[go: up one dir, main page]

 
 
Sign in to use this feature.

Years

Between: -

Subjects

remove_circle_outline
remove_circle_outline
remove_circle_outline
remove_circle_outline
remove_circle_outline
remove_circle_outline
remove_circle_outline
remove_circle_outline
remove_circle_outline

Journals

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

Article Types

Countries / Regions

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

Search Results (127,973)

Search Parameters:
Keywords = proteins

Order results
Result details
Results per page
Select all
Export citation of selected articles as:
33 pages, 1571 KiB  
Review
Malnutrition and Allergies: Tipping the Immune Balance Towards Health
by Emilia Vassilopoulou, Carina Venter and Franziska Roth-Walter
J. Clin. Med. 2024, 13(16), 4713; https://doi.org/10.3390/jcm13164713 (registering DOI) - 11 Aug 2024
Abstract
Malnutrition, which includes macro- and micronutrient deficiencies, is common in individuals with allergic dermatitis, food allergies, rhinitis, and asthma. Prolonged deficiencies of proteins, minerals, and vitamins promote Th2 inflammation, setting the stage for allergic sensitization. Consequently, malnutrition, which includes micronutrient deficiencies, fosters the [...] Read more.
Malnutrition, which includes macro- and micronutrient deficiencies, is common in individuals with allergic dermatitis, food allergies, rhinitis, and asthma. Prolonged deficiencies of proteins, minerals, and vitamins promote Th2 inflammation, setting the stage for allergic sensitization. Consequently, malnutrition, which includes micronutrient deficiencies, fosters the development of allergies, while an adequate supply of micronutrients promotes immune cells with regulatory and tolerogenic phenotypes. As protein and micronutrient deficiencies mimic an infection, the body’s innate response limits access to these nutrients by reducing their dietary absorption. This review highlights our current understanding of the physiological functions of allergenic proteins, iron, and vitamin A, particularly regarding their reduced bioavailability under inflamed conditions, necessitating different dietary approaches to improve their absorption. Additionally, the role of most allergens as nutrient binders and their involvement in nutritional immunity will be briefly summarized. Their ability to bind nutrients and their close association with immune cells can trigger exaggerated immune responses and allergies in individuals with deficiencies. However, in nutrient-rich conditions, these allergens can also provide nutrients to immune cells and promote health. Full article
(This article belongs to the Special Issue New Clinical Advances in Pediatric Allergic Diseases)
Show Figures

Figure 1

Figure 1
<p>Risk factors for malnutrition.</p>
Full article ">Figure 2
<p>Nutritional Immunity promotes malabsorption. While in the normal steady state, water-and fat-soluble compounds cross the epithelial barrier and enter the body via the blood system and/or the lacteals, inflammation will trigger nutritional immunity. This results in impaired absorption of minerals and vitamins, particularly in those following the blood route. In contrast the “lymph route” remains accessible as it still allows monitoring of nutrients for potential pathogens.</p>
Full article ">Figure 3
<p>Protein- and micronutrient-poor conditions promote type 2 inflammation. Micronutrientrich conditions foster a regulatory and anti-inflammatory phenotype in lymphocytes, macrophages, and mast cells, while nutrient-poor conditions prime the immune system. A lack of micronutrients, particularly of iron and vitamin A, initially mounts a Th1/Th17-dominated immune response, which results in B cells transforming into plasma cells and secreting IgG-antibodies. When nutrient-poor conditions persevere for longer time periods, the immune response shifts toward Th2 (due to the more nutrient-sensitive nature of Th1 cells) and promotes eosinophils, as well as class switch toward IgE antibodies. M2: regulatory macrophage, Treg: regulatory T cells, B: naïve B-cells, EOS: eosinophils, MC: mast cells, PC: plasma cells.</p>
Full article ">
8 pages, 1940 KiB  
Article
Hsa-miR-874-3p Reduces Endogenous Expression of RGS4-1 Isoform In Vitro
by Feng-Ling Xu and Bao-Jie Wang
Genes 2024, 15(8), 1057; https://doi.org/10.3390/genes15081057 (registering DOI) - 11 Aug 2024
Abstract
Background: The level of the regulator of G-protein signaling 4-1 (RGS4-1) isoform, the longest RGS4 isoform, is significantly reduced in the dorsolateral prefrontal cortex (DLPFC) of people with schizophrenia. However, the mechanism behind this has not been clarified. The 3′untranslated regions (3′UTRs) are [...] Read more.
Background: The level of the regulator of G-protein signaling 4-1 (RGS4-1) isoform, the longest RGS4 isoform, is significantly reduced in the dorsolateral prefrontal cortex (DLPFC) of people with schizophrenia. However, the mechanism behind this has not been clarified. The 3′untranslated regions (3′UTRs) are known to regulate the levels of their mRNA splice variants. Methods: We constructed recombinant pmir-GLO vectors with a truncated 3′ regulatory region of the RGS4 gene (3R1, 3R2, 3R3, 3R4, 3R5, and 3R6). The dual-luciferase reporter assay was conducted to find functional regions in HEK-293, SK-N-SH, and U87cells and then predicted miRNA binding to these regions. We performed a dual-luciferase reporter assay and a Western blot analysis after transiently transfecting the predicted miRNAs. Results: The dual-luciferase reporter assay found that regions +401–+789, +789–+1152, and +1562–+1990 (with the last base of the termination codon being +1) might be functional regions. Hsa-miR-874-3p, associated with many psychiatric disorders, might target the +789–+1152 region in the 3′UTR of the RGS4 gene. In the dual-luciferase reporter assay, the hsa-miR-874-3p mimic, co-transfected with 3R1, down-regulated the relative fluorescence intensities. However, this was reversed when the hsa-miR-874-3p mimic was co-transfected with m3R1 (deletion of +853–+859). The hsa-miR-874-3p mimic significantly decreased the endogenous expression of the RGS4-1 isoform in HEK-293 cells. Conclusions: Hsa-miR-874-3p inhibits the expression of the RGS4-1 isoform by targeting +853–+859. Full article
(This article belongs to the Section Molecular Genetics and Genomics)
Show Figures

Figure 1

Figure 1
<p>The truncated 3′ regulatory region of the <span class="html-italic">RGS4</span> gene recombined into pmir-GLO vectors.</p>
Full article ">Figure 2
<p>The predicted miR-874-3p binding site in the RGS4 mRNA 3′-UTR and the deletion sequences in m3R1. The red bases are the deleted ones in m3R1.</p>
Full article ">Figure 3
<p>The relative fluorescence intensities of recombined pmir-GLO vectors (3R1–3R6) in HEK-293, U87, and SK-N-SH cells. * <span class="html-italic">p</span> ≤ 0.05; ** <span class="html-italic">p</span> ≤ 0.01; *** <span class="html-italic">p</span> ≤ 0.001.</p>
Full article ">Figure 4
<p>The relative fluorescence intensities of 3R1 and m3R1 co-transfected with hsa-miR-874-3p mimic or NC. * <span class="html-italic">p</span> ≤ 0.05; ** <span class="html-italic">p</span> ≤ 0.01; *** <span class="html-italic">p</span> ≤ 0.001.</p>
Full article ">Figure 5
<p>Western blot assay measuring the endogenous protein levels of the RGS4-1 inform after transfection of NC (1, 2, 3), the hsa-miR-874-3p mimic (4, 5, 6), an NC inhibitor (7, 8, 9), and an hsa-miR-874-3p inhibitor (10, 11, 12), in HEK-293 (<b>A</b>,<b>D</b>), SK (<b>B</b>,<b>E</b>), and U87 (<b>C</b>,<b>F</b>). ** <span class="html-italic">p</span> ≤ 0.01; *** <span class="html-italic">p</span> ≤ 0.001, (<b>A</b>–<b>C</b>) each panel summarizes data from two separate western blots.</p>
Full article ">
16 pages, 14227 KiB  
Article
Optimizing the Extraction of Protein from Defatted Schizochytrium Cell Residues and Studying the Emulsification Characteristics of Protein
by Yingying Yang, Xiangying Zhao, Liping Liu, Xinyu Wang, Ruiguo Li and Jiaxiang Zhang
Fermentation 2024, 10(8), 416; https://doi.org/10.3390/fermentation10080416 (registering DOI) - 11 Aug 2024
Abstract
In this study, proteins were prepared from Schizochytrium pombe residue after oil extraction using isoelectric point precipitation, and their physicochemical and emulsifying properties were investigated. Our objective was to assess the suitability of these proteins for functional ingredient applications. Through a one-way experiment and [...] Read more.
In this study, proteins were prepared from Schizochytrium pombe residue after oil extraction using isoelectric point precipitation, and their physicochemical and emulsifying properties were investigated. Our objective was to assess the suitability of these proteins for functional ingredient applications. Through a one-way experiment and optimization using response surface design, the effects of time, temperature, pH, and the material–liquid ratio of NaOH alkaline extraction were explored. The isoelectric point is verified by isoelectric point precipitation; the results revealed that crude protein from Schizochytrium (SCP) is minimally soluble at pH 4.2. Compared with whey protein (WP), it promotes better emulsion stability through the emulsification test. This study suggests that Schizochytrium oil-processing byproducts represent a promising source of protein, with potential applications as functional ingredients, offering implications for the usage of these byproducts in various industries. Full article
(This article belongs to the Section Microbial Metabolism, Physiology & Genetics)
Show Figures

Figure 1

Figure 1
<p>Optimization of single-factor experiments for alkaline leaching of DOS proteins: (<b>a</b>) effect of pH on protein extraction rate; (<b>b</b>) effect of temperature on protein extraction rate; (<b>c</b>) effect of solid–liquid ratio on protein extraction rate; (<b>d</b>) effect of time on protein extraction rate. Different lowercase letters in the diagram suggest significant differences (<span class="html-italic">p</span> &lt; 0.05).</p>
Full article ">Figure 2
<p>Response surface and contour plots for each factor. Temperature (A), pH (B), and time (C) response surface plots (<b>a</b>,<b>c</b>,<b>e</b>) and contour plots (<b>b</b>,<b>d</b>,<b>f</b>).</p>
Full article ">Figure 3
<p>Effect of acid precipitation conditions on SCP precipitation rate and content: (<b>a</b>) effect of acid precipitation pH on SCP precipitation rate and content; (<b>b</b>) effect of acid precipitation temperature on precipitation rate and content of SCP.</p>
Full article ">Figure 4
<p>Effect of SCP solution concentration and pH on the interfacial tension of oil and water: (<b>a</b>) variation in interfacial tension at the oil–water interface with time for different SCP protein concentrations (0.1–2% <span class="html-italic">w</span>/<span class="html-italic">v</span>) and WP (0.6% and 1.5% <span class="html-italic">w</span>/<span class="html-italic">v</span>); (<b>b</b>) interfacial tension of SCP at pH 3.0–9.0.</p>
Full article ">Figure 5
<p>EAI and ESI: (<b>a</b>) EAI and ESI at different concentrations of SCP (0.1–2% <span class="html-italic">w</span>/<span class="html-italic">v</span>) and WP (0.6% and 1.5% <span class="html-italic">w</span>/<span class="html-italic">v</span>); (<b>b</b>) EAI and ESI of SCP at different pH (3–9) conditions. WP 1 refers to a WP emulsion with a protein concentration of 0.6%, and WP 2 refers to a WP emulsion with a protein concentration of 1.5%.</p>
Full article ">Figure 6
<p>Mean emulsion droplet diameters (D<sub>4,3</sub>): (<b>a</b>) Mean droplet diameters of emulsions (D<sub>4,3</sub>) with different concentrations of SCP (0.1–2% <span class="html-italic">w</span>/<span class="html-italic">v</span>) and WP (0.6% and 1.5% <span class="html-italic">w</span>/<span class="html-italic">v</span>). (<b>b</b>) Mean droplet diameters of SCP emulsions (D<sub>4,3</sub>) at different pH (3–9). WP 1 refers to a WP emulsion with a protein concentration of 0.6% and WP 2 refers to a WP emulsion with a protein concentration of 1.5%.</p>
Full article ">Figure 7
<p>Micrographs of SCP emulsions at 0 and 30 days after storage: (<b>a</b>) micrographs of different concentrations of SCP (0.1–2% <span class="html-italic">w</span>/<span class="html-italic">v</span>) and WP (0.6% and 1.5% <span class="html-italic">w</span>/<span class="html-italic">v</span>) emulsions at 0 and 30 days after placement; (<b>b</b>) micrographs of SCP emulsions at different pH (3–7) at 0 and 30 days after placement.</p>
Full article ">Figure 8
<p>Creaming index of emulsions and appearance of SCP emulsion (after 0 h and 30 d of storage): (<b>a</b>) creaming index of emulsions and appearance of SCP (0.1–2% <span class="html-italic">w</span>/<span class="html-italic">v</span>) and WP (0.6% and 1.5% <span class="html-italic">w</span>/<span class="html-italic">v</span>) emulsions at different concentrations over a 30-day period of placement; (<b>b</b>) creaming index of emulsions and appearance of SCP emulsions at different pH (3–9) during 30 days of placement.</p>
Full article ">
18 pages, 5690 KiB  
Article
Transcription Factors Sox8 and Sox10 Contribute with Different Importance to the Maintenance of Mature Oligodendrocytes
by Lisa Mirja Jörg, Ursula Schlötzer-Schrehardt, Véronique Lefebvre, Elisabeth Sock and Michael Wegner
Int. J. Mol. Sci. 2024, 25(16), 8754; https://doi.org/10.3390/ijms25168754 (registering DOI) - 11 Aug 2024
Abstract
Myelin-forming oligodendrocytes in the vertebrate nervous system co-express the transcription factor Sox10 and its paralog Sox8. While Sox10 plays crucial roles throughout all stages of oligodendrocyte development, including terminal differentiation, the loss of Sox8 results in only mild and transient perturbations. Here, we [...] Read more.
Myelin-forming oligodendrocytes in the vertebrate nervous system co-express the transcription factor Sox10 and its paralog Sox8. While Sox10 plays crucial roles throughout all stages of oligodendrocyte development, including terminal differentiation, the loss of Sox8 results in only mild and transient perturbations. Here, we aimed to elucidate the roles and interrelationships of these transcription factors in fully differentiated oligodendrocytes and myelin maintenance in adults. For that purpose, we conducted targeted deletions of Sox10, Sox8, or both in the brains of two-month-old mice. Three weeks post-deletion, none of the resulting mouse mutants exhibited significant alterations in oligodendrocyte numbers, myelin sheath counts, myelin ultrastructure, or myelin protein levels in the corpus callosum, despite efficient gene inactivation. However, differences were observed in the myelin gene expression in mice with Sox10 or combined Sox8/Sox10 deletion. RNA-sequencing analysis on dissected corpus callosum confirmed substantial alterations in the oligodendrocyte expression profile in mice with combined deletion and more subtle changes in mice with Sox10 deletion alone. Notably, Sox8 deletion did not affect any aspects of the expression profile related to the differentiated state of oligodendrocytes or myelin integrity. These findings extend our understanding of the roles of Sox8 and Sox10 in oligodendrocytes into adulthood and have important implications for the functional relationship between the paralogs and the underlying molecular mechanisms. Full article
(This article belongs to the Special Issue The Function of Glial Cells in the Nervous System)
Show Figures

Graphical abstract

Graphical abstract
Full article ">Figure 1
<p>Efficient gene inactivation in mutant mice. (<b>a</b>–<b>l</b>) Co-immunohistochemical stainings of the corpus callosum from the control mice (ctrl, (<b>a</b>–<b>c</b>)), Sox8 mutant mice (Sox8cko, (<b>d</b>–<b>f</b>)), Sox10 mutant mice (Sox10cko, (<b>g</b>–<b>i</b>)), and Sox8/Sox10 double-mutant mice (dcko, (<b>j</b>–<b>l</b>)) at 21 dpi, with antibodies directed against Sox8 (<b>a</b>,<b>d</b>,<b>g</b>,<b>j</b>), Sox10 (<b>b</b>,<b>e</b>,<b>h</b>,<b>k</b>), and Sox9 (<b>c</b>,<b>f</b>,<b>i</b>,<b>l</b>). Scale bar: 50 µm. Microscope enlargement: 200×. (<b>m</b>) Quantification of the percentage of Sox8-, Sox10-, or Sox9-expressing cells among all cells in the corpus callosum (identified by DAPI stain) in the various mouse lines. (<b>n</b>) The absolute numbers of cells expressing Sox8, Sox10, or Sox9 per mm<sup>2</sup> in the corpus callosum of the various mouse lines. Bar graphs show the mean ± standard error of the mean (n = 4). The statistical significance relative to the controls was determined for each Sox protein by Student’s <span class="html-italic">t</span>-test (***, <span class="html-italic">p</span> ≤ 0.001).</p>
Full article ">Figure 2
<p>Unchanged oligodendroglial cell numbers in mutant mice with Sox8 and Sox10 deletions. (<b>a</b>–<b>l</b>) Immunohistochemical stainings of the corpus callosum from the control mice (ctrl, (<b>a</b>,<b>e</b>,<b>i</b>)), Sox8 mutant mice (Sox8cko, (<b>b</b>,<b>f</b>,<b>j</b>)), Sox10 mutant mice (Sox10cko, (<b>c</b>,<b>g</b>,<b>k</b>)), and Sox8/Sox10 double-mutant mice (dcko, (<b>d</b>,<b>h</b>,<b>l</b>)) at 21 dpi, with antibodies directed against Olig2 (<b>a</b>–<b>d</b>), Pdgfra (<b>e</b>–<b>h</b>), and Sox6 (<b>i</b>–<b>l</b>). Scale bar: 50 µm. Microscope enlargement: 200×. (<b>m</b>–<b>o</b>) Quantification of the percentage of cells expressing Olig2 (<b>m</b>), Pdgfra (<b>n</b>), or Sox6 (<b>o</b>) among all the cells in the corpus callosum (identified by DAPI stain) of the various mouse lines. Bar graphs show the mean ± standard error of the mean (n = 4). No statistical significance was determined for any of the mutants or markers by Student’s <span class="html-italic">t</span>-test.</p>
Full article ">Figure 3
<p>Mild reductions in the myelin protein amounts in mutant mice with Sox8 and Sox10 deletions. (<b>a</b>) Quantification of signal intensities for Bcas1, Mbp, Plp1, and Mog following immunohistochemical staining of the corpus callosum of the control (ctrl), Sox8 mutant (Sox8cko), Sox10 mutant (Sox10cko), and Sox8/Sox10 double-mutant (dcko) mouse lines at 21 dpi. Bar graphs show the mean ± standard error of the mean (n = 4) for each genotype with intensities for the control animals set to 1. (<b>b</b>) Quantification of the percentage of the Aspa-expressing cells among all cells (identified by DAPI stain) in the corpus callosum of the various mouse lines. Statistical significance relative to the controls was determined for each marker by Student’s <span class="html-italic">t</span>-test (*, <span class="html-italic">p</span> ≤ 0.05; ***, <span class="html-italic">p</span> ≤ 0.001). (<b>c</b>–<b>j</b>) Exemplary immunohistochemical stainings of the corpus callosum from the control (<b>c</b>,<b>g</b>), Sox8cko (<b>d</b>,<b>h</b>), Sox10cko (<b>e</b>,<b>i</b>), and dcko (<b>f</b>,<b>j</b>) mice with antibodies directed against Mbp (<b>c</b>–<b>f</b>) and Aspa (<b>g</b>–<b>j</b>). Scale bar: 50 µm. Microscope enlargement: 200×.</p>
Full article ">Figure 4
<p>Unaffected myelin sheaths in mutant mice with Sox8 and Sox10 deletions. (<b>a</b>–<b>d</b>) Electron microscopic pictures of the sagittally cut corpus callosum from the control mice (ctrl, (<b>a</b>)), Sox8 mutant mice (Sox8cko, (<b>b</b>)), Sox10 mutant mice (Sox10cko, (<b>c</b>)), and Sox8/Sox10 double-mutant mice (dcko, (<b>d</b>)) at 21 dpi. Electron microscope enlargement: 3600×. (<b>e</b>) Quantification of the absolute number of unmyelinated axons per 100 µm<sup>2</sup> of sagittally cut corpus callosum in the various mouse mutants. (<b>f</b>–<b>i</b>) Representation of the g-ratios in the corpus callosum of the controls and the various mouse mutants, as determined over all axons (<b>f</b>) after binning of axons according to diameter size (<b>g</b>) or presented for single axons in a scatter plot (<b>h</b>,<b>i</b>). (<b>j</b>–<b>r</b>) Immunohistochemical stainings of the corpus callosum from the control (<b>j</b>,<b>n</b>), Sox8cko (<b>k</b>,<b>o</b>), Sox10cko (<b>l</b>,<b>p</b>), and dcko (<b>m</b>,<b>q</b>) mice at 21 dpi, with antibodies directed against Iba1 (<b>j</b>–<b>m</b>) and Gfap (<b>n</b>–<b>q</b>), as well as the corresponding quantification of signal intensities (<b>r</b>). Bar graphs show the mean ± standard error of the mean (n = 4) for each genotype, with intensities for the control animals set to 1. No statistical significance relative to the controls was determined for any of the parameters or markers by Student’s <span class="html-italic">t</span>-test. Scale bar: 50 µm. Microscope enlargement: 200×.</p>
Full article ">Figure 5
<p>Altered expression of the myelin genes in mutant mice with Sox8 and Sox10 deletions. (<b>a</b>–<b>i</b>) In situ hybridization of the corpus callosum from the control (ctrl, <b>a</b>,<b>e</b>), Sox8 mutant (Sox8cko, <b>b</b>,<b>f</b>), Sox10 mutant (Sox10cko, <b>c</b>,<b>g</b>), and Sox8/Sox10 double-mutant (dcko, <b>d</b>,<b>h</b>) mice at 21 dpi with antisense probes directed against <span class="html-italic">Plp1</span> (<b>a</b>–<b>d</b>) and <span class="html-italic">Mog</span> (<b>e</b>–<b>h</b>), as well as corresponding quantifications (<b>i</b>). The bar graph shows the absolute numbers of the transcript-positive cells per mm<sup>2</sup> in the corpus callosum of the various mouse lines as the mean ± standard error of the mean (n = 4) for each genotype. Scale bar: 50 µm. Microscope enlargement: 200×. (<b>j</b>) Expression of the <span class="html-italic">Sox8</span>, <span class="html-italic">Sox10</span>, <span class="html-italic">Mbp</span>, <span class="html-italic">Mog</span>, <span class="html-italic">Plp1</span>, and <span class="html-italic">Fa2h</span> in the corpus callosum from the control and mutant mice, as determined by quantitative RT-PCR. Values represent the mean relative expression levels ± standard error of the mean (n = 4 per genotype), with transcript levels of each marker set to 1 in the controls. Statistical significance relative to the controls was determined for each marker by Student’s <span class="html-italic">t</span>-test (***, <span class="html-italic">p</span> ≤ 0.001).</p>
Full article ">Figure 6
<p>Changed expression profiles in the corpus callosum of mice with oligodendrocyte-specific deletions of Sox8 and Sox10. (<b>a</b>–<b>f</b>) PCA plots showing pairwise comparisons of the results from the control (ctrl), Sox8 mutant (Sox8cko), Sox10 mutant (Sox10cko), and Sox8/Sox10 double-mutant (dcko) mice obtained by RNA sequencing of corpus callosum samples. (<b>g</b>–<b>i</b>) Pie charts summarizing the number of upregulated (red) or downregulated (blue) differentially expressed genes (DEGs, as defined by a log2-fold change of ≥±0.75, <span class="html-italic">p</span>-value of ≤0.05, and a mean base count of &gt;20 gene-specific transcripts per million transcripts) in the corpus callosum of the Sox8cko (<b>g</b>), Sox10cko, (<b>h</b>) and dcko mice. (<b>j</b>–<b>l</b>) Graphical representation of the DEGs in Sox8cko (<b>j</b>), Sox10cko (<b>k</b>), and dcko (<b>l</b>) mice according to the <span class="html-italic">p</span>-value and log2-fold change (logfc) in Volcano plots.</p>
Full article ">Figure 7
<p>Altered cellular processes in the corpus callosum of mice with oligodendrocyte-specific deletion of Sox8 and Sox10. (<b>a</b>–<b>i</b>) Gene set enrichment analyses (GSEA; (<b>a</b>,<b>d</b>,<b>g</b>)) and gene ontology (GO) studies on the upregulated (<b>b</b>,<b>e</b>,<b>h</b>) and downregulated (<b>c</b>,<b>f</b>,<b>i</b>) differentially expressed genes (as defined by a log2-fold change of ≥±0.75, <span class="html-italic">p</span>-value of ≤0.05, and a mean base count of &gt;20 gene-specific transcripts per million transcripts) according to the RNA-sequencing data from the corpus callosum of the Sox8 mutant (Sox8cko, (<b>a</b>–<b>c</b>)), Sox10 mutant (Sox10cko, (<b>d</b>–<b>f</b>)), and Sox8/Sox10 double-mutant (dcko, (<b>g</b>–<b>i</b>)) mice. The top 15 terms (if present) according to adjusted <span class="html-italic">p</span>-values are listed. Terms related to glia are labeled in dark red, those related to lipid and cholesterol metabolism are in light red; all others are in gray.</p>
Full article ">Figure 8
<p>Altered expression of select genes in the corpus callosum of the Sox8 mutant (Sox8cko), Sox10 mutant (Sox10cko), and Sox8/Sox10 double-mutant (dcko) mice. (<b>a</b>–<b>l</b>) Expression levels of <span class="html-italic">Mbp</span> (<b>a</b>), <span class="html-italic">Plp1</span> (<b>b</b>), <span class="html-italic">Mog</span> (<b>c</b>), <span class="html-italic">Mag</span> (<b>d</b>), <span class="html-italic">Mal</span> (<b>e</b>), <span class="html-italic">Opalin</span> (<b>f</b>), <span class="html-italic">Aspa</span> (<b>g</b>), <span class="html-italic">Fa2h</span> (<b>h</b>), <span class="html-italic">Elovl1</span> (<b>i</b>), <span class="html-italic">Enpp6</span> (<b>j</b>), <span class="html-italic">Gal3st1</span> (<b>k</b>), and <span class="html-italic">Ugt8a</span> (<b>l</b>) in the control, Sox8cko, Sox10cko, and dcko mice according to RNA-sequencing data and are represented as gene-specific transcript per million transcripts (TPM). Statistical significance relative to the controls was determined for each marker by Student’s <span class="html-italic">t</span>-test (*, <span class="html-italic">p</span> ≤ 0.05; **, <span class="html-italic">p</span> ≤ 0.01; and ***, <span class="html-italic">p</span> ≤ 0.001).</p>
Full article ">Figure 9
<p>Comparison of the expression changes in mice with oligodendrocyte-specific deletion of Sox8 and Sox10. (<b>a</b>–<b>f</b>) Venn diagrams showing the overlap of upregulated (<b>a</b>–<b>c</b>) and downregulated (<b>d</b>–<b>f</b>) genes between the Sox8 mutant (Sox8cko) and Sox10 mutant (Sox10cko) mice (<b>a</b>,<b>b</b>); Sox8cko and Sox8/Sox10 double-mutant (dcko) mice (<b>c</b>,<b>d</b>); or the Sox10cko and dcko (<b>e</b>,<b>f</b>) mice. Jointly deregulated genes were defined as genes that are similarly up or downregulated in both mutant mice with a <span class="html-italic">p</span>-value ≤ 0.05 and achieving a log2fold deregulation of ≥±0.75 in at least one mutant.</p>
Full article ">
14 pages, 2604 KiB  
Article
The Mitochondrial Genome of Ylistrum japonicum (Bivalvia, Pectinidae) and Its Phylogenetic Analysis
by Yida Han, Yaoyu Xie, Zhenlin Hao, Junxia Mao, Xubo Wang, Yaqing Chang and Ying Tian
Int. J. Mol. Sci. 2024, 25(16), 8755; https://doi.org/10.3390/ijms25168755 (registering DOI) - 11 Aug 2024
Abstract
The Ylistrum japonicum is a commercially valuable scallop known for its long-distance swimming abilities. Despite its economic importance, genetic and genomic research on this species is limited. This study presents the first complete mitochondrial genome of Y. japonicum. The mitochondrial genome [...] Read more.
The Ylistrum japonicum is a commercially valuable scallop known for its long-distance swimming abilities. Despite its economic importance, genetic and genomic research on this species is limited. This study presents the first complete mitochondrial genome of Y. japonicum. The mitochondrial genome is 19,475 bp long and encompasses 13 protein-coding genes, three ribosomal RNA genes, and 23 transfer RNA genes. Two distinct phylogenetic analyses were used to explore the phylogenetic position of the Y. japonicum within the family Pectinidae. Based on one mitochondrial phylogenetic analysis by selecting 15 Pectinidae species and additional outgroup taxa and one single gene phylogenetic analysis by 16S rRNA, two phylogenetic trees were constructed to provide clearer insights into the evolutionary placement of Y. japonicum within the family Pectinidae. Our analysis reveals that Ylistrum is a basal lineage to the Pectininae clade, distinct from its previously assigned tribe, Amusiini. This study offers critical insights into the genetic makeup and evolutionary history of Y. japonicum, enhancing our knowledge of this economically vital species. Full article
(This article belongs to the Section Molecular Biology)
Show Figures

Figure 1

Figure 1
<p>Mitochondrial genome map of <span class="html-italic">Ylistrum japonicum</span>.</p>
Full article ">Figure 2
<p>Gene orders of <span class="html-italic">Ylistrum japonicum</span>, <span class="html-italic">Argopecten irradians irradians</span>, <span class="html-italic">Amusium pleuronectes</span>, and <span class="html-italic">Chlamys farreri</span>, with newly annotated atp8 genes. The same color indicates identical gene junctions (excluding the tRNA genes).</p>
Full article ">Figure 3
<p>Gene collinearity analysis of 5 Pectininae species. The level of similarity at each position is shown in the blocks. The white, red, and green boxes represent protein-coding, rRNA, and tRNA genes.</p>
Full article ">Figure 4
<p>Phylogenetic tree derived from Maximum likelihood (ML) and Bayesian inference (BI) based on the sequences of mitochondrial protein-coding genes (PCGs). Numbers above the branches indicate bootstrap support; numbers below branches are Bayesian posterior probability. A dash indicates no support for that node.</p>
Full article ">Figure 5
<p>Phylogenetic tree of genus <span class="html-italic">Ylistrum</span> and some species from three tribes of Pectininae inferred by Maximum likelihood (ML) of 16S rRNA sequences. Numbers indicate bootstrap support.</p>
Full article ">Figure 6
<p>Sampling location of <span class="html-italic">Ylistrum japonicum</span> (modified from d-maps: <a href="https://d-maps.com" target="_blank">https://d-maps.com</a>, accessed on 10 May 2024).</p>
Full article ">
12 pages, 468 KiB  
Article
Effect of Adding the Antimicrobial L-Carnitine to Growing Rabbits’ Drinking Water on Growth Efficiency, Hematological, Biochemical, and Carcass Aspects
by Mohamed I. Hassan, Naela Abdel-Monem, Ayman Moawed Khalifah, Saber S. Hassan, Hossam Shahba, Ahmad R. Alhimaidi, In Ho Kim and Hossam M. El-Tahan
Antibiotics 2024, 13(8), 757; https://doi.org/10.3390/antibiotics13080757 (registering DOI) - 11 Aug 2024
Abstract
The current study was designed to assess the impact of L-carnitine (LC) supplementation in the drinking water of growing Alexandria-line rabbits on performance and physiological parameters. Two hundred eighty-eight 35-day-old rabbits were divided into four groups of twenty-four replicates each (seventy-two rabbits/treatment). The [...] Read more.
The current study was designed to assess the impact of L-carnitine (LC) supplementation in the drinking water of growing Alexandria-line rabbits on performance and physiological parameters. Two hundred eighty-eight 35-day-old rabbits were divided into four groups of twenty-four replicates each (seventy-two rabbits/treatment). The treatment groups were a control group without LC and three groups receiving 0.5, 1, and 1.5 g/L LC in the drinking water intermittently. The results showed that the group receiving 0.5 g LC/L exhibited significant improvements in final body weight, body weight gain, feed conversion ratio, and performance index compared to the other groups. The feed intake remained unaffected except for the 1.5 g LC/L group, which had significantly decreased intake. Hematological parameters improved in all supplemented groups. Compared with those in the control group, the 0.5 g LC/L group showed significant increases in serum total protein and high-density lipoprotein, along with decreased cholesterol and low-density lipoprotein. Compared to other supplemented groups, this group also demonstrated superior carcass traits (carcass, dressing, giblets, and percentage of nonedible parts). In conclusion, intermittent supplementation of LC in the drinking water, particularly at 0.5 g/L twice a week, positively influenced the productivity, hematology, serum lipid profile, and carcass traits of Alexandria-line growing rabbits at 84 days of age. Full article
(This article belongs to the Special Issue Natural Compounds as Antimicrobial Agents, 2nd Edition)
24 pages, 9689 KiB  
Article
Genome-Wide Identification, Evolution, and Expression Analysis of the Dirigent Gene Family in Cassava (Manihot esculenta Crantz)
by Mingchao Li, Kai Luo, Wenke Zhang, Man Liu, Yunfei Zhang, Huling Huang, Yinhua Chen, Shugao Fan and Rui Zhang
Agronomy 2024, 14(8), 1758; https://doi.org/10.3390/agronomy14081758 (registering DOI) - 11 Aug 2024
Abstract
Dirigent (DIR) genes play a pivotal role in plant development and stress adaptation. Manihot esculenta Crantz, commonly known as cassava, is a drought-resistant plant thriving in tropical and subtropical areas. It is extensively utilized for starch production, bioethanol, and animal feed. [...] Read more.
Dirigent (DIR) genes play a pivotal role in plant development and stress adaptation. Manihot esculenta Crantz, commonly known as cassava, is a drought-resistant plant thriving in tropical and subtropical areas. It is extensively utilized for starch production, bioethanol, and animal feed. However, a comprehensive analysis of the DIR family genes remains unexplored in cassava, a crucial cash and forage crop in tropical and subtropical regions. In this study, we characterize a total of 26 cassava DIRs (MeDIRs) within the cassava genome, revealing their uneven distribution across 13 of the 18 chromosomes. Phylogenetic analysis classified these genes into four subfamilies: DIR-a, DIR-b/d, DIR-c, and DIR-e. Comparative synteny analysis with cassava and seven other plant species (Arabidopsis (Arabidopsis thaliana), poplar (Populus trichocarpa), soybean (Glycine max), tomato (Solanum lycopersicum), rice (Oryza sativa), maize (Zea mays), and wheat (Triticum aestivum)) provided insights into their likely evolution. We also predict protein interaction networks and identify cis-acting elements, elucidating the functional differences in MeDIR genes. Notably, MeDIR genes exhibited specific expression patterns across different tissues and in response to various abiotic and biotic stressors, such as pathogenic bacteria, cadmium chloride (CdCl2), and atrazine. Further validation through quantitative real-time PCR (qRT-PCR) confirmed the response of DIR genes to osmotic and salt stress. These findings offer a comprehensive resource for understanding the characteristics and biological functions of MeDIR genes in cassava, enhancing our knowledge of plant stress adaptation mechanisms. Full article
21 pages, 6311 KiB  
Article
Investigation of Antioxidant Activity of Protein Hydrolysates from New Zealand Commercial Low-Grade Fish Roes
by Shuxian Li, Alan Carne and Alaa El-Din Ahmed Bekhit
Mar. Drugs 2024, 22(8), 364; https://doi.org/10.3390/md22080364 (registering DOI) - 11 Aug 2024
Abstract
The objective of this study was to investigate the nutrient composition of low-grade New Zealand commercial fish (Gemfish and Hoki) roe and to investigate the effects of delipidation and freeze-drying processes on roe hydrolysis and antioxidant activities of their protein hydrolysates. Enzymatic hydrolysis [...] Read more.
The objective of this study was to investigate the nutrient composition of low-grade New Zealand commercial fish (Gemfish and Hoki) roe and to investigate the effects of delipidation and freeze-drying processes on roe hydrolysis and antioxidant activities of their protein hydrolysates. Enzymatic hydrolysis of the Hoki and Gemfish roe homogenates was carried out using three commercial proteases: Alcalase, bacterial protease HT, and fungal protease FP-II. The protein and lipid contents of Gemfish and Hoki roes were 23.8% and 7.6%; and 17.9% and 10.1%, respectively. The lipid fraction consisted mainly of monounsaturated fatty acid (MUFA) in both Gemfish roe (41.5%) and Hoki roe (40.2%), and docosahexaenoic (DHA) was the dominant polyunsaturated fatty acid (PUFA) in Gemfish roe (21.4%) and Hoki roe (18.6%). Phosphatidylcholine was the main phospholipid in Gemfish roe (34.6%) and Hoki roe (28.7%). Alcalase achieved the most extensive hydrolysis, and its hydrolysate displayed the highest 2,2-dipheny1-1-picrylhydrazyl (DPPH)˙ and 2,2′-azino-bis(3-ethylbenzothiazoline-6-sulfonic acid) (ABTS) radical scavenging activities and ferric reducing antioxidant power (FRAP). The combination of defatting and freeze-drying treatments reduced DPPH˙ scavenging activity (by 38%), ABTS˙ scavenging activity (by 40%) and ferric (Fe3+) reducing power by18% (p < 0.05). These findings indicate that pre-processing treatments of delipidation and freeze-drying could negatively impact the effectiveness of enzymatic hydrolysis in extracting valuable compounds from low grade roe. Full article
(This article belongs to the Special Issue The Bioactive Potential of Marine-Derived Peptides and Proteins)
Show Figures

Figure 1

Figure 1
<p>Time course hydrolysis of frozen-thawed Hoki and Gemfish roes with lipid present. Homogenates of fresh Hoki and Gemfish roe were prepared by adding 7.3 g and 5.5 g of roe to 200 mL potassium phosphate buffer (pH 7) to achieve 6.5 mg/mL protein concentration, and aliquots of 20 mL were subject to hydrolysis with three different concentrations (2%, 6%, and 10%) of either microbial protease Alcalase (<span class="html-italic">v</span>/<span class="html-italic">w</span>), bacterial protease HT (<span class="html-italic">w</span>/<span class="html-italic">w</span>), or fungal protease FP-II (<span class="html-italic">w</span>/<span class="html-italic">w</span>), respectively, at 45 °C. Alcalase is commercially available as a solution, and HT and FP-II are commercially available as powders. The degree of hydrolysis was determined based on the L-serine equivalent. The data were obtained from three independent hydrolyses for each roe and protease combination. Analysis of variance (ANOVA) was carried out. Letters a–c indicate significant differences among different samples prepared with varying enzyme concentrations at the same time point. (<span class="html-italic">p</span> &lt; 0.05).</p>
Full article ">Figure 2
<p>Time course hydrolysis of fresh Hoki and Gemfish roes after removing the lipid fraction. Before hydrolysis, the Hoki roe and Gemfish roe homogenates were delipidated by ethanol and hexane using the ETHEX lipid extraction method. For hydrolysis, homogenates of delipidated frozen-thawed Hoki and Gemfish roes were prepared by adding 7.3 g and 5.5 g of thawed delipidation roe to 200 mL of potassium phosphate buffer (pH 7) to achieve 6.5 mg/mL protein concentration and aliquots of 20 mL were subject to hydrolysis at three different concentrations (2%, 6%, and 10%) of either Alcalase (<span class="html-italic">v/w</span>), bacterial protease HT (<span class="html-italic">w/w</span>), or fungal protease FP-II(<span class="html-italic">w/w</span>) at 45 °C. The degree of hydrolysis of the samples was determined based on the L-serine equivalent method. The data were obtained from three independent hydrolyses for each roe and protease combination. Analysis of variance (ANOVA) was carried out. Letters a–c indicate significant differences among different samples prepared with varying enzyme concentrations at the same time point (<span class="html-italic">p</span> &lt; 0.05).</p>
Full article ">Figure 3
<p>Time course hydrolysis of freeze-dried Hoki and Gemfish roe without lipid extraction. Before hydrolysis, the Hoki roe and Gemfish roe homogenates were freeze-dried. For hydrolysis, homogenates of freeze-dried Hoki and Gemfish roe were prepared by adding 2.76 g and 1.96 g of freeze-dried roe powder to 200 mL of potassium phosphate buffer (pH 7) to achieve the protein concentration 6.5 mg/mL, and aliquots of 20 mL were subjected to hydrolysis with three different concentrations (2%, 6%, and 10%) of either Alcalase (v/w), bacterial protease HT (<span class="html-italic">w/w</span>), or fungal protease FP-II (<span class="html-italic">w/w</span>) at 45 °C. The data were obtained from three independent hydrolyses for each roe and protease combination. Analysis of variance (ANOVA) was carried out. Letters a–c indicate significant differences among different samples prepared with varying enzyme concentrations at the same time point (<span class="html-italic">p</span> &lt; 0.05).</p>
Full article ">Figure 4
<p>Time course hydrolysis of freeze-dried Hoki and Gemfish roe without lipid present. Before hydrolysis, Hoki roe and Gemfish roe homogenate delipidated and freeze-dried by the ETHEX method and the freeze drier. For hydrolysis, homogenates of freeze-dried Hoki and Gemfish roe were prepared by adding 2.07 g and 2.12 g of delipidated freeze-dried roe powder to 200 mL of potassium phosphate buffer (pH 7), and aliquots of 20 mL were subject to hydrolysis with three different concentrations (2%, 6%, and 10%) of either plant-based protease Alcalase (<span class="html-italic">v/w</span>), bacterial protease HT (<span class="html-italic">w/w</span>), or fungal protease FP-II (<span class="html-italic">w/w</span>), respectively, at 45 °C. Analysis of variance (ANOVA) was carried out. Letters a–c indicate significant differences among different samples prepared with varying enzyme concentrations at the same time point (<span class="html-italic">p</span> &lt; 0.05).</p>
Full article ">Figure 5
<p>Time course hydrolysis of freeze-dried Hoki and Gemfish roe without lipid present. 1D SDS-PAGE of protein hydrolysis profiles of three proteases (Alcalase, HT, and FP-II) with the concentration [2%, 6%, and 10% (<span class="html-italic">w</span>/<span class="html-italic">v</span> or <span class="html-italic">v</span>/<span class="html-italic">v</span>)] control before 24 h incubation (<b>a</b>), control after 24 h incubation (<b>b</b>), and the sample control of Hoki and Gemfish roe homogenate with 4 treatments before hydrolysis (<b>c</b>). F = fresh, dL = delipidation, FD = freeze-dried, and FD-dL = freeze-dried with delipidation. A darker blue band indicates a higher concentration of protein.</p>
Full article ">Figure 6
<p>1D SDS-PAGE of protein hydrolysis profiles of Hoki roe (<b>a</b>,<b>c</b>,<b>e</b>) and Gemfish roe (<b>b</b>,<b>d</b>,<b>f</b>) homogenate. Alcalase (<b>a</b>,<b>b</b>), HT (<b>c</b>,<b>d</b>), and FP-II (<b>e</b>,<b>f</b>) all with the same amount of protease concentration (10% <span class="html-italic">v</span>/<span class="html-italic">w</span> or <span class="html-italic">w</span>/<span class="html-italic">w</span>) and effect of 4 treatments (F, dL, FD, and dL-FD) on protein hydrolysis of Hoki and Gemfish roe homogenate. F = fresh, dL = delipidation, FD = freeze-dried, and FD-dL: freeze-dried with delipidation. A darker blue band indicates a higher concentration of protein.</p>
Full article ">Figure 7
<p>DPPH radical scavenging activity (<b>a</b>), ABTS radical scavenging activity (<b>b</b>), and Ferric (Fe<sup>3+</sup>) reducing power (<b>c</b>). Different enzyme treatments and fish roe homogenates (<b>I</b>), different enzyme treatments and fish roe homogenates (<b>II</b>), and different treatments and enzyme treatments (<b>III</b>). (a–f) indicating a significant difference in antioxidant activity among different treatments (<span class="html-italic">p</span> &lt; 0.05). F = fresh, dL = delipidation, FD = freeze-dried, and FD-dL: freeze-dried with delipidation.</p>
Full article ">
18 pages, 4884 KiB  
Article
Genome-Wide Identification of B-Box Family Genes and Their Potential Roles in Seed Development under Shading Conditions in Rapeseed
by Si Chen, Yushan Qiu, Yannong Lin, Songling Zou, Hailing Wang, Huiyan Zhao, Shulin Shen, Qinghui Wang, Qiqi Wang, Hai Du, Jiana Li and Cunmin Qu
Plants 2024, 13(16), 2226; https://doi.org/10.3390/plants13162226 (registering DOI) - 11 Aug 2024
Abstract
B-box (BBX) proteins, a subfamily of zinc-finger transcription factors, are involved in various environmental signaling pathways. In this study, we conducted a comprehensive analysis of BBX family members in Brassica crops. The 482 BBX proteins were divided into five groups based on gene [...] Read more.
B-box (BBX) proteins, a subfamily of zinc-finger transcription factors, are involved in various environmental signaling pathways. In this study, we conducted a comprehensive analysis of BBX family members in Brassica crops. The 482 BBX proteins were divided into five groups based on gene structure, conserved domains, and phylogenetic analysis. An analysis of nonsynonymous substitutions and (Ka)/synonymous substitutions (Ks) revealed that most BBX genes have undergone purifying selection during evolution. An analysis of transcriptome data from rapeseed (Brassica napus) organs suggested that BnaBBX3d might be involved in the development of floral tissue-specific RNA-seq expression. We identified numerous light-responsive elements in the promoter regions of BnaBBX genes, which were suggestive of participation in light signaling pathways. Transcriptomic analysis under shade treatment revealed 77 BnaBBX genes with significant changes in expression before and after shading treatment. Of these, BnaBBX22e showed distinct expression patterns in yellow- vs. black-seeded materials in response to shading. UPLC-HESI-MS/MS analysis revealed that shading influences the accumulation of 54 metabolites, with light response BnaBBX22f expression correlating with the accumulation of the flavonoid metabolites M46 and M51. Additionally, BnaBBX22e and BnaBBX22f interact with BnaA10.HY5. These results suggest that BnaBBXs might function in light-induced pigment accumulation. Overall, our findings elucidate the characteristics of BBX proteins in six Brassica species and reveal a possible connection between light and seed coat color, laying the foundation for further exploring the roles of BnaBBX genes in seed development. Full article
(This article belongs to the Special Issue Molecular Genetics and Breeding of Oilseed Crops—2nd Edition)
Show Figures

Figure 1

Figure 1
<p>Phylogenetic tree of B-box genes from <span class="html-italic">Arabidopsis</span> and six <span class="html-italic">Brassica</span> U-triangle species. The phylogenetic tree, constructed using the protein matrix with iq-tree, is grouped into five clades (I–V) labeled with different colors. The phylogenetic tree was visualized using B-box genes from Arabidopsis, <span class="html-italic">B. rapa</span>, <span class="html-italic">B. nigra</span>, <span class="html-italic">B. oleracea</span>, <span class="html-italic">B. juncea</span>, <span class="html-italic">B. napus</span>, and <span class="html-italic">B. carinata</span>.</p>
Full article ">Figure 2
<p>WebLogos of conserved domains of BBX family members in six <span class="html-italic">Brassica</span> plants. (<b>A</b>–<b>C</b>) Typical WebLogos of B-box1, B-box2, and CCT domains in the A, B, and C subgenomes. The <span class="html-italic">x</span>-axis depicts the conserved sequences of the structural domains, where the height of each letter signifies the degree of conservation of each residue across all proteins. The <span class="html-italic">y</span>-axis indicates the relative entropy scale, which represents the conservation degree of each amino acid.</p>
Full article ">Figure 3
<p>Number of BBX family gene pairs in the six <span class="html-italic">Brassicaceae</span> species. (<b>A</b>) Number of BBX genes in <span class="html-italic">B. rapa</span> (Bra), <span class="html-italic">B. nigra</span> (Bni), <span class="html-italic">B. oleracea</span> (Bol), <span class="html-italic">B. napus</span> (Bna), <span class="html-italic">B. juncea</span> (Bju), and <span class="html-italic">B. carinata</span> (Bca). (<b>B</b>) Number of BBX family gene pairs in the A subgenome. (<b>C</b>) Number of BBX family gene pairs in the B subgenome. (<b>D</b>) Number of BBX family gene pairs in the C subgenome.</p>
Full article ">Figure 4
<p>Heatmap of the expression patterns of <span class="html-italic">BnaBBXs</span> across different tissues and organs. The expression profiles of each <span class="html-italic">BnaBBX</span> gene are based on log2-transformed values (FPKM value + 1). FPKM, fragments per kilobase of exon model per million mapped fragments; DAF, days after flowering.</p>
Full article ">Figure 5
<p>Expression patterns of BnaBBXs under shading conditions. (<b>A</b>,<b>B</b>) Phenotypes of L1262 (<b>A</b>) and L1263 (<b>B</b>) seeds under shading conditions (TR) and control conditions (CK, seeds under normal conditions) at different time points. D, days after shading. Scale bars, 2 mm. (<b>C</b>) Phenotypes of mature L1262 and L1263 seeds under shading conditions and the corresponding controls. Scale bars, 2 mm. (<b>D</b>) Heatmap of RNA-seq data of BnaBBX genes of seeds under shading. Three biological replicates for each type of sample were used. The transcriptome data of each <span class="html-italic">BnaBBX</span> gene are based on log<sub>2</sub>-transformed values (FPKM value + 1); (<b>D</b>) days after shading. (<b>E</b>) Venn diagram of genes with significant changes in expression (TR/CK, fold change ≥ 2 or fold change ≤ 0.5) before and after shading. (<b>F</b>) qRT-PCR analysis of BnaBBX family genes. The expression level of BnaActin7 was used to normalize the qRT-PCR data. <span class="html-italic">p</span>-values were calculated using multiple Student <span class="html-italic">t</span>-tests, comparing the levels in L1262CK (black-seeded material) and L1263CK (yellow-seeded material). * <span class="html-italic">p</span> &lt; 0.05; ** <span class="html-italic">p</span> &lt; 0.01; *** <span class="html-italic">p</span> &lt; 0.001; ****<span class="html-italic">p</span> &lt; 0.0001; ns, no difference. Error bars denote the standard deviation (SD) from three independent biological replicates. D, days after shading; TR, shading treatment.</p>
Full article ">Figure 6
<p>Validation of protein–protein interactions between BnaBBX22e and BnaA10.HY5. (<b>A</b>) Changes in metabolite contents before and after shading. Black blocks indicate that data were not available. The color scale represents the fold change in metabolite contents before and after shading, with red, blue, and yellow blocks indicating fold change ≥ 2, fold change ≤ 0.5, and fold change &gt; 0.5 and &lt;2, respectively. (<b>B</b>) The relationships among BnaBBX proteins and metabolites (|R value| ≥ 0.8, <span class="html-italic">p</span>-value ≤ 0.05). (<b>C</b>) Protein–protein interaction networks among BnaBBX proteins and BnaA10.HY5 proteins. The two red arrows point to BnaBBX22e and BnaBBX22f. (<b>D</b>) Yeast point-to-point validation between BnaBBX22e and BnaA10.HY5 as well as between BnaBBX22e and BnaA10.HY5. pGBKT7-53 and pGADT7-T, positive control; pGBKT7-lam and pGADT7-T, negative control; empty pGBKT7 and prey vector, testing for autoactivation and toxicity; bait vector and empty pGADT7, testing for autoactivation and toxicity.</p>
Full article ">
14 pages, 10544 KiB  
Article
Bioinformatic Evaluation of KLF13 Genetic Variant: Implications for Neurodevelopmental and Psychiatric Symptoms
by Mirella Vinci, Donatella Greco, Simone Treccarichi, Valeria Chiavetta, Maria Grazia Figura, Antonino Musumeci, Vittoria Greco, Concetta Federico, Francesco Calì and Salvatore Saccone
Genes 2024, 15(8), 1056; https://doi.org/10.3390/genes15081056 (registering DOI) - 11 Aug 2024
Abstract
The Krüppel-like factor (KLF) family represents a group of transcription factors (TFs) performing different biological processes that are crucial for proper neuronal function, including neuronal development, synaptic plasticity, and neuronal survival. As reported, genetic variants within the KLF family have been associated with [...] Read more.
The Krüppel-like factor (KLF) family represents a group of transcription factors (TFs) performing different biological processes that are crucial for proper neuronal function, including neuronal development, synaptic plasticity, and neuronal survival. As reported, genetic variants within the KLF family have been associated with a wide spectrum of neurodevelopmental and psychiatric symptoms. In a patient exhibiting attention deficit hyperactivity disorder (ADHD) combined with both neurodevelopmental and psychiatric symptoms, whole-exome sequencing (WES) analysis revealed a de novo heterozygous variant within the Krüppel-like factor 13 (KLF13) gene, which belongs to the KLF family and regulates axonal growth, development, and regeneration in mice. Moreover, in silico analyses pertaining to the likely pathogenic significance of the variant and the impact of the mutation on the KLF13 protein structure suggested a potential deleterious effect. In fact, the variant was localized in correspondence to the starting residue of the N-terminal domain of KLF13, essential for protein–protein interactions, DNA binding, and transcriptional activation or repression. This study aims to highlight the potential involvement of the KLF13 gene in neurodevelopmental and psychiatric disorders. Nevertheless, we cannot rule out that excluded variants, those undetectable by WES, or the polygenic risk may have contributed to the patient’s phenotype given ADHD’s high polygenic risk. However, further functional studies are required to validate its potential contribution to these disorders. Full article
(This article belongs to the Section Human Genomics and Genetic Diseases)
Show Figures

Figure 1

Figure 1
<p>Detection of c.20T&gt;G within <span class="html-italic">KLF13</span> gene. (<b>a</b>) Depiction of the nucleotide sequence corresponding to the region where the mutation was identified within the <span class="html-italic">KLF13</span> gene. Furthermore, the chromosomal localization of this gene is illustrated. Figure was modified from the UCSC genome database. The asterisk indicates the precise variant site. (<b>b</b>) Whole-exome sequencing (WES) results are presented using the Integrative Genomics Viewer (IGV) visualization tool. As shown in the picture, WES was carried out for the examined patient and both healthy parents. (<b>c</b>) Conventional Sanger sequencing was performed to highlight the c.20T&gt;G variant identified by WES. In the electropherograms, the black, blue, green, and red profiles indicate nucleotides G, C, A, and T.</p>
Full article ">Figure 2
<p>Structure prediction analysis and functional domains related to KLF13 protein. (<b>a</b>) Protein structure prediction related to the wild-type KLF13. Each functional domain is marked by different colors. (<b>b</b>) Focus on the wild-type valine residue at position 7, which did not engage in hydrogen bonds with other amino acids. (<b>c</b>) Mutated KLF13 protein. As predicted, the different structural protein folding as result of the mutation is evident. (<b>d</b>) Close-up of the mutated residue as a result of the missense mutation p.Val7Gly. (<b>e</b>) Domain organization patterns related to the KLF13 protein. The specific mutation site is indicated by the black arrow. The light blue asterisk in (<b>b</b>,<b>d</b>,<b>e</b>) indicates the precise position of the missense mutation. (<b>a</b>–<b>d</b>) were generated by UCSF ChimeraX software, while (<b>e</b>) was modified from Uniprot database.</p>
Full article ">Figure 3
<p>Structure prediction analysis of the KLF13 protein, focusing on the amino acid residues from positions 3 to 19, revealed significant structural variation from the primary structure to an α helix. Notably, the N-terminal domain of Krüppel-like factor 13 (from residues 7 to 168) begins at the specific mutation site at amino acid 7. The colors used are consistent with the domain organization patterns shown in <a href="#genes-15-01056-f002" class="html-fig">Figure 2</a>. (<b>a</b>) Wild-type KLF13 protein. (<b>b</b>) Close-up of the wild-type protein segment from alanine 3 to serine 19. (<b>c</b>) Mutated KLF13 protein, with visibly different predicted protein folding compared to the wild type. (<b>d</b>) Close-up of the segment from alanine 3 to serine 19, highlighting the mutated glycine 7 (marked with an asterisk). The missense mutation is predicted to result in the formation of an α helix containing 17 hydrogen bonds. (<b>a</b>–<b>d</b>) were generated by UCSF ChimeraX software and subsequently modified.</p>
Full article ">Figure 4
<p>Line plots generated with PONDR tool with VLXT score to assess the impact of the mutation on protein stability and flexibility. (<b>a</b>) The VLXT score (blue line) from PONDR analysis for the wild-type KLF13 protein indicates a high rate of structural order at the specific site (green line indicating valine at position 7), with scores lower than 0.5 (orange line). (<b>b</b>) The VLXT score from PONDR analysis for the mutated KLF13 protein shows a higher rate of disorder as a result of the mutation. This is evidenced by the values of the residues before the mutation site being higher than 0.5 (orange line).</p>
Full article ">
18 pages, 1844 KiB  
Article
The Use of Compost and Arbuscular Mycorrhizal Fungi and Their Combination to Improve Tomato Tolerance to Salt Stress
by Fadoua Mekkaoui, Mohamed Ait-El-Mokhtar, Nada Zaari Jabri, Ilham Amghar, Soukaina Essadssi and Abdelaziz Hmyene
Plants 2024, 13(16), 2225; https://doi.org/10.3390/plants13162225 (registering DOI) - 11 Aug 2024
Abstract
Salinity poses a significant challenge to tomato plant development and metabolism. This study explores the use of biostimulants as eco-friendly strategies to enhance tomato plant tolerance to salinity. Conducted in a greenhouse, the research focuses on the Solanum lycopersicum L. behavior under saline [...] Read more.
Salinity poses a significant challenge to tomato plant development and metabolism. This study explores the use of biostimulants as eco-friendly strategies to enhance tomato plant tolerance to salinity. Conducted in a greenhouse, the research focuses on the Solanum lycopersicum L. behavior under saline conditions. Tomato seeds were treated with arbuscular mycorrhizal fungi (AMF), compost, and their combination under both non-saline and saline conditions (0 and 150 mM NaCl). Plant height, number of flowers and fruits, shoot fresh weight, and root dry weight were negatively impacted by salt stress. The supplementation with compost affected the colonization of AMF, but the application of stress had no effect on this trait. However, the use of compost and AMF separately or in combination showed positive effects on the measured parameters. At the physiological level, compost played a beneficial role in increasing photosynthetic efficiency, whether or not plants were subjected to salinity. In addition, the application of these biostimulants led to an increase in nitrogen content in the plants, irrespective of the stress conditions. AMF and compost, applied alone or in combination, showed positive effects on photosynthetic pigment concentrations and protein content. Under salt stress, characterized by an increase in lipid peroxidation and H2O2 content, the application of these biostimulants succeeded in reducing both these parameters in affected plants through exhibiting an increase in antioxidant enzyme activity. In conclusion, incorporating compost, AMF, or their combined application emerges as a promising approach to alleviate the detrimental impacts of salt stress on both plant performances. These findings indicate optimistic possibilities for advancing sustainable and resilient agricultural practices. Full article
(This article belongs to the Special Issue Advances in Soil Fertility Management for Sustainable Crop Production)
Show Figures

Figure 1

Figure 1
<p>Photosynthetic pigment ((<b>a</b>) chlorophyll <span class="html-italic">a</span>, (<b>b</b>) chlorophyll <span class="html-italic">b</span>, (<b>c</b>) total chlorophyll, (<b>d</b>) carotenoids) content of tomato plants under saline and non-saline conditions after the application of compost and AMF alone or in combination. Ct: control; C: compost; M: arbuscular mycorrhizal fungi; CM: combination of compost and arbuscular mycorrhizal fungi. Data are means of three replicates (n = 3) ± standard error (SE). The bars of each parameter labeled by different letters indicate significant differences assessed by Duncan’s test after performing a three-way ANOVA (<span class="html-italic">p</span> &lt; 0.05).</p>
Full article ">Figure 2
<p>Nitrogen content of tomato leaves (<b>a</b>) and roots (<b>b</b>) under saline and non-saline conditions after the application of compost and AMF alone or in combination. Ct: control; C: compost; M: arbuscular mycorrhizal fungi; CM: combination of compost and arbuscular mycorrhizal fungi. Data are means of three replicates (n = 3) ± standard error (SE). The bars of each parameter labeled by different letters indicate significant differences assessed by Duncan’s test after performing a three-way ANOVA (<span class="html-italic">p</span> &lt; 0.05).</p>
Full article ">Figure 3
<p>Hydrogen peroxide (<b>a</b>,<b>b</b>) and malondialdehyde (<b>c</b>,<b>d</b>) contents in the shoot and the root of tomato plants under saline and non-saline conditions after the application of compost and AMF alone or in combination. Ct: control; C: compost; M: arbuscular mycorrhizal fungi; CM: combination of compost and arbuscular mycorrhizal fungi. Data are means of three replicates (n = 3) ± standard error (SE). The bars of each parameter labeled by different letters indicate significant differences assessed by Duncan’s test after performing three-way ANOVA (<span class="html-italic">p</span> &lt; 0.05).</p>
Full article ">Figure 4
<p>Protein (<b>a</b>,<b>b</b>) and soluble sugar (<b>c</b>,<b>d</b>) contents in the shoots and the roots of tomato plants under saline and non-saline conditions after the application of compost and AMF alone or in combination. Ct: control; C: compost; M: arbuscular mycorrhizal fungi; CM: combination of compost and arbuscular mycorrhizal fungi. Data are means of three replicates (n = 3) ± standard error (SE). The bars of each parameter labeled by different letters indicate significant differences assessed by Duncan’s test after performing three-way ANOVA (<span class="html-italic">p</span> &lt; 0.05).</p>
Full article ">Figure 5
<p>Principal component analysis of the studied traits of different biostimulant treatments in tomato plants growing under saline and non-saline conditions. S: salinity; Ct: control; C: compost; M: arbuscular mycorrhizal fungi; CM: combination of compost and arbuscular mycorrhizal fungi; Ct-S: control+stress conditions; C-S: compost+stress conditions; M-S: arbuscular mycorrhizal fungi+stress conditions, CM-S: combination of compost and arbuscular mycorrhizal fungi+stress conditions; PH: plant height; NL: number of leaves; SDW: shoot dry weight; RDW: root dry weight; Fa: AMF infection frequency; Ma: AMF infection intensity; gs: stomatal conductance; Fv/Fm: photosynthetic efficiency; Chl <span class="html-italic">a</span>: chlorophyll <span class="html-italic">a</span>; Chl <span class="html-italic">b</span>: chlorophyll <span class="html-italic">b</span>; Chl T: total chlorophyll; Carot: carotenoids; H<sub>2</sub>O<sub>2</sub>-L: leaves H<sub>2</sub>O<sub>2</sub> content; H<sub>2</sub>O<sub>2</sub>-R: root H<sub>2</sub>O<sub>2</sub> content; MDA-L: leaves MDA content; MDA-R: root MDA content; Sugar-L: leaf sugar content; Sugar-R: root sugar content; APX-L: leaf ascorbate peroxidase activity; APX-R: root ascorbate peroxidase activity; CAT-L: leaf catalase activity; CAT-R: root catalase activity; POX-L: leaf peroxidase activity; POX-R: root peroxidase activity; PPO-L: leaf polyphenol oxidase activity; PPO-R: root polyphenol oxidase activity; Prot-L: leaf protein content; Prot-R: root protein content.</p>
Full article ">Figure 6
<p>Schematic representation of various mechanisms induced by AMF and compost application in tomato plants under salt stress.</p>
Full article ">
15 pages, 2945 KiB  
Article
Morphology, Glycan Pattern, Heat Shock Proteins, and Sex Steroid Receptors Expression in the Tubal Fimbria Epithelium of the Baboon Papio hamadryas during the Menstrual Cycle
by Salvatore Desantis, Mario Cinone, Luca Lacitignola, Pietro Laricchiuta, Roberta Rossi, Antonio Ciro Guaricci, Leonardo Resta and Maria Albrizio
Animals 2024, 14(16), 2321; https://doi.org/10.3390/ani14162321 (registering DOI) - 11 Aug 2024
Abstract
The oviductal fimbria is the first extraovarian anatomical structure that the cumulus–oocyte complex (COC) encounters, and is sensitive to sex hormone changes. The morphology, glycan pattern, expression of heat shock proteins (HSPs), estradiol receptor (ER), and progesterone receptor (PR) were investigated in the [...] Read more.
The oviductal fimbria is the first extraovarian anatomical structure that the cumulus–oocyte complex (COC) encounters, and is sensitive to sex hormone changes. The morphology, glycan pattern, expression of heat shock proteins (HSPs), estradiol receptor (ER), and progesterone receptor (PR) were investigated in the oviductal fimbria epithelium of the baboon (Papio hamadryas) during the menstrual cycle. The morphology was investigated by light and scanning electron microscopy; the glycopattern was characterized using conventional and lectin histochemistry; HSPs (60, −70, −90), ER, and PR were localized immunohistochemically. Well-differentiated ciliated and nonciliated cells were present only during the preovulatory phase. The nonciliated cells contained small apical protrusions and thin microvilli. During the preovulatory phase (1) the luminal surface of the fimbria displayed acidic glycans, complex N-glycans containing fucose, and oligolactosamine residues; (2) nonciliated cells expressed HSP60 and HSP90 in the apical blebs, HSP70 in the nucleus and cytoplasm, as well as nuclear ERα and PR; (3) ciliated cells showed HSP70 in the nucleus, cytoplasm, and cilia that also expressed HSP90 and PR. These results are related to the function of the fimbria where the early COC–oviduct crosstalk occurs and may represent a benchmark for translational studies of other primates. Full article
(This article belongs to the Section Animal Reproduction)
Show Figures

Figure 1

Figure 1
<p>Macroscopic (<b>A</b>) and histological view (<b>B</b>) of baboon <span class="html-italic">Papio hamadryas</span> oviductal fimbriae stained with Hematoxylin-Eosin. In B, note the opening of the infundibulum (oi). oi, ostium of the infundibulum; mi, muscular of the infundibulum; arrow, mucosal folds of the fimbriae. Scale bar: (<b>A</b>) 2 mm; (<b>B</b>) 500 µm.</p>
Full article ">Figure 2
<p>Light micrographs showing the morphological changes in the epithelium of baboon <span class="html-italic">Papio hamadryas</span> oviductal fimbriae during the menstrual cycle. Hematoxylin-Eosin staining. Scale bar: 10 µm.</p>
Full article ">Figure 3
<p>Scanning electron micrographs of the epithelium of baboon <span class="html-italic">Papio hamadryas</span> oviductal fimbriae during the menstrual cycle. ci, cilia. Scale bar: 8 µm.</p>
Full article ">Figure 4
<p>Conventional histochemical staining of fimbria epithelium of baboon <span class="html-italic">Papio hamadryas</span> oviduct during the menstrual cycle. Note the absence of PAS staining (<b>A</b>–<b>C</b>) and the presence of AB 2.5 staining (azur staining) during the follicular and preovulatory phase (<b>D</b>,<b>E</b>). HID staining (brown staining) was detected only during the preovulatory phase (<b>E</b>). In (<b>A</b>–<b>C</b>), nuclei were stained with Hematoxylin. In (<b>D</b>–<b>F</b>), nuclei were stained with nuclear fast red. lp, lamina propria; arrow, luminal surface of the epithelium; asterisk, blood vessel. Scale bar: 20 µm.</p>
Full article ">Figure 5
<p>AAL, GNL, LCA, and RCA<sub>120</sub> binding sites in the mucosal epithelium of baboon <span class="html-italic">Papio hamadryas</span> oviductal fimbriae during the menstrual cycle. (<b>A</b>–<b>C</b>), expression of aL-Fuc terminating glycans revealed with LTA. (<b>D</b>–<b>F</b>), presence of terminal α1-3mannose residue detected by GNL. (<b>G</b>–<b>I</b>), localization of complex N-linked glycans by means of LCA. (<b>J</b>–<b>L</b>), lactosamine terminating glycans identified by RCA<sub>120</sub>. Note that the investigated glycans were mainly expressed during the preovulatory phase compared to other menstrual cycle phases. bv, blood vessel; e, epithelium; lp, lamina propria; arrow, luminal surface of the epithelium. Scale bar: 40 µm. AAL, GNL, and LCA were FITC-conjugated lectins. RCA was TRITC-conjugated lectin.</p>
Full article ">Figure 6
<p>Immunostaining pattern of the HSP60 (<b>A</b>–<b>C</b>), 70 (<b>D</b>–<b>F</b>), 90 (<b>G</b>–<b>I</b>) in the mucosal epithelium of baboon <span class="html-italic">Papio hamadryas</span> oviductal fimbriae during the menstrual cycle. The inset images display the absence of immunoreactivity in negative controls. cc, ciliated cell; e, epithelium; lp, lamina propria; nc, nonciliated cell; arrowhead, cilia; *, apical bleb. Scale bar: 20 µm; insets, 10 µm.</p>
Full article ">Figure 7
<p>ERα and PR immunostaining pattern in the baboon <span class="html-italic">Papio hamadryas</span> oviductal fimbriae during the menstrual cycle. (<b>A</b>–<b>C</b>), ERα was present only in the nuclei. (<b>D</b>–<b>F</b>), PR immunoreactivity was observed unevenly in the cytoplasm and nucleus of epithelial cells during the follicular and luteal phases, whereas it was present in the nuclei of the nonciliated cells and in the cilia during the preovulatory phase. The inset images display the absence of immunoreactivity in negative controls. cc, ciliated cell; e, epithelium; lp, lamina propria; nc, nonciliated cell; arrowhead, cilia; asterisk, negative nucleus. Scale bar: 10 µm.</p>
Full article ">
10 pages, 5611 KiB  
Communication
A Newly Developed Anti-L1CAM Monoclonal Antibody Targets Small Cell Lung Carcinoma Cells
by Miki Yamaguchi, Sachie Hirai, Masashi Idogawa, Toshiyuki Sumi, Hiroaki Uchida and Yuji Sakuma
Int. J. Mol. Sci. 2024, 25(16), 8748; https://doi.org/10.3390/ijms25168748 (registering DOI) - 11 Aug 2024
Viewed by 86
Abstract
Few effective treatments are available for small cell lung cancer (SCLC), indicating the need to explore new therapeutic options. Here, we focus on an antibody–drug conjugate (ADC) targeting the L1 cell adhesion molecule (L1CAM). Several publicly available databases reveal that (1) L1CAM is [...] Read more.
Few effective treatments are available for small cell lung cancer (SCLC), indicating the need to explore new therapeutic options. Here, we focus on an antibody–drug conjugate (ADC) targeting the L1 cell adhesion molecule (L1CAM). Several publicly available databases reveal that (1) L1CAM is expressed at higher levels in SCLC cell lines and tissues than in those of lung adenocarcinoma and (2) the expression levels of L1CAM are slightly higher in SCLC tissues than in adjacent normal tissues. We conducted a series of in vitro experiments using an anti-L1CAM monoclonal antibody (termed HSL175, developed in-house) and the recombinant protein DT3C, which consists of diphtheria toxin lacking the receptor-binding domain but containing the C1, C2, and C3 domains of streptococcal protein G. Our HSL175-DT3C conjugates theoretically kill cells only when the conjugates are internalized by the target (L1CAM-positive) cells through antigen–antibody interaction. The conjugates (an ADC analog) were effective against two SCLC-N (NEUROD1 dominant) cell lines, Lu-135 and STC-1, resulting in decreased viability. In addition, L1CAM silencing rendered the two cell lines resistant to HSL175-DT3C conjugates. These findings suggest that an ADC consisting of a humanized monoclonal antibody based on HSL175 and a potent anticancer drug would be effective against SCLC-N cells. Full article
(This article belongs to the Section Molecular Pathology, Diagnostics, and Therapeutics)
Show Figures

Figure 1

Figure 1
<p>Small cell lung carcinoma (SCLC) cell lines and tissues express <span class="html-italic">L1CAM</span> mRNA. (<b>A</b>) <span class="html-italic">L1CAM</span> and <span class="html-italic">SYP</span> mRNA expression in SCLC and lung adenocarcinoma (LUAD) cell lines. The RNA-seq data of 29 SCLC and 36 LUAD cell lines are presented. (<b>B</b>) Correlation between <span class="html-italic">L1CAM</span> and <span class="html-italic">SYP</span> mRNA expression in SCLC cell lines (<span class="html-italic">n</span> = 52). Pearson’s correlation coefficient (R) = 0.541. (<b>C</b>) <span class="html-italic">L1CAM</span> mRNA expression in normal lung, SCLC, LUAD, and LUSC tissues. Expression data from 7 normal lung and 79 SCLC tissues were obtained from a previous report [<a href="#B13-ijms-25-08748" class="html-bibr">13</a>]. Expression data from 534 LUAD and 502 LUSC samples were derived from the Cancer Genome Atlas database [<a href="#B14-ijms-25-08748" class="html-bibr">14</a>,<a href="#B15-ijms-25-08748" class="html-bibr">15</a>]. *** <span class="html-italic">p</span> &lt; 0.001.</p>
Full article ">Figure 2
<p>Two SCLC cell lines (Lu-135 and STC-1) express L1CAM. (<b>A</b>) Phase contrast images of Lu-135 and STC-1 cells with L1CAM silencing. Cells were reverse-transfected with NC siRNA, L1CAM siRNA #1, or L1CAM siRNA #2 (10 nM each) and cultured for 48 h. Scale bars: 200 μm (left; low magnification) and 50 μm (right; high magnification). (<b>B</b>) Conventional RT-PCR for the expression of <span class="html-italic">L1CAM</span>, <span class="html-italic">SYP</span>, <span class="html-italic">NEUROD1</span>, and <span class="html-italic">ACTB</span> mRNA in Lu-135 and STC-1 cells. Cells were treated as described in (<b>A</b>). (<b>C</b>) Western blot analysis of the cells treated as described in (<b>A</b>). Of note, NCI-H69 was used as a positive control for ASCL1. (<b>D</b>) Western blot analysis of HuL cells and SCLC-N cells for L1CAM. (<b>E</b>) Correlation between <span class="html-italic">L1CAM</span> and <span class="html-italic">NEUROD1</span> or <span class="html-italic">ASCL1</span> mRNA expression in SCLC cell lines (<span class="html-italic">n</span> = 52). (<b>F</b>) Regulated Gene Ontology results for the top 100 protein-coding DEGs in L1CAM-silenced Lu-135 cells compared with control cells.</p>
Full article ">Figure 3
<p>Lu-135 cells are highly sensitive to HSL175-DT3C conjugates. (<b>A</b>) Effects of HSL175-DT3C conjugates on the viability of Lu-135 cells. Cells were transfected with NC siRNA, L1CAM siRNA #1, or L1CAM siRNA #2 (10 nM each), cultured for 48 h, and then incubated with HSL175-DT3C conjugates (0–10 μg/mL each) for another 72 h. Results are presented as mean ± SD. *** <span class="html-italic">p</span> &lt; 0.001. (<b>B</b>) Phase contrast images of Lu-135 cells. Cells were transfected with NC siRNA, L1CAM siRNA #1, or L1CAM siRNA #2 (10 nM each), cultured for 48 h, and then incubated with HSL175-DT3C conjugates (1 μg/mL each) for another 72 h. Scale bars: 200 μm (left; low magnification) and 50 μm (right; high magnification). (<b>C</b>) Effects of HSL175-DT3C conjugates on the viability of Lu-135 cells. Cells were cultured with control IgG-DT3C conjugates or HSL175-DT3C conjugates (0.1 μg/mL each) for 96 h. Results are presented as mean ± SD. *** <span class="html-italic">p</span> &lt; 0.001. (<b>D</b>) Effects of HSL175-DT3C conjugates on levels of the apoptosis marker cleaved PARP in Lu-135 cells. Cells were cultured with control IgG-DT3C conjugates or HSL175-DT3C conjugates (1 μg/mL each) for 72 h.</p>
Full article ">Figure 4
<p>HSL175-DT3C conjugates are also effective against STC-1 cells. (<b>A</b>) Effects of HSL175-DT3C conjugates on the viability of STC-1 cells. Cells were transfected with NC siRNA, L1CAM siRNA #1, or L1CAM siRNA #2 (10 nM each), cultured for 48 h, and then incubated with HSL175-DT3C conjugates (0–10 μg/mL each) for another 72 h. Results are presented as mean ± SD. *** <span class="html-italic">p</span> &lt; 0.001. (<b>B</b>) Phase contrast images of STC-1 cells. Cells were transfected with NC siRNA, L1CAM siRNA #1, or L1CAM siRNA #2 (10 nM each), cultured for 48 h, and then incubated with HSL175-DT3C conjugates (1 μg/mL each) for another 48 h. Scale bars: 200 μm (left; low magnification) and 50 μm (right; high magnification). (<b>C</b>) Effects of HSL175-DT3C conjugates on the viability of STC-1 cells. Cells were cultured with control IgG-DT3C conjugates or HSL175-DT3C conjugates (1 μg/mL each) for 96 h. Results are presented as mean ± SD. *** <span class="html-italic">p</span> &lt; 0.001. (<b>D</b>) Effects of HSL175-DT3C conjugates on levels of the apoptosis marker cleaved PARP in STC-1 cells. Cells were cultured with control IgG-DT3C conjugates or HSL175-DT3C conjugates (1 μg/mL each) for 48 h.</p>
Full article ">
15 pages, 2851 KiB  
Article
Characterization and Phylogenetic Analysis of the Complete Mitochondrial Genome of Triplophysa microphthalma
by Ping Yang, Wei Guo, Chao Wei, Xin Wang, Yixuan Wang and Jia Wang
Biology 2024, 13(8), 608; https://doi.org/10.3390/biology13080608 (registering DOI) - 11 Aug 2024
Viewed by 97
Abstract
The complete mitochondrial genome has been extensively utilized in studies related to phylogenetics, offering valuable perspectives on evolutionary relationships. The mitochondrial genome of the fine-eyed plateau loach, Triplophysa microphthalma, has not attracted much attention, although this species is endemic to China. In [...] Read more.
The complete mitochondrial genome has been extensively utilized in studies related to phylogenetics, offering valuable perspectives on evolutionary relationships. The mitochondrial genome of the fine-eyed plateau loach, Triplophysa microphthalma, has not attracted much attention, although this species is endemic to China. In this study, we characterized the mitochondrial genome of T. microphthalma and reassessed the classification status of its genus. The complete mitochondrial genome of T. microphthalma was 16,591 bp and contained thirty-seven genes, including thirteen protein-coding genes (PCGs), two ribosomal RNA genes (rRNAs), and twenty-two transfer RNA genes (tRNAs). All but one of the thirteen PCGs had the regular start codon ATG; the gene cox1 started with GTG. Six PCGs had incomplete stop codons (T--). These thirteen PCGs are thought to have evolved under purifying selection, and the mitogenome shared a high degree of similarity with the genomes of species within the genus Leptobotia. All tRNA genes exhibited the standard clover-shaped structure, with the exception of the trnS1 gene, which lacked a DHU stem. A phylogenetic analysis indicated that T. microphthalma was more closely related to species within the genus Triplophysa than to those in Barbatula. The present study contributes valuable genomic information for T. microphthalma, and offers new perspectives on the phylogenetic relationships among species of Triplophysa and Barbatula. The findings also provide essential data that can inform the management and conservation strategies for T. microphthalma and other species of Triplophysa and Barbatula. Full article
(This article belongs to the Special Issue Internal Defense System and Evolution of Aquatic Animals)
Show Figures

Figure 1

Figure 1
<p>Mapping of the mitochondrial genome of <span class="html-italic">Triplophysa microphthalma</span>.</p>
Full article ">Figure 2
<p>The predicted secondary structures of the 22 tRNAs within the mitochondrial genome of <span class="html-italic">Triplophysa microphthalma</span>. Each tRNA gene is identified by its standard abbreviations in the upper left corner, alongside its corresponding amino acid in parentheses for translation purposes.</p>
Full article ">Figure 3
<p>The relative synonymous codon usage (RSCU) values for PCGs in the complete mitochondrial genome of <span class="html-italic">Triplophysa microphthalma</span>. Distinct colors signify various codon families that correspond to amino acids.</p>
Full article ">Figure 4
<p>A maximum likelihood (ML) phylogenetic tree, bolstered by bootstrap support values at the nodes, was constructed using nucleotide sequences from 43 Cyprinidae species, encompassing 13 PCGs, 22 tRNAs, and 2 rRNAs.</p>
Full article ">Figure 5
<p>Co-linear scatter plot. The horizontal and vertical axes represent two different genomes, and the connected portion of the line represents the region of covariance conserved between the two genomes. The more the lines are connected, the more similarities exist between the two genomes. (<b>a</b>) Co-linear scatter plot of <span class="html-italic">Triplophysa microphthalma</span> and Barbatula barbatula. (<b>b</b>) Co-linear scatter plot of <span class="html-italic">Triplophysa microphthalma</span> and <span class="html-italic">Triplophysa ulacholica</span>. (<b>c</b>) Co-linear scatter plot of <span class="html-italic">Triplophysa microphthalma</span> and Triplophysa strauchii. (<b>d</b>) Co-linear scatter plot of <span class="html-italic">Triplophysa microphthalma</span> and Triplophysa dorsalis.</p>
Full article ">Figure 5 Cont.
<p>Co-linear scatter plot. The horizontal and vertical axes represent two different genomes, and the connected portion of the line represents the region of covariance conserved between the two genomes. The more the lines are connected, the more similarities exist between the two genomes. (<b>a</b>) Co-linear scatter plot of <span class="html-italic">Triplophysa microphthalma</span> and Barbatula barbatula. (<b>b</b>) Co-linear scatter plot of <span class="html-italic">Triplophysa microphthalma</span> and <span class="html-italic">Triplophysa ulacholica</span>. (<b>c</b>) Co-linear scatter plot of <span class="html-italic">Triplophysa microphthalma</span> and Triplophysa strauchii. (<b>d</b>) Co-linear scatter plot of <span class="html-italic">Triplophysa microphthalma</span> and Triplophysa dorsalis.</p>
Full article ">
15 pages, 1723 KiB  
Article
Human Serum Albumin Protein Corona in Prussian Blue Nanoparticles
by Chiara Colombi, Giacomo Dacarro, Yuri Antonio Diaz Fernandez, Angelo Taglietti, Piersandro Pallavicini and Lavinia Doveri
Nanomaterials 2024, 14(16), 1336; https://doi.org/10.3390/nano14161336 (registering DOI) - 11 Aug 2024
Viewed by 106
Abstract
Prussian Blue nanoparticles (PBnps) are now popular in nanomedicine thanks to the FDA approval of PB. Despite the numerous papers suggesting or describing the in vivo use of PBnps, no studies have been carried out on the formation of a protein corona on [...] Read more.
Prussian Blue nanoparticles (PBnps) are now popular in nanomedicine thanks to the FDA approval of PB. Despite the numerous papers suggesting or describing the in vivo use of PBnps, no studies have been carried out on the formation of a protein corona on the PBnp surface and its stabilizing role. In this paper, we studied qualitatively and quantitatively the corona formed by the most abundant protein of blood, human serum albumin (HSA). Cubic PBnps (41 nm side), prepared in citric acid solution at PB concentration 5 × 10−4 M, readily form a protein corona by redissolving ultracentrifuged PBnp pellets in HSA solutions, with CHSA ranging from 0.025 to 7.0 mg/mL. The basic decomposition of PBnp@HSA was studied in phosphate buffer at the physiological pH value of 7.4. Increased stability with respect to uncoated PBnps was observed at all concentrations, but a minimum CHSA value of 3.0 mg/mL was determined to obtain stability identical to that observed at serum-like HSA concentrations (35–50 mg/mL). Using a modified Lowry protocol, the quantity of firmly bound HSA in the protein corona (hard corona) was determined for all the CHSA used in the PBnp@HSA synthesis, finding increasing quantities with increasing CHSA. In particular, an HSA/PBnp number in the 1500–2300 range was found for CHSA 3.0–7.0 mg/mL, largely exceeding the 180 HSA/PBnp value calculated for an HSA monolayer on a PBnp. Finally, the stabilization brought by the HSA corona allowed us to carry out pH-spectrophotometric titrations on PBnp@HSA in the 3.5-9-0 pH range, revealing a pKa value of 6.68 for the water molecules bound to the Fe3+ centers on the PBnp surface, whose deprotonation is responsible for the blue-shift of the PBnp band from 706 nm (acidic solution) to 685 nm (basic solution). Full article
(This article belongs to the Section Inorganic Materials and Metal-Organic Frameworks)
Show Figures

Figure 1

Figure 1
<p>(<b>A</b>) Absorption spectrum of citrate-coated PBnps (black) and of PBnp@HSA obtained with C<sub>HSA</sub> 1 mg/mL (orange); (<b>B</b>) PBnp and albumin (modelized as a rectangular object) sketched maintaining the authentic proportions between their dimensions; (<b>C</b>) TEM image of citrate PBnps; (<b>D</b>) same, coated with C<sub>HSA</sub> 1 mg/mL.</p>
Full article ">Figure 2
<p>(<b>A</b>) Blue spectrum: 2.0 mL PBnp@HSA solution (prepared with C<sub>HSA</sub> 1 mg/mL) diluted with 1 mL water, pH 5.8; red spectrum, 2.0 mL of the same PBnp@HSA solution, diluted with 1 mL phosphate buffer at pH 7.4; (<b>B</b>) absorption spectra recorded during the titration with microadditions of 0.05 M NaOH of an acidified PBnp@HSA solution (C<sub>HSA</sub> 1 mg/mL); the first spectrum (red, pH 3.51) and the final one (blue, pH 9.03) has been colored to stress the trend of the titration; the spectrum obtained at pH 4.81, affected by scattering due to turbidity, has been colored in green, with dashed line. (<b>C</b>) Plots of zeta-potential (black circles) and Abs<sub>780</sub> (blue triangles) vs. pH; the dashed lines are plots of the sigmoidal fittings of the data (R<sup>2</sup> = 0.989 and 0.993, respectively), with calculated inflection points at pH 5.18 (zeta-potential) and 6.68 (Abs<sub>780</sub>). The red triangle data (Abs<sub>780</sub> at pH 4.81) has been excluded from data fitting.</p>
Full article ">Figure 3
<p>(<b>A</b>) Series of spectra recorded at 2 h intervals for citrate-coated PBnps at pH 7.4 (phosphate buffer); the spectrum at t = 0 and 24 h are in blue and pink color, respectively, while all the spectra at intermediate times are in black; (<b>B</b>) same, for PBnp@HSA prepared with C<sub>HSA</sub> = 5 mg/mL; (<b>C</b>) percentage of residual absorbance after 2, 6, 12, and 24 h for untreated (citrate-coated) PBnps and for PBnp@HSA prepared with all C<sub>HSA</sub> concentrations; different symbols correspond to the classification groups within the same analyzed time (i.e., 2 h, 6 h, 12 h, or 24 h) for the ANOVA Tukey’s test (<span class="html-italic">p</span> &lt; 0.05).</p>
Full article ">Figure 4
<p>(<b>A</b>) Calibration points (black circles) and fitting curve (dashed curve) using the Lowry method with different concentrations of HSA and PBnp solutions as the background; the red circles are points obtained using bidistilled water as the background; (<b>B</b>) (<b>i</b>) % HSA bound to PBnp (p-C<sub>HSA</sub>) vs. total HSA added in the synthesis (C<sub>HSA</sub>); (<b>ii</b>) concentration (μg/mL) of HSA bound to the pellet vs. total HSA added in the synthesis (C<sub>HSA</sub>); data in panel (<b>i</b>,<b>ii</b>) are significantly different at the ANOVA Tukey’s test (<span class="html-italic">p</span> &lt; 0.05) except the couples marked with NS.</p>
Full article ">
Back to TopTop