[go: up one dir, main page]
More Web Proxy on the site http://driver.im/
You seem to have javascript disabled. Please note that many of the page functionalities won't work as expected without javascript enabled.
 
 
Sign in to use this feature.

Years

Between: -

Subjects

remove_circle_outline
remove_circle_outline
remove_circle_outline
remove_circle_outline
remove_circle_outline
remove_circle_outline
remove_circle_outline
remove_circle_outline
remove_circle_outline

Journals

Article Types

Countries / Regions

Search Results (139)

Search Parameters:
Keywords = aspartic proteases

Order results
Result details
Results per page
Select all
Export citation of selected articles as:
14 pages, 952 KiB  
Article
Effects of Bacillus subtilis KG109 on Growth Performance, Carcass Quality, Serum Indicators, Intestinal Morphology, and Digestive Enzymes in Broilers
by Hong Chen, Weixin Liu, Hao Zhang, Yibo Yan, Meiqi Chen, Xiaoling Ding, Cheng Zhang, Runsheng Jiang and Zaigui Wang
Animals 2024, 14(24), 3650; https://doi.org/10.3390/ani14243650 - 18 Dec 2024
Viewed by 354
Abstract
The purpose of this experiment is to investigate how different doses of Bacillus subtilis KG109 powder affect the growth performance, carcass quality, serum biochemical indexes, serum antioxidant and immunological index, intestinal morphology, and digestive enzyme activity of broilers. Four hundred chicks of a [...] Read more.
The purpose of this experiment is to investigate how different doses of Bacillus subtilis KG109 powder affect the growth performance, carcass quality, serum biochemical indexes, serum antioxidant and immunological index, intestinal morphology, and digestive enzyme activity of broilers. Four hundred chicks of a similar weight (1 day old) are randomly assigned to four groups of five replicates of 20 chicks each (half males and half females). The control group is fed a basal ration, and the experimental groups T1, T2, and T3 are supplemented with 6.0 × 108 CFU/kg, 1.2 × 109 CFU/kg, and 1.8 × 109 CFU/kg of Bacillus subtilis KG109 bacterial powder, respectively, in the basal ration. The feeding cycle is 52 d. Compared with the control group, Bacillus subtilis KG109 powder (1) increases the broiler feed conversion ratio (FCR) (p < 0.05), (2) improves the carcass quality (slaughter rate, cooking loss, L* and b* values) (p < 0.05), (3) enhances the serum biochemical indexes (alanine transaminase (ALT), alkaline phosphatase (ALP), aspartate transaminase (AST), albumin (ALB), and triglycerides (TG)) (p < 0.05), (4) improves the serum antioxidant capacity (total an-tioxidant capacity (T-AOC), total superoxide dismutase (T-SOD), and glutathione peroxidase (GSH-PX)) and immunoglobulins (lg A, lg G, lg M) (p < 0.05), (5) improves the intestinal morphology (villus height and villus height to crypt depth (VCR)) (p < 0.05), and (6) increases the intestinal digestive enzyme activities (amylase, protease, and lipase) (p < 0.05). In summary, adding Bacillus subtilis KG109 to broiler diets can result in a significant decrease in broilers’ FCR, an increase in their slaughtering rate, a decrease in their serum ALT, ALP, and AST activities, an increase in their serum TG content, an improvement of their immune and antioxidant capacity, an improvement of their intestinal morphology, and an improvement of their intestinal digestive enzyme activity. It is recommended to add 1.8 × 109 CFU/kg of bacteria. Full article
(This article belongs to the Section Poultry)
Show Figures

Figure 1

Figure 1
<p>Slice atlases of duodenum and jejunum (40×). (<b>A1</b>–<b>D1</b>) represent the duodenum in the control, T1, T2, and T3 groups, respectively; (<b>A2</b>–<b>D2</b>) represent the jejunum in the control, T1, T2, T3 groups, respectively.</p>
Full article ">
16 pages, 6782 KiB  
Article
Functional Characterization of FgAsp, a Gene Coding an Aspartic Acid Protease in Fusarium graminearum
by Ping Li, Zhizhen Fu, Mengru Wang, Tian Yang, Yan Li and Dongfang Ma
J. Fungi 2024, 10(12), 879; https://doi.org/10.3390/jof10120879 - 17 Dec 2024
Viewed by 342
Abstract
Aspartic proteases (APs), hydrolases with aspartic acid residues as catalytic active sites, are closely associated with processes such as plant growth and development and fungal and bacterial pathogenesis. F. graminearum is the dominant pathogenic fungus that causes Fusarium head blight (FHB) in wheat. [...] Read more.
Aspartic proteases (APs), hydrolases with aspartic acid residues as catalytic active sites, are closely associated with processes such as plant growth and development and fungal and bacterial pathogenesis. F. graminearum is the dominant pathogenic fungus that causes Fusarium head blight (FHB) in wheat. However, the relationship of APs to the growth, development, and pathogenesis of F. graminearum is not clear. Therefore, we selected the FGSG_09558 gene, whose function annotation is aspartate protease, for further study. In this study, FGSG_09558 was found to contain a conserved structural domain and signal peptide sequence of aspartic acid protease and was therefore named FgAsp. The function of FgAsp in F. graminearum was investigated by constructing the knockout and complementation mutants of this gene. The results showed that with respect to the wild type (PH-1), the knockout mutant showed a significant reduction in mycelial growth, asexual spore production, and sexual spore formation, highlighting the key role of FgAsp in the growth and development of F. graminearum. In addition, the mutants showed a significant reduction in the virulence and accumulation level of deoxynivalenol (DON) content on maize whiskers, wheat germ sheaths, and wheat ears. DON, as a key factor of virulence, plays an important role in the F. graminearum infection of wheat ears, suggesting that FgAsp is involved in the regulation of F. graminearum pathogenicity by affecting the accumulation of the DON toxin. FgAsp had a significant effect on the ability of F. graminearum to utilize various sugars, especially arabinose. In response to the stress, hydrogen peroxide inhibited the growth of the mutant most significantly, indicating the important function of FgAsp in the strain’s response to environmental stress. Finally, FgAsp plays a key role in the regulation of F. graminearum growth and development, pathogenicity, and environmental stress response. Full article
(This article belongs to the Special Issue Growth and Virulence of Plant Pathogenic Fungi)
Show Figures

Figure 1

Figure 1
<p>The <span class="html-italic">FgAsp</span> gene deletion and complementation strategies.</p>
Full article ">Figure 2
<p>Description of <span class="html-italic">FgAsp</span>. (<b>A</b>) Conserved functional domain. (<b>B</b>) Identification of transmembrane domains. (<b>C</b>) Three-dimensional homology modeling. (<b>D</b>) Signal peptide prediction results. (<b>E</b>) Gene expression level of <span class="html-italic">FgAsp</span> in <span class="html-italic">F. graminearum</span>.</p>
Full article ">Figure 3
<p>(<b>A</b>) Colony morphology of PH-1, Δ<span class="html-italic">FgAsp</span>, and CΔ<span class="html-italic">FgAsp</span>. (<b>B</b>) Growth rates of wild-type PH-1, Δ<span class="html-italic">FgAsp</span>, and CΔ<span class="html-italic">FgAsp</span> strains. (<b>C</b>) PH-1, Δ<span class="html-italic">FgAsp</span>, and CΔ<span class="html-italic">FgAsp</span> hyphal edge morphology. Scale bar = 20 μm. Means and standard errors were calculated using <span class="html-italic">t</span>-tests based on data from three independent biological replicates. Different letters indicate significant difference at the level of 0.05.</p>
Full article ">Figure 4
<p>(<b>A</b>) Conidiophores of PH-1, Δ<span class="html-italic">FgAsp</span>, and CΔ<span class="html-italic">FgAsp</span>. The red arrows indicate the attached conidia on the conidial peduncle of each strain. Scale bar = 25 μm. (<b>B</b>) The sporogenesis rates of PH-1, Δ<span class="html-italic">FgAsp</span>, and CΔ<span class="html-italic">FgAsp</span>. Different lowercase letters a and b represent significant differences. (<b>C</b>) Statistics of the number of septa in conidia of PH-1, Δ<span class="html-italic">FgAsp</span>, and CΔ<span class="html-italic">FgAsp</span>. Scale bar = 25 μm. (<b>D</b>) Conidia germination statistics of PH-1, Δ<span class="html-italic">FgAsp</span>, and CΔ<span class="html-italic">FgAsp</span>.</p>
Full article ">Figure 5
<p>Pathogenicity and lesion length of PH-1, Δ<span class="html-italic">FgAsp</span>, and CΔ<span class="html-italic">FgAsp</span>: (<b>A1</b>,<b>A2</b>) Wheat coleoptiles, (<b>B1</b>,<b>B2</b>) wheat leaves, (<b>C1</b>,<b>C2</b>) wheat ears, (<b>D1</b>,<b>D2</b>) corn silks. The images above show the pathogenicity and lesion pictures, and the violin plot of lesion length is displayed below.</p>
Full article ">Figure 6
<p>Sexual reproduction of PH-1, Δ<span class="html-italic">FgAsp</span>, and CΔ<span class="html-italic">FgAsp</span>: (<b>A</b>) Number of ascospores produced by sexual reproduction of PH-1, Δ<span class="html-italic">FgAsp</span>, and CΔ<span class="html-italic">FgAsp</span>. (<b>B</b>) Eruption of ascocarp primordia. Scale bar = 2000 μm. (<b>C</b>) Ascospores. Scale bar = 50 μm. (<b>D</b>) Number of ascospores per asci (individuals). Different lowercase letters a and b represent significant differences.</p>
Full article ">Figure 7
<p>(<b>A</b>) DON toxin content in TBI medium of PH-1, Δ<span class="html-italic">FgAsp</span>, and CΔ<span class="html-italic">FgAsp</span>. (<b>B</b>) DON toxin content in wheat kernels of PH-1, Δ<span class="html-italic">FgAsp</span>, and CΔ<span class="html-italic">FgAsp</span>. (<b>C</b>) Expression levels of TRI gene clusters in PH-1, Δ<span class="html-italic">FgAsp</span>, and CΔ<span class="html-italic">FgAsp</span> after 6 days of TBI culture. Means and standard errors were calculated using <span class="html-italic">t</span>-tests based on data from three independent biological replicates. Different letters indicate significant difference at the level of 0.05.</p>
Full article ">Figure 8
<p>(<b>A</b>) Colony morphology of PH-1, Δ<span class="html-italic">FgAsp</span>, and CΔ<span class="html-italic">FgAsp</span> on PSA medium containing NaCl, KCl, MgCl<sub>2</sub>, CaCl<sub>2</sub>, and H<sub>2</sub>O<sub>2</sub>. (<b>B</b>) Stress growth inhibition rate analysis. Means and standard errors were calculated using <span class="html-italic">t</span>-tests based on data from three independent biological replicates. An asterisk (*) indicates a <span class="html-italic">p</span> value of less than 0.05, that is, the difference is significant at the 5% significance level. Two asterisks (**) indicate a <span class="html-italic">p</span> value of less than 0.01, that is, significant at the 1% significance level. Three asterisks (***) indicate a <span class="html-italic">p</span> value of less than 0.001, which is extremely significant at the 0.1% significance level. ns indicates no difference.</p>
Full article ">Figure 9
<p>(<b>A</b>) Colony morphology of PH-1, Δ<span class="html-italic">FgAsp</span>, and CΔ<span class="html-italic">FgAsp</span> on PSA medium containing sucrose, arabinose, mannose, glucose, and galactose. (<b>B</b>) Analysis of different glycogen inhibition rates of wild-type PH-1, Δ<span class="html-italic">FgAsp</span>, and CΔ<span class="html-italic">FgAsp</span> strains. Data were tested by <span class="html-italic">t</span>-test, and error bars represent the standard deviation (SD). Different letters indicate a significant difference at the level of 0.05.</p>
Full article ">
22 pages, 2950 KiB  
Article
Egg Protein Compositions over Embryonic Development in Haemaphysalis hystricis Ticks
by Qiwu Tang, Tianyin Cheng and Wei Liu
Animals 2024, 14(23), 3466; https://doi.org/10.3390/ani14233466 - 30 Nov 2024
Viewed by 472
Abstract
Tick eggs contain a series of proteins that play important roles in egg development. A thorough characterization of egg protein expression throughout development is essential for understanding tick embryogenesis and for screening candidate molecules to develop novel interventions. In this study, eggs at [...] Read more.
Tick eggs contain a series of proteins that play important roles in egg development. A thorough characterization of egg protein expression throughout development is essential for understanding tick embryogenesis and for screening candidate molecules to develop novel interventions. In this study, eggs at four developmental stages (0, 7, 14, and 21 incubation days) were collected, and their protein extraction was profiled using sodium dodecyl sulfate–polyacrylamide gel electrophoresis (SDS-PAGE). On the first day of egg protein extraction, protein bands from day-1 eggs were re-collected and subsequently analyzed using liquid chromatography–tandem mass spectrometry (LC-MS/MS). The dynamic changes in forty egg proteins during development were further investigated using LC-parallel reaction monitoring (PRM)/MS analysis. A total of 108 transcripts were detected in day-1 eggs. Based on protein functions and families, these transcripts were classified into eight categories: transporters, enzymes, immunity and antimicrobial proteins, proteinase inhibitors, cytoskeletal proteins, heat shock proteins, secreted proteins, and uncharacterized proteins. Identification of the protein bands revealed that nine bands predominantly consisted of vitellogenin and vitellin-A, while other notable proteins included cathepsins and Kunitz domain-containing proteins. LC-PRM/MS analysis indicated that 28 transcripts increased significantly in abundance, including 13/18 enzymes, 1/1 antimicrobial peptide, 2/2 neutrophil elastase inhibitors, 3/4 vitellogenins, 3/3 heat shock proteins, 3/3 cytoskeletal proteins, 1/1 elongation factor-1, and 1/1 uncharacterized protein. Conversely, five transcripts showed a decrease significantly, including 1/1 Kunitz domain-containing protein, 2/6 aspartic proteases, and 2/5 serpins. This research provides a comprehensive overview of egg proteins and highlights the dynamic changes in protein expression during embryonic development, which may be pivotal for understanding protein functions and selecting potential candidates for further study. Full article
Show Figures

Figure 1

Figure 1
<p>SDS-PAGE analysis of protein extract from <span class="html-italic">H. hystricis</span> eggs. A total of 80 μg of protein extract was loaded per sample. Lane M represents the 15–250 kDa molecular weight marker. D1, D7, D14, and D21 correspond to eggs incubated for 1, 7, 14, and 21 days, respectively. Bands 1–12 indicate the protein bands excised for further analysis.</p>
Full article ">Figure 2
<p>Top 4 most prominent proteins identified in 12 bands from day-one eggs.</p>
Full article ">Figure 3
<p>Analysis of protein dynamics in <span class="html-italic">H. hystricis</span> tick eggs across different incubation days. Each sample was spiked with the stable isotope iRT KIT peptide as an internal standard. Tryptic peptides were analyzed using the nLC-1200 system. Protein abundances at 7, 14, and 21 days of incubation were normalized to the levels observed at day. D1, D7, D14, and D21 correspond to eggs incubated for 1, 7, 14, and 21 days, respectively.</p>
Full article ">
18 pages, 2801 KiB  
Article
From Beer to Cheese: Characterization of Caseinolytic and Milk-Clotting Activities of Proteases Derived from Brewer’s Spent Grain (BSG)
by Maximiliano M. Villegas, Johana N. Silva, Florencia R. Tito, Claudia V. Tonón, Fernando F. Muñoz, Alfonso Pepe and María G. Guevara
Foods 2024, 13(22), 3658; https://doi.org/10.3390/foods13223658 - 17 Nov 2024
Viewed by 1029
Abstract
This study explores the extraction and characterization of proteolytic enzymes from brewer’s spent grain (BSG) and their potential as sustainable coagulants in the dairy industry. BSG samples from various beer types (Blonde Ale, IPA, Kölsch, Honey, and Porter) were obtained from two artisanal [...] Read more.
This study explores the extraction and characterization of proteolytic enzymes from brewer’s spent grain (BSG) and their potential as sustainable coagulants in the dairy industry. BSG samples from various beer types (Blonde Ale, IPA, Kölsch, Honey, and Porter) were obtained from two artisanal breweries in Mar del Plata, Argentina. Optimization of caseinolytic activity (CA) and protein extraction was conducted using a Plackett–Burman design, followed by a Box–Behnken design. Optimal protein concentration was achieved at intermediate pH and high temperature, while CA peaked at pH 8.0. The specific caseinolytic activity (SCA) varied among the extracts, with BSG3 showing the highest activity (99.6 U mg−1) and BSG1 the lowest (60.4 U mg−1). Protease inhibitor assays suggested the presence of aspartic, serine, metallo, and cysteine proteases. BSG3 and BSG4 showed the highest hydrolysis rates for α-casein (70% and 78%). For κ-casein, BSG1, BSG2, and BSG3 demonstrated moderate activity (56.5%, 49%, and 55.8), while BSG4 and BSG5 exhibited the lowest activity. Additionally, the milk-clotting activity (MCA) of BSG extracts was comparable to plant-based coagulants like Cynara cardunculus and Ficus carica. These findings highlight the potential of BSG-derived proteases as alternative coagulants for cheese production, offering a sustainable link between the brewing and dairy industries. Full article
(This article belongs to the Section Food Analytical Methods)
Show Figures

Figure 1

Figure 1
<p>Pareto (<b>A</b>,<b>C</b>) and main effect plots (<b>B</b>,<b>D</b>) from the Plackett–Burman design analysis. The six independent variables examined were pH, temperature, homogenization time, and the concentrations of Triton X-100, DTT, and CaCl<sub>2</sub>. The blue plots represent CA, while the green plots correspond to protein concentration. Variables with values exceeding the threshold indicated by the dashed red line are considered significant.</p>
Full article ">Figure 2
<p>RSM analysis of the BBD. Two response variables were evaluated: CA and protein concentration. The independent variables identified as significant in the screening design (pH, temperature, and CaCl<sub>2</sub> concentration) were selected for the BBD. Subpanels (<b>A</b>,<b>E</b>) display the significant terms for each model, with values exceeding the threshold indicated by the dashed red line considered significant. (<b>B</b>–<b>D</b>) depict the 3D CA response surfaces, showing the interactions between two independent variables while holding the third at its optimal level. (<b>F</b>–<b>H</b>) depict the 3D protein concentration response surfaces, showing the interactions between two independent variables while holding the third at its optimal level.</p>
Full article ">Figure 3
<p>Biochemical characterization of CA of BSG1. (<b>A</b>) pH dependence: Relative activity of the BSG extract was measured across different pH values using citrate buffer (pH 3, 4, 5), phosphate buffer (pH 6, 7, 8), and Tris-HCl buffer (pH 9). (<b>B</b>) Temperature dependence: The relative activity of the extract was assessed at various temperatures using 100 mM phosphate buffer at pH 6.5. (<b>C</b>) Endopeptidases inhibition profile: Inhibitory effects on CA of BSG protein extracts of 40 mM Pepstatin A (aspartic peptidase inhibitor), 1.5 mM PMSF (serine peptidase inhibitor), 5 mM EDTA (metallopeptidase inhibitor), and 4 mM Iodoacetamide (cysteine peptidase inhibitor) were analyzed. CA values for each condition were calculated relative to BSG without inhibitors (control), as described in <a href="#sec2-foods-13-03658" class="html-sec">Section 2</a> Materials and Methods. (<b>D</b>) Specific caseinolytic activity (SCA): The SCA of BSG endopeptidases on bovine α-, β- and κ-casein subunits was determined as indicated in the <a href="#sec2-foods-13-03658" class="html-sec">Section 2</a> Materials and Method. Hydrolysis percentages were calculated using the SDS-PAGE densitometric analysis, considering 100% hydrolysis as the complete disappearance of the casein bands. Bars represent the standard deviation of three independent assays. Different letters indicate statistically significant differences (<span class="html-italic">p</span> &lt; 0.05).</p>
Full article ">Figure 4
<p>Milk-clotting activity of BSG proteases. Commercial skim milk was dissolved in a phosphate buffer (pH 6.5) containing 40 µL of 10 mM CaCl<sub>2</sub>, in a final volume of 2 mL, and incubated at 37 °C for 24 h. The following conditions were tested: (<b>A</b>) chymosin served as a positive control, (<b>B</b>) phosphate buffer with a pH of 6.5 was used as a negative control, (<b>C</b>) commercial skim milk in phosphate buffer with a pH of 6.5 with 10 mM CaCl<sub>2</sub> and BSG extract at a 1:2 ratio, and (<b>D</b>) 1:1 ratio.</p>
Full article ">Figure 5
<p>Comparative analysis of BSG extracts: protein concentration, specific caseinolytic activity, and activity on different casein subunits. (<b>A</b>) Protein concentration (mg mL<sup>−1</sup>) of different BSG extracts evaluated. (<b>B</b>) Specific caseinolytic activity of BSG extracts from different styles of beer. (<b>C</b>) Proteolytic activity on different casein subunits of BSG extracts from different styles of beer. Bars represent the standard deviation of three independent assays. Different letters indicate statistically significant differences (<span class="html-italic">p</span> &lt; 0.05).</p>
Full article ">Figure 6
<p>MCA of BSG extracts from different types of beer: 1 mL of extract was added to 1 mL of 12 % <span class="html-italic">w</span>/<span class="html-italic">v</span> commercial skim milk dissolved in 100 mM of the phosphate buffer with a pH of 6.5 containing 40 μL of 10 mM CaCl<sub>2</sub>, in a final volume of 2 mL, and incubated at 37 °C for 24 h. Chymosin was used as the positive control, and 100 mM of the phosphate buffer with a pH of 6.5 was used as the negative control.</p>
Full article ">
6163 KiB  
Proceeding Paper
Phytochemical Constituents from Globimetula oreophila as Plasmepsin I and II Inhibitors in Antimalarial Drug Discovery: An In Silico Approach
by Dauda Garba, Bila Hassan Ali, Bashar Bawa, Abdullahi Maryam, Hamza Asmau Nasiru, Yahaya Mohammed Sani, Muhammad Garba Magaji, Musa Isma’il Abdullahi, Aliyu Muhammad Musa and Hassan Halimatu Sadiya
Chem. Proc. 2024, 16(1), 42; https://doi.org/10.3390/ecsoc-28-20220 - 14 Nov 2024
Viewed by 61
Abstract
Malaria remains a critical global health challenge, particularly affecting Sub-Saharan Africa. Plasmepsins, vital in hydrolyzing peptide bonds within proteins, present promising targets for antimalarial drugs. Plasmepsins I and II, key aspartic proteases, are crucial in various parasite processes. This study investigates the inhibitory [...] Read more.
Malaria remains a critical global health challenge, particularly affecting Sub-Saharan Africa. Plasmepsins, vital in hydrolyzing peptide bonds within proteins, present promising targets for antimalarial drugs. Plasmepsins I and II, key aspartic proteases, are crucial in various parasite processes. This study investigates the inhibitory properties of quercetin, quercetrin, dihydrostilbene, 4′-methoxy-isoliquiritigenin, and stigmasterol from Globimetula oreophila on plasmepsins through in silico techniques, including ADME predictions and molecular docking. Results reveal strong interactions of these compounds with active site residues, with quercetrin and stigmasterol displaying notable binding affinities. These findings suggest the potential of G. oreophila metabolites as potent plasmepsin inhibitors, offering prospects for malaria treatment and prevention. Full article
Show Figures

Figure 1

Figure 1
<p>Three-dimensional molecular pose (<b>left</b>) and 2D (<b>right</b>) interactions of DG1 on binding cavity of plasmepsin I.</p>
Full article ">Figure 2
<p>Three-dimensional molecular pose (<b>left</b>) and 2D (<b>right</b>) interactions of DG2 on binding cavity of plasmepsin I.</p>
Full article ">Figure 3
<p>Three-dimensional molecular pose (<b>left</b>) and 2D (<b>right</b>) interactions of DG3 on binding cavity of plasmepsin I.</p>
Full article ">Figure 4
<p>Three-dimensional molecular pose (<b>left</b>) and 2D (<b>right</b>) interactions of DG4 on binding cavity of plasmepsin I.</p>
Full article ">Figure 5
<p>Three-dimensional molecular pose (<b>left</b>) and 2D (<b>right</b>) interactions of DG5 on binding cavity of plasmepsin I.</p>
Full article ">Figure 6
<p>Three-dimensional molecular pose (<b>left</b>) and 2D (<b>right</b>) interactions of DG1 on binding cavity of plasmepsin II.</p>
Full article ">Figure 7
<p>Three-dimensional molecular pose (<b>left</b>) and 2D (<b>right</b>) interactions of DG2 on binding cavity of plasmepsin II.</p>
Full article ">Figure 8
<p>Three-dimensional molecular pose (<b>left</b>) and 2D (<b>right</b>) interactions of DG3 on binding cavity of plasmepsin II.</p>
Full article ">Figure 9
<p>Three-dimensional molecular pose (<b>left</b>) and 2D (<b>right</b>) interactions of DG4 on binding cavity of plasmepsin II.</p>
Full article ">Figure 10
<p>Three-dimensional molecular pose (<b>left</b>) and 2D (<b>right</b>) interactions of DG5 on binding cavity of plasmepsin II.</p>
Full article ">
12 pages, 2995 KiB  
Article
Caspase-8-and Gasdermin D (GSDMD)-Dependent PANoptosis Participate in the Seasonal Atrophy of Scented Glands in Male Muskrats
by Xiaofeng Tong, Xuefei Zhao, Yue Ma, Haimeng Li, Jinpeng Zhang, Zuoyang Zhang, Sirui Hua, Bo Li, Wei Zhang, Yu Zhang and Suying Bai
Animals 2024, 14(22), 3194; https://doi.org/10.3390/ani14223194 - 7 Nov 2024
Viewed by 740
Abstract
The muskrat (Ondatra zibethicus) is an animal with special economic significance whose scented glands rapidly atrophy during the non-breeding season, but the mechanism of atrophy is not clear, with significant differences in apoptotic and pyroptotic signaling pathway expression according to transcriptome [...] Read more.
The muskrat (Ondatra zibethicus) is an animal with special economic significance whose scented glands rapidly atrophy during the non-breeding season, but the mechanism of atrophy is not clear, with significant differences in apoptotic and pyroptotic signaling pathway expression according to transcriptome sequencing. During the non-breeding season, key apoptosis-related genes such as Tnfr1 (TNF Receptor Superfamily Member 1A), TRADD (TNFRSF1A Associated via Death Domain), FADD (Fas Associated via Death Domain), Casp-8 (Cysteine-aspartic proteases-8), and Bax (Bcl-associated X protein) were upregulated in the scented glands, while Bcl2 (B-cell lymphoma-2) expression was downregulated. In the classical pyroptosis pathway, the mRNA expression levels of key genes including Nlrp3 (the Nod-like receptor family pyrin domain-containing 3), ASC (the apoptosis-associated speck-like protein), Casp-1 (Cysteine-aspartic proteases-1), Gsdmd (Gasdermin D), and IL-1β (Interleukin 1 Beta) were higher during the non-breeding season, similar to the transcription level of Ripk1 (Receptor Interacting Serine/Threonine Kinase 1) in the non-canonical pyroptosis pathway, while TAK1 (transforming growth factor kinase) expression was downregulated in this latter pathway. TUNEL assays and immunofluorescence analysis indicated increased apoptosis and GSDMD and Caspase-8 protein levels during the non-breeding season. Indeed, the protein levels of GSDMD-N, Caspase-8 p43, and Caspase-8 p18 were significantly higher during the non-breeding season, while the GSDMD levels were significantly lower compared to the secretion season. These results suggest that apoptosis and pyroptosis play regulatory roles in scented gland atrophy and that there is an interplay between them during this process. Full article
(This article belongs to the Section Mammals)
Show Figures

Figure 1

Figure 1
<p>Transcriptomic analyses of scented glands during breeding and non-breeding seasons. (<b>a</b>) Venn diagram of shared and unique transcripts. (<b>b</b>,<b>c</b>) Clustering analysis heatmap and volcano plot of the differentially expressed genes. (<b>d</b>,<b>e</b>) GO and KEGG enrichment analyses.</p>
Full article ">Figure 2
<p>Real-time quantitative PCR was used to detect the mRNA expression levels of the following genes in muskrats’ scented glands during the breeding and non-breeding seasons. (<b>a</b>) TNFR1, TRADD, FADD, Caspase-8, BAX, and BCL2. (<b>b</b>) NLRP3, ASC, Caspase-1, GSDMD, and IL-1β. (<b>c</b>) TAK1, RIPK1, Caspase-8, GSDMD, and FADD. (<b>d</b>,<b>e</b>) Also shown are the protein expression results of GSDMD and Caspase-8. (<b>f</b>–<b>i</b>) The grayscale analysis of GSDMD, GSDMD-N, Caspase-8 p18, and Caspase-8 p43. B—breeding season; NB—non-breeding season. The error bars represent the means ± SEM (<span class="html-italic">n</span> = 3, each stage). * Statistical significance (* <span class="html-italic">p</span> &lt; 0.05; ** <span class="html-italic">p</span> &lt; 0.01).</p>
Full article ">Figure 3
<p>Immunofluorescence results in muskrats’ scented glands during the breeding (<b>a</b>–<b>d</b>) and non-breeding seasons (<b>e</b>–<b>h</b>). The green (<b>a</b>,<b>e</b>) and red (<b>b</b>,<b>f</b>) fluorescence signals represent GSDMD and Caspase-8, respectively. Scale bar = 100 μm.</p>
Full article ">Figure 4
<p>TUNEL results in scented glands during the breeding (<b>a</b>–<b>c</b>) and non-breeding seasons (<b>d</b>–<b>f</b>); scale bar = 100 μm. Apoptosis index during the breeding and non-breeding seasons is shown (<b>g</b>). The error bars represent the means ± SEM (<span class="html-italic">n</span> = 5, each stage). * Statistical significance (*** <span class="html-italic">p</span> &lt; 0.001).</p>
Full article ">Figure 5
<p>Sketch of apoptosis and pyroptosis involved in muskrats’ scented glands.</p>
Full article ">
20 pages, 4369 KiB  
Article
Cathepsin B- and L-like Protease Activities Are Induced During Developmental Barley Leaf Senescence
by Igor A. Schepetkin and Andreas M. Fischer
Plants 2024, 13(21), 3009; https://doi.org/10.3390/plants13213009 - 28 Oct 2024
Viewed by 586
Abstract
Leaf senescence is a developmental process allowing nutrient remobilization to sink organs. Previously cysteine proteases have been found to be highly expressed during leaf senescence in different plant species. Using biochemical and immunoblotting approaches, we characterized developmental senescence of barley (Hordeum vulgare [...] Read more.
Leaf senescence is a developmental process allowing nutrient remobilization to sink organs. Previously cysteine proteases have been found to be highly expressed during leaf senescence in different plant species. Using biochemical and immunoblotting approaches, we characterized developmental senescence of barley (Hordeum vulgare L. var. ‘GemCraft’) leaves collected from 0 to 6 weeks after the onset of flowering. A decrease in total protein and ribulose-1,5-bisphosphate carboxylase/oxygenase (Rubisco) large subunits occurred in parallel with an increase in proteolytic activity measured using the fluorogenic substrates Z-RR-AMC, Z-FR-AMC, and casein labeled with fluorescein isothiocyanate (casein-FITC). Aminopeptidase activity detected with R-AMC peaked at week 3 and then decreased, reaching a low level by week 6. Maximal proteolytic activity with Z-FR-AMC and Z-RR-AMC was detected from pH 4.0 to pH 5.5 and pH 6.5 to pH 7.4, respectively, while two pH optima (pH 3.6 to pH 4.5 and pH 6.5 to pH 7.4) were found for casein-FITC. Compound E-64, an irreversible cysteine protease inhibitor, and CAA0225, a selective cathepsin L inhibitor, effectively inhibited proteolytic activity with IC50 values in the nanomolar range. CA-074, a selective cathepsin B inhibitor, was less potent under the same experimental conditions, with IC50 in the micromolar range. Inhibition by leupeptin and phenylmethylsulfonyl fluoride (PMSF) was weak, and pepstatin A, an inhibitor of aspartic acid proteases, had no effect at the concentrations studied (up to 0.2 mM). Maximal proteolytic activity with the aminopeptidase substrate R-AMC was detected from pH 7.0 to pH 8.0. The pH profile of DCG-04 (a biotinylated activity probe derived from E-64) binding corresponded to that found with Z-FR-AMC, suggesting that the major active proteases are related to cathepsins B and L. Moreover, immunoblotting detected increased levels of barley SAG12 orthologs and aleurain, confirming a possible role of these enzymes in senescing leaves. Full article
(This article belongs to the Special Issue Barley: A Versatile Crop for Sustainable Food Production)
Show Figures

Figure 1

Figure 1
<p>Total soluble protein content was followed over a timespan of 6 weeks, starting with the onset of flowering (0 weeks). Second leaves below the ear were analyzed. The data are presented as the mean values ± S.D. of four biological replicates. Statistically significant differences (* <span class="html-italic">p</span> &lt; 0.01) with “week-0” samples are indicated.</p>
Full article ">Figure 2
<p>Rubisco LSU contents were followed over a timespan of 6 weeks, starting with the onset of flowering (0 weeks). Second leaves below the ear were analyzed. Rubisco LSU was detected using antibodies raised against the N-terminus (panel (<b>A</b>)) or the C-terminus (panel (<b>B</b>)). Each lane was loaded with extract obtained from an equal amount of fresh weight (corresponding to 20 μg protein at week 0). Panel (<b>C</b>): Densitometry calculations of the 53 kD LSU bands are based on 3 blots for each data point, with bands normalized to week 0 (100%).</p>
Full article ">Figure 3
<p>Total proteolytic activity in senescing barley leaves. Panel (<b>A</b>): Activity measured with the fluorogenic substrate casein-FITC was followed over a timespan of 6 weeks, starting with the onset of flowering (0 weeks). Second leaves below the ear were analyzed. The data show mean values ± S.D. of four biological replicates. Panel (<b>B</b>): Concentration-dependent inhibition of casein-FITC activity by the cysteine protease inhibitor E-64. The experiment was performed three times, with one replicate shown. The activity assays (panels (<b>A</b>,<b>B</b>)) were performed at pH 7.4. Statistically significant differences (* <span class="html-italic">p</span> &lt; 0.01) with “week-0” samples are indicated.</p>
Full article ">Figure 4
<p>Proteolytic activities in senescing barley leaves. Activities measured with fluorogenic substrates were followed over a timespan of 6 weeks, starting with the onset of flowering (0 weeks). Second leaves below the ear were analyzed. Protease activities were measured with Z-RR-AMC (panel (<b>A</b>)), Z-FR-AMC (panel (<b>B</b>)), or R-AMC (panel (<b>C</b>)). All activity assays were performed at pH 7.4. The data in the panels are presented as mean values ± S.D. of four biological replicates. Statistically significant differences (* <span class="html-italic">p</span> &lt; 0.01) with “week-0” samples are indicated.</p>
Full article ">Figure 5
<p>Influence of pH on proteolytic activities in senescing barley leaves. Proteases were extracted from senescing leaves (second leaves below the ear) at 1 week (open circles) and 3 or 4 weeks (closed circles) after the onset of flowering. Activities were measured using the fluorogenic substrates casein-FITC (panel (<b>A</b>)), Z-RR-AMC (panel (<b>B</b>)), Z-FR-AMC (panel (<b>C</b>)), and R-AMC (panel (<b>D</b>)). Three different buffers were used with increments of 0.4 or 0.5, including Na-citrate buffer, pH: 3.6–5.5 (blue line); Na-phosphate buffer, pH: 5.5–7.4 (red line); and Tris–HCl buffer, pH: 8–10.5 (green line). The data in the panels are presented as mean values ± S.D. of three technical replicates.</p>
Full article ">Figure 6
<p>Concentration-dependent inhibition of R-AMC cleavage activity by cysteine protease inhibitor E-64. Proteases were extracted from senescing leaves (second leaves below the ear) at 3 weeks after the onset of flowering. Aminopetidase activity was measured using the fluorogenic substrate R-AMC at pH 5.5 or 7.4. The data show one experiment that is representative of three independent experiments.</p>
Full article ">Figure 7
<p>Influence of pH on binding of DCG-04, a cysteine peptidase-specific probe, in barley leaf extracts. Proteins were extracted from second leaves below the ear at 4 weeks after the onset of flowering, with equal amounts of fresh weight loaded in each lane. Proteins were labeled with 2.5 μM DCG-04 for 3 h. Panel (<b>A</b>): DCG-04 labeling was detected using streptavidin-HRP and a luminescent substrate. Panel (<b>B</b>): Densitometric analysis based on 3 blots (mean values ± S.D.); data were normalized to the 43 kDa band at pH 4.0. Buffers were used with increments of 0.4 or 0.5, including pH: 3.6, 4.0, 4.5, 5.0, 5.5, 6.0, 6.5, 7.0, 7.4, 8.0, 8.5, 9.0, 9.5, 10.0, and 10.5.</p>
Full article ">Figure 8
<p>Inhibition of DCG-04 binding by the inhibitors E-64, CAA0225, and CA-074. The protein extracts of barley leaves obtained 4 weeks after flowering were preincubated with different concentrations of the inhibitors and proteins were labeled with 2.5 μM DCG-04 at pH 5.5 for 3 h. DMSO (solvent for inhibitors) was used as a control.</p>
Full article ">Figure 9
<p>Violin-type plots showing the distribution of molecular weights for full-length cathepsin L- and B-like proteases with reported gene upregulation in transcriptomic analyses of barley leaf senescence [<a href="#B13-plants-13-03009" class="html-bibr">13</a>,<a href="#B14-plants-13-03009" class="html-bibr">14</a>,<a href="#B15-plants-13-03009" class="html-bibr">15</a>,<a href="#B16-plants-13-03009" class="html-bibr">16</a>,<a href="#B17-plants-13-03009" class="html-bibr">17</a>] (see <a href="#plants-13-03009-t001" class="html-table">Table 1</a>).</p>
Full article ">Figure 10
<p>Concentration-dependent effect of the protease inhibitors E-64 (panel (<b>A</b>)), CAA0225 (panel (<b>B</b>)), and CA-074 (panel (<b>C</b>)) on DCG-04 binding to the 38 and 43 kDa bands and IC<sub>50</sub> values, calculated from the densitometric analysis of blots. Densitometric analysis was based on 2 blots (mean values ± S.D.); data were normalized to the control value (DMSO solvent without inhibitor). The chemical structures of the inhibitors and calculated values of iLogP are shown on the right side of the figure.</p>
Full article ">Figure 11
<p>Cysteine peptidase activities in senescing barley leaves. Activities detected by labeling with 2.5 μM DCG-04 (3 h) were followed over a time span of 6 weeks, starting with the onset of flowering (0 weeks). Second leaves below the ear were analyzed, with equal amounts of fresh weight loaded in each lane. Panel (<b>A</b>): DCG-04 labeling. Panel (<b>B</b>): Densitometric analysis. Calculations were based on 2 blots (mean values ± S.D.).</p>
Full article ">Figure 12
<p>Immunoblot analysis of SAG12 in senescing barley leaves. Panel (<b>A</b>): Protein levels were followed over a timespan of 6 weeks, starting with the onset of flowering (0 weeks). Second leaves below the ear were analyzed, with equal amounts of fresh weight loaded in each lane. Panel (<b>B</b>): SAG12 densitometric analysis (38 kDa band). Densitometric analyses were performed on 2 blots (mean values ± S.D. are shown).</p>
Full article ">Figure 13
<p>Immunoblot analysis of aleurain in senescing barley leaves. Panel (<b>A</b>): Protein levels were followed over a timespan of 6 weeks, starting with the onset of flowering (0 weeks). Second leaves below the ear were analyzed, with equal amounts of fresh weight loaded in each lane. Panel (<b>B</b>): Aleurain densitometric analysis. Densitometric analyses were performed on 2 blots (mean values ± S.D.).</p>
Full article ">
14 pages, 8117 KiB  
Article
Ferulic Acid Interferes with Radioactive Intestinal Injury Through the DJ-1-Nrf2 and Sirt1-NF-κB-NLRP3 Pathways
by Xuemei Zhang, Haoyu Zhang, Mingyue Huang, Yu Mei, Changkun Hu, Congshu Huang, Huiting Zhang, Xue Wei, Yue Gao and Zengchun Ma
Molecules 2024, 29(21), 5072; https://doi.org/10.3390/molecules29215072 - 26 Oct 2024
Viewed by 969
Abstract
Radiation-induced intestinal injury is a common complication of radiotherapy for abdominal and pelvic malignancies. Due to its rapid proliferation, the small intestine is particularly sensitive to radiation, making it a critical factor limiting treatment. Ferulic acid (FA), a derivative of cinnamic acid, exhibits [...] Read more.
Radiation-induced intestinal injury is a common complication of radiotherapy for abdominal and pelvic malignancies. Due to its rapid proliferation, the small intestine is particularly sensitive to radiation, making it a critical factor limiting treatment. Ferulic acid (FA), a derivative of cinnamic acid, exhibits antioxidant, anti-inflammatory, and anti-radiation properties. In this study, we established a mouse model of radiation-induced intestinal injury using a dose of 11 Gy at a rate of 96.62 cGy/min. Our findings indicate that FA’s protective effects against radiation-induced intestinal injury may be mediated through the parkinsonism-associated deglycase (DJ-1) nuclear factor erythroid 2-related factor 2 (Nrf2) and silent mating type information regulation 2 homolog 1 (Sirt1) nuclear Factor kappa-light-chain-enhancer of activated B cells (NF-κB) NOD-like receptor family, pyrin domain containing 3 (NLRP3). FA was found to mitigate changes in oxidative stress indices and inflammatory factors induced by radiation, as well as to attenuate radiation-induced pathological alterations in the small intestine. Furthermore, FA enhanced the expression of DJ-1 and Nrf2 at both the transcriptional and protein levels, inhibited NLRP3 protein fluorescence intensity, and reduced the expression of NLRP3, interleukin-18 (IL-18), and interleukin-1 beta (IL-1β). Additionally, FA suppressed the transcription and translation of NF-κB, NLRP3, cysteine-aspartic acid protease-1 (Caspase-1), IL-18, and IL-1β by upregulating Sirt1, thereby alleviating radiation-induced inflammatory injury in the small intestine. Thus, FA holds promise as an effective therapeutic agent for ameliorating radiation-induced intestinal injury. Full article
Show Figures

Graphical abstract

Graphical abstract
Full article ">Figure 1
<p>Effect of FA on oxidative stress index of intestines after radiation. (<b>A</b>) Effect of FA on MDA content in the intestines of irradiated mice. (<b>B</b>) Effect of FA on SOD content in intestines of irradiated mice. Data are expressed as mean ± SD (<span class="html-italic">n</span> = 3). Compared with control group, *** <span class="html-italic">p</span> &lt; 0.001; compared with the radiation group, ## <span class="html-italic">p</span> &lt; 0.05 and ### <span class="html-italic">p</span> &lt; 0.001 (C: control; R: radiation; FA: ferulic acid; Res: resveratrol).</p>
Full article ">Figure 2
<p>Effect of FA on the expression of inflammatory factors in the intestines of mice after radiation. (<b>A</b>) Effect of FA on TNF-α content in the intestines of irradiated mice. (<b>B</b>) Effect of FA on IL-6 content in the intestines of irradiated mice. Data are expressed as mean ± SD (n = 3). Compared with control group, *** <span class="html-italic">p</span> &lt; 0.001; compared with the radiation group, # <span class="html-italic">p</span> &lt; 0.05 and ### <span class="html-italic">p</span> &lt; 0.001.</p>
Full article ">Figure 3
<p>H&amp;E stained images showing the intestinal morphology in cross-sections of the intestine. Comparisons of histopathological characteristics of hippocampal tissue among the seven group. Scale bar = 500 μm.</p>
Full article ">Figure 4
<p>FA activates the DJ-1-Nrf2 pathway in the intestines of mice after radiation. (<b>A</b>) The mRNA level of DJ-1 and Nrf2 were detected by qRT-PCR. Data are expressed as mean ± SD (n = 3). (<b>B</b>) The protein expression of DJ-1 and Nrf2 were detected by Western blot assay. Compared with control group, *** <span class="html-italic">p</span> &lt; 0.001, * <span class="html-italic">p</span> &lt; 0.05; compared with radiation group, # <span class="html-italic">p</span> &lt; 0.05, ## <span class="html-italic">p</span> &lt; 0.05, and ### <span class="html-italic">p</span> &lt; 0.05.</p>
Full article ">Figure 5
<p>FA inhibited the expression of NLRP3 protein in the intestines of radiation mice by immunofluorescence assay. The expression and localization of NLRP3 in intestine were analyzed by immunofluorescence. All magnifications: ×400.</p>
Full article ">Figure 6
<p>Effects of FA on related protein in the intestines of mice after radiation. (<b>A</b>) Effect of FA on NLRP3 protein expression after radiation. (<b>B</b>) Effect of FA on IL-18 protein expression after radiation. (<b>C</b>) Effect of FA on IL-1β protein expression after radiation. Compared with control group, *** <span class="html-italic">p</span> &lt; 0.001; compared with radiation group, # <span class="html-italic">p</span> &lt; 0.05, ## <span class="html-italic">p</span> &lt; 0.05, and ### <span class="html-italic">p</span> &lt; 0.05.</p>
Full article ">Figure 7
<p>FA activates the Sirt1-NF-κB-NLRP3 pathway in the intestines of mice after radiation. (<b>A</b>) The mRNA level of Sirt1, NF-κB, NLRP3, Caspase-1, and IL-18, IL-1β were detected by qRT-PCR. Data are expressed as mean ± SD (n = 3). (<b>B</b>) The protein expression of Sirt1, NF-κB, NLRP3, Caspase-1, and IL-18, IL-1β were detected by Western blot assay. Compared with control group, * <span class="html-italic">p</span> &lt; 0.05, ** <span class="html-italic">p</span> &lt; 0.1, and *** <span class="html-italic">p</span> &lt; 0.001; compared with radiation group, # <span class="html-italic">p</span> &lt; 0.05, ## <span class="html-italic">p</span> &lt; 0.01, and ### <span class="html-italic">p</span> &lt; 0.001.</p>
Full article ">
15 pages, 15159 KiB  
Article
Apoptosis, Mitochondrial Autophagy, Fission, and Fusion Maintain Mitochondrial Homeostasis in Mouse Liver Under Tail Suspension Conditions
by Lu-Fan Li, Jiao Yu, Rui Li, Shan-Shan Li, Jun-Yao Huang, Ming-Di Wang, Li-Na Jiang, Jin-Hui Xu and Zhe Wang
Int. J. Mol. Sci. 2024, 25(20), 11196; https://doi.org/10.3390/ijms252011196 - 18 Oct 2024
Viewed by 1049
Abstract
Microgravity can induce alterations in liver morphology, structure, and function, with mitochondria playing an important role in these changes. Tail suspension (TS) is a well-established model for simulating the effects of microgravity on muscles and bones, but its impact on liver function remains [...] Read more.
Microgravity can induce alterations in liver morphology, structure, and function, with mitochondria playing an important role in these changes. Tail suspension (TS) is a well-established model for simulating the effects of microgravity on muscles and bones, but its impact on liver function remains unclear. In the current study, we explored the regulatory mechanisms of apoptosis, autophagy, fission, and fusion in maintaining liver mitochondrial homeostasis in mice subjected to TS for 2 or 4 weeks (TS2 and TS4). The results showed the following: (1) No significant differences were observed in nuclear ultrastructure or DNA fragmentation between the control and TS-treated groups. (2) No significant differences were detected in the mitochondrial area ratio among the three groups. (3) Cysteine aspartic acid-specific protease 3 (Caspase3) activity and the Bcl-2-associated X protein (bax)/B-cell lymphoma-2 (bcl2) ratio were not higher in the TS2 and TS4 groups compared to the control group. (4) dynamin-related protein 1 (DRP1) protein expression was increased, while mitochondrial fission factor (MFF) protein levels were decreased in the TS2 and TS4 groups compared to the control, suggesting stable mitochondrial fission. (5) No significant differences were observed in the optic atrophy 1 (OPA1), mitofusin 1 and 2 (MFN1 and MFN2) protein expression levels across the three groups. (6) Mitochondrial autophagy vesicles were present in the TS2 and TS4 groups, with a significant increase in Parkin phosphorylation corresponding to the duration of the TS treatment. (7) ATP synthase and citrate synthase activities were significantly elevated in the TS2 group compared to the control group but were significantly reduced in the TS4 group compared to the TS2 group. In summary, the coordinated regulation of apoptosis, mitochondrial fission and fusion, and particularly mitochondrial autophagy preserved mitochondrial morphology and contributed to the restoration of the activities of these two key mitochondrial enzymes, thereby maintaining liver mitochondrial homeostasis in mice under TS conditions. Full article
Show Figures

Figure 1

Figure 1
<p>Influence of TS on morphological data in mice. Numerical values are mean ± standard deviation. n = 8. CON, control group; TS2, tail suspension 2-week group; and TS4, tail suspension 4-week group. ** <span class="html-italic">p</span> &lt; 0.01 and * <span class="html-italic">p</span> &lt; 0.05.</p>
Full article ">Figure 2
<p>Histological morphology of mouse liver under different TS treatments. (<b>a</b>) Scale = 100 μm. (<b>b</b>) Scale = 20 μm. Arrows point to hepatocytes. Dashed arrows point to sinusoids. # shows vasculature. Cells within square boxes are necrotic liver cells. CON, control group; TS2, tail suspension 2-week group; and TS4, tail suspension 4-week group.</p>
Full article ">Figure 3
<p>Ultrastructure of mouse liver tissue and analysis of mitochondrial number and area under different TS treatments. (<b>a</b>) Scale = 1 μm. * shows mitochondrial autophagic vesicles. Arrows point to mitochondria. (<b>b</b>) Scale = 10 μm. # shows liver nucleus. Arrows point to mitochondria. (<b>c</b>) Number of mitochondria. (<b>d</b>) Mitochondrial cross-sectional area. (<b>e</b>) Mitochondrial area ratio. Numerical values are mean ± standard deviation. Fifteen pictures were analyzed in each group. CON, control group; TS2, tail suspension 2-week group; and TS4, tail suspension 4-week group. ** <span class="html-italic">p</span> &lt; 0.01 and * <span class="html-italic">p</span> &lt; 0.05.</p>
Full article ">Figure 4
<p>Activities of ATP synthase (<b>a</b>), CS (<b>b</b>), and Caspase3 (<b>c</b>) in mouse liver under different TS treatments. Numerical values are mean ± standard deviation. n = 8. CON, control group; TS2, tail suspension 2-week group; and TS4, tail suspension 4-week group. ** <span class="html-italic">p</span> &lt; 0.01.</p>
Full article ">Figure 5
<p>TUNEL staining of mouse liver under different TS treatments. (<b>a</b>) Scale = 20 μm. Arrows point to DNA fragmentation. Blue fluorescence indicates nuclei; green fluorescence indicates DNA fragmentation. (<b>b</b>) Negative control for TUNEL staining of mouse liver. Scale = 20 μm. CON, control group; TS2, tail suspension 2-week group; and TS4, tail suspension 4-week group.</p>
Full article ">Figure 6
<p>Expression levels of liver apoptosis-associated proteins in mice under different TS treatments. (<b>a</b>) Representative Western blot gels. (<b>b</b>) Polyacrylamide gel of total protein. (<b>c</b>) Apoptosis-associated protein level. Numerical values are mean ± standard deviation. n = 8. CON, control group; TS2, tail suspension 2-week group; and TS4, tail suspension 4-week group. * <span class="html-italic">p</span> &lt; 0.05.</p>
Full article ">Figure 7
<p>Expression levels of mitochondrial fission- and fusion-associated proteins in mice liver mitochondria under different TS treatments. (<b>a</b>) Representative Western blot gels. (<b>b</b>) Polyacrylamide gel of total protein in the liver. (<b>c</b>) Mitochondrial fission-associated protein levels. (<b>d</b>) Expression levels of mitochondrial fusion-associated proteins. Numerical values are mean ± standard deviation. n = 8. CON, control group; TS2, tail suspension 2-week group; and TS4, tail suspension 4-week group. * <span class="html-italic">p</span> &lt; 0.05.</p>
Full article ">Figure 8
<p>Expression levels of autophagy-associated proteins in mouse liver mitochondria under different TS treatments. (<b>a</b>) Representative Western blot gels. (<b>b</b>) Polyacrylamide gel of total protein. (<b>c</b>) Mitochondrial autophagy-associated protein levels. Numerical values are mean ± standard deviation. n = 8. ** <span class="html-italic">p</span> &lt; 0.01 and * <span class="html-italic">p</span> &lt; 0.05.</p>
Full article ">Figure 9
<p>Summary of impact of TS on liver mitochondrial homeostasis in mice. bcl2, B-cell lymphoma-2; bax, Bcl-2-associated X protein; caspase3, cysteine aspartic acid-specific protease 3; MFN1, mitofusin 1; MFN2, mitofusin 2; OPA1, optic atrophy 1; MFF, mitochondrial fission factor; DRP1, dynamin-related protein 1; Parkin, Parkinson disease protein 2; P-Parkin, phosphorylated Parkin; ATP synthase, adenosine triphosphate synthase; and CS, citrate synthase. Yellow represents apoptosis-related proteins. Gray represents mitochondrial fusion-related proteins. Pink represents mitochondrial fission-related proteins. Green represents mitochondrial autophagy-related proteins. Blue represents oxidative phosphorylation-related proteins. Red arrows represent up- or down-regulation in TS2 group. Blue arrows represent up- or down-regulation in TS4 group.</p>
Full article ">
18 pages, 4999 KiB  
Article
Screening, Growing, and Validation by Catalog: Using Synthetic Intermediates from Natural Product Libraries to Discover Fragments for an Aspartic Protease Through Crystallography
by Franziska U. Huschmann, Janis Mueller, Alexander Metz, Moritz Ruf, Johanna Senst, Serghei Glinca, Johannes Schiebel, Andreas Heine and Gerhard Klebe
Crystals 2024, 14(9), 755; https://doi.org/10.3390/cryst14090755 - 25 Aug 2024
Viewed by 847
Abstract
Fragment screening directly on protein crystals has been applied using AnalytiCon’s collection of intermediates that have been utilized to generate libraries of larger synthetic natural product-like molecules. The fragments with well-balanced physicochemical properties show an impressively high hit rate for a screen using [...] Read more.
Fragment screening directly on protein crystals has been applied using AnalytiCon’s collection of intermediates that have been utilized to generate libraries of larger synthetic natural product-like molecules. The fragments with well-balanced physicochemical properties show an impressively high hit rate for a screen using the aspartic protease endothiapepsin. The subsequent validation and expansion of the discovered fragment hits benefits from AnalytiCon’s comprehensive library design. Since the screened fragments are intermediates that share a common core with larger and closely related analogs with modulated substitution patterns, they allow for the retrieval of off-the-shelf follow-up compounds, which enable the development of design strategies for fragment optimization. A promising bicyclic core scaffold found in several fragment hits could be validated by selecting a set of enlarged follow-up compounds. Due to unexpected changes in binding mode and no significant improvement in ligand efficiency, this series was quickly deemed unsuitable and therefore discontinued. The structures of follow-up compounds of two other fragments helped to evaluate a putative fusion of two overlapping fragment hits. A design concept on how to fuse the two fragments could be proposed and helps to plan a suitable substitution pattern and promising central bridging element. Full article
Show Figures

Figure 1

Figure 1
<p>(<b>a</b>–<b>o</b>). Crystallographically determined binding modes of the 15 fragments bound in complex with EP. The fragments are shown as yellow sticks, with the fragment ID in bold and the chemical formulas, contoured by the unbiased (F<sub>o</sub>−F<sub>c</sub>)-difference electron density, shown as green meshes (contour level at 2 σ). EP residues of the final refined EP complex structure are shown as gray sticks. The atoms are labeled with atom type colors (PyMOL default). The hydrogen bonds up to a donor-acceptor distance of 3.2 Å are shown as yellow dashed lines and involve interstitial water molecules. Waters are depicted as red spheres and labeled with numbers according to the corresponding PDB files (see <a href="#app1-crystals-14-00755" class="html-app">Table S1</a>). The spatial position of the binding sites, S1–S6, S1’, and S2’, are approximately marked as magenta areas. Fragment 175 is found at the protein surface distal from any of the substrate binding sites.</p>
Full article ">Figure 2
<p>The three fragments, <b>56</b> (<b>a</b>), <b>80</b> (<b>b</b>), and <b>81</b> (<b>c</b>), share a bicyclic parent scaffold but adopt different binding positions with the protein. While <b>56</b> and <b>81</b> have rather similar orientations and form a direct interaction with both aspartates of the catalytic dyad via their presumably protonated and charged amino group, <b>80</b> interacts with the dyad via the catalytic water molecule. The latter is displaced by the other two fragments. Comparing the binding modes of seven follow-up compounds with the same central parent scaffold, three different binding poses are observed. For <b>F147</b> and <b>F283</b> (<b>d</b>), the basic nitrogen of the ligands forms a water-mediated interaction with the aspartate residues in the catalytic center. <b>F144</b>, <b>F151</b>, <b>F976</b>, and <b>F985</b> (<b>e</b>) adopt a second binding mode and bind slightly further away from the dyad. A weak polar interaction is formed by the bridging ether oxygen in the central bicycle of the ligands to the aspartates mediated through the catalytic water molecule. A third binding mode is found for <b>F290</b> (<b>f</b>), which binds even further away (&gt;3.8 Å) from the catalytic water and the dyad. This ligand does not have basic nitrogen that could serve as an H-bond donor to the catalytic center.</p>
Full article ">Figure 3
<p>(<b>a</b>–<b>j</b>). Crystallographically determined binding modes of ten follow-up compounds in EP. The follow-up compounds are shown as yellow sticks, with their IDs in bold (“<b>F</b>” stands for “follow-up”), contoured by the unbiased (F<sub>o</sub>−F<sub>c</sub>)-difference electron density, shown as green meshes (contour level at 2 σ). EP residues of the final refined EP complex structure are shown as gray sticks. The atoms are labeled with atom type colors (PyMOL default). The hydrogen bonds up to a donor-acceptor distance of 3.2 Å are shown as yellow dashed lines and involve interstitial water molecules. Waters are depicted as red spheres and labeled with numbers according to the corresponding PDB files (see <a href="#app1-crystals-14-00755" class="html-app">Table S1</a>). The spatial position of the binding sites S1–S6 and S1’, S2’ are approximately marked as magenta areas.</p>
Full article ">Figure 4
<p>(<b>a</b>) In a structural superposition, the two fragments, <b>75</b> (<span class="html-italic">brown</span>) and <b>283</b> (<span class="html-italic">gray</span>), formally overlap in the center by a pyrrolidine and a piperidine ring. Both rings use their polar nitrogen to interact directly with the aspartates of the catalytic dyad. (<b>b</b>) The follow-up ligand <b>F903</b> (<span class="html-italic">pink</span>) confirms the binding mode of <b>75</b> (<span class="html-italic">brown</span>). Surprisingly, <b>F717</b> (<span class="html-italic">yellow</span>), which is structurally related to <b>283</b> (<span class="html-italic">gray</span>), adopts a completely different binding mode at a distant position. It adopts two slightly different conformations at the remote site. (<b>c</b>) The follow-up compound <b>F717</b> (<span class="html-italic">green</span>) shows the terminal piperidin-4-ol moiety in a twist boat conformation, whereas a chair conformation is observed for <b>283</b> (<span class="html-italic">gray</span>). Only in the chair conformation can a favorable H-bond be formed to the aspartates of the dyad. To demonstrate this difference, <b>F717</b> has been fitted onto the binding mode of fragment <b>283</b>. (<b>d</b>) The protein is shown with a <span class="html-italic">purple</span> surface. In <b>283</b> (<span class="html-italic">yellow</span> surface), there is a propargyl substituent that fits into a small crevice between loops 79–81 and loops 300–302. When <b>F717<sup>fit</sup></b> (<span class="html-italic">green</span>) is artificially superimposed on the position of <b>283</b> (<span class="html-italic">yellow</span>), its terminal pyridine moiety (<span class="html-italic">green</span> surface) would sterically collide with the loop formed by residues 300–302 of the protein. This could explain its altered binding mode found experimentally and suggest a propargyl substituent to be used in a putative ligand fused through a central pyrrolidine ring.</p>
Full article ">
14 pages, 2332 KiB  
Article
Effects of Dietary L-glutamic acid on the Growth Performance, Gene Expression Associated with Muscle Growth-Related Gene Expression, and Intestinal Health of Juvenile Largemouth Bass (Micropterus salmoides)
by Feifan Jiang, Wenqing Huang, Meng Zhou, Hongyan Gao, Xiaozhou Lu, Zhoulin Yu, Miao Sun and Yanhua Huang
Fishes 2024, 9(8), 312; https://doi.org/10.3390/fishes9080312 - 6 Aug 2024
Viewed by 824
Abstract
The present research examined the impact of L-glutamic acid (Glu) supplementation on the growth performance, muscle composition, gene expression correlated with muscle growth, and intestinal health of largemouth bass. There were 525 fish in total, which were distributed randomly into five groups. Each [...] Read more.
The present research examined the impact of L-glutamic acid (Glu) supplementation on the growth performance, muscle composition, gene expression correlated with muscle growth, and intestinal health of largemouth bass. There were 525 fish in total, which were distributed randomly into five groups. Each group had three replicates, and each replicate consisted of 35 fish. Groups with control and experimental diets were assigned glutamic acid amounts of 0.2%, 0.4%, 0.6%, and 0.8%. The findings demonstrated that glutamic acid supplementation enhanced growth performance, feed intake (FI), and condition factor (CF), with the best value being attained at 0.4% Glu. The mean muscle fiber area was increased and the muscle fiber density was decreased in the 0.6% Glu group. The levels of total amino acids and specific amino acids, such as glutamic acid, aspartic acid, leucine, valine, alanine, and glycine, were shown to be higher in the 0.6% Glu group. In the 0.6% Glu group, the mRNA expression levels of atrogin-1, murf-1, foxo3a, and 4e-bp1 were decreased compared to the control group. Conversely, the mRNA expression levels of myf5, myog, myod, s6k1, tor, akt, and pi3k were increased in the 0.6% Glu group compared to the control group. The 0.4% Glu group had higher intestinal amylase, lipase, and protease activities and greater villus height, villus width, and muscle thickness. In summary, Glu can support largemouth bass growth, muscular development, intestinal digestion, and absorption. Full article
Show Figures

Figure 1

Figure 1
<p>Effects of dietary Glu on muscle histology of largemouth bass. Means in the same row with different superscripts are significantly different (mean ± SD; ANOVA, <span class="html-italic">p</span> &lt; 0.05; <span class="html-italic">n</span> = 3).</p>
Full article ">Figure 2
<p>Effects of dietary Glu supplementation on the relative mRNA expression levels of phosphatidyl inositol 3 kinase (<span class="html-italic">pi3k</span>), protein kinase B (<span class="html-italic">akt</span>), and target of rapamycin (<span class="html-italic">tor</span>) in the muscle of largemouth bass. Means in the same row with different superscripts are significantly different (mean ± SD; ANOVA, <span class="html-italic">p</span> &lt; 0.05; <span class="html-italic">n</span> = 3).</p>
Full article ">Figure 3
<p>Relative mRNA expression level of 70 kDa ribosomal protein S6 kinase 1 (<span class="html-italic">S6K1</span>), myoblast determination protein (<span class="html-italic">myod</span>), myogenin (<span class="html-italic">myog</span>), myogenic factor 5 (<span class="html-italic">myf5</span>) in the muscle of largemouth bass fed Glu diets. Means in the same row with different superscripts are significantly different (mean ± SD; ANOVA, <span class="html-italic">p</span> &lt; 0.05; <span class="html-italic">n</span> = 3).</p>
Full article ">Figure 4
<p>Relative mRNA expression level of eukaryotic translation initiation factor 4E-binding protein 1 (<span class="html-italic">4e-bp1</span>), forkhead boxO3a (<span class="html-italic">foxo3a</span>), muscle-specific RING finger protein 1 (<span class="html-italic">murfF-1</span>), and atrogin-1 in the muscle of largemouth bass fed Glu diets. Means in the same row with different superscripts are significantly different (mean ± SD; ANOVA, <span class="html-italic">p</span> &lt; 0.05; <span class="html-italic">n</span> = 3).</p>
Full article ">Figure 5
<p>Effects of dietary Glu supplementation on intestinal morphology of largemouth bass.</p>
Full article ">
22 pages, 20873 KiB  
Article
Phytochelatin Synthase: An In Silico Comparative Analysis in Cyanobacteria and Eukaryotic Microalgae
by Michele Ferrari, Matteo Marieschi, Radiana Cozza and Anna Torelli
Plants 2024, 13(15), 2165; https://doi.org/10.3390/plants13152165 - 5 Aug 2024
Viewed by 1030
Abstract
Phytochelatins (PCs) are small cysteine-rich peptides involved in metal detoxification, not genetically encoded but enzymatically synthesized by phytochelatin synthases (PCSs) starting from glutathione. The constitutive PCS expression even in the absence of metal contamination, the wide phylogenetic distribution and the similarity between PCSs [...] Read more.
Phytochelatins (PCs) are small cysteine-rich peptides involved in metal detoxification, not genetically encoded but enzymatically synthesized by phytochelatin synthases (PCSs) starting from glutathione. The constitutive PCS expression even in the absence of metal contamination, the wide phylogenetic distribution and the similarity between PCSs and the papain-type cysteine protease catalytic domain suggest a wide range of functions for PCSs. These proteins, widely studied in land plants, have not been fully analyzed in algae and cyanobacteria, although these organisms are the first to cope with heavy-metal stress in aquatic environments and can be exploited for phytoremediation. To fill this gap, we compared the features of the PCS proteins of different cyanobacterial and algal taxa by phylogenetic linkage. The analyzed sequences fall into two main, already known groups of PCS-like proteins. Contrary to previous assumptions, they are not classed as prokaryotic and eukaryotic sequences, but rather as sequences characterized by the alternative presence of asparagine and aspartic/glutamic acid residues in proximity of the catalytic cysteine. The presence of these enzymes with peculiar features suggests differences in their post-translational regulation related to cell/environmental requirements or different cell functions rather than to differences due to their belonging to different phylogenetic taxa. Full article
(This article belongs to the Special Issue Heavy Metal Tolerance in Plants and Algae)
Show Figures

Figure 1

Figure 1
<p>The evolutionary history of PCSs. The maximum likelihood tree is shown. The PCS sequences are split into two main clusters indicated as branches 1 and 2. The PCS sequences are split into two main clusters indicated as branches 1 and 2.The percentage of trees in which the associated taxa clustered together is shown below the branches. The tree is drawn to scale, with branch lengths measured in the number of substitutions per site. This analysis involved 220 amino acid sequences (<a href="#app1-plants-13-02165" class="html-app">Table S1</a>) for a total of 3208 positions in the final dataset.</p>
Full article ">Figure 2
<p>Partial representation of multiple-sequence alignment of PCS proteins showing discriminant residues N, E and D (excerpt of <a href="#app1-plants-13-02165" class="html-app">Figure S2</a>). Partial representation of multiple-sequence alignment of PCS protein sequences referred to in <a href="#app1-plants-13-02165" class="html-app">Table S1</a> and <a href="#app1-plants-13-02165" class="html-app">Figure S2</a>. The selected sequences were chosen as representatives of different taxa. Alignment was conducted with ClustalW; identical and similar residues are shaded in black and gray, and consensus sequence is shown below alignment. Sequences of group 1 are characterized by one asparagine residue (N, magenta), often followed by a glutamine (Q, orange) four amino acids upstream of the catalytic cysteine (C, green). In the sequences of group 2, the asparagine residue is substituted by residue of glutamic acid (E, red), or by aspartic acid (D, cyan) in a sub-group of diatoms, followed by a proline (P, lilac). In yellow, the threonine (T) residue is a possible target of phosphorylation.</p>
Full article ">Figure 3
<p>The WebLogo of the sequence preceding the conserved Cys residue of the catalytic triad in different taxa. The sequences start from the initial Thr residue presumably involved in post-translational regulation through phosphorylation and the alternative residues characterizing the “N”, “E” and “D” isoforms (this position is marked with an asterisk). (<b>a</b>) Cyanobacteria; (<b>b</b>) diatoms; (<b>c</b>) red algae; (<b>d</b>) Chlamydomonadales; (<b>e</b>) Trebouxiophyceae; (<b>f</b>) Sphaeropleales.</p>
Full article ">Figure 4
<p>Multiple-sequence alignment of PCS proteins of cyanobacteria possessing both “N” and “E” isoforms. Alignment was conducted with ClustalX 2.0; identical and similar residues are shaded in black and gray, and the consensus sequence is shown below the alignment. Catalytic triad residues (Cys, His and Asp) are highlighted in green and cysteines in yellow. In orange is shown the Thr residue passible of phosphorylation, and in magenta and red are highlighted the Asn and Glu residues characteristics, respectively, of group 1 or 2 in the phylogenetic tree represented in <a href="#plants-13-02165-f001" class="html-fig">Figure 1</a>.</p>
Full article ">Figure 5
<p>Partial representation of multiple-sequence alignment of PCS proteins of diatoms possessing “N”, “E” and “D” isoforms. Alignment was conducted with ClustalW; identical and similar residues are shaded in black and gray, and consensus sequence is shown below alignment. Catalytic triad residues (Cys, His and Asp) are highlighted in green and cysteines in yellow. In red, cyan and magenta are highlighted the Glu, Asp, (characteristic of sub-branches “E” and “D” of group 2) and Asn residues of group 1, respectively, in the phylogenetic tree in <a href="#plants-13-02165-f001" class="html-fig">Figure 1</a>. In orange, is indicated the Thr residue, a possible target of phosphorylation.</p>
Full article ">Figure 6
<p>Partial representation of multiple-sequence alignment of PCS proteins of the model PCS sequences referred to in <a href="#plants-13-02165-t001" class="html-table">Table 1</a>. Alignment was conducted with ClustalX 2.0; identical and similar residues are shaded in black and gray, and consensus sequence is shown below alignment. Cys residues of the catalytic triad (Cys, His and Asp) is highlighted in green, and other conserved cysteines are highlighted in yellow. Asp, Glu (characteristic of the sub-branches “D” and “E” of cluster 2) and Asn residues of cluster 1 of the phylogenetic tree in <a href="#plants-13-02165-f001" class="html-fig">Figure 1</a> are highlighted in red, cyan and magenta, respectively.</p>
Full article ">Figure 7
<p>Predicted 3D model of catalytic triad (Cys in yellow, His in lilac and Asp in red) and residues potentially involved in PCS post-translational regulation (Thr in orange, Arg in violet, Glu in dark red and Asp in red) (see text). Alternative residues occupying the same positions are Asn (in pink), Lys (in lilac) and Ala (in green). (<b>a</b>) <span class="html-italic">Scytonema</span> sp. WP_073628507.1 “N” isoform; (<b>b</b>) <span class="html-italic">Scytonema</span> sp. WP_155743291.1 “E” isoform; (<b>c</b>) <span class="html-italic">T. pseudonana</span> AGE13359.1 “N” isoform; (<b>d</b>) <span class="html-italic">T. pseudonana</span> AGE13358.1 “D” isoform; (<b>e</b>) <span class="html-italic">T. pseudonana</span> Thaps_257216 “E” isoform. Models were generated by using Alphafold.</p>
Full article ">Figure 8
<p>Predicted 3D model showing the different arrangements in the four conserved cysteine residues (represented in ball and stick) described in <span class="html-italic">At</span>PCS1 (highlighted in yellow in <a href="#plants-13-02165-f007" class="html-fig">Figure 7</a>). (<b>a</b>) <span class="html-italic">Scytonema</span> sp. WP_073628507.1 “N” isoform; (<b>b</b>) <span class="html-italic">Scytonema</span> sp. WP_155743291.1 “E” isoform; (<b>c</b>) <span class="html-italic">T. pseudonana</span> AGE13359.1 “N” isoform; (<b>d</b>) <span class="html-italic">T. pseudonana</span> AGE13358.1 “D” isoform; (<b>e</b>) <span class="html-italic">T. pseudonana</span> Thaps_257216 “E” isoform. Models were generated by using Alphafold.</p>
Full article ">Figure 9
<p>Phylogenetic tree showing 38 protein sequences of PCSs from 6 cyanobacteria and 15 eukaryotic algae belonging to different taxa and chosen as models (<a href="#plants-13-02165-t001" class="html-table">Table 1</a>). The tree was constructed by MEGA11 with the ML method, and the bootstrap consensus tree was generated with 1000 replicates. The bootstrap percentage is represented by circles on each branch. The block diagram representation of the most conserved 15 motifs in the PCS protein sequences obtained with the MEME tool [<a href="#B53-plants-13-02165" class="html-bibr">53</a>]. The catalytic domains are distributed closely together in the <span class="html-italic">N</span>-terminal domain. The less conserved <span class="html-italic">C</span>-terminal domain is present only in the “E” PCS sequences belonging to green algae and cyanobacteria closely related to them (the lowest sub-branch of the tree).</p>
Full article ">
15 pages, 1716 KiB  
Article
Aspergillus Fumigatus Spore Proteases Alter the Respiratory Mucosa Architecture and Facilitate Equine Herpesvirus 1 Infection
by Joren Portaels, Eline Van Crombrugge, Wim Van Den Broeck, Katrien Lagrou, Kathlyn Laval and Hans Nauwynck
Viruses 2024, 16(8), 1208; https://doi.org/10.3390/v16081208 - 27 Jul 2024
Cited by 1 | Viewed by 1025
Abstract
Numerous Aspergillus fumigatus (Af) airborne spores are inhaled daily by humans and animals due to their ubiquitous presence. The interaction between the spores and the respiratory epithelium, as well as its impact on the epithelial barrier function, remains largely unknown. The epithelial barrier [...] Read more.
Numerous Aspergillus fumigatus (Af) airborne spores are inhaled daily by humans and animals due to their ubiquitous presence. The interaction between the spores and the respiratory epithelium, as well as its impact on the epithelial barrier function, remains largely unknown. The epithelial barrier protects the respiratory epithelium against viral infections. However, it can be compromised by environmental contaminants such as pollen, thereby increasing susceptibility to respiratory viral infections, including alphaherpesvirus equine herpesvirus type 1 (EHV-1). To determine whether Af spores disrupt the epithelial integrity and enhance susceptibility to viral infections, equine respiratory mucosal ex vivo explants were pretreated with Af spore diffusate, followed by EHV-1 inoculation. Spore proteases were characterized by zymography and identified using mass spectrometry-based proteomics. Proteases of the serine protease, metalloprotease, and aspartic protease groups were identified. Morphological analysis of hematoxylin-eosin (HE)-stained sections of the explants revealed that Af spores induced the desquamation of epithelial cells and a significant increase in intercellular space at high and low concentrations, respectively. The increase in intercellular space in the epithelium caused by Af spore proteases correlated with an increase in EHV-1 infection. Together, our findings demonstrate that Af spore proteases disrupt epithelial integrity, potentially leading to increased viral infection of the respiratory epithelium. Full article
(This article belongs to the Special Issue Animal Herpesvirus)
Show Figures

Figure 1

Figure 1
<p>Both <span class="html-italic">Aspergillus fumigatus</span> strains show proteolytic activities. (<b>A</b>) The zymography with gelatin substrate was stained with Coomassie blue, whereby the proteolytic activity by proteases appears as white bands on a blue background. The zymogram shows, from left to right, a protein ladder, a plasmin marker, the Af wild type strain, and the Af clinical isolate strain. (<b>B</b>) Plot profiling of the lanes of both <span class="html-italic">Aspergillus fumigatus</span> strains, made in imageJ on an 8-bit grey-scaled image. Plot shows value peaks (grey) at the corresponding white proteolytic bands (white), as indicated by red arrows.</p>
Full article ">Figure 2
<p>Protein (<b>left</b>) and proteolytic (<b>right</b>) profiles of the <span class="html-italic">Aspergillus fumigatus</span> clinical isolate spore diffusate. Three regions of the protein profile, corresponding to the three regions of proteolytic activity on the zymogram (A–C, red boxes), were excised of the SDS-PAGE and subjected to mass spectrometry (MS)-based proteomics. The zymography was adapted from <a href="#viruses-16-01208-f001" class="html-fig">Figure 1</a>A.</p>
Full article ">Figure 3
<p>The spore diffusate alters the epithelium integrity of respiratory horse explants. (<b>A</b>) The explants were treated with EGTA for 1 h (positive control), serum-free medium (negative control), and the spore diffusate of the Af clinical isolate at a dilution of 1:25 and 1:250 for 24 h. (<b>B</b>) The percentage of intercellular space, upon Af spore diffusate treatment at a 1:250 dilution, was determined by measuring the blank space in ImageJ software after manually determining the region of interest (ROI, the respiratory epithelium) and the threshold value. (<b>C</b>) The height of the epithelium after treatment was measured simultaneously using ImageJ. Data are represented as a mean of three replicates with standard deviation. Statistical significance is indicated by asterisks (** = <span class="html-italic">p</span>-value &lt; 0.01, *** = <span class="html-italic">p</span>-value &lt; 0.001 and non-significance (ns)). Scale bar measures 50 µm.</p>
Full article ">Figure 4
<p>EHV-1 infection of the respiratory epithelium after spore diffusate treatment. (<b>A</b>) Explants were treated with SFM (negative control), EGTA treatment (positive control), or the spore diffusate of the clinical isolate <span class="html-italic">Aspergillus fumigatus</span> at a dilution of 1:250 for 24 h. Cryosections were made and stained for EHV-1 late glycoproteins to assess the EHV-1 infection by counting the number of new plaques in 50 consequent cryosections. (<b>B</b>) The maximum plaque latitude of each new EHV-1 plaque was also determined. Data are represented as a mean of three replicates with standard deviation. No statistical significance was observed, indicated by non-significance (ns). (<b>C</b>) Representative confocal images of EHV-1 plaques for the different pretreated mucosal explants are shown. EHV-1 late glycoprotein was stained in green, basement membrane in red, and nuclei in blue. Scale bar measures 100 µm.</p>
Full article ">
13 pages, 3419 KiB  
Article
Preparation, Characterization and Stability of Calcium-Binding Peptides Derived from Chicken Blood
by Jing Yang, Jing Shi, Ying Zhou, Ye Zou, Weimin Xu, Xiudong Xia and Daoying Wang
Foods 2024, 13(15), 2368; https://doi.org/10.3390/foods13152368 - 26 Jul 2024
Cited by 2 | Viewed by 915
Abstract
Calcium-binding peptides have gained significant attention due to their potential applications in various fields. In this study, we aimed to prepare, characterize, and evaluate the stability of calcium-binding peptides derived from chicken blood. Chicken hemoglobin peptides (CPs) were obtained by protease hydrolysis and [...] Read more.
Calcium-binding peptides have gained significant attention due to their potential applications in various fields. In this study, we aimed to prepare, characterize, and evaluate the stability of calcium-binding peptides derived from chicken blood. Chicken hemoglobin peptides (CPs) were obtained by protease hydrolysis and were applied to prepare chicken hemoglobin peptide–calcium chelate (CP-Ca). The preparation conditions were optimized, and the characteristics and stability of CP-Ca were analyzed. The optimal chelating conditions were determined by single-factor and response surface tests, and the maximum calcium ion chelating rate was 77.54%. Amino acid analysis indicated that glutamic acid and aspartic acid motifs played an important role in the chelation of the calcium ions and CP. According to the characterization analysis, CP-Ca was a different substance compared with CP; calcium ions chelated CPs via the sites of carbonyl oxygen, carboxyl oxygen, and amino nitrogen groups; and after the chelation, the structure changed from a smooth homogeneous plate to compact granular. The stability analysis showed that CP-Ca was stable at different temperatures, pH, and gastrointestinal conditions. The study indicates that chicken blood is a promising source of peptide–calcium chelates, providing a theoretical basis for application in functional foods and improving the utilization value of chicken blood. Full article
Show Figures

Figure 1

Figure 1
<p>Effects of mass ratio of peptides/CaCl<sub>2</sub> for (<b>A</b>), time (<b>B</b>), temperature (<b>C</b>), and pH (<b>D</b>) on calcium-binding capacity. The different letters mean the significant difference within the groups (<span class="html-italic">p</span> &lt; 0.05).</p>
Full article ">Figure 2
<p>The characterization of CP-Ca. (<b>A</b>) Ultraviolet spectra of CP and CP-Ca. (<b>B</b>) Fluorescence spectra of CP and CP-Ca. (<b>C</b>) Fourier-transform infrared (FT-IR) spectra of CP and CP-Ca in the region from 4000 to 500 cm<sup>−1</sup>. (<b>D</b>) X-ray diffraction pattern of CP and CP-Ca. (<b>E</b>) Scanning electron microscope images of CP and CP-Ca at ×1500 magnification. (<b>F</b>) Typical TG-DSC thermograms of CP and CP-Ca.</p>
Full article ">Figure 3
<p>Stability of CP-Ca. (<b>A</b>) pH stability. (<b>B</b>) Temperature stability. (<b>C</b>) Simulated digestion <span class="html-italic">in vitro</span>. Different lowercase letters at the top of the bars indicate a significant difference in values among samples (<span class="html-italic">p</span> &lt; 0.05).</p>
Full article ">
12 pages, 2061 KiB  
Article
The Role of Erythropoietin in Bovine Sperm Physiology
by Vasiliki G. Sapanidou, Byron Asimakopoulos, Theodoros Lialiaris, Sophia N. Lavrentiadou, Konstantinos Feidantsis, Georgios Kourousekos and Maria P. Tsantarliotou
Animals 2024, 14(15), 2175; https://doi.org/10.3390/ani14152175 - 26 Jul 2024
Cited by 1 | Viewed by 696
Abstract
Erythropoietin (EPO), a hormone secreted mainly by the kidney, exerts its biological function by binding to its cell-surface receptor (EpoR). The presence of EPO and EpoR in the male and female reproductive system has been verified. Therefore, some of the key properties of [...] Read more.
Erythropoietin (EPO), a hormone secreted mainly by the kidney, exerts its biological function by binding to its cell-surface receptor (EpoR). The presence of EPO and EpoR in the male and female reproductive system has been verified. Therefore, some of the key properties of EPO, such as its antioxidant and antiapoptotic effects, could improve the fertilizing capacity of spermatozoa. In the present study, the effect of two different concentrations of EPO (10 mIU/μL and 100 mIU/μL) on bovine sperm-quality parameters was evaluated during a post-thawing 4-h incubation at 37 °C. EPO had a positive effect on sperm motility, viability, and total antioxidant capacity. Moreover, EPO inhibited apoptosis, as it reduced both BCL2-associated X apoptosis regulator (Bax)/B-cell lymphoma 2 (Bcl-2) ratio and cleaved cysteine-aspartic proteases (caspases) substrate levels in a dose-dependent manner. In addition, EPO induced sperm capacitation and acrosome reaction in spermatozoa incubated in capacitation conditioned medeia. These results establish a foundation for the physiological role of EPO in reproductive processes and hopefully will provide an incentive for further research in order to fully decipher the role of EPO in sperm physiology and reproduction. Full article
(This article belongs to the Section Animal Physiology)
Show Figures

Figure 1

Figure 1
<p>The effect of EPO (10 mΙU/μL and 100 mIU/μL) supplementation on the percentage of live spermatozoa with intact acrosome during a 240 min incubation. Data are presented as mean ± SD. Asterisk (*) depicts statistically significant difference between the control and the treated groups for each given time point (<span class="html-italic">p</span> &lt; 0.05, <span class="html-italic">n</span> = 6).</p>
Full article ">Figure 2
<p>Intracellular superoxide production of frozen/thawed bovine spermatozoa supplemented with two different concentrations of EPO (10 mΙU/μL and 100 mIU/μL) after 240 min incubation in the presence of a negative control, which was set as 100%. Data are presented as mean ± SD (<span class="html-italic">p</span> &lt; 0.05, <span class="html-italic">n</span> = 6).</p>
Full article ">Figure 3
<p>Total antioxidant capacity determined as % of reduced DPPH<sup>•</sup> in frozen/thawed bovine spermatozoa supplemented with two different concentrations of EPO (10 mΙU/μL and 100 mIU/μL) after 240 min incubation in the presence of a negative control, which was set as 100%. Data are presented as mean ± SD. Asterisk (*) denotes statistically significant differences (<span class="html-italic">p</span> &lt; 0.05) compared to the control group (<span class="html-italic">p</span> &lt; 0.05, <span class="html-italic">n</span> = 6).</p>
Full article ">Figure 4
<p>The effect of EPO (10 mΙU/μL and 100 mIU/μL) supplementation on the percentage of live spermatozoa with acrosome reaction. Data are presented as mean ± SD. Asterisk (*) denotes statistically significant differences (<span class="html-italic">p</span> &lt; 0.05) compared to the control group, cross (+) denotes statistically significant differences between the two concentrations of EPO, while caret (^) indicates statistically significant differences compared to the positive control (heparin) (<span class="html-italic">p</span> &lt; 0.05, <span class="html-italic">n</span> = 6).</p>
Full article ">Figure 5
<p>Analysis of apoptotic markers Bax, Bcl-2 (normalized to β-actin), and cleaved caspases, as shown by representative blots (<b>A</b>). Bax, Bcl-2 (<b>B</b>), Bax/Bcl-2 ratio (<b>C</b>), and cleaved caspase (<b>D</b>) levels of frozen/thawed bovine spermatozoa under the effect of 10 mIU/μL (ΕPO10) and 100 mIU/μL (ΕPO100) EPO. Spermatozoa extracts from control, Epo 10, and Epo 100 groups were immunoblotted for Bax, Bcl-2, cleaved caspases, and β-actin to verify equal loading. Asterisk (*) denotes statistically significant differences (<span class="html-italic">p</span> &lt; 0.05) compared to the control group, while cross (+) denotes statistically significant differences (<span class="html-italic">p</span> &lt; 0.05) between the two concentrations of EPO (Panels B, C, and D). Values constitute means ± SD (<span class="html-italic">n</span> = 6).</p>
Full article ">
Back to TopTop