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

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

Years

Between: -

Subjects

remove_circle_outline
remove_circle_outline
remove_circle_outline
remove_circle_outline
remove_circle_outline
remove_circle_outline
remove_circle_outline
remove_circle_outline
remove_circle_outline

Journals

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

Article Types

Countries / Regions

remove_circle_outline
remove_circle_outline
remove_circle_outline
remove_circle_outline

Search Results (2,308)

Search Parameters:
Keywords = helix

Order results
Result details
Results per page
Select all
Export citation of selected articles as:
19 pages, 3632 KiB  
Article
The Effect of Cellulose Nanocrystals on the Molecular Organization, Thermomechanical, and Shape Memory Properties of Gelatin-Matrix Composite Films
by Cristina Padilla, Marzena Pępczyńska, Cristian Vizueta, Franck Quero, Paulo Díaz-Calderón, William Macnaughtan, Tim Foster and Javier Enrione
Gels 2024, 10(12), 766; https://doi.org/10.3390/gels10120766 (registering DOI) - 25 Nov 2024
Viewed by 156
Abstract
Gelatin is a natural hydrocolloid with excellent film-forming properties, high processability, and tremendous potential in the field of edible coatings and food packaging. However, its reinforcing by materials such as cellulose nanocrystals (CNC) is often necessary to improve its mechanical behavior, including shape [...] Read more.
Gelatin is a natural hydrocolloid with excellent film-forming properties, high processability, and tremendous potential in the field of edible coatings and food packaging. However, its reinforcing by materials such as cellulose nanocrystals (CNC) is often necessary to improve its mechanical behavior, including shape memory properties. Since the interaction between these polymers is complex and its mechanism still remains unclear, this work aimed to study the effect of low concentrations of CNC (2, 6, and 10 weight%) on the molecular organization, thermomechanical, and shape memory properties in mammalian gelatin-based composite films at low moisture content (~10 weight% dry base). The results showed that the presence of CNCs (with type I and type II crystals) interfered with the formation of the gelatin triple helix, with a decrease from 21.7% crystallinity to 12% in samples with 10% CNC but increasing the overall crystallinity (from 21.7% to 22.6% in samples with 10% CNC), which produced a decrease in the water monolayer in the composites. These changes in crystallinity also impacted significantly their mechanical properties, with higher E’ values (from 1 × 104 to 1.3 × 104 Pa at 20 °C) and improved thermal stability at higher CNC content. Additionally, the evaluation of their shape memory properties indicated that while molecular interactions between the two components occur, CNCs negatively impacted the magnitude and kinetics of the shape recovery of the composites (more particularly at 10 weight% CNC, reducing shape recovery from 90% to 70%) by reducing the netting point associated with the lower crystallinity of the gelatin. We believe that our results contribute in elucidating the interactions of gelatin–CNC composites at various structural levels and highlights that even though CNC acts as a reinforcement material on gelatin matrices, their interaction are complex and do not imply synergism in their properties. Further investigation is, however, needed to understand CNC–gelatin interfacial interactions with the aim of modulating their interactions depending on their desired application. Full article
(This article belongs to the Special Issue Design and Development of Gelatin-Based Materials)
Show Figures

Graphical abstract

Graphical abstract
Full article ">Figure 1
<p>The powder X-ray diffraction pattern showing the characteristic diffraction planes of CNCs.</p>
Full article ">Figure 2
<p>FT-IR spectra of gelatin, CNC, and gelatin–CNC composite films equilibrated to 33% RH, ~10 wt.% water content d.b. The highlighted absorption peak located at 1055 cm<sup>−1</sup> is related to the vibrational motions of the C-O stretching of primary alcohols that belong to the molecular structure of cellulose.</p>
Full article ">Figure 3
<p>Gelatin’s and CNCs’ surface distribution at the surface of gelatin–CNC composite films equilibrated to 33% RH, ~10wt.% water content d.b. Raman images show gelatin’s (green) and CNCs’ (red) distribution at the surface of composite films. (<b>A</b>) BG 2 wt.% CNC; (<b>B</b>) BG 6 wt.% CNC; (<b>C</b>) BG 10 wt.% CNC.</p>
Full article ">Figure 4
<p>X-ray diffraction patterns of gelatin, CNC, and gelatin–CNC composite films equilibrated to 33% RH, ~10 wt.% water content d.b with various CNC weight percentages.</p>
Full article ">Figure 5
<p>Sorption isotherms at 20 °C of gelatin, CNC, and gelatin–CNC composite films with different CNC weight fractions. Lines in the graphs represent the fitting to the experimental data using the GAB equation. The inset shows isotherms at low moisture content.</p>
Full article ">Figure 6
<p>DSC thermograms (first heating scans) of gelatin and gelatin–CNC composite films with different CNC weight fractions equilibrated under 33% RH, ~10wt.% water content d.b. Black arrows indicate the glass transition temperature (Tg) shift with the presence of CNC.</p>
Full article ">Figure 7
<p>Storage modulus (E’) of gelatin and gelatin–CNC composite films with different CNC weight fractions equilibrated at RH 33% at 20 and 50 °C. Data from 4 independent experiments.</p>
Full article ">Figure 8
<p>Thermograms of gelatin, CNC (inset), and gelatin–CNC composite films with various CNC weight percentages equilibrated at RH 33%.</p>
Full article ">Figure 9
<p>The percentage of the shape recovery of gelatin and gelatin–CNC composite films with different CNC weight fractions at 30 °C as a function of time.</p>
Full article ">
14 pages, 3567 KiB  
Article
Effect of Adding Konjac Glucomannan on the Physicochemical Properties of Indica Rice Flour and the Quality of Its Product of Instant Dry Rice Noodles
by Chunmiao Lu, Ying Yang, Xin Zhao, Zhiyu Liu, Xiaoyan Liao, Yingying Zhang, Dailin Wu, Jing Li and Jiangtao Li
Foods 2024, 13(23), 3749; https://doi.org/10.3390/foods13233749 - 22 Nov 2024
Viewed by 249
Abstract
Instant dry rice noodles have a broad market prospect due to their advantages of long shelf life, convenient transportation, and convenient eating, but there are still quality problems such as long rehydration times and poor eating quality. In order to improve the quality [...] Read more.
Instant dry rice noodles have a broad market prospect due to their advantages of long shelf life, convenient transportation, and convenient eating, but there are still quality problems such as long rehydration times and poor eating quality. In order to improve the quality of instant dry rice noodles, the effects of konjac glucomannan (KGM) on the gelatinization characteristics, pasting properties, and rheological properties of Indica rice flour and the structure, food quality, and starch digestibility of instant dry rice noodles made of Indica rice flour were studied. The results showed that the starch gelatinization conclusion temperature and endothermic enthalpy of Indica rice flour were reduced by adding ≤ 3% KGM, the peak viscosity, valley viscosity, final viscosity, and setback value of Indica rice flour in the pasting process decreased with the increase in the KGM addition amount, and the pseudoplasticity, viscosity, and elasticity of Indica rice flour paste were reduced by adding > 1% KGM. When the KGM addition amount was 2%, the endothermic enthalpy, final viscosity, and setback value of Indica rice flour were 2.74 J/g, 2379.5 cp, and 961.5 cp, respectively. The instant dry rice noodles made of Indica rice flour had a looser microstructure after adding KGM, and its short-range ordered structure and double helix content were reduced by adding 1~3% KGM. When the KGM addition amount was 2%, the rehydration time of instant dry rice noodles was 290 s, which was shortened by 14.7%, while the texture and sensory quality remained unchanged, and the SDS content was reduced by 16.4% while the RS content was increased by 28.8%. Therefore, the physicochemical properties of Indica rice flour and the quality of its instant dry rice noodles can be improved by adding an appropriate amount of KGM. This study can promote the application of KGM in improving the quality of rice products. Full article
(This article belongs to the Section Grain)
15 pages, 1912 KiB  
Article
Cell Wall-Mediated Antifungal Activity of the Aqueous Extract of Hedera helix L. Leaves Against Diplodia corticola
by Christina Crisóstomo, Luara Simões, Lillian Barros, Tiane C. Finimundy, Ana Cunha and Rui Oliveira
Antibiotics 2024, 13(12), 1116; https://doi.org/10.3390/antibiotics13121116 - 22 Nov 2024
Viewed by 266
Abstract
Background/Objectives: Cork oak forests have been declining due to fungal pathogens such as Diplodia corticola. However, the preventive fungicides against this fungus have restricted use due to the deleterious effects on human health and the environment, prompting the need for sustainable alternatives. [...] Read more.
Background/Objectives: Cork oak forests have been declining due to fungal pathogens such as Diplodia corticola. However, the preventive fungicides against this fungus have restricted use due to the deleterious effects on human health and the environment, prompting the need for sustainable alternatives. Here, we describe the antifungal activity of an aqueous extract of Hedera helix L. leaves (HAE) against D. corticola and the possible mechanism of action. Results/Methods: The chemical analysis revealed compounds like the saponin hederacoside C, quinic acid, 5-O-caffeoylquinic acid, rutin, and glycoside derivatives of quercetin and kaempferol, all of which have been previously reported to possess antimicrobial activity. Remarkable in vitro antifungal activity was observed, reducing radial mycelial growth by 70% after 3 days of inoculation. Saccharomyces cerevisiae mutants, bck1 and mkk1/mkk2, affected the cell wall integrity signaling pathway were more resistant to HAE than the wild-type strain, suggesting that the extract targets kinases of the signaling pathway, which triggers toxicity. The viability under osmotic stress with 0.75 M NaCl was lower in the presence of HAE, suggesting the deficiency of osmotic protection by the cell wall. Conclusions: These results suggest that ivy extracts can be a source of new natural antifungal agents targeting the cell wall, opening the possibility of preventing fungal infections in cork oaks and improving the cork production sector using safer and more sustainable approaches. Full article
(This article belongs to the Section Plant-Derived Antibiotics)
Show Figures

Figure 1

Figure 1
<p>Percentage of growth inhibition of <span class="html-italic">D. corticola</span> by <span class="html-italic">H. helix</span> aqueous extract (HAE). Radial growth of <span class="html-italic">D. corticola</span> mycelium was measured in Petri dishes with PDA medium containing 50, 100, 500, 1000, or 1500 µg/mL of HAE. Representative images of HAE antifungal activity against <span class="html-italic">D. corticola</span> after 6 days of incubation (<b>A</b>). In the negative control (C-), the highest volume of extract used on the assays was replaced by sterilized deionized water (extract solvent). The diameter was measured after 3 (black bars; (<b>B</b>)) and 6 (grey bars; (<b>B</b>)) days of incubation at 25 °C in the dark. Each bar represents the mean ± SD of three independent experiments. A two-way ANOVA was used for the analysis and different letters represent statistical significance. Capital letters were used for comparisons of extract concentrations within the same time of incubation and lowercase letters were used for comparisons of different times of incubation for each concentration tested.</p>
Full article ">Figure 2
<p>Viability of <span class="html-italic">S. cerevisiae</span> BY4741 (<b>A</b>) and the mutant strains <span class="html-italic">erg2</span> (<b>B</b>), <span class="html-italic">bck1</span> (<b>C</b>), and <span class="html-italic">mkk1/mkk2 (</span><b>D</b>) in the presence of aqueous extract of <span class="html-italic">H. helix</span> (HAE). Cells from exponentially growing cultures were exposed to 10 µg/mL (white squares), 50 µg/mL (black triangles), 75 µg/mL (white inverted triangles), 100 µg/mL (black diamonds), or 250 µg/mL (white circles) of HAE and incubated at 30 °C, 200 rpm. Viability was assessed using CFU after 0, 30, 60, and 90 min of incubation upon spreading 10<sup>−4</sup> dilutions on YPDA plates and incubation at 30 °C, 200 rpm. The negative control (black circles) was prepared by replacing the extract by the solvent at the highest volume of extract used in the assays. The data represent the mean ± SD of three independent experiments. A one-way ANOVA and Dunnett’s post hoc test were used for the analysis, and concentrations were compared at each time-point. For clarity of graphical representation, significance is shown only for the 90 min timepoint, where ** means 0.001 &lt; <span class="html-italic">p</span> ≤ 0.01 and **** means <span class="html-italic">p</span> &lt; 0.0001; the absence of significance is not marked with any symbol. Other relevant significances are presented in the text.</p>
Full article ">Figure 3
<p>Viability of <span class="html-italic">S. cerevisiae</span> BY4741 in the presence of aqueous extract of <span class="html-italic">H. helix</span> (HAE) and osmotic stress. Cells from exponentially growing cultures were exposed to 10 µg/mL (grey bars) or 50 µg/mL (dark grey bars) of HAE at 30 °C, 200 rpm, for 30 min. The viability was assessed using CFU by plating 10<sup>−4</sup> dilutions on YPDA plates containing different concentrations of NaCl (0, 0.25, 0.5, or 0.75 M) and further incubation at 30 °C for 48 h. The negative control (black bars) was prepared by replacing the extract with the solvent at the highest volume of extract used in the assays. The data represent the mean ± SD of three independent experiments. A two-way ANOVA and Tukey’s post hoc test were used for the analysis. A letter code was used. Lowercase letters are used for comparisons of extract concentration effects within each salt concentration (<span class="html-italic">p</span> &lt; 0.0001) and capital letters are used for comparisons between salt concentrations for each extract concentration (<span class="html-italic">p</span> &lt; 0.01). Mean values followed by the same letters are not statistically different.</p>
Full article ">Figure 4
<p>Percentage of growth inhibition of <span class="html-italic">D. corticola</span> by <span class="html-italic">H. helix</span> aqueous extract (HAE) in the absence and presence of osmotic stress. Radial growth of <span class="html-italic">D. corticola</span> mycelium was measured in Petri dishes with PDA medium containing 50, 500, or 1500 µg/mL HAE in the absence or presence of 0.4 M of NaCl. Representative images of HAE antifungal activity against <span class="html-italic">D. corticola</span> after 6 days of incubation (<b>A</b>). In the negative control (C-), the highest volume of extract used on the assays was replaced by sterilized deionized water (extract solvent). The diameter was measured after 6 days of incubation at 25 °C in the dark and the percentage of growth inhibition was calculated, taking C- as a reference (<b>B</b>). Each bar represents the mean ± SD of three independent experiments. A one-way ANOVA and Dunnett’s post hoc test were used for the analysis, where *** means 0.0001 &lt; <span class="html-italic">p</span> ≤ 0.001 and **** means <span class="html-italic">p</span> &lt; 0.0001.</p>
Full article ">
14 pages, 3326 KiB  
Article
The bHLH Transcription Factor PubHLH66 Improves Salt Tolerance in Daqing Poplar (Populus ussuriensis)
by Dandan Li, Jindan Wang, Yuxin Pan, Hui Wang, Xinyao Dang, Shihao Zhao and Yucheng Wang
Forests 2024, 15(11), 2051; https://doi.org/10.3390/f15112051 - 20 Nov 2024
Viewed by 295
Abstract
Elevated salinity negatively impacts plant growth and yield, presenting substantial challenges to agricultural and forestry productivity. The bHLH transcription factor family is vital for plants to cope with various abiotic stresses. However, it remains uncertain whether bHLH transcription factors can regulate salt stress [...] Read more.
Elevated salinity negatively impacts plant growth and yield, presenting substantial challenges to agricultural and forestry productivity. The bHLH transcription factor family is vital for plants to cope with various abiotic stresses. However, it remains uncertain whether bHLH transcription factors can regulate salt stress in Populus ussuriensis. In the following study, a salt-induced bHLH transcription factor PubHLH66 was identified from P. ussuriensis. PubHLH66 has a typical and conserved bHLH domain. Subcellular localization and yeast two-hybrid (Y2H) assays confirmed that it is a nucleus-localized transactivator and the activation region is located at the N-terminus. PubHLH66-OE and PubHLH66-SRDX transgenic P. ussuriensis were obtained through Agrobacterium-mediated leaf disc transformation. Morphological and physiological results demonstrated that PubHLH66-OE enhanced salinity tolerance, as indicated by reduced electrolyte leakage (EL), malondialdehyde (MDA), and H2O2 levels, along with increased proline contents and activities of peroxidase (POD) and superoxide dismutase (SOD). In contrast, PuHLH66-SRDX poplar showed decreased salt tolerance. Quantitative real-time PCR (RT-qPCR) confirmed that PubHLH66 enhanced salt tolerance by regulating the expression of genes such as PuSOD, PuPOD, and PuP5CS, resulting in reduced reactive oxygen species (ROS) accumulation and an improved osmotic potential. Thus, PubHLH66 could be a candidate gene for molecular breeding to enhance salt tolerance in plants. These results laid a foundation for exploring the mechanisms of salt tolerance in P. ussuriensis, facilitating the development of more salt-tolerant trees to combat the increasing issue of soil salinization globally. Full article
(This article belongs to the Special Issue Abiotic and Biotic Stress Responses in Trees Species)
Show Figures

Figure 1

Figure 1
<p>Sequence analysis and gene expression pattern of <span class="html-italic">PubHLH66</span>. (<b>A</b>) Multiple sequence alignments and analysis of PubHLH66 with homologous bHLH proteins from different plant species. The accession numbers corresponding to these proteins in NCBI ID are OsbHLH66 (XP_015627343), AtbHLH66 (BAD44153), MdbHLH66 (XP_028964276), and CtbHLH66 (XP_006473971). (<b>B</b>) Analysis of the phylogenetic tree constructed using the neighbor-joining (NJ) method, with a bootstrap test performed using 1000 iterations in MEGA. The black dot represents PubHLH66. The scale bar represents 0.1 substitutions per site. (<b>C</b>) Relative expression level of <span class="html-italic">PubHLH66</span> in the root, stem and leaves of plants under 150 mM NaCl stress determined using RT-qPCR. Error bars represent the variability among three biological replicates. The <span class="html-italic">x</span>-axis represents the time points following treatment with 150 mM NaCl.</p>
Full article ">Figure 2
<p>Subcellular localization and transactivation activity of PubHLH66. (<b>A</b>) Subcellular localization of PubHLH66. The 35S::GFP (control) and 35S::PubHLH66-GFP translational fusion constructs were transiently introduced into onion epidermal cells, and DAPI was utilized as a marker for the nucleus. Bar = 50 μm. (<b>B</b>) pGBKT7-PubHLH66, pGBKT7-PubHLH66<sup>N241</sup>, pGBKT7-PubHLH66<sup>C214</sup>, and pGBKT7 (negative control) were transformed in the Y2H Gold yeast strain. Yeast transformants were cultured in either SD/-Trp or SD/-Trp/-His/-Ade/X-α-Gal media. LacZ activity was measured in the presence of X-α-Gal with pGBKT7. The gray bars represent BD, and blue bars represent gene segment of PubHLH66.</p>
Full article ">Figure 3
<p>Salt tolerance analysis of the <span class="html-italic">PubHLH66</span> transgenic <span class="html-italic">P. ussuriensis</span>. (<b>A</b>) Phenotype of <span class="html-italic">PubHLH66</span> transgenic and WT poplars during salinity treatments. Bars, 10 cm. (<b>B</b>) Survival rates, (<b>C</b>) dry weights, (<b>D</b>) chlorophyll contents, (<b>E</b>) electrolyte leakage (EL), (<b>F</b>) malondialdehyde (MDA) contents, and (<b>G</b>) proline contents of the poplars after growth under normal and NaCl stress conditions for 7 d. (<b>H</b>) The expression pattern of <span class="html-italic">PuP5CS1</span> and <span class="html-italic">PuP5CS2</span>. WT and <span class="html-italic">PubHLH66</span> transgenic <span class="html-italic">P. ussuriensis</span> were subjected to 150 mM NaCl for 24 h. The WT line was used as a control and set to 1. The <span class="html-italic">PuActin</span> gene served as a housekeeping gene. Asterisks (* <span class="html-italic">p</span> &lt; 0.05 and ** <span class="html-italic">p</span> &lt; 0.01) indicate significant differences determined using Student’s <span class="html-italic">t</span>-test compared to WT plants.</p>
Full article ">Figure 4
<p>Analysis of the antioxidant capacity of <span class="html-italic">PubHLH66</span>. (<b>A</b>,<b>B</b>) DAB and NBT staining. Poplar leaves subjected to NaCl treatment were infiltrated with DAB (<b>A</b>) for hydrogen peroxide (H<sub>2</sub>O<sub>2</sub>) detection and with NBT (<b>B</b>) for superoxide (O<sub>2</sub><sup>−</sup>) detection. (<b>C</b>) H<sub>2</sub>O<sub>2</sub> content assay (<b>D</b>,<b>E</b>) measurement of POD and SOD activities in the poplars after growth under normal and NaCl stress conditions for 7 d. Data are presented as the means and SDs of three independent experiments. Asterisks (*) represent <span class="html-italic">p</span> &lt; 0.05 and (**) <span class="html-italic">p</span> &lt; 0.01.</p>
Full article ">Figure 5
<p>PubHLH66 regulates the expression of <span class="html-italic">PuPOD</span>s and <span class="html-italic">PuSOD</span>s. (<b>A</b>,<b>B</b>) The expression pattern of <span class="html-italic">PuPOD</span>s, (<b>C</b>,<b>D</b>) the expression pattern of <span class="html-italic">PuSOD</span> genes after exposure to 150 mM NaCl for 24 h. Wild type (WT) plants were utilized as a control and normalized to 1, while the <span class="html-italic">PuActin</span> gene served as the internal control. The asterisk (*) represents <span class="html-italic">p</span> &lt; 0.05.</p>
Full article ">
15 pages, 5198 KiB  
Article
Study on the Structural Changes of Boneless Chicken Claw Collagen and Its Effect on Water Retention Performance
by Zheng Tang, Yiguo He, Jing Zhang, Zhifeng Zhao, Yiming Nie and Xingxiu Zhao
Foods 2024, 13(22), 3682; https://doi.org/10.3390/foods13223682 - 19 Nov 2024
Viewed by 427
Abstract
The purpose of this study was to explore the water retention mechanism of chicken claws by detecting the structural changes in collagen in boneless chicken claws under different expansion rates. Firstly, boneless chicken claw collagen with different expansion rates (0%, 10%, 20%, 30%, [...] Read more.
The purpose of this study was to explore the water retention mechanism of chicken claws by detecting the structural changes in collagen in boneless chicken claws under different expansion rates. Firstly, boneless chicken claw collagen with different expansion rates (0%, 10%, 20%, 30%, 40%, 50%) was extracted by the acid–enzyme complex method, and the changes in collagen were determined by scanning electron microscopy (SEM), ultraviolet spectroscopy (UV), Fourier transform infrared spectroscopy (FTIR), circular dichroism (CD), low-field nuclear magnetic resonance LF-NMR) and surface hydrophobicity to explore the mechanism that leads to changes in the water retention performance. The results of scanning electron microscopy showed that with the increase in the expansion rate, collagen molecules showed curling, shrinking, breaking and crosslinking, forming a loose and irregular pore-like denatured collagen structure. UV analysis showed that the maximum absorption wavelength of chicken claw collagen was blue shifted under different expansion rates, and the maximum absorption peak intensity increased first and then decreased with the increase in expansion rate. The FTIR results showed that collagen had obvious characteristic absorption peaks in the amide A, B, I, II and III regions under different expansion rates, and that the intensity and position of the characteristic absorption peaks changed with the expansion rate. The results of the CD analysis showed that collagen at different expansion rates had obvious positive absorption peaks at 222 nm, and that the position of negative absorption peaks was red shifted with the increase in expansion rate. This shows that the expansion treatment makes the collagen of chicken claw partially denatured, and that the triple helix structure becomes relaxed or unwound, which provides more space for the combination of water molecules, thus enhancing the water absorption capacity of boneless chicken claw. The results of the surface hydrophobicity test showed that the surface hydrophobicity of boneless chicken claw collagen increased with the increase in expansion rate and reached the maximum at a 30% expansion rate, and then decreased with the further increase in the expansion rate. The results of LF-NMR showed that the water content of boneless chicken claws increased significantly after the expansion treatment, and that the water retention performance of chicken claws was further enhanced with the increase in the expansion rate. In this study, boneless chicken claws were used as raw materials, and the expansion process of boneless chicken claws was optimized by acid combined with a water-retaining agent, which improved the expansion rate of boneless chicken claws and the quality of boneless chicken claws. The effects of the swelling degree on the collagen structure, water absorption and water retention properties of boneless chicken claws were revealed by structural characterization. These findings explain the changes in the water retention of boneless chicken claws after expansion. By optimizing the expansion treatment process, the water retention performance and market added value of chicken feet products can be significantly improved, which is of great economic significance. Full article
Show Figures

Figure 1

Figure 1
<p>Collagen extraction process.</p>
Full article ">Figure 2
<p>Scanning electron microscope images of chicken claw collagen under different expansion rates. Note: (<b>A1</b>–<b>F1</b>): Boneless chicken claw collagen (×1000); (<b>A2</b>–<b>F2</b>): boneless chicken claw collagen (×500); among them, (<b>A1</b>,<b>A2</b>): 0%; (<b>B1</b>,<b>B2</b>): 10%; (<b>C1</b>,<b>C2</b>): 20%; (<b>D1</b>,<b>D2</b>): 30%; (<b>E1</b>,<b>E2</b>): 40%; (<b>F1</b>,<b>F2</b>): 50%.</p>
Full article ">Figure 2 Cont.
<p>Scanning electron microscope images of chicken claw collagen under different expansion rates. Note: (<b>A1</b>–<b>F1</b>): Boneless chicken claw collagen (×1000); (<b>A2</b>–<b>F2</b>): boneless chicken claw collagen (×500); among them, (<b>A1</b>,<b>A2</b>): 0%; (<b>B1</b>,<b>B2</b>): 10%; (<b>C1</b>,<b>C2</b>): 20%; (<b>D1</b>,<b>D2</b>): 30%; (<b>E1</b>,<b>E2</b>): 40%; (<b>F1</b>,<b>F2</b>): 50%.</p>
Full article ">Figure 3
<p>UV–visible spectra of collagen under different expansion rates.</p>
Full article ">Figure 4
<p>Infrared spectra of collagen under different expansion rates.</p>
Full article ">Figure 5
<p>CD spectra of collagen under different expansion rates.</p>
Full article ">Figure 6
<p>Surface hydrophobicity of collagen under different expansion rates. Note: ** means extremely significant (<span class="html-italic">p</span> &lt; 0.01), ns means not significant (<span class="html-italic">p</span> &gt; 0.05).</p>
Full article ">Figure 7
<p>Moisture migration of chicken claws under different expansion rates.</p>
Full article ">
25 pages, 4113 KiB  
Article
Rabbit and Human Angiotensin-Converting Enzyme-2: Structure and Electric Properties
by Svetlana H. Hristova, Trifon T. Popov and Alexandar M. Zhivkov
Int. J. Mol. Sci. 2024, 25(22), 12393; https://doi.org/10.3390/ijms252212393 - 19 Nov 2024
Viewed by 352
Abstract
The angiotensin-converting enzyme-2 (ACE2) is a transmembrane glycoprotein, consisting of two segments: a large carboxypeptidase catalytic domain and a small transmembrane collectrin-like segment. This protein plays an essential role in blood pressure regulation, transforming the peptides angiotensin-I and angiotensin-II (vasoconstrictors) into angiotensin-1-9 and [...] Read more.
The angiotensin-converting enzyme-2 (ACE2) is a transmembrane glycoprotein, consisting of two segments: a large carboxypeptidase catalytic domain and a small transmembrane collectrin-like segment. This protein plays an essential role in blood pressure regulation, transforming the peptides angiotensin-I and angiotensin-II (vasoconstrictors) into angiotensin-1-9 and angiotensin-1-7 (vasodilators). During the COVID-19 pandemic, ACE2 became best known as the receptor of the S-protein of SARS-CoV-2 coronavirus. The purpose of the following research is to reconstruct the 3D structure of the catalytic domain of the rabbit enzyme rACE2 using its primary amino acid sequence, and then to compare it with the human analog hACE2. For this purpose, we have calculated the electric properties and thermodynamic stability of the two protein globules employing computer programs for protein electrostatics. The analysis of the amino acid content and sequence demonstrates an 85% identity between the two polypeptide chains. The 3D alignment of the catalytic domains of the two enzymes shows coincidence of the α-helix segments, and a small difference in two unstructured segments of the chain. The electric charge of the catalytic domain of rACE2, determined by 70 positively chargeable amino acid residues, 114 negatively chargeable ones, and two positive charges of the Zn2+ atom in the active center exceeds that of hACE2 by one positively and four negatively chargeable groups; however, in 3D conformation, their isoelectric points pI 5.21 coincide. The surface electrostatic potential is similarly distributed on the surface of the two catalytic globules, but it strongly depends on the pH of the extracellular medium: it is almost positive at pH 5.0 but strongly negative at pH 7.4. The pH dependence of the electrostatic component of the free energy discloses that the 3D structure of the two enzymes is maximally stable at pH 6.5. The high similarity in the 3D structure, as well as in the electrostatic and thermodynamic properties, suggests that rabbit can be successfully used as an animal model to study blood pressure regulation and coronavirus infection, and the results can be extrapolated to humans. Full article
Show Figures

Figure 1

Figure 1
<p>Renin-angiotensin system. <span class="html-italic">Subsection</span> (<b>a</b>) on the left: Renin and angiotensin-converting enzyme (ACE) action. The precursor α-2-globulin angiotensinogen is produced by hepatocytes. The renal enzyme renin cleaves the covalent peptide bond after the first 10 amino acids from the N-terminus of angiotensinogen, leading to the formation of angiotensin-I (first reaction). This decapeptide is converted by the pulmonary angiotensin-converting enzyme (ACE) to the octapeptide angiotensin-II by the cleavage of the last two amino acid residues, resulting in the emergence of a carboxylic group on the C-terminus (second reaction). <span class="html-italic">Subsection</span> (<b>b</b>) on the right: Angiotensin-converting enzyme-2 (ACE2) action. The angiotensin-converting enzyme-2 (ACE2) cleaves one amino acid residue from the C-terminus of both angiotensin peptides, which leads to the formation of the nonapeptide angiotensin-1–9 and the heptapeptide angiotensin-1–7, respectively. The cleaved peptide bonds (in the two subsections) are shown by dotted lines colored according to the corresponding enzyme: blue (ACE) or red (ACE2). The amino acid residues are colored according to their charge and hydrophilicity: green (uncharged hydrophilic), blue (positively charged hydrophilic), red (negatively charged hydrophilic), and yellow (uncharged hydrophobic); the charges are determined at neutral pH. The end-side residues are marked by double color considering the protonated α-amino group (NH<sub>3</sub><sup>+</sup>–) on the N-terminus and the deprotonated carboxylic group (–COO<sup>−</sup>) on the C-terminus. The N-end aspartic acid (Asp) residue bears one positive (NH<sub>3</sub><sup>+</sup>–) and one negative (COO<sup>−</sup>–) charge. The C-end histidine (His) residue of the angiotensin-1-9 bears one negative charge (the deprotonated carboxyl group) and one positive charge at acid pH (the protonated imidazole group), which disappears at basic pH. The C-ends of the remaining three peptides have only negative charge (–COO<sup>−</sup>).</p>
Full article ">Figure 2
<p>Primary amino acid sequence of the polypeptide chains of human (the rows beginning with <b>h</b>) and rabbit (<b>r</b>) ACE2. The numbers at the end of every row indicate the first and the last amino acid residue on the corresponding row. The amino acid residues are denoted according to the standard one-letter code. The cells of the different residues are colored according to the electric charge at pH 7.0 and the hydrophilicity of the given residue: red (negatively charged hydrophilic), blue (positively charged hydrophilic), green (uncharged hydrophilic), and yellow (uncharged hydrophobic). The polypeptide chain is divided (indicated by vertical lines) into three segments: signal peptide (amino acid residues 1–18, colored in bright orange), catalytic domain (19–615, colored in bright gray), and transmembrane segment (616–805 colored in bright purple). The five residues included in the zinc-binding motif HEMGH of the active center of the enzyme are denoted by the red rectangle.</p>
Full article ">Figure 2 Cont.
<p>Primary amino acid sequence of the polypeptide chains of human (the rows beginning with <b>h</b>) and rabbit (<b>r</b>) ACE2. The numbers at the end of every row indicate the first and the last amino acid residue on the corresponding row. The amino acid residues are denoted according to the standard one-letter code. The cells of the different residues are colored according to the electric charge at pH 7.0 and the hydrophilicity of the given residue: red (negatively charged hydrophilic), blue (positively charged hydrophilic), green (uncharged hydrophilic), and yellow (uncharged hydrophobic). The polypeptide chain is divided (indicated by vertical lines) into three segments: signal peptide (amino acid residues 1–18, colored in bright orange), catalytic domain (19–615, colored in bright gray), and transmembrane segment (616–805 colored in bright purple). The five residues included in the zinc-binding motif HEMGH of the active center of the enzyme are denoted by the red rectangle.</p>
Full article ">Figure 3
<p>(<b>a</b>–<b>d</b>) 3D structure models of the human (red, hACE2) and rabbit (green, rACE2) catalytic domains of the angiotensin-converting enzyme-2 (the upper two models (<b>a</b>,<b>b</b>)); the α-helix segments are depicted as ribbon spirals. The model of the rabbit ACE2 (<b>b</b>) is reconstructed by replacement of the different amino acid residues in the hACE2 (PDB: 1r42) model (<b>a</b>). The violet spherical object (Zn<sup>2+</sup>) is the zinc atom in the enzyme active center. The low two models (<b>c</b>,<b>d</b>) present the aliment (hACE2+rACE2) of the human and rabbit ACE2 in two different projections: the 3D volume of the protein globules (the ribbon model on the left) (<b>c</b>) and the 2D surface of the globules (the atomic model on the right) (<b>d</b>). The right bottom model (<b>d</b>) presents the atoms exposed on the surface of the aligned two protein globules; the atoms are colored fully in red (hACE2) or green (rACE2) when they are entirely protruded above the others, or in mixed color when their coordinates partially coincide. The brightness, shade, and color nuance of the atomic images give the impression for a quasi 3D surface of the protein globules.</p>
Full article ">Figure 4
<p>pH dependences of the net electric charge <span class="html-italic">nz</span> + 2 of the globular catalytic domain of human (PDB: 1r42, hACE2, red curve 1) and reconstructed rabbit (rACE2, green curve 2) angiotensin-converting enzyme-2 in 3D conformation of the polypeptide chain in aqueous medium. The net charge of the two globular domains is the algebraic sum of the positive and negative coulombic charges of the polypeptide chain with the addition of two positive charges of the Zn<sup>2+</sup> atom in the catalytic center. <span class="html-italic">Insert</span>: pH dependence of hACE2 (red curve 1) and rACE2 (green curve 2) with denoted isoelectric point (<span class="html-italic">nz</span> = 0): pI 5.21 (human) and pI 5.21 (rabbit) ACE2.</p>
Full article ">Figure 5
<p>Electrostatic potential on the 3D surface of the catalytic domain of human (hACE2, two upper models, PDB: 1r42) and reconstructed rabbit (rACE2, two lower models) of angiotensin-converting enzyme-2 at pH 5.0 (two left models) and at pH 7.0 (two right models). The surfaces of the models are colored according to the electrostatic potential (negative—red, positive—blue), computed at pH 5.0 or pH 7.0, ionic strength 0.0001 mol/L, and temperature 20 °C, and visualized in the range <span class="html-italic">kT</span>/<span class="html-italic">e</span> = ±6 J/C (the scale on the right); 1 <span class="html-italic">kT</span>/<span class="html-italic">e</span> = 25.26 mV at 20 °C or 26.73 mV at 37 °C.</p>
Full article ">Figure 6
<p>Amino acid sequence of the polypeptide chains of human (rows beginning with <b>h</b>) and rabbit (<b>r</b>) ACE2. The numbers at the end of every row indicate the first and the last amino acid residue on the corresponding row. The cells of the amino acid residues (denoted by the standard one-letter code) are colored according to their affinity to the water molecules: hydrophilic (green) or hydrophobic (yellow). The vertical lines and the red rectangle denote the beginning and the end of the catalytic domain and the amino acid residues from the zinc-binding motif included in the enzyme active center, respectively.</p>
Full article ">Figure 6 Cont.
<p>Amino acid sequence of the polypeptide chains of human (rows beginning with <b>h</b>) and rabbit (<b>r</b>) ACE2. The numbers at the end of every row indicate the first and the last amino acid residue on the corresponding row. The cells of the amino acid residues (denoted by the standard one-letter code) are colored according to their affinity to the water molecules: hydrophilic (green) or hydrophobic (yellow). The vertical lines and the red rectangle denote the beginning and the end of the catalytic domain and the amino acid residues from the zinc-binding motif included in the enzyme active center, respectively.</p>
Full article ">Figure 7
<p>pH dependences of the electrostatic component Δ<span class="html-italic">G</span><sub>el</sub> of the folding energy Δ<span class="html-italic">G</span><sub>fold</sub> of the human (hACE2, PDB: 1r42, curve 1) and reconstructed rabbit (rACE2, curve 2) polypeptide chains of angiotensin-converting enzyme-2 at the transformation of the polypeptide chain from fully unfolded (random coil) to folded (globular 3D structure) conformation. The two 3D models are optimized by the program YASARA.</p>
Full article ">Figure 8
<p>Amino acid sequence of the polypeptide chains of transmembrane collectrin-like segment of the human ACE2 (the rows denoted by <b>A</b>) and human collectrin (<b>C</b>). The numbers at the end of every row indicate the first and the last amino acid residue on the corresponding row (the numbering corresponds to that in the hACE2 shown in <a href="#ijms-25-12393-f006" class="html-fig">Figure 6</a>). The cells of the amino acid residues (denoted by the standard one-letter code) are colored according to their affinity to water: hydrophilic (green) or hydrophobic (yellow). The absent amino acid residues are denoted by dashes.</p>
Full article ">Figure 9
<p>pH dependences of the electrostatic component Δ<span class="html-italic">G</span><sub>el</sub> of the folding energy of the catalytic domain of rabbit ACE2 polypeptide chain at transformation from random coil to 3D structure for four models: reconstructed on the base of hACE2 (PDB: 1r42, curve 1) and optimized by Chimera (curve 2) or YASARA (curve 3), and created by AlphaFold2 on the base of amino acid sequence of rACE2 (curve 4).</p>
Full article ">Figure 10
<p>pH dependences of the net electric charge <span class="html-italic">nz</span> of the catalytic domain of human (hACE2, curves 1, 2, and 3) and rabbit (rACE2, curves 4, 5, and 6) angiotensin-converting enzyme-2 in aqueous medium when the polypeptide chain is in unfolded conformation (random coil, curves 1 and 4), folded in 3D globule without Zn<sup>2+</sup> (curves 2 and 5) and when the Zn<sup>2+</sup> cation is bound in the enzyme active center (curves 3 and 6). The folded 3D conformations correspond to the crystallographic model of hACE2 (PDB:1r42, red curve 3) and to the reconstructed model of rACE2 (green curve 6). The isoelectric points are denoted by open cycles. The net charge <span class="html-italic">nz</span> is the algebraic sum of the positive and negative coulombic charges of the polypeptide chain without (curves 1, 2, 4, 5) or with the attached Zn<sup>2+</sup> cation (curves 3 and 6).</p>
Full article ">Figure 11
<p>pH dependences of the electrostatic component Δ<span class="html-italic">G</span><sub>el</sub> of the folding energy (the main figure) and the net electric charge <span class="html-italic">nz</span> (the inserted figure) of horse myoglobine according to the original 3D model (PDB: 1AZI, curves 1) and its reconstructed analog (curves 2). The reconstruction of the 3D model of the horse myoglobine is performed on the base of human myoglobine (PDB: 3RGK) considering the difference in the amino acid sequences of the human and horse myoglobine without optimization.</p>
Full article ">
15 pages, 18440 KiB  
Article
Evaluation of Continuous GMA Welding Characteristics Based on the Copper-Plating Method of Solid Wire Surfaces
by Dong-Yoon Kim and Jiyoung Yu
Metals 2024, 14(11), 1300; https://doi.org/10.3390/met14111300 - 18 Nov 2024
Viewed by 375
Abstract
Gas metal arc welding (GMAW) is widely used in various industries, such as automotive and heavy equipment manufacturing, because of its high productivity and speed, with solid wires being selected based on the mechanical properties required for welded joints. GMAW consists of various [...] Read more.
Gas metal arc welding (GMAW) is widely used in various industries, such as automotive and heavy equipment manufacturing, because of its high productivity and speed, with solid wires being selected based on the mechanical properties required for welded joints. GMAW consists of various components, among which consumables such as the contact tip and continuously fed solid wire have a significant impact on the weld quality. In particular, the copper-plating method can affect the conductivity and arc stability of the solid wire, causing differences in the continuous welding performance. This study evaluated the welding performance during 60 min continuous GMAW using an AWS A5.18 ER70S-3 solid wire, which was copper-plated using chemical plating (C-wire) and electroplating (E-wire). The homogeneity and adhesion of the copper-plated surface of the E-wire were superior to those of the C-wire. The E-wire exhibited better performance in terms of arc stability. The wear rate of the contact tip was approximately 45% higher when using the E-wire for 60 min of welding compared with the C-wire, which was attributed to the larger variation rate in the cast and helix in the E-wire. Additionally, the amount of spatter adhered to the nozzle during 60 min, with the E-wire averaging 5.9 g, approximately half that of the C-wire at 12.9 g. The E-wire exhibits superior arc stability compared with the C-wire based on the spatter amount adhered to the nozzle. This study provides an important reference for understanding the impact of copper plating methods and wire morphology on the replacement cycles of consumable welding parts in automated welding processes such as continuous welding and wire-arc additive manufacturing. Full article
(This article belongs to the Special Issue Welding and Joining of Advanced High-Strength Steels (2nd Edition))
Show Figures

Figure 1

Figure 1
<p>Appearance and specifications of the welding workability evaluation system.</p>
Full article ">Figure 2
<p>Schematic of the welding current and voltage measurement.</p>
Full article ">Figure 3
<p>CT results of solid wire contact condition inside the contact tip during wire feeding.</p>
Full article ">Figure 4
<p>Surface analysis results of the solid wire according to the plating method; drawing direction: (<b>a</b>) C-wire; (<b>b</b>) E-wire; transverse direction; (<b>c</b>) C-wire; (<b>d</b>) E-wire.</p>
Full article ">Figure 5
<p>Analysis of the copper surface homogeneity of the solid wire based on the plating method: (<b>a</b>) C-wire; (<b>b</b>) E-wire.</p>
Full article ">Figure 6
<p>Results of coating adhesion: (<b>a</b>) C-wire; (<b>b</b>) E-wire.</p>
Full article ">Figure 7
<p>Arc stability evaluation for C-wire during 1 h of welding (average current/voltage and standard deviation): (<b>a</b>) C-wire #1; (<b>b</b>) C-wire #2.</p>
Full article ">Figure 8
<p>Actual welding current/voltage waveform of C-wire (10 kHz, 10 s): (<b>a</b>) stable welding section (point a); (<b>b</b>) unstable welding section (point b).</p>
Full article ">Figure 9
<p>Analysis of the contact tip and wire after 48.5 min of welding with C-wire #1: (<b>a</b>) the contact tip inside; (<b>b</b>) surface of C-wire after welding.</p>
Full article ">Figure 10
<p>Arc stability evaluation for the E-wire during 1 h of welding (average current/voltage and standard deviation): (<b>a</b>) C-wire #1; (<b>b</b>) C-wire #2.</p>
Full article ">Figure 11
<p>Weld appearance during continuous welding: (<b>a</b>) C-wire #1; (<b>b</b>) C-wire #2; (<b>c</b>) E-wire #1; (<b>d</b>) E-wire #2.</p>
Full article ">Figure 12
<p>Appearance of the contact tip end during continuous welding.</p>
Full article ">Figure 13
<p>Nozzle spatter adhesion state by welding time for each solid wire during continuous welding.</p>
Full article ">
20 pages, 5686 KiB  
Article
Genome-Wide Identification and Characterization of bHLH Gene Family in Hevea brasiliensis
by Zheng Wang, Yuan Yuan, Fazal Rehman, Xin Wang, Tingkai Wu, Zhi Deng and Han Cheng
Forests 2024, 15(11), 2027; https://doi.org/10.3390/f15112027 - 18 Nov 2024
Viewed by 385
Abstract
The basic helix-loop-helix (bHLH) transcription factors play crucial roles in plant growth, development, and stress responses. However, their identification and insights into the understanding of their role in rubber trees remain largely uncovered. In this study, the bHLH gene family was explored and [...] Read more.
The basic helix-loop-helix (bHLH) transcription factors play crucial roles in plant growth, development, and stress responses. However, their identification and insights into the understanding of their role in rubber trees remain largely uncovered. In this study, the bHLH gene family was explored and characterized in rubber trees using systematic bioinformatics approaches. In total, 180 bHLH genes were identified in the rubber tree genome, distributed unevenly across 18 chromosomes, and phylogenetic analysis classified these genes into 23 distinct subfamilies. Promoter regions revealed a high density of cis-elements responsive to light and hormones. Enrichment analysis indicated involvement in numerous biological processes, including growth, development, hormone responses, abiotic stress resistance, and secondary metabolite biosynthesis. Protein interaction network analysis identified extensive interactions between HbbHLH genes and other functional genes, forming key clusters related to iron homeostasis, plant growth, and stomatal development. Expression profiling of HbbHLH genes have demonstrated varied responses to endogenous and environmental changes. RT-qPCR of eleven HbbHLH genes in different tissues and under ethylene, jasmonic acid, and cold treatments revealed tissue-specific expression patterns and significant responses to these stimuli, highlighting the roles of these genes in hormone and cold stress responses. These findings establish a framework for exploring the molecular functions of bHLH transcription factors in rubber trees. Full article
Show Figures

Figure 1

Figure 1
<p>Phylogenetic tree of diverse species showing the number of bHLH families. The phylogenetic tree reflects the evolutionary relationships and divergence times of various plant species as determined using the TimeTree database (<a href="http://www.timetree.org" target="_blank">http://www.timetree.org</a>, accessed on 28 March 2024). Different colored nodes represent various classifications, including <span class="html-italic">Chlorophyta</span> (yellow), <span class="html-italic">Spermatophyta</span> (green), <span class="html-italic">Solanaceae</span> (red), <span class="html-italic">Brassicaceae</span> (purple), <span class="html-italic">Euphorbiaceae</span> (blue), and <span class="html-italic">Rosaceae</span> (orange), among others. A linear scale of time in MYA (millions of years ago) and a geological timescale are shown at the bottom of the tree.</p>
Full article ">Figure 2
<p>The bHLH domain is highly conserved across all the <span class="html-italic">HbbHLH</span> proteins. The overall height of the stack indicates the sequence conservation at that position. Capital letters indicate amino acids with more than 50% conservation, whereas asterisks indicate amino acids with more than 75% conservation across the 180 <span class="html-italic">HbbHLH</span> domains.</p>
Full article ">Figure 3
<p>Phylogenetic analysis of bHLH gene families. The phylogenetic tree was generated using MEGA 7.0 with 1000 bootstrap replicates. Different colors indicate different subgroups. Red triangles represent <span class="html-italic">AtbHLH</span> proteins and blue triangles represent <span class="html-italic">HbbHLH</span> proteins.</p>
Full article ">Figure 4
<p>Gene structure, <span class="html-italic">cis</span>-regulatory elements, and chromosomal localization of <span class="html-italic">HbbHLH</span> genes. (<b>a</b>) Gene structure and domain positions of <span class="html-italic">HbbHLH</span> IIId and IIIe subfamilies. (<b>b</b>) Statistics of the three categories of <span class="html-italic">cis</span>-regulatory elements in <span class="html-italic">HbbHLH</span> genes. (<b>c</b>) Chromosomal localization of <span class="html-italic">HbbHLH</span> genes, with tandemly duplicated genes marked in green.</p>
Full article ">Figure 5
<p>Collinear analysis of <span class="html-italic">HbbHLH</span> genes. Gray lines in the background indicate all collinear blocks within the rubber tree genome, whereas red lines indicate collinear gene pairs of <span class="html-italic">bHLH</span> genes.</p>
Full article ">Figure 6
<p>The Gene Ontology enrichment analysis of <span class="html-italic">HbbHLH</span> genes in rubber trees. Categorized into biological processes (BP), (only the top ten processes are shown), molecular functions (MF), and cellular components (CC).</p>
Full article ">Figure 7
<p>Protein interaction network of <span class="html-italic">HbbHLH</span> genes mapped to <span class="html-italic">Arabidopsis</span> genes. The circle size indicates the number of interaction partners, with larger circles representing more extensive interaction networks. The thickness of the connecting lines reflects the combined interaction scores, with thicker lines denoting stronger interactions. Orange circles highlight rubber tree bHLH genes, with black text indicating rubber tree gene ID and white text (in parentheses) showing their corresponding <span class="html-italic">Arabidopsis</span> homologues. Non-orange circles represent non-bHLH proteins labelled with <span class="html-italic">Arabidopsis</span> homologue names. Different colors denote distinct functional classifications.</p>
Full article ">Figure 8
<p>Temporal and spatial expression patterns of <span class="html-italic">HbbHLH</span> genes in 16 rubber tree varieties. (<b>a</b>) Heatmap of <span class="html-italic">HbbHLH</span> gene expression patterns in rubber trees. Each row represents an <span class="html-italic">HbbHLH</span> gene and the column names are formatted as variety_tissue_treatment. The variety numbers represent the following rubber tree varieties: 1: BT3410, 2: CATAS7-20-59, 3: CATAS7-33-97, 4: CATAS8-79, 5: CATAS88-13, 6: CATAS93-114, 7: FX3864, 8: GT1, 9: PR107, 10: PR255, 11: REKEN501, 12: RRΙΙ105, 13: RRIM600, 14: RRIM928, 15: TB1, and 16: Wencang11. (<b>b</b>,<b>c</b>) The series of diagrams illustrates the patterns of dynamic changes in <span class="html-italic">HbbHLH</span> DEGs during ethylene treatment and cold exposure, respectively, using Mfuzz.</p>
Full article ">Figure 9
<p>Transcriptional analysis of five <span class="html-italic">HbbHLH</span> genes across different tissues in the rubber tree with error bars representing the standard deviation of three technical replicates. Bk, bark; Lf, leaf; Lx, latex; FF, female flower; MF, male flower. Statistical significance was determined using one-way ANOVA and Tukey’s multiple comparison test, with differences denoted by lowercase letters.</p>
Full article ">Figure 10
<p>Transcriptional analysis of 11 <span class="html-italic">HbbHLH</span> genes in latex following treatment with ethylene (<b>a</b>) and methyl jasmonate (<b>b</b>). The x-axis labels denote ethylene (ET) and methyl jasmonate (JA). Error bars represent the standard deviation of three technical replicates. Statistical significance was assessed using one-way ANOVA and Tukey’s multiple comparison test, with differences indicated by lowercase letters.</p>
Full article ">Figure 11
<p>Transcriptional analysis of 12 <span class="html-italic">HbbHLH</span> genes in leaves at low temperatures (4 °C). Error bars represent the standard deviations of three technical replicates. Statistical significance was assessed using one-way ANOVA and Tukey’s multiple comparison test, with differences indicated by lowercase letters.</p>
Full article ">
21 pages, 3854 KiB  
Article
Optimization of a Gorlov Helical Turbine for Hydrokinetic Application Using the Response Surface Methodology and Experimental Tests
by Juan Camilo Pineda, Ainhoa Rubio-Clemente and Edwin Chica
Energies 2024, 17(22), 5747; https://doi.org/10.3390/en17225747 - 17 Nov 2024
Viewed by 363
Abstract
The work presents an analysis of the Gorlov helical turbine (GHT) design using both computational fluid dynamics (CFD) simulations and response surface methodology (RSM). The RSM method was applied to investigate the impact of three geometric factors on the turbine’s power coefficient (C [...] Read more.
The work presents an analysis of the Gorlov helical turbine (GHT) design using both computational fluid dynamics (CFD) simulations and response surface methodology (RSM). The RSM method was applied to investigate the impact of three geometric factors on the turbine’s power coefficient (CP): the number of blades (N), helix angle (γ), and aspect ratio (AR). Central composite design (CCD) was used for the design of experiments (DOE). For the CFD simulations, a three-dimensional computational domain was established in the Ansys Fluent software, version 2021R1 utilizing the k-ω SST turbulence model and the sliding mesh method to perform unsteady flow simulations. The objective function was to achieve the maximum CP, which was obtained using a high-correlation quadratic mathematical model. Under the optimum conditions, where N, γ, and AR were 5, 78°, and 0.6, respectively, a CP value of 0.3072 was achieved. The optimal turbine geometry was validated through experimental testing, and the CP curve versus tip speed ratio (TSR) was determined and compared with the numerical results, which showed a strong correlation between the two sets of data. Full article
(This article belongs to the Section B: Energy and Environment)
Show Figures

Figure 1

Figure 1
<p>Geometric factors involved in the Gorlov helical turbine (GHT) design.</p>
Full article ">Figure 2
<p>Dimensions of the computational domain and the setup of boundary conditions for the Ansys Fluent simulation.</p>
Full article ">Figure 3
<p>Mesh of computational domain dimensions.</p>
Full article ">Figure 4
<p>Experimental setup of the recirculating water channel. (1) Motor of 14.9 kW, (2) impeller, (3) water inlet value, (4) channel, (5) gate, (6) model vertical-axis hydrokinetic turbine, (7) connection axis to the sensor, (8) weir assembly, and (9) feed tank.</p>
Full article ">Figure 5
<p>Response surface plots for the power coefficient (<math display="inline"><semantics> <msub> <mi>C</mi> <mi>P</mi> </msub> </semantics></math>) by using the regression model. (<b>a</b>) Effects of N and <math display="inline"><semantics> <mi>γ</mi> </semantics></math>; (<b>b</b>) effects of N and AR; (<b>c</b>) effects of <math display="inline"><semantics> <mi>γ</mi> </semantics></math> and AR. The other factors were set at the optimal values.</p>
Full article ">Figure 6
<p>(<b>a</b>) Frequency distribution and (<b>b</b>) normal probability plots for the power coefficient (<math display="inline"><semantics> <msub> <mi>C</mi> <mi>P</mi> </msub> </semantics></math>).</p>
Full article ">Figure 7
<p>Numerical and experimental comparison of the power coefficient (<math display="inline"><semantics> <msub> <mi>C</mi> <mi>P</mi> </msub> </semantics></math>) vs. tip speed ratio (<math display="inline"><semantics> <mi>λ</mi> </semantics></math>) curves.</p>
Full article ">
19 pages, 1430 KiB  
Article
Broad Neutralization Capacity of an Engineered Thermostable Three-Helix Angiotensin-Converting Enzyme 2 Polypeptide Targeting the Receptor-Binding Domain of SARS-CoV-2
by Davide Cavazzini, Elisabetta Levati, Saveria Germani, Bao Loc Ta, Lara Monica, Angelo Bolchi, Gaetano Donofrio, Valentina Garrapa, Simone Ottonello and Barbara Montanini
Int. J. Mol. Sci. 2024, 25(22), 12319; https://doi.org/10.3390/ijms252212319 - 16 Nov 2024
Viewed by 447
Abstract
The mutational drift of SARS-CoV-2 and the appearance of multiple variants, including the latest Omicron variant and its sub-lineages, has significantly reduced (and in some cases abolished) the protective efficacy of Wuhan spike-antigen-based vaccines and therapeutic antibodies. One of the most functionally constrained [...] Read more.
The mutational drift of SARS-CoV-2 and the appearance of multiple variants, including the latest Omicron variant and its sub-lineages, has significantly reduced (and in some cases abolished) the protective efficacy of Wuhan spike-antigen-based vaccines and therapeutic antibodies. One of the most functionally constrained and thus largely invariable regions of the spike protein is the one involved in the interaction with the ACE2 receptor mediating the cellular entry of SARS-CoV-2. Engineered ACE2, both as a full-length protein or as an engineered polypeptide fragment, has been shown to be capable of preventing the host-cell binding of all viral variants and to be endowed with potent SARS-CoV-2 neutralization activity both in vitro and in vivo. Here, we report on the biochemical and antiviral properties of rationally designed ACE2 N-terminal, three-helix fragments that retain a native-like conformation. One of these fragments, designated as PRP8_3H and produced in recombinant form, bears structure-stabilizing and binding-affinity enhancing mutations in α-helix-I and in both α-helix I and II, respectively. While the native-like, unmodified three α-helices ACE2 fragment proved to be thermally unstable and without any detectable pseudovirion neutralization capacity, PRP8_3H was found to be highly thermostable and capable of binding to the SARS-CoV-2 spike receptor-binding domain with nanomolar affinity and to neutralize both Wuhan and Omicron spike-expressing pseudovirions at (sub)micromolar concentrations. PRP8_3H thus lends itself as a highly promising ACE2 decoy prototype suitable for a variety of formulations and prophylactic applications. Full article
(This article belongs to the Section Biochemistry)
Show Figures

Figure 1

Figure 1
<p>Amino acid sequences and predicted structures of the three ACE2 polypeptide fragments investigated in this study. (<b>A</b>) Amino acid sequences of the engineered ACE2 polypeptides aligned with the corresponding sequence of the human ACE2 receptor (WT). In WT ACE2 (upper sequence), the amino acid residues involved in SARS-CoV-2 RBD binding are shown in violet. In the engineered ACE2 polypeptide fragment sequences (indicated as PRP8_3H, WTL_3H, and PRP8_1H, for the three-helical engineered and wild-type-like, and the mono-helical P8 polypeptides, respectively; see text for details), invariant, wild-type identical amino acids are in black, while mutated residues are shown in red, blue, or green in the case of P8, V2.4 affinity-enhancing, and helix-III mutated residues, respectively. (<b>B</b>) Three-dimensional structure of the PRP8_3H decoy predicted by the SWISS-MODEL server, represented as a light blue ribbon (see <a href="#sec4-ijms-25-12319" class="html-sec">Section 4</a> for details). The positions of mutated amino acid residues are indicated on the ribbon with the same color-code as in (<b>A</b>). PRP8_3H structure is superimposed on the corresponding structure of human ACE2 (ACE2-RBD crystal structure, PDB code 6M17), with the SARS-CoV-2 RBD surface shown in red. (<b>C</b>) Same as (<b>B</b>) for the WTL_3H (upper image) and PRP8_1H (lower image) ACE2 fragment polypeptide derivatives.</p>
Full article ">Figure 2
<p>Biochemical characterization of the engineered PRP8_3H ACE2 polypeptide. (<b>A</b>) SDS-PAGE profile of the purified P8PE_3H protein after the metal-affinity chromatography (lane 1) and SEC (lane 2) purification steps (see <a href="#sec4-ijms-25-12319" class="html-sec">Section 4</a> for details); the migration positions of molecular size standards (bovine serum albumin, carbonic anhydrase, and lysozyme) are shown in lane “M”. (<b>B</b>) SEC analysis of purified PRP8_3H (see <a href="#sec4-ijms-25-12319" class="html-sec">Section 4</a> for details); the sizes and elution positions of molecular mass standards (thyroglobulin, bovine serum albumin, and lysozyme, from left to right) are indicated; blue line: absorbance at 280nm, red line: absorbance at 260 nm. (<b>C</b>) Far-UV circular dichroism spectra (190–250 nm) of PRP8_3H recorded at different temperature, as indicated. (<b>D</b>) Thermal stability of PRP8_3H (blue circles), PRP8_1H (red circles), WTL_3H (green circles), and the LCB1 mini protein (cyan circles) derived from CD spectrum transitions at 222 nm measured at temperatures ranging from 20 to 98 °C, as indicated.</p>
Full article ">Figure 3
<p>Binding of P8PR_3H, other ACE2 fragments, and the LCB1 artificial mini-protein to SARS-CoV-2 Wuhan spike RBD measured by ELISA. Concentration-dependent binding of P8PR_3H (red), WTL_3H (green), P8PR_1H (blue), and the LCB1 positive control (black) analyzed by ELISA, using Wuhan SARS-CoV-2 spike RBD as capture agent; data, expressed as maximum absorbance at 415 nm, are the mean of triplicate measurements (error bars represent the standard deviation). The EC50 estimated values are 32 nm, 253 nm, 420 nm, and 0.3 nM for P8PR_3H, WTL_3H, P8PR_1H, and LCB1, respectively.</p>
Full article ">Figure 4
<p>Cell viability and SARS-CoV-2 pseudovirion neutralization assays. (<b>A</b>) A cell viability assay was performed with the MTT dye on HEK293T cells incubated for different times at 37 °C with PRP8_3H or WTL_3H at the concentration indicated (from 0.6 to 20 μM; see <a href="#sec4-ijms-25-12319" class="html-sec">Section 4</a> for details). As a reference, phosphate buffer was used instead of protein. Data, expressed as percent survival, relative to the ‘vehicle only’ control, were collected at the time indicated after treatment and represent the mean ± SD of three technical replicates. (<b>B</b>) Neutralization of SARS-CoV-2 spike Wuhan-Hu-1 pseudovirions by different engineered ACE2-derived fragments and the LCB1 mini-protein. (<b>C</b>) Same as (<b>B</b>) with Omicron BA.2 spike-expressing pseudovirions. (<b>D</b>) same as (<b>B</b>) with Omicron BQ 1.1 spike-expressing pseudovirions. CC (cell control, i.e., cells incubated without any added pseudovirus) and VC (virus control, i.e., cells incubated in the absence of any ACE2-mimicking polypeptide) represent the positive and negative controls, respectively; data are the mean (±SD) of two independent replicates (see <a href="#sec4-ijms-25-12319" class="html-sec">Section 4</a> for details).</p>
Full article ">
18 pages, 6312 KiB  
Article
Identification and Analysis of Anticancer Therapeutic Targets from the Polysaccharide Krestin (PSK) and Polysaccharopeptide (PSP) Using Inverse Docking
by Carlos Iván López-Gil, Alejandro Téllez-Jurado, Marco Antonio Velasco-Velázquez and Miguel Angel Anducho-Reyes
Molecules 2024, 29(22), 5390; https://doi.org/10.3390/molecules29225390 - 15 Nov 2024
Viewed by 369
Abstract
The natural compounds PSK and PSP have antitumor and immunostimulant properties. These pharmacological benefits have been documented in vitro and in vivo, although there is no information in silico which describes the action mechanisms at the molecular level. In this study, the inverse [...] Read more.
The natural compounds PSK and PSP have antitumor and immunostimulant properties. These pharmacological benefits have been documented in vitro and in vivo, although there is no information in silico which describes the action mechanisms at the molecular level. In this study, the inverse docking method was used to identify the interactions of PSK and PSP with two local databases: BPAT with 66 antitumor proteins, and BPSIC with 138 surfaces and intracellular proteins. This led to the identification interactions and similarities of PSK and the AB680 inhibitor in the active site of CD73. It was also found that PSK binds to CD59, interacting with the amino acids APS22 and PHE23, which coincide with the rlLYd4 internalization inhibitor. With the isoform of the K-RAS protein, PSK bonded to the TYR32 amino acid at switch 1, while with BAK it bonded to the region of the α1 helix, while PSP bonded to the activation site and the C-terminal and N-terminal ends of that helix. In Bcl-2, PSK interacted at the binding site of the Venetoclax inhibitor, showing similarities with the amino acids ASP111, VAL133, LEU137, MET115, PHE112, and TYR108, while PSP had similarities with THR132, VAL133, LEU137, GLN118, MET115, APS111, PHE112, and PHE104. Full article
(This article belongs to the Section Computational and Theoretical Chemistry)
Show Figures

Graphical abstract

Graphical abstract
Full article ">Figure 1
<p>(<b>a</b>). Bar graphs show the hits found in the database of antitumor proteins (BPAT) and the database of intracellular and surface proteins (BPSIC) with the PSK ligand. (<b>b</b>) Bar graphs show the hits found for BPAT and BPSIC with the PSP ligand.</p>
Full article ">Figure 2
<p>General diagram of the interaction of the PSP and PSK ligands with diverse extracellular and intracellular proteins. The action mechanisms with the hits or target proteins are shown in different colors: blue = immunostimulatory; orange = apoptosis resistance; green = proliferation; red = apoptosis.</p>
Full article ">Figure 3
<p>(<b>a</b>) Model 1, PSK bonded at the Bcl-2 activation site and Venetoclax interactions bonded at the Bcl-2 activation site. (<b>b</b>) Model 1, interactions of PSK at the Bcl-2activation site.</p>
Full article ">Figure 4
<p>Model 1, CD59-PSK complex and its hydrogen bonding interactions (blue) and Van der Waals interactions (green).</p>
Full article ">Figure 5
<p>Models 6–8 of PSK interactions with the crystalline structure of the K-RAS isoform (5USJ); model 8, PSK interactions at the GTP binding site. The surface of switch 1 is highlighted in yellow and the surface of switch 2 is highlighted in purple.</p>
Full article ">Figure 6
<p>Models 6 and 7, superposition of the 3144 and PSK molecules at switch 1 of K-RAS. (<b>a</b>) Interactions of the K-RAS-3144 compound. (<b>b</b>) Model 6, interactions of the K-RAS-PSK compound. (<b>c</b>) Model 7, interactions of the K-RAS-PSK compound.</p>
Full article ">Figure 7
<p>PSK-Bak binding complex in α1 helix.</p>
Full article ">Figure 8
<p>(<b>a</b>) Model 9, PSP-Bak interaction complex at the activator site. (<b>b</b>). Model 1, PSP-Bak binding complex in α1 helix. The surface of the activator site is highlighted in yellow and the PSP molecule is shown in blue.</p>
Full article ">Figure 9
<p>(<b>a</b>) Image of the 6Z9B-AB680 complex; the inhibitor binds to the active site, hydrogen bonding interactions are shown in green, Van der Waals interactions in yellow, and π-π interactions in red. (<b>b</b>). Image of the 6Z9B-PSK complex; PSK binds to the active site, interactions with amino acids are shown in green. (<b>c</b>). Image of the 6Z9B-AB680 complex; the inhibitor binds to an allosteric site, hydrogen bonding interactions are shown in green and π-π interactions in red (<b>d</b>). Image of the 6Z9B-PSK complex; PSK binds to an allosteric site, hydrogen bonding interactions are shown in green.</p>
Full article ">Figure 10
<p>Model 1, superimposition of the AB680 inhibitor and PSK in the active site of the transformation of adenosine phosphate to adenosine of the CD73 protein. Amino acid interaction surface (green).</p>
Full article ">
10 pages, 2739 KiB  
Article
The Inhibitory Effect of Hedera helix and Coptidis Rhizome Mixture in the Pathogenesis of Laryngopharyngeal Reflux: Cleavage of E-Cadherin in Acid-Exposed Primary Human Pharyngeal Epithelial Cells
by Nu-Ri Im, Byoungjae Kim, You Yeon Chung, Kwang-Yoon Jung, Yeon Soo Kim and Seung-Kuk Baek
Int. J. Mol. Sci. 2024, 25(22), 12244; https://doi.org/10.3390/ijms252212244 - 14 Nov 2024
Viewed by 312
Abstract
Laryngopharyngeal reflux disease (LPRD) is a prevalent upper airway disorder characterized by inflammation and epithelial damage due to the backflow of gastric contents. Current treatments, primarily proton pump inhibitors (PPIs), often show variable efficacy, necessitating the exploration of alternative or adjunctive therapies. This [...] Read more.
Laryngopharyngeal reflux disease (LPRD) is a prevalent upper airway disorder characterized by inflammation and epithelial damage due to the backflow of gastric contents. Current treatments, primarily proton pump inhibitors (PPIs), often show variable efficacy, necessitating the exploration of alternative or adjunctive therapies. This study investigates the therapeutic potential of a mixture of Hedera helix and Coptidis rhizome (HHCR) in mitigating the pathophysiological mechanisms of LPRD. Using an in vitro model of human pharyngeal epithelial cells exposed to acidic conditions, we observed that acid exposure significantly increased the expression of adenosine A3 receptor (adenosine A3) and matrix metalloproteinase-7 (MMP-7), leading to E-cadherin cleavage and compromised epithelial integrity. Treatment with the HHCR mixture effectively suppressed adenosine A3 expression and MMP-7 activity, thereby reducing E-cadherin cleavage and preserving cellular cohesion. These results highlight the HHCR mixture’s ability to modulate the adenosine A3–MMP-7–E-cadherin pathway, suggesting its potential as a valuable adjunctive therapy for LPRD, particularly for patients unresponsive to conventional PPI treatment. This study provides new insights into the molecular interactions involved in LPRD and supports further clinical evaluation of HHCR as a complementary treatment option. Full article
(This article belongs to the Special Issue Functional Roles of Epithelial and Endothelial Cells)
Show Figures

Figure 1

Figure 1
<p>Changes in adenosine A3 expression following acid exposure: (<b>A</b>) qPCR analysis revealed an increase in the mRNA levels of adenosine A3 with the increase in the duration of acid exposure 1 or 5 min (<span class="html-italic">p</span> &lt; 0.05); (<b>B</b>) ELISA revealed no changes in the cyclic AMP expression following treatment with the adenosine A3 receptor antagonist; (<b>C</b>) adenosine A3 expression increased with the increase in the duration of acid exposure time, whereas these changes were not observed following treatment with an adenosine A3 antagonist. Statistical analysis was performed using one-way ANOVA, and error bars represent [SD]. Asterisks indicate statistically significant differences (<span class="html-italic">p</span> &lt; 0.05).</p>
Full article ">Figure 2
<p>Role of adenosine A3 in E-cadherin cleavage following acid exposure: (<b>A</b>) decrease in the intracellular (cell) and increase in the extracellular (media) levels of the active form of MMP-7 depending on the duration of the acid exposure time. These changes were not observed following treatment with the adenosine A3 antagonist. (<b>B</b>) While MMP7 enzyme activity was statistically significantly increased in proportion to acid exposure time, it was confirmed that there was no change when treated with adenosine A3 antagonist. (<b>C</b>) The intracellular cleavage (cell) and extracellular secretion (media) of E-cadherin following acid exposure time. These changes were not observed following treatment with the adenosine A3 antagonist. (<b>D</b>) The immunocytochemical staining images show MMP-7 and E-cadherin expression. Scale bar = 200 µm. The lower row of panel D indicates no decrease in the intracellular levels of MMP-7 or E-cadherin cleavage following the treatment with the adenosine A3 antagonist. Statistical analysis was performed using one-way ANOVA, and error bars represent [SD]. Asterisks indicate statistically significant differences (<span class="html-italic">p</span> &lt; 0.05).</p>
Full article ">Figure 3
<p>Effect of the treatment with the HHCR mixture on E-cadherin cleavage following acid exposure in human pharyngeal mucosal epithelial cells: (<b>A</b>) It was confirmed that adenosine A3, which increased statistically significantly in proportion to acid exposure time, was suppressed when treated with the HHCR mixture. (<b>B</b>) Adenosine A3 protein expression in epithelial cells was significantly suppressed following the treatment with the HHCR mixture. (<b>C</b>) Immunofluorescence images show the effect of treatment with adenosine A3 antagonist for 24 h with or without acid exposure (green: adenosine A3, blue: DAPI). Scale bar = 50 µm. (<b>D</b>) Low intracellular expression of MMP-7 and high extracellular expression of MMP-7 (media) after acid exposure were significantly reversed by the HHCR mixture. (<b>E</b>) Increased MMP-7 enzymatic activity in an acidic environment was inhibited by the HHCR mixture. (<b>F</b>) E-cadherin cleavage was inhibited in the cells treated with the HHCR mixture. Statistical analysis was performed using one-way ANOVA, and error bars represent [SD]. Asterisks indicate statistically significant differences (<span class="html-italic">p</span> &lt; 0.05).</p>
Full article ">
19 pages, 5407 KiB  
Article
OsbHLH5 Synergically Regulates Phenolamide and Diterpenoid Phytoalexins Involved in the Defense of Rice Against Pathogens
by Shen Zhou, Ran Zhang, Qiming Wang, Jinjin Zhu, Junjie Zhou, Yangyang Sun, Shuangqian Shen and Jie Luo
Int. J. Mol. Sci. 2024, 25(22), 12152; https://doi.org/10.3390/ijms252212152 - 12 Nov 2024
Viewed by 383
Abstract
Rice (Oryza sativa) produces phenolamides and diterpenoids as major phytoalexins. Although the biosynthetic pathways of phenolamides and diterpenoids in plants have been revealed, knowledge of their accumulation regulatory mechanisms remains limited, and, in particular, no co-regulatory factor has been identified to [...] Read more.
Rice (Oryza sativa) produces phenolamides and diterpenoids as major phytoalexins. Although the biosynthetic pathways of phenolamides and diterpenoids in plants have been revealed, knowledge of their accumulation regulatory mechanisms remains limited, and, in particular, no co-regulatory factor has been identified to date. Here, using a combined co-expression and evolutionary analysis, we identified the basic helix–loop–helix (bHLH) transcription factor OsbHLH5 as a positive bifunctional regulator of phenolamide and diterpenoid biosynthesis in rice. Metabolomic analysis revealed that OsbHLH5 significantly increased the content of phenolamides (such as feruloyl tryptamine (Fer-Trm) and p-coumaroyl tyramine (Cou-Tyr)) and diterpenoid phytoalexins (such as momilactones A, momilactones B) in the overexpression lines, while their content was reduced in the OsbHLH5 knockout lines. Gene expression and dual-luciferase assays revealed that OsbHLH5 activates phenolamide biosynthetic genes (including putrescine hydroxycinnamoyltransferase 3 (OsPHT3), tyramine hydroxycinnamoyltransferases 1/2 (OsTHT1/2), and tryptamine benzoyltransferase 2 (OsTBT2)) as well as diterpenoid biosynthetic genes (including copalyl diphosphate synthase 4 (OsCPS4) and kaurene synthase-like 4/7/10/11 (OsKSL4/7/10/11)). Furthermore, we have demonstrated that OsbHLH5 is induced by jasmonic acid (JA), while pathogen inoculation assays indicated that the overexpression of OsbHLH5 in transgenic rice plants leads to enhanced resistance to Xanthomonas oryzae pv. oryzae (Xoo). Overall, we have identified a positive regulator of phenolamide and diterpenoid biosynthesis and have demonstrated that biotic stress induces phytoalexin accumulation partly in an OsbHLH5-dependent manner, providing new insights into the metabolic interactions involved in pathogen response and offering valuable gene resources for the development, through genetic improvement, of new rice varieties that are resistant to diseases. Full article
(This article belongs to the Special Issue Molecular and Metabolic Regulation of Plant Secondary Metabolism)
Show Figures

Figure 1

Figure 1
<p>Analysis of candidate target genes regulated by OsbHLH5. (<b>A</b>) Networks established from correlations among genes and transcription factors (TFs) of the phenolamide and DP biosynthetic pathways. Pearson correlation coefficient values were calculated for each pair of genes. (<b>B</b>) An unrooted phylogenetic tree was constructed as described in the Methods section. The red marking indicates the transcription factor involved in this study–OsbHLH5. Bootstrap values &gt; 70% (based on 1000 replications) are indicated at each node (bar: 0.1 amino acid substitutions per site). (<b>C</b>) Promoter analysis of phenolamide and diterpenoid synthesis genes in rice.</p>
Full article ">Figure 2
<p>Subcellular localization and expression pattern analysis of <span class="html-italic">OsbHLH5</span>. (<b>A</b>) Subcellular localization pattern of OsbHLH5. Transient expression of OsbHLH5 fused to green fluorescent protein (GFP) in rice leaf protoplasts. OsGhd7-RFP as a nuclear marker (scale bar, 10 µm), repeated three times. (<b>B</b>) Expression pattern analysis of <span class="html-italic">OsbHLH5</span>. All tissues were sampled from a widely used rice variety, Zhonghua11, grown in a paddy field. The expression level was determined using RT–qPCR. Data are means ± SD (<span class="html-italic">n</span> = 3).</p>
Full article ">Figure 3
<p>Metabolite analysis of <span class="html-italic">OsbHLH5</span> transgenic individuals in vivo. (<b>A</b>) Expression analysis of <span class="html-italic">OsbHLH5</span> in the overexpression (ox) of transgenic lines; the rice <span class="html-italic">OsUbc13</span> gene was used as the internal control. The data are presented as mean ± SD, <span class="html-italic">n</span> = 3. (<b>B</b>) CRISPR/Cas9 target PAM sequence and edited types for <span class="html-italic">OsbHLH5</span>. CR, CRISPR. (<b>C</b>,<b>D</b>) Bar plots for the content of phenolamides in the OX (<b>C</b>) and mutant lines (<b>D</b>) of <span class="html-italic">OsbHLH5</span>. Fer-Put, feruloyl-putrescine; Cou-Spd, <span class="html-italic">p</span>-coumaroyl-spermidine; Fer-Agm, feruloyl-agmatine; Fer-Trm, feruloyl-tryptamine; Cou-Tyr, <span class="html-italic">p</span>-coumaroyl-tyramine; Fer-Sen, feruloyl-serotonin. The data are presented as mean ± SD, <span class="html-italic">n</span> = 3. (<b>E</b>,<b>F</b>) Bar plots for the content of diterpenoids in the OX (<b>E</b>) and mutant lines (<b>F</b>) of <span class="html-italic">OsbHLH5</span>. The data are presented as mean ± SD, <span class="html-italic">n</span> = 3, * <span class="html-italic">p</span> &lt; 0.05, ** <span class="html-italic">p</span> &lt; 0.01, *** <span class="html-italic">p</span> &lt; 0.001, Student’s <span class="html-italic">t</span> tests.</p>
Full article ">Figure 4
<p>Expression pattern analysis of genes involved in phenolamide and diterpenoid biosynthetic pathway in the overexpression (OX) and mutant lines (CR) of <span class="html-italic">OsbHLH5</span>. (<b>A</b>,<b>B</b>) Expression analysis of phenolamide biosynthetic genes in the OX (<b>A</b>) and mutant lines (<b>B</b>) of OsbHLH5. The data are presented as mean ± SD, <span class="html-italic">n</span> = 3. (<b>C</b>,<b>D</b>) Expression analysis of diterpenoid biosynthetic genes in the OX (<b>C</b>) and mutant lines (<b>D</b>) of OsbHLH5. The data are presented as mean ± SD, <span class="html-italic">n</span> = 3. The rice OsUbc13 gene was used as the internal control. Asterisks indicate values that are significantly different from the control; * <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, Student’s <span class="html-italic">t</span> test.</p>
Full article ">Figure 5
<p>OsbHLH5 positively regulates genes involved in phenolamides and diterpenoid biosynthesis in rice. (<b>A</b>) Schematic diagram of the effector and reporter plasmids used in the transient assay in leaf epidermal cells of <span class="html-italic">N. benthamiana</span>. Ren, Renilla luciferase; LUC, firefly luciferase. (<b>B</b>) OsbHLH5 activates the transcription of biosynthetic genes for phenolamides (<span class="html-italic">Os4CL5</span>, <span class="html-italic">OsPHT</span>s, and <span class="html-italic">OsTHT</span>s) and diterpenoids (<span class="html-italic">OsCPS4</span>, <span class="html-italic">OsKCL4</span>, <span class="html-italic">OsKCL7</span>, and <span class="html-italic">OsKCL10</span>). <span class="html-italic">N. benthamiana</span> leaves were infiltrated with different combinations of effectors and reporters. The LUC activity was normalized to the REN activity as an internal control. The <span class="html-italic">p</span>-value was calculated using Student’s <span class="html-italic">t</span> test, <span class="html-italic">n</span> = 3. * <span class="html-italic">p</span> &lt; 0.05, ** <span class="html-italic">p</span> &lt; 0.01, *** <span class="html-italic">p</span> &lt; 0.001.</p>
Full article ">Figure 6
<p>Functional validations of <span class="html-italic">OsbHLH5</span> in plants. (<b>A</b>) Gene expression profiling of <span class="html-italic">OsbHLH5</span> under low temperature (4 °C), 20% PEG, 200 mM NaCl, UV-B, 100 µM ABA, 10 µM BR, 100 µM IAA, and 100 µM JA treatment. The rice <span class="html-italic">OsUbc13</span> gene was used as the internal control. h, hour. The data are presented as mean ± SD, <span class="html-italic">n</span> = 3. (<b>B</b>,<b>C</b>) Function of <span class="html-italic">OsbHLH5</span> transgenic individuals on <span class="html-italic">Xoo</span> pathogen interaction. Typical leaves were photographed (<b>B</b>) and the lesion length (<b>C</b>) was counted after infection with the <span class="html-italic">Xoo</span>-PXO99 strains for 14 days. Ten replicates for lesion length. Bars represent mean ± SD, <span class="html-italic">n</span> = 3; * <span class="html-italic">p</span> &lt; 0.05, ** <span class="html-italic">p</span> &lt; 0.01, Student’s <span class="html-italic">t</span> test.</p>
Full article ">Figure 7
<p>A proposed working model of the role of OsbHLH5, which synergistically activates the biosynthesis of phenolamides and diterpenoids, conferring pathogen resistance in rice.</p>
Full article ">
19 pages, 8146 KiB  
Article
Computational Insights into Acrylamide Fragment Inhibition of SARS-CoV-2 Main Protease
by Ping Chen, Liyuan Wu, Bo Qin, Haodong Yao, Deting Xu, Sheng Cui and Lina Zhao
Curr. Issues Mol. Biol. 2024, 46(11), 12847-12865; https://doi.org/10.3390/cimb46110765 - 12 Nov 2024
Viewed by 491
Abstract
The pathogen of COVID-19, SARS-CoV-2, has caused a severe global health crisis. So far, while COVID-19 has been suppressed, the continuous evolution of SARS-CoV-2 variants has reduced the effectiveness of vaccines such as mRNA-1273 and drugs such as Remdesivir. To uphold the effectiveness [...] Read more.
The pathogen of COVID-19, SARS-CoV-2, has caused a severe global health crisis. So far, while COVID-19 has been suppressed, the continuous evolution of SARS-CoV-2 variants has reduced the effectiveness of vaccines such as mRNA-1273 and drugs such as Remdesivir. To uphold the effectiveness of vaccines and drugs prior to potential coronavirus outbreaks, it is necessary to explore the underlying mechanisms between biomolecules and nanodrugs. The experimental study reported that acrylamide fragments covalently attached to Cys145, the main protease enzyme (Mpro) of SARS-CoV-2, and occupied the substrate binding pocket, thereby disrupting protease dimerization. However, the potential mechanism linking them is unclear. The purpose of this work is to complement and validate experimental results, as well as to facilitate the study of novel antiviral drugs. Based on our experimental studies, we identified two acrylamide fragments and constructed corresponding protein-ligand complex models. Subsequently, we performed molecular dynamics (MD) simulations to unveil the crucial interaction mechanisms between these nanodrugs and SARS-CoV-2 Mpro. This approach allowed the capture of various binding conformations of the fragments on both monomeric and dimeric Mpro, revealing significant conformational dissociation between the catalytic and helix domains, which indicates the presence of allosteric targets. Notably, Compound 5 destabilizes Mpro dimerization and acts as an effective inhibitor by specifically targeting the active site, resulting in enhanced inhibitory effects. Consequently, these fragments can modulate Mpro’s conformational equilibrium among extended monomeric, compact, and dimeric forms, shedding light on the potential of these small molecules as novel inhibitors against coronaviruses. Overall, this research contributes to a broader understanding of drug development and fragment-based approaches in antiviral covalent therapeutics. Full article
(This article belongs to the Collection Feature Papers in Current Issues in Molecular Biology)
Show Figures

Figure 1

Figure 1
<p>(<b>a</b>) Transient structural conformation of SARS-CoV-2 monomeric Mpro bound to Compounds 2 and 5. Transient conformation of SARS-CoV-2 dimeric Mpro modified by (<b>b</b>) Compound <b>5</b> and (<b>c</b>) Compound <b>2</b>. Superposition of the compact state, active protomer of the dimer, and the extended Mpro, colored navy blue and red. Light blue and pink represent the active dimeric Mpro without Compounds. Red arrow indicates the position of the small molecule at the active site.</p>
Full article ">Figure 2
<p>This RMSD of (<b>a</b>) dimeric Mpro and (<b>b</b>) monomeric Mpro in the apo form, both without Compounds and covalently attached to Compounds 2 and 5.</p>
Full article ">Figure 3
<p>RMSD of each residue on dimeric Mpro covalently attached to (<b>a</b>) Compound <b>2</b> and (<b>b</b>) Compound <b>5</b>, and RMSD of each residue on monomeric Mpro covalently bound with (<b>c</b>) Compound <b>2</b> and (<b>d</b>) Compound <b>5</b>. The blue amino acid residues represent those with the smallest motion amplitude, while the red residues correspond to those with the largest motion amplitude.</p>
Full article ">Figure 4
<p>Amino acids within 0.50 nm of the active site cysteine Cys145 on dimeric Mpro with (<b>a</b>) Compound <b>2</b> and (<b>b</b>) Compound <b>5</b>, and on monomeric Mpro with (<b>c</b>) Compound <b>2</b> and (<b>d</b>) Compound <b>5</b>.</p>
Full article ">Figure 5
<p>Interactions between chain A (red) and chain B (navy blue) of the helical domain connected with (<b>a</b>) Compound <b>2</b> and (<b>b</b>) Compound <b>5</b>. The interactions between chain A and chain B of the catalytic domain connected with (<b>c</b>) Compound <b>2</b> and (<b>d</b>) Compound <b>5</b>, respectively.</p>
Full article ">Figure 6
<p>Time-series analysis of allosteric effect for SARS-CoV-2 dimeric or monomeric Mpro. (<b>a</b>) The interaction energy of chain A-B in Mpro dimers after bound to Compound <b>2</b> and Compound <b>5</b>. (<b>b</b>) Solvent accessible surface area (SASA) values of Mpro dimers. (<b>c</b>) Radius of gyration (Rg) analysis of Mpro dimers. (<b>d</b>) Rg values of Mpro monomers.</p>
Full article ">Figure 7
<p>(<b>a</b>,<b>b</b>) Electrostatic surface potential (ESP) analysis for different Compounds. (<b>c</b>,<b>d</b>) Average local ionization energy (ALIE) mapped onto the van der Waals surface of the Compounds. The blue regions, indicating weaker electron density, suggest higher reactivity of the electrons in the acrylamide fragments, making these areas more susceptible to electrophilic reactions.</p>
Full article ">Figure 8
<p>Comparative analysis of Mpro from PCA and covariance matrix. The conformational space sampled of MD trajectories projected onto subspaces spanned by PC1-2 of the Mpro complex with (<b>a</b>) Compound <b>2</b> and (<b>b</b>) Compound <b>5</b>, respectively. Covariance matrix calculations in atomic coordinates corresponding to Mpro complexed with (<b>c</b>) Compound <b>2</b> and (<b>d</b>) Compound <b>5</b>, respectively.</p>
Full article ">
19 pages, 4739 KiB  
Article
The Effect of Hydrodynamic Cavitation on the Structural and Functional Properties of Soy Protein Isolate–Lignan/Stilbene Polyphenol Conjugates
by Ning Hua, Xian’e Ren, Feng Yang, Yongchun Huang, Fengyan Wei and Lihui Yang
Foods 2024, 13(22), 3609; https://doi.org/10.3390/foods13223609 - 12 Nov 2024
Viewed by 624
Abstract
In this study, hydrodynamic cavitation technology was utilized to prepare conjugates of soy protein isolate (SPI) with polyphenols, including resveratrol (RA) and polydatin (PD) from the stilbene category, as well as arctiin (AC) and magnolol (MN) from the lignan category. To investigate the [...] Read more.
In this study, hydrodynamic cavitation technology was utilized to prepare conjugates of soy protein isolate (SPI) with polyphenols, including resveratrol (RA) and polydatin (PD) from the stilbene category, as well as arctiin (AC) and magnolol (MN) from the lignan category. To investigate the effects of hydrodynamic cavitation treatment on the interactions between SPI and these polyphenols, the polyphenol binding capacity with SPI was measured and the changes in the exposed sulfhydryl and free amino contents were analyzed. Various methods, including ultraviolet–visible spectroscopy, fluorescence spectroscopy, Fourier transform infrared spectroscopy, and circular dichroism spectroscopy, were also used to characterize the structural properties of the SPI–polyphenol conjugates. The results showed that compared to untreated SPI, SPI treated with hydrodynamic cavitation exposed more active groups, facilitating a greater binding capacity with the polyphenols. After the hydrodynamic cavitation treatment, the ultraviolet–visible absorption of the SPI–polyphenol conjugates increased while the fluorescence intensity decreased. Additionally, the content of exposed sulfhydryl and free amino groups declined, and changes in the secondary structure were observed, characterized by an increase in the α-helix and random coil content accompanied by a decrease in the β-sheet and β-turn content. Furthermore, the SPI–polyphenol conjugates treated with hydrodynamic cavitation demonstrated improved emulsifying characteristics and antioxidant activity. As a result, hydrodynamic cavitation could be identified as an innovative technique for the preparation of protein–phenolic conjugates. Full article
(This article belongs to the Section Food Physics and (Bio)Chemistry)
Show Figures

Figure 1

Figure 1
<p>Structures of stilbene polyphenols and lignan polyphenols.</p>
Full article ">Figure 2
<p>Schematic diagram of hydrodynamic cavitation device structure.</p>
Full article ">Figure 3
<p>The polyphenol binding capacity (<b>A</b>), free amino groups (<b>B</b>), and exposed sulfhydryl groups (<b>C</b>) of SPI and the SPI–polyphenol conjugates with or without hydrodynamic cavitation. SPI means soy protein isolate; UH means the samples without hydrodynamic cavitation; H means the samples with hydrodynamic cavitation. There were significant differences (<span class="html-italic">p</span> &lt; 0.05) among different uppercase letters with hydrodynamic cavitation, and the same uppercase letter means there was no significant difference (<span class="html-italic">p</span> &gt; 0.05). There were significant differences (<span class="html-italic">p</span> &lt; 0.05) among different lowercase letters without hydrodynamic cavitation, and the same lowercase letter means there was no significant difference (<span class="html-italic">p</span> &gt; 0.05). * represents a significant difference between the protein–polyphenol conjugate groups of the same polyphenol before and after the cavitation treatment.</p>
Full article ">Figure 4
<p>UV spectra (<b>A</b>) and endogenous fluorescence spectra (<b>B</b>) of SPI and SPI–polyphenol conjugates with or without hydrodynamic cavitation.</p>
Full article ">Figure 5
<p>FT-IR spectra of H group (<b>A</b>) and UH group (<b>B</b>) of SPI and SPI–polyphenol conjugates.</p>
Full article ">Figure 6
<p>CD spectroscopy of UH group (<b>A</b>) and H group (<b>B</b>) of SPI and SPI–polyphenol conjugates.</p>
Full article ">Figure 7
<p>The EAI (<b>A</b>), ESI (<b>B</b>), and H<sub>0</sub> (<b>C</b>) of SPI and the SPI–polyphenol conjugates with or without hydrodynamic cavitation. SPI means soy protein isolate; UH means the samples without hydrodynamic cavitation; H means the samples with hydrodynamic cavitation.There were significant differences (<span class="html-italic">p</span> &lt; 0.05) among different uppercase letters with hydrodynamic cavitation, and the same uppercase letter means there was no significant difference (<span class="html-italic">p</span> &gt; 0.05). There were significant differences (<span class="html-italic">p</span> &lt; 0.05) among different lowercase letters without hydrodynamic cavitation, and the same lowercase letter means there was no significant difference (<span class="html-italic">p</span> &gt; 0.05). * represents a significant difference between the protein–polyphenol conjugate groups of the same polyphenol before and after the cavitation treatment.</p>
Full article ">Figure 8
<p>The DPPH radical scavenging capacity (<b>A</b>), ABTS radical scavenging capacity (<b>B</b>), and iron ion reduction capacity (<b>C</b>) of SPI and the SPI–polyphenol conjugates with or without hydrodynamic cavitation. SPI means soy protein isolate; UH means the samples without hydrodynamic cavitation; H means the samples with hydrodynamic cavitation. There were significant differences (<span class="html-italic">p</span> &lt; 0.05) among different uppercase letters with hydrodynamic cavitation, and the same uppercase letter means there was no significant difference (<span class="html-italic">p</span> &gt; 0.05). There were significant differences (<span class="html-italic">p</span> &lt; 0.05) among different lowercase letters without hydrodynamic cavitation, and the same lowercase letter means there was no significant difference (<span class="html-italic">p</span> &gt; 0.05). * represents a significant difference between the protein–polyphenol conjugate groups of the same polyphenol before and after the cavitation treatment.</p>
Full article ">
Back to TopTop