">n</span> = 29, CAD<sup>low</sup>]; II, 30–48 [<span class="html-italic">n</span> = 13]; III, 49–67.5 [<span class="html-italic">n</span> = 11]; IV, 77–150 [<span class="html-italic">n</span> = 8]. Mean +/− S.D.; Mann–Whitney test; one-tail; * <span class="html-italic">p</span> &lt; 0.05; ** <span class="html-italic">p</span> &lt; 0.01; *** <span class="html-italic">p</span> &lt; 0.001.</p>
Full article ">Figure 2
<p>NK cell <span class="html-italic">SPON2</span> expression is higher in pro-homeostatic plaque tissue. (<b>A</b>) NK cell gene expression between carotid plaques (pro-inflammatory, <span class="html-italic">n</span> =3) relative to the femoral plaques (pro-homeostatic, <span class="html-italic">n</span> =7)<b>.</b> Gene expression is displayed as mean z-score. (<b>B</b>) Mean NK cell <span class="html-italic">SPON2</span> expression in femoral relative to carotid plaque tissue. (<b>C</b>) Pearson correlation (one-tail) between NK cell <span class="html-italic">SPON2</span> and <span class="html-italic">IFNG</span> expression in atherosclerosis plaques (<span class="html-italic">n</span> = 10). (<b>D</b>) NK cell <span class="html-italic">SPON2</span>/<span class="html-italic">IFNG</span> ratio in carotid or femoral plaques. (<b>E</b>) Pearson correlation (one-tail) between NK cell <span class="html-italic">SPON2</span> and <span class="html-italic">IFNG</span> expression in CAD patients. NK cell <span class="html-italic">SPON2</span>/<span class="html-italic">IFNG</span> ratio in (<b>F</b>) CAD<sup>low</sup> vs. CAD<sup>high</sup> patients or (<b>G</b>) relative to stenosis severity. (<b>H</b>) Pearson correlation (one-tail) between NK cell <span class="html-italic">IFNG</span> expression and hs-CRP levels in CAD patients. <sup>#</sup> To avoid misinterpretation of the data, one patient outlier (hs-CRP (mg/L) = 150) was removed from the analysis. (<b>I</b>) NK cell <span class="html-italic">SPON2</span>/ hs-CRP ratio relative to stenosis severity. (<b>J</b>) Schematic representation of findings. Mean +/− S.D.; Mann–Whitney test; one-tail; * <span class="html-italic">p</span> &lt; 0.05; ** <span class="html-italic">p</span> &lt; 0.01; dot = patient.</p>
Full article ">Figure 3
<p>HCMV serostatus and diabetes impact on NK cell <span class="html-italic">SPON2</span> expression. Mean NK cell <span class="html-italic">SPON2</span> expression per patient grouped by CAD<sup>low</sup> vs. CAD<sup>high</sup>; (<b>A</b>) HCMV seronegative (HCMV<sup>−</sup>) or (<b>B</b>) HCMV seropositive (HCMV<sup>+</sup>). NK cell <span class="html-italic">SPON2</span> expression relative to stenosis severity in (<b>C</b>) HCMV<sup>−</sup> or (<b>D</b>) HCMV<sup>+</sup> patients. (<b>E</b>) NK cell <span class="html-italic">SPON2</span> expression in patients, grouped by CAD, diabetes, and HCMV status (black: HCMV<sup>−</sup>; red: HCMV<sup>+</sup>); (<b>E.i</b>) all patients, (<b>E.ii</b>) HCMV<sup>−</sup> patients, or (<b>E.iii</b>) HCMV<sup>+</sup> patients. (<b>F</b>) Schematic representation of figure findings. Mean+/− S.D.; Mann–Whitney test; one-tail; * <span class="html-italic">p</span> &lt; 0.05; ** <span class="html-italic">p</span> &lt; 0.01; dot = patient.</p>
Full article ">Figure 4
<p><span class="html-italic">SPON2<sup>low</sup></span> HCMV-induced NK cell frequencies do not decrease with stenosis severity. (<b>A</b>) UMAP of NK cell clusters. The NK cell cluster (Cluster 4, <a href="#ijms-26-01369-f001" class="html-fig">Figure 1</a>A) was re-clustered based on CITE-seq protein expression of CD56, CD16, CD25, CD27, CD2, or HLA-DR. (<b>B</b>) Heatmap of mean NK cell gene expression for the indicated genes. (<b>C</b>) density plots of NK cell clusters in HCMV<sup>−</sup> or HCMV<sup>+</sup> patients. C2 frequencies (%) between (<b>D</b>) HCMV<sup>−</sup> and HCMV<sup>+</sup> patients, or patients grouped by (<b>E</b>) CAD low vs. high status, (<b>F</b>) CAD and HCMV serostatus, or (<b>G</b>) CAD, diabetes, and HCMV status. (<b>H</b>) Density plot of NK cell clusters in patients grouped by CAD, diabetes, and HCMV status. (<b>I</b>) Schematic representation of findings. (<b>J</b>) Blood NK cell frequencies in patients grouped by CAD, diabetes, and HCMV status. (<b>K</b>) C2 frequencies (%) (upper panels), C3/C2 ratio (middle panels), or NKG2C<sup>−</sup>/NKG2C<sup>+</sup> ratio (lower panels) relative to stenosis severity. Patient groups (not all patients) are shown as black: HCMV<sup>−</sup> or red: HCMV<sup>+</sup>). Mean+/− S.D.; Mann–Whitney test; one-tail; * <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; dot = patient.</p>
Full article ">Figure 5
<p>Higher NK cell <span class="html-italic">SPON2</span> expression corresponds to lower C2 frequencies and higher PBMC <span class="html-italic">TGFβ</span> expression in CAD<sup>high</sup> patients. (<b>A</b>) C2 frequencies in CAD<sup>high</sup> patients with low (C2<sup>low</sup>) or high (C2<sup>high</sup>) frequencies (left), and relative to diabetes status (right: black, diabetes<sup>−</sup>; blue, diabetes<sup>+</sup>). (<b>B</b>) Number of HCMV<sup>−</sup> or HCMV<sup>+</sup> cases in C2<sup>low</sup> vs. C2<sup>high</sup> patients: left, diabetes<sup>−</sup>; right, diabetes<sup>+</sup>. (<b>C</b>) Mean NK cell <span class="html-italic">SPON2</span> expression in CAD<sup>high</sup> C2<sup>low</sup> or C2<sup>high</sup> patients (left), and relative to diabetes status (right: black, diabetes<sup>−</sup>; blue, diabetes<sup>+</sup>). (<b>D</b>) PBMC <span class="html-italic">TGFβ</span> expression in CAD low vs. high patients. Right: expression of PBMC <span class="html-italic">TGFβ</span> in patients grouped by CAD, diabetes, and HCMV status, or (<b>E</b>) in CAD<sup>high</sup> C2<sup>low</sup> or C2<sup>high</sup> patients, and relative to diabetes status (right: black, diabetes<sup>−</sup>; blue, diabetes<sup>+</sup>). (<b>F</b>) PBMC <span class="html-italic">IL15</span>/<span class="html-italic">TGFβ</span> ratio in CAD<sup>high</sup> C2<sup>low</sup> or C2<sup>high</sup> patients (left), and relative to diabetes status (right: black, diabetes<sup>−</sup>; blue, diabetes<sup>+</sup>). (<b>G</b>) PBMC <span class="html-italic">IL15</span>/<span class="html-italic">TGFβ</span> ratio (<b>G.i</b>) or mean NK cell <span class="html-italic">SPON2</span> expression (<b>G.ii</b>), relative to stenosis severity in (left to right) diabetes<sup>−</sup>HCMV<sup>−</sup>, diabetes<sup>−</sup>HCMV<sup>+</sup>, diabetes<sup>+</sup>HCMV<sup>−</sup>, or diabetes<sup>+</sup>HCMV<sup>+</sup> patients. Patient groups are shown as black: HCMV<sup>−</sup> or red: HCMV<sup>+</sup>). Mean +/− S.D.; Mann–Whitney test; one-tail; * <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; dot = patient.</p>
Full article ">Figure 6
<p>Differential upregulation of Spondin-2 by IL-15 and TGFβ in primary NK cells. (<b>A</b>) Dot plots of intracellular Spondin-2 (X-axis) vs. IFNγ (Y-axis) expression in isolated human primary NK cells (<span class="html-italic">n</span> = 2) following 24 h of CD16 stimulation with media without cytokines or with IL-12 (1 ng/mL), IL-2 (300 U/mL), IL-15 (50 ng/mL), or IFNα (50 ng/mL). (<b>B</b>) Histograms of Spondin-2 intracellular expression (red) relative to media without IL-15 control (black) in isolated primary NK cells following 3 days of stimulation with IL-15 (50 ng/mL). (<b>C</b>) Histograms of Spondin-2 intracellular expression between the defined NK cell subsets (color-coded) after 3 days of stimulation with IL-15 (50 ng/mL) or media without cytokines. The right dot plot represents the gating of mature CD56<sup>dim</sup>CD16<sup>+</sup>NK cells to identify adaptive NK cell subsets by NKG2C vs. FcεR1γ protein expression. Right panel: IL-2 or IL-15 concentration-dependent expression of Spondin-2 between the defined NK cell subsets (<span class="html-italic">n</span> = 5). IL-2 (300, 30, and 3 U/mL), IL-15 (50 and 0.5 ng/mL), or media (media without cytokines). (<b>D</b>) FcεR1γ (left) or Spondin-2 (right) intracellular expression or (<b>E</b>) cell trace violet (left) or relative NK cell numbers/well (right) in purified primary non-adaptive NK cells (<span class="html-italic">n</span> = 4) stimulated for 6 days with IL-15 (10 ng/mL), with or without glucose (16 or 4 g/L) or TGFβ (5 ng/mL) relative to media without cytokines. (<b>F</b>) Expression of FcεR1γ (left) or Spondin-2 (right) in purified primary non-adaptive NK cells stimulated for 6 days with IL-15 (50 ng/mL) with or without rapamycin (RAPA, 10 nM) or FOXO1 inhibitor (50 nM) relative to media without cytokines. (<b>G</b>) Schematic representation of findings. Mean+/− S.D.; Mann–Whitney test; one-tail; * <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; dot = donor.</p>
Full article ">
17 pages, 9918 KiB  
Article
Aspirin Foliar Spray-Induced Changes in Light Energy Use Efficiency, Chloroplast Ultrastructure, and ROS Generation in Tomato
by Julietta Moustaka, Ilektra Sperdouli, Emmanuel Panteris, Ioannis-Dimosthenis S. Adamakis and Michael Moustakas
Int. J. Mol. Sci. 2025, 26(3), 1368; https://doi.org/10.3390/ijms26031368 - 6 Feb 2025
Viewed by 380
Abstract
Aspirin (Asp) is extensively used in human health as an anti-inflammatory, antipyretic, and anti-thrombotic drug. In this study, we investigated if the foliar application of Asp on tomato plants has comparable beneficial effects on photosynthetic function to that of salicylic acid (SA), with [...] Read more.
Aspirin (Asp) is extensively used in human health as an anti-inflammatory, antipyretic, and anti-thrombotic drug. In this study, we investigated if the foliar application of Asp on tomato plants has comparable beneficial effects on photosynthetic function to that of salicylic acid (SA), with which it shares similar physiological characteristics. We assessed the consequences of foliar Asp-spray on the photosystem II (PSII) efficiency of tomato plants, and we estimated the reactive oxygen species (ROS) generation and the chloroplast ultrastructural changes. Asp acted as an osmoregulator by increasing tomato leaf water content and offering antioxidant protection. This protection kept the redox state of plastoquinone (PQ) pull (qp) more oxidized, increasing the fraction of open PSII reaction centers and enhancing PSII photochemistry (ΦPSII). In addition, Asp foliar spray decreased reactive oxygen species (ROS) formation, decreasing the excess excitation energy on PSII. This resulted in a lower singlet oxygen (1O2) generation and a lower quantum yield for heat dissipation (ΦNPQ), indicating the photoprotective effect provided by Asp, especially under excess light illumination. Simultaneously, we observed a decrease in stomatal opening by Asp, which reduced the transpiration. Chloroplast ultrastructural data revealed that Asp, by offering a photoprotective effect, decreased the need for the photorespiration process, which reduces photosynthetic performance. It is concluded that Asp shares similar physiological characteristics with SA, having an equivalent beneficial impact to SA by acting as a biostimulant of the photosynthetic function for an enhanced crop yield. Full article
(This article belongs to the Special Issue Molecular Mechanisms of Plant Abiotic Stress Tolerance)
Show Figures

Figure 1

Figure 1
<p>The chlorophyll content of water (WA)-sprayed and Aspirin (Asp)-sprayed leaves 24- and 96-h after the spray, expressed in relative units (<span class="html-italic">n</span> = 10 ± SD). Significant differences are shown by different lower-case letters (<span class="html-italic">p</span> &lt; 0.05).</p>
Full article ">Figure 2
<p>The effective quantum yield of PSII photochemistry (Φ<span class="html-italic"><sub>PSII</sub></span>) at the growth light intensity (GLI) (<b>a</b>) and at the high light intensity (HLI) (<b>b</b>) of WA-sprayed and Asp-sprayed leaves 24- and 96-h after the spray (<span class="html-italic">n</span> = 6 ± SD). Significant differences are shown by different lower-case letters (<span class="html-italic">p</span> &lt; 0.05).</p>
Full article ">Figure 3
<p>The quantum yield of regulated non-photochemical energy loss in PSII (Φ<span class="html-italic"><sub>NPQ</sub></span>) at the GLI (<b>a</b>) and at the HLI (<b>b</b>) of WA-sprayed and Asp-sprayed leaves 24- and 96-h after the spray (<span class="html-italic">n</span> = 6 ± SD). Significant differences are shown by different lower-case letters (<span class="html-italic">p</span> &lt; 0.05).</p>
Full article ">Figure 4
<p>The quantum yield of non-regulated energy loss in PSII (Φ<span class="html-italic"><sub>NO</sub></span>) at the GLI (<b>a</b>) and at the HLI (<b>b</b>) of WA-sprayed and Asp-sprayed leaves 24- and 96-h after the spray (<span class="html-italic">n</span> = 6 ± SD). Significant differences are shown by different lower-case letters (<span class="html-italic">p</span> &lt; 0.05).</p>
Full article ">Figure 5
<p>The non-photochemical quenching (NPQ), at the GLI (<b>a</b>) and at the HLI (<b>b</b>) of WA-sprayed and Asp-sprayed leaves 24- and 96-h after the spray (<span class="html-italic">n</span> = 6 ± SD). Significant differences are shown by different lower-case letters (<span class="html-italic">p</span> &lt; 0.05).</p>
Full article ">Figure 6
<p>The fraction of open PSII reaction centers (RCs) (q<span class="html-italic">p</span>), at the GLI (<b>a</b>) and at the HLI (<b>b</b>) of WA-sprayed and Asp-sprayed leaves 24- and 96-h after the spray (<span class="html-italic">n</span> = 6 ± SD). Significant differences are shown by different lower-case letters (<span class="html-italic">p</span> &lt; 0.05).</p>
Full article ">Figure 7
<p>The efficiency of the open PSII RCs (F<span class="html-italic">v</span>’/F<span class="html-italic">m</span>’) at the GLI (<b>a</b>) and at the HLI (<b>b</b>) of WA-sprayed and Asp-sprayed leaves 24- and 96-h after the spray (<span class="html-italic">n</span> = 6 ± SD). Significant differences are shown by different lower-case letters (<span class="html-italic">p</span> &lt; 0.05).</p>
Full article ">Figure 8
<p>The electron transport rate (ETR) at the GLI (<b>a</b>) and at the HLI (<b>b</b>) of WA-sprayed and Asp-sprayed leaves 24- and 96-h after the spray (<span class="html-italic">n</span> = 6 ± SD). Significant differences are shown by different lower-case letters (<span class="html-italic">p</span> &lt; 0.05).</p>
Full article ">Figure 9
<p>The excitation pressure at PSII (1 − q<span class="html-italic">L</span>), measured at the GLI (<b>a</b>) and at the HLI (<b>b</b>) of WA-sprayed and Asp-sprayed leaves 24- and 96-h after the spray (<span class="html-italic">n</span> = 6 ± SD). Significant differences are shown by different lower-case letters (<span class="html-italic">p</span> &lt; 0.05).</p>
Full article ">Figure 10
<p>The excess excitation energy at PSII (EXC), at the GLI (<b>a</b>) and at the HLI (<b>b</b>) of WA- sprayed and Asp-sprayed leaves 24- and 96-h after the spray (<span class="html-italic">n</span> = 6 ± SD). Significant differences are shown by different lower-case letters (<span class="html-italic">p</span> &lt; 0.05).</p>
Full article ">Figure 11
<p>The relationship between the excess excitation energy (EXC) and the excitation pressure at PSII (1 − q<span class="html-italic">L</span>) at the GLI (<b>a</b>) and at the HLI (<b>b</b>) of WA- sprayed and Asp-sprayed leaves 24- and 96-h after the spray (based on the data of <a href="#ijms-26-01368-f009" class="html-fig">Figure 9</a>a,b and <a href="#ijms-26-01368-f010" class="html-fig">Figure 10</a>a,b). Each blue dot represents the paired measurement of the variables, while the red line is the regression line that shows the relationship between the two variables.</p>
Full article ">Figure 12
<p>ROS production in tomato leaflets of WA-sprayed (<b>a</b>) and Asp-sprayed <b>(b</b>) leaves 24-h after the spray. The light green color indicates ROS generation. Scale bar, 200 μm.</p>
Full article ">Figure 13
<p>Transmission electron micrographs of mesophyll cells of tomato leaves sprayed with water (<b>a</b>,<b>b</b>) or with Asp (<b>c</b>,<b>d</b>). Note the peroxisomes (arrows in (<b>a</b>)), which include electron-dense crystals (asterisk in (<b>b</b>)) in cells of WA-sprayed leaves. Starch grains (sg) can be observed in chloroplasts of Asp-sprayed leaves (<b>c</b>,<b>d</b>) but not in those of WA-sprayed leaves. m: mitochondrion. Scale bars as indicated on the micrographs.</p>
Full article ">
28 pages, 7769 KiB  
Article
Impact of African-Specific ACE2 Polymorphisms on Omicron BA.4/5 RBD Binding and Allosteric Communication Within the ACE2–RBD Protein Complex
by Victor Barozi and Özlem Tastan Bishop
Int. J. Mol. Sci. 2025, 26(3), 1367; https://doi.org/10.3390/ijms26031367 - 6 Feb 2025
Viewed by 594
Abstract
Severe acute respiratory symptom coronavirus 2 (SARS-CoV-2) infection occurs via the attachment of the spike (S) protein’s receptor binding domain (RBD) to human ACE2 (hACE2). Natural polymorphisms in hACE2, particularly at the interface, may alter RBD–hACE2 interactions, potentially affecting viral infectivity across populations. [...] Read more.
Severe acute respiratory symptom coronavirus 2 (SARS-CoV-2) infection occurs via the attachment of the spike (S) protein’s receptor binding domain (RBD) to human ACE2 (hACE2). Natural polymorphisms in hACE2, particularly at the interface, may alter RBD–hACE2 interactions, potentially affecting viral infectivity across populations. This study identified the effects of six naturally occurring hACE2 polymorphisms with high allele frequency in the African population (S19P, K26R, M82I, K341R, N546D and D597Q) on the interaction with the S protein RBD of the BA.4/5 Omicron sub-lineage through post-molecular dynamics (MD), inter-protein interaction and dynamic residue network (DRN) analyses. Inter-protein interaction analysis suggested that the K26R variation, with the highest interactions, aligns with reports of enhanced RBD binding and increased SARS-CoV-2 susceptibility. Conversely, S19P, showing the fewest interactions and largest inter-protein distances, agrees with studies indicating it hinders RBD binding. The hACE2 M82I substitution destabilized RBD–hACE2 interactions, reducing contact frequency from 92 (WT) to 27. The K341R hACE2 variant, located distally, had allosteric effects that increased RBD–hACE2 contacts compared to WThACE2. This polymorphism has been linked to enhanced affinity for Alpha, Beta and Delta lineages. DRN analyses revealed that hACE2 polymorphisms may alter the interaction networks, especially in key residues involved in enzyme activity and RBD binding. Notably, S19P may weaken hACE2–RBD interactions, while M82I showed reduced centrality of zinc and chloride-coordinating residues, hinting at impaired communication pathways. Overall, our findings show that hACE2 polymorphisms affect S BA.4/5 RBD stability and modulate spike RBD–hACE2 interactions, potentially influencing SARS-CoV-2 infectivity—key insights for vaccine and therapeutic development. Full article
(This article belongs to the Section Biochemistry)
Show Figures

Figure 1

Figure 1
<p>Structural representation of the hACE2 (PDB ID: 7L7F [<a href="#B38-ijms-26-01367" class="html-bibr">38</a>]) protein in a complex with the S protein RBD (PDB ID: 6M0J [<a href="#B18-ijms-26-01367" class="html-bibr">18</a>]). (<b>A</b>) Cartoon representation of the hACE2 sub-domains: the collectrin domain (brown) and the peptidase sub-domains (sub-domain I: blue and II: yellow) in complex with S RBD (grey). The RBD receptor binding motif (RBM) is shown in light brown. (<b>B</b>) A zoomed in view of the hACE2 active site pocket showing zinc (orange) and chloride (green) ions. The ion-coordinating residues are shown as sticks in yellow and orange, respectively. (<b>C</b>) A detailed view of the interface between hACE2 and S RBD, with the interacting interface residues shown as sticks (blue for hACE2 and brown for RBD).</p>
Full article ">Figure 2
<p>Three-dimensional structural representation of the RBD–hACE2 complex showing the BA.4/BA.5 Omicron RBD mutations (red spheres) and the hACE2 polymorphisms (red sticks). The hACE2 sub-domains I and II are shown in sky blue and pale yellow, respectively, whereas S RBD is in grey. The hACE2 substituted residue interactions within 3 Å are highlighted in yellow, purple and blue for hydrogen bonds (H-bonds), salt bridges and π–π stacking interactions, respectively.</p>
Full article ">Figure 3
<p>(<b>A</b>,<b>B</b>) Violin plots showing the RMSD distribution for the RBD and hACE2, respectively, in the WThACE2 and hACE2 variant systems. The plots are arranged in ascending order of the median RMSD, obtained by comparing median values of each system RMSD value. WThACE2 is shown as green and the hACE2 variants in orange. (<b>C</b>,<b>D</b>) The RMSF for the RBD and hACE2 protein systems, respectively. The same color scheme is used as in (<b>A</b>,<b>B</b>). Mutation positions in both the RBD and hACE2 proteins are indicated by red markers.</p>
Full article ">Figure 4
<p>Heatmaps of the top 5% and 4% high <span class="html-italic">BC</span> residues in the RBD and hACE2 proteins, respectively, at the individual protein level and global level. Residues are on the x-axis and protein systems on the y-axis. The color scale from white to dark red shows the degree of centrality. The centrality color scale is different for the local and global residues because, in the case of local analysis, centrality calculation is based on the individual systems as opposed to the whole ensemble under the global analysis. The residues with high centrality (hubs) are annotated with metric values in the heatmap, while homologous residues from other systems remain unannotated.</p>
Full article ">Figure 5
<p>The distribution of the top 5% (RBD) and 4% (hACE2) <span class="html-italic">BC</span> hubs in the RBD–hACE2 complexes. For each system, the RBD is shown as a grey cartoon and hACE2 sub-domains I and II as sky blue and pale yellow, respectively. Zinc and chloride ions are shown as orange and green spheres, respectively, whereas the hACE2 mutations are dark green spheres. Hubs (high centrality residues) are shown as sky blue spheres (hACE2) and grey spheres (RBD). The same colors are used for <span class="html-italic">BC</span> hubs in the hACE2 variant systems. The five highest centrality <span class="html-italic">BC</span> hubs in the RBD and the hACE2 are shown as dark grey and dark blue spheres, respectively. The gains and losses in interface residue hubs in the hACE2 variant systems compared to the WThACE2 are annotated with + and − symbols, respectively. The hACE2 mutation positions are shown as red spheres.</p>
Full article ">Figure 6
<p>Heatmaps of the top 5% and 4% <span class="html-italic">EC</span> residues in the RBD and hACE2, respectively. Residue numbers are on the x-axis and systems on the y-axis. The color scale from white to dark red shows the degree of centrality. The centrality color scale is different for the local and global residues, because in the case of local analysis, centrality calculation is based on the individual systems as opposed to the whole ensemble under the global analysis.</p>
Full article ">Figure 7
<p>The distribution of the top 5% and 4% RBD and hACE2 <span class="html-italic">EC</span> hubs in the RBD–hACE2 complexes. The RBD is shown as a grey cartoon and hACE2 sub-domains I and II as sky blue and pale yellow, respectively. Zinc and chloride ions are shown as orange and green spheres, respectively, whereas the hACE2 mutations are dark green spheres. Hubs are shown as sky blue spheres (hACE2) and grey spheres (RBD). The five highest centrality <span class="html-italic">EC</span> hubs in the RBD and the hACE2 are shown as dark grey and dark blue spheres, respectively. The gains and losses in interface residue hubs in the hACE2 variant systems are indicated with + and − symbols, respectively. hACE2 mutations are shown as raspberry spheres.</p>
Full article ">Figure 8
<p>The network of the hACE2 <span class="html-italic">EC</span> hubs in each system. The hubs are shown as nodes and their connectedness as edges. The color scale from light to dark green shows the degree of residue <span class="html-italic">eigenvector centrality (EC)</span>. The hACE2 SNP bearing systems generally have more connected <span class="html-italic">EC</span> hubs, especially involving zinc coordinating residues, signifying the increased influence of this region in protein signal communication.</p>
Full article ">Figure 9
<p>The distribution of the top 5% and 4% <span class="html-italic">CC</span> hubs in the RBD and hACE2, respectively. For each system, the RBD is shown as a grey cartoon and hACE2 sub-domains I and II as sky blue and yellow, respectively. The zinc and chloride ions are shown as orange and green spheres, respectively, whereas the hACE2 mutations are dark green spheres. Hubs are shown as sky blue spheres (hACE2) and grey spheres (RBD). The five highest centrality <span class="html-italic">CC</span> hubs in the RBD and hACE2 are shown as dark grey and dark blue spheres, respectively. The gains and losses in interface residue hubs in the hACE2 variant systems compared to the WThACE2 are annotated with + and −, respectively. Gained and lost hubs in the hACE2 and RBD variants are annotated in blue and grey, respectively. RBD mutation positions are shown as red spheres. The line subplots in each system show the comparative inter-protein COM distance for the WThACE2 (green) and hACE2 variant systems (orange) over 100 ns.</p>
Full article ">Figure 10
<p>(<b>A</b>) The differences between WThACE2 and each hACE2 variant system inter-protein residue interaction frequencies are shown as a heatmap. The differences were calculated by subtracting the variant interface residue pair frequencies from that of WThACE2. The color gradient from red through white to blue shows the delta frequency of residue interactions between the RBD and hACE2. Red means more contact frequency in WthACE2 and vice versa for blue. White shows no difference between the hACE2 variants and WThACE2. (<b>B</b>) Bipartite graphs of the weighted pairwise residue interaction frequencies between the RBD residues (sky blue nodes) and hACE2 residues (orange nodes).</p>
Full article ">
16 pages, 1841 KiB  
Article
AgNP-Containing Niosomes Functionalized with Fucoidan Potentiated the Intracellular Killing of Mycobacterium abscessus in Macrophages
by Nereyda Niño-Martínez, Kayla Audreyartha, Kaitlyn Cheung, Sol Melchor Parra, Gabriel Martínez-Castañón and Horacio Bach
Int. J. Mol. Sci. 2025, 26(3), 1366; https://doi.org/10.3390/ijms26031366 - 6 Feb 2025
Viewed by 415
Abstract
Intracellular pathogens represent a challenge for therapy because the antibiotics used need to diffuse into the cytoplasm to target the pathogens. The situation is more complicated in the mycobacteria family because members of this family infect and multiply within macrophages, the cells responsible [...] Read more.
Intracellular pathogens represent a challenge for therapy because the antibiotics used need to diffuse into the cytoplasm to target the pathogens. The situation is more complicated in the mycobacteria family because members of this family infect and multiply within macrophages, the cells responsible for clearing microorganisms in the body. In addition, mycobacteria members are enclosed inside pathogen-containing vesicles or phagosomes. The treatments of these pathogens are aggravated when these pathogens acquire resistance to antibiotic molecules. As a result, new antimicrobial alternatives are needed. Niosomes are vesicles composed of cholesterol and nonionic surfactants that can be used for antibiotic encapsulation and delivery. The current study developed a systematic formulation of niosomes to determine the best option for niosome functionalizing for precise delivery to the intracellular pathogen Mycobacterium abscessus. Silver nanoparticles (AgNPs) were synthesized using gallic acid as an antibacterial agent. Then, niosomes were prepared and characterized, following the encapsulation of AgNPs functionalized with a single-chain antibody screened against the cell wall glycopeptidolipid of Mycobacterium abscessus. For a precise delivery of the cargo into macrophages, the niosomes were also functionalized with the polysaccharide fucoidan, taken specifically by the scavenger receptor class A expressed on the surface of macrophages. Results of the study showed a steady decrease in the intracellular pathogen load after 48 h post-infection. In conclusion, this system could be developed into a platform to target other types of intracellular pathogens and as an option for antimicrobial therapy. Full article
(This article belongs to the Special Issue Advances in Antimicrobial Nanomaterials 2.0)
Show Figures

Figure 1

Figure 1
<p>(<b>A</b>) Structure of GPL. (<b>B</b>). The antibody scFv recognizes <span class="html-italic">M. abscessus</span>. Bacteria were labeled with Rhodamine B (10 μg/mL, MilliporeSigma, Burlington, MA USA) and exposed to <span class="html-italic">M. abscessus,</span> which had previously flamed on a microscope slide. The scFv-51 (not co-expressed with the red fluorescent protein) and bacteria were incubated for 30 min, washed with PBS (×3), and incubated with FITC-mouse-anti-his tag (Thermo Fisher, Waltham, MA USA) for 30 min. The slides were analyzed by fluorescence microscopy after being washed with PBS (×3). The yellow color indicates the co-localization of scFv-51 to the pathogen. ABS, <span class="html-italic">M. abscessus</span>; BCG, <span class="html-italic">M. bovis</span> strain BCG; MAA, <span class="html-italic">M. avium avium</span>.</p>
Full article ">Figure 2
<p>Effect of the sonication on the niosome size. NA4 niosomes were subjected to sonication during their last step of synthesis. (<b>A</b>) Impact of the sonicator power on the niosome sizes. (<b>B</b>) Impact of the sonication time on the niosome size using a probe power between 12 and 15 W. The numbers at each time point represent the highest percentage of the main peak, as measured by PDI in the DLS device.</p>
Full article ">Figure 3
<p>Scheme showing the synthesis of AgNP-scFv-51 with aminated fucoidan. Created with <a href="http://Biorender.com" target="_blank">Biorender.com</a>.</p>
Full article ">Figure 4
<p>Cytotoxicity of niosomes. Differentiated THP-1 was treated with different concentrations of NC3 AgNP-scFv-51 and evaluated by MTT. Untreated and SDS-treated cells were used as negative and positive controls, respectively. OD, optical density. * <span class="html-italic">p</span>-value &lt; 0.05. Experiments were performed in triplicate.</p>
Full article ">Figure 5
<p>Survival of <span class="html-italic">M. abscessus</span> within macrophages. Differentiated THP-1 cells were infected with <span class="html-italic">M. abscessus</span> at a multiplicity of infection of 1. The macrophages were treated with the NB5 5 μg/mL AgNP-scFv-51 + 10% cholesteryl acetate:fucoidan. Untreated cells were used as a negative control, whereas AgNPs were used as a positive control. Shown is the median ± SD of three independent experiments. CFU, colony-forming units. * <span class="html-italic">p</span>-value &lt;0.05.</p>
Full article ">
14 pages, 7571 KiB  
Article
Sterol Regulatory Element-Binding Protein Sre1 Mediates the Development and Pathogenicity of the Grey Mould Fungus Botrytis cinerea
by Ye Yuan, Shengnan Cao, Jiao Sun, Jie Hou, Mingzhe Zhang, Qingming Qin and Guihua Li
Int. J. Mol. Sci. 2025, 26(3), 1365; https://doi.org/10.3390/ijms26031365 - 6 Feb 2025
Viewed by 498
Abstract
The grey mould fungus Botrytis cinerea is a dangerous plant pathogen responsible for substantial agricultural losses worldwide. The pathogenic mechanisms still have many unclear aspects, and numerous new pathogenic genes remain to be identified. Here, we show that the sterol regulatory element-binding protein [...] Read more.
The grey mould fungus Botrytis cinerea is a dangerous plant pathogen responsible for substantial agricultural losses worldwide. The pathogenic mechanisms still have many unclear aspects, and numerous new pathogenic genes remain to be identified. Here, we show that the sterol regulatory element-binding protein Sre1 plays an important role in the development and pathogenicity of B. cinerea. We identified a homologue of gene SRE1 in the B. cinerea genome and utilized a reverse genetics approach to create the knockout mutant Δsre1. Our results demonstrate that SRE1 is essential for conidiation, as Δsre1 produced only 3% of the conidia compared to the wild-type strain. Conversely, Δsre1 exhibited increased sclerotium production, indicating a negative regulatory role of SRE1 in sclerotium formation. Furthermore, ergosterol biosynthesis was significantly reduced in the Δsre1 mutant, correlating with increased sensitivity to low-oxygen conditions. Pathogenicity assays revealed that Δsre1 had significantly reduced virulence, although it maintained normal infection cushion formation and penetration capabilities. Additionally, SRE1 was found to be crucial for hypoxia adaptation, as Δsre1 showed abnormal germination and reduced growth under low-oxygen conditions. These findings suggest that SRE1 mediates the development and pathogenicity of B. cinerea by regulating lipid homeostasis and facilitating adaptation to host tissue environments. Full article
(This article belongs to the Special Issue Plant Responses to Biotic and Abiotic Stresses)
Show Figures

Figure 1

Figure 1
<p>Generations of <span class="html-italic">Botrytis cinerea SRE1</span> knockout mutants and its genetic complemented strains. (<b>a</b>) Strategy for generation of <span class="html-italic">SRE1</span> knockout strain Δ<span class="html-italic">sre1</span> via gene replacement approach. WT, the wild-type strain B05.10; pSRE1-ko, <span class="html-italic">SRE1</span> knockout vector. <span class="html-italic">HPH</span>, the hygromycin resistance gene. (<b>b</b>) Screening of Δ<span class="html-italic">sre1</span> strains. Numbers 1–23 indicate partial selected transformants. PCR amplifications were used for detecting <span class="html-italic">HPH</span> recombination (rec) and <span class="html-italic">SRE1</span> loss in transformants, respectively, with indicated primers. Up-rec, upstream recombination; down-rec, downstream recombination. (<b>c</b>) Verification of the complemented strain Δ<span class="html-italic">sre1</span>-c. (<b>d</b>) Relative <span class="html-italic">SRE1</span> expression level in the indicated strains determined by quantitative reverse transcription PCR. M, DNA marker D2000. ND, not detected. Data represent means ± standard deviations (SD) from at least three independent experiments. *, ***, significance at <span class="html-italic">p</span> &lt; 0.05, 0.001, respectively.</p>
Full article ">Figure 2
<p><span class="html-italic">SRE1</span> is required for <span class="html-italic">B. cinerea</span> conidiation but dispensable for conidial morphogenesis and germination. (<b>a</b>) Colony of tested strains cultured on PDA at 3 days post inoculation (DPI). (<b>b</b>) Quantification of the colony sizes (determined by the relative colony diameter). (<b>c</b>) Conidiation on CM plates at 12 DPI. (<b>d</b>) Quantification of the relative conidiation of the indicated strains (determined by the relative conidial number per plate). (<b>e</b>) Conidial morphology of tested strains. Bar = 10 μm. (<b>f</b>) Conidial germination of tested strains at 4 h post inoculation (HPI). Bar = 10 μm. Data represent means ± standard deviations (SD) from at least three independent experiments. ***, significance at <span class="html-italic">p</span> &lt; 0.001.</p>
Full article ">Figure 3
<p><span class="html-italic">SRE1</span> mediates sclerotium production and ergosterol biosynthesis in <span class="html-italic">B. cinerea</span>. (<b>a</b>) Deletion of <span class="html-italic">SRE1</span> in <span class="html-italic">B. cinerea</span> increases sclerotial production. The tested strains were cultured on CM at 20 °C in darkness and observed at 15 DPI. (<b>b</b>) Quantification of sclerotium production (determined by the relative sclerotial number per plate). (<b>c</b>) Quantification of ergosterol content (determined by the relative ergosterol content per gram of mycelium fresh weight). Data represent means ± standard deviations (SDs) from at least three independent experiments. *, **, ***, significance at <span class="html-italic">p</span> &lt; 0.05, 0.01, 0.001, respectively.</p>
Full article ">Figure 4
<p>Deletion of <span class="html-italic">SRE1</span> increases <span class="html-italic">B. cinerea</span> resistance to antifungal drug imidazole. Conidial suspensions (containing 50 mM glucose) of each strain were treated with 0, 2, 5, and 10 mM imidazole, respectively. Conidial germinations were observed at 4 h post inoculation (HPI).</p>
Full article ">Figure 5
<p><span class="html-italic">SRE1</span> is required for <span class="html-italic">B. cinerea</span> virulence but dispensable for its infection cushion formation and penetration. (<b>a</b>) Mycelial plugs of each strain were inoculated on green bean leaves and the lesions were photographically documented at 3 DPI. (<b>b</b>) Conidial suspensions of each strain were inoculated on green bean leaves and the lesions were photographically documented at 3 DPI. (<b>c</b>) Infection cushion formation of tested strains at 20 HPI. (<b>d</b>) Onion epidermis penetration of the test strains at 12 HPI. Successful penetrations are indicated by red arrows. (<b>e</b>) Quantification of lesion size (determined by the relative lesion area per inoculation) caused by the indicated strains on green bean leaves shown in (<b>a</b>). (<b>f</b>) Quantification of penetration (determined per conidium) by the indicated strains on onion epidermis shown in (<b>d</b>). Data represent means ± standard deviations (SDs) from at least three independent experiments. *, ***, significance at <span class="html-italic">p</span> &lt; 0.05, 0.001, respectively.</p>
Full article ">Figure 6
<p><span class="html-italic">SRE1</span> is involved in hypoxia adaptation. (<b>a</b>) Conidial suspensions (containing 50 mM glucose) of each strain were treated with 100 or 200 μM cobalt chloride (CoCl<sub>2</sub>), a hypoxia-mimicking agent. Conidial germinations were observed at 4 HPI. Germ tubes are indicated by red arrows. (<b>b</b>) Quantification of abnormal conidial germination in hypoxia-mimic with 100 μM CoCl<sub>2</sub>. (<b>c</b>) Quantification of mycelial biomass cultured in normoxia or in hypoxia-mimic with 100 μM CoCl<sub>2</sub>. Data represent means ± standard deviations (SD) from at least three independent experiments. **, ***, significance at <span class="html-italic">p</span> &lt; 0.01, 0.001, respectively.</p>
Full article ">
Previous Issue
Back to TopTop