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

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

Years

Between: -

Subjects

remove_circle_outline
remove_circle_outline
remove_circle_outline
remove_circle_outline
remove_circle_outline
remove_circle_outline
remove_circle_outline
remove_circle_outline
remove_circle_outline

Journals

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

Article Types

Countries / Regions

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

Search Results (27,790)

Search Parameters:
Keywords = bioactivity

Order results
Result details
Results per page
Select all
Export citation of selected articles as:
37 pages, 5914 KiB  
Review
Integrating Physical and Biochemical Cues for Muscle Engineering: Scaffolds and Graft Durability
by Farbod Yousefi, Lauren Ann Foster, Omar A. Selim and Chunfeng Zhao
Bioengineering 2024, 11(12), 1245; https://doi.org/10.3390/bioengineering11121245 (registering DOI) - 9 Dec 2024
Viewed by 101
Abstract
Muscle stem cells (MuSCs) are essential for skeletal muscle regeneration, influenced by a complex interplay of mechanical, biochemical, and molecular cues. Properties of the extracellular matrix (ECM) such as stiffness and alignment guide stem cell fate through mechanosensitive pathways, where forces like shear [...] Read more.
Muscle stem cells (MuSCs) are essential for skeletal muscle regeneration, influenced by a complex interplay of mechanical, biochemical, and molecular cues. Properties of the extracellular matrix (ECM) such as stiffness and alignment guide stem cell fate through mechanosensitive pathways, where forces like shear stress translate into biochemical signals, affecting cell behavior. Aging introduces senescence which disrupts the MuSC niche, leading to reduced regenerative capacity via epigenetic alterations and metabolic shifts. Transplantation further challenges MuSC viability, often resulting in fibrosis driven by dysregulated fibro-adipogenic progenitors (FAPs). Addressing these issues, scaffold designs integrated with pharmacotherapy emulate ECM environments, providing cues that enhance graft functionality and endurance. These scaffolds facilitate the synergy between mechanotransduction and intracellular signaling, optimizing MuSC proliferation and differentiation. Innovations utilizing human pluripotent stem cell-derived myogenic progenitors and exosome-mediated delivery exploit bioactive properties for targeted repair. Additionally, 3D-printed and electrospun scaffolds with adjustable biomechanical traits tackle scalability in treating volumetric muscle loss. Advanced techniques like single-cell RNA sequencing and high-resolution imaging unravel muscle repair mechanisms, offering precise mapping of cellular interactions. Collectively, this interdisciplinary approach fortifies tissue graft durability and MuSC maintenance, propelling therapeutic strategies for muscle injuries and degenerative diseases. Full article
Show Figures

Figure 1

Figure 1
<p>Understanding muscle stem cell states and PAX7 distribution. Understanding Muscle Stem Cell States: Quiescence, Activation, Proliferation, Differentiation, and Self-Renewal. Panel (<b>a</b>) illustrates the dynamic transition of muscle stem cells through various states, starting from quiescence, progressing through activation and proliferation, leading to differentiation, and the potential for self-renewal [<a href="#B29-bioengineering-11-01245" class="html-bibr">29</a>]. The schematic emphasizes the directional flow and sequential nature of these states, with each phase contributing to muscle regeneration and maintenance. Panel (<b>b</b>) presents PAX7 distribution curves within the niche, highlighting the variability and shifts in PAX7 expression levels during different states. The distribution curves illustrate how PAX7 expression peaks and shifts as muscle stem cells transition between quiescence, activation, and other states, reflecting their changing roles and regulatory mechanisms within the cellular environment [<a href="#B35-bioengineering-11-01245" class="html-bibr">35</a>]. Together, these panels provide a comprehensive overview of muscle stem cell state transitions and the critical role of PAX7 as a marker modulating these processes.</p>
Full article ">Figure 2
<p>Cellular interactions and regulatory mechanisms in the muscle stem cell niche. This figure depicts the complex interplay among various cellular components of the muscle stem cell niche, including fibro-adipogenic progenitors (FAPs), pericytes, fibroblasts, and macrophages (M1 and M2 phenotypes). The top section illustrates the role of macrophages, showing the transition from pro-inflammatory M1 macrophages, which secrete cytokines like TNF-α and IL-1β, to anti-inflammatory M2 macrophages involved in tissue repair and regeneration, producing factors such as IL-10 and TGF-β. Adjacent panels highlight interactions with T cells and their influence on the local inflammatory response. The central portion emphasizes cellular and extracellular matrix (ECM) interactions mediated by FAPs, pericytes, and fibroblasts, influencing the structural and regenerative dynamics of the muscle niche. The figure also outlines cell cycle regulation, detailing checkpoints and the role of cyclins (Cyclin A, B, D, E) and cyclin-dependent kinases (CDKs) in governing progression through G1, S, G2, and M phases. The accompanying graph on cyclin and CDK concentration during the cell cycle phases provides a visual representation of how cyclin–CDK complexes fluctuate, coordinating MuSC proliferation and differentiation [<a href="#B47-bioengineering-11-01245" class="html-bibr">47</a>]. Specific cyclin–CDK complexes’ elevated levels promote cell cycle progression, while their inhibition leads to cell cycle arrest, thus tightly regulating MuSC proliferation within the niche [<a href="#B48-bioengineering-11-01245" class="html-bibr">48</a>]. Together, these elements underscore the regulatory pathways and cellular interactions that maintain muscle homeostasis and regeneration within the niche.</p>
Full article ">Figure 3
<p>Heatmap of molecular pathways influencing muscle stem cell fates. This heatmap provides a visualization of the correlation between various molecular pathways and muscle stem cell fates, based on a qualitative exploratory review of relevant literature. The pathways considered include IL-6/JAK/STAT, TNF-α/NF-κB, TGF-β/SMAD, Notch, ROS/antioxidant mechanisms, and others. Muscle stem cell fates<tt>—</tt>quiescence, activation, proliferation, differentiation, and self-renewal<tt>—</tt>are represented along the <span class="html-italic">x</span>-axis, while molecular pathways are listed on the <span class="html-italic">y</span>-axis. The color gradient reflects the influence score of each pathway, with red hues indicating a positive correlation and blue hues representing a negative correlation. This visualization highlights how different pathways modulate specific cellular states, providing insight into the complex regulatory networks governing muscle stem cell behavior. The heatmap underscores the critical interplay of signaling cascades that dictate stem cell responses and tissue regeneration, as referenced in <a href="#bioengineering-11-01245-t001" class="html-table">Table 1</a>.</p>
Full article ">Figure 4
<p>Scaffold characteristics. The figure provides an overview of critical scaffold properties that influence muscle stem cell function and tissue engineering outcomes. The properties are categorized into structural and functional aspects, including vascularity, architecture, pore size, degradation, mechanical properties, and nano-specific modifications. The upper left panels illustrate how cells interact with scaffolds, emphasizing effects on cellular organization, polarity, and matrix binding, driven by diffusion gradients and cell–extracellular matrix (ECM) interactions. The central portion of the figure demonstrates mechanical properties, ranging from viscoelasticity (depicting a strain-stress curve) to stiffness levels (soft to stiff matrices) that provide mechanical cues to cells, critical for their behavior and differentiation. Adjacent panels depict architectural considerations such as pore size and overall scaffold structure, influencing cell infiltration, tissue integration, and mass transport. Degradation profiles are shown, emphasizing scaffold dissolution kinetics that match tissue regeneration rates. On the right side, nano-specific modifications, including particles (solid lipid, polymeric vesicles, quantum dots, etc.), highlight applications such as drug delivery and enhanced cellular interactions. The bottom panels explore complex architectures, like double-network gels that combine rigid and flexible networks, enhancing mechanical resilience and adaptability in dynamic tissue environments. The integration of nanoparticles with radioconjugates illustrates potential therapeutic and imaging applications in regenerative medicine and diagnostics. The progressive narrative of the figure emphasizes the interplay of scaffold properties with biological, mechanical, and nanoscale features for optimizing tissue regeneration outcomes.</p>
Full article ">Figure 5
<p>Mechanisms and pathways in muscle stem cell regulation. This figure encompasses multiple schematic representations that outline critical regulatory mechanisms influencing muscle stem cell behavior and fate. Panel (<b>a</b>) illustrates major signaling pathways, including Wnt, BMP, TGF-β, Notch, and PI3K/AKT pathways, emphasizing the journey from receptor activation at the cell membrane to downstream nuclear transcription factors that modulate gene expression, impacting muscle stem cell states and responses. Panel (<b>b</b>) provides detailed pathway diagrams, showcasing interactions among key regulatory molecules in muscle stem cell signaling. This includes cross-talk between different pathways that guide cell proliferation, differentiation, quiescence, and self-renewal, depicting an integrated regulatory network. Panel (<b>c</b>) highlights the role of mechanical forces, such as shear stress and tensile stress, in triggering cellular mechanotransduction processes, with a specific focus on mechanical receptors and their influence on intracellular pathways. Panel (<b>d</b>) illustrates epigenetic regulatory mechanisms, including DNA methylation, histone modification, and non-coding RNAs, which fine-tune muscle stem cell states by modifying chromatin structure and gene expression, contributing to the maintenance of stem cell quiescence, activation, or differentiation. Together, these panels underscore the complexity and integration of biochemical, mechanical, and epigenetic influences on muscle stem cell regulation.</p>
Full article ">Figure 6
<p>Network diagram of pharmacotherapy and muscle stem cell pathways. This figure visualizes the integration of pharmacotherapy interventions, such as growth factors, small molecules, and other modulators, with muscle stem cell pathways and markers. Nodes represent key elements of pharmacotherapy and stem cell signaling, while edges illustrate their interactions based on a qualitative exploratory review of the relevant literature. The color-coded nodes indicate distinct muscle stem cell states influenced by pharmacotherapy: blue for quiescence maintenance, green for activation, yellow for proliferation, red for differentiation, and purple for self-renewal. The network emphasizes the complex, multi-faceted approach to muscle regeneration, showcasing how targeted pharmacotherapy influences specific pathways, such as Wnt signaling modulators, SIRT1 activators, BMP signaling, and more. This comprehensive view of pharmacological modulation highlights the interplay between small molecules, growth factors, and muscle stem cell regulatory networks, facilitating targeted therapeutic strategies for muscle repair and regeneration, as referenced in <a href="#bioengineering-11-01245-t003" class="html-table">Table 3</a>.</p>
Full article ">Figure 7
<p>Heatmap of scaffold characteristics impact on muscle stem cell markers. This heatmap visually represents the estimated impact of various scaffold characteristics on muscle stem cell markers, based on a non-systematic literature review. The scaffold properties examined include stiffness, porosity, mechanical loading, alignment, and others. Muscle stem cell markers (e.g., Pax7, Myf5, Myod1, Myogenin) are displayed along the <span class="html-italic">x</span>-axis, while scaffold characteristics are on the <span class="html-italic">y</span>-axis. Color intensity and direction indicate the correlation strength, with red hues representing positive correlations and blue hues showing negative correlations. These interactions reflect the influence of scaffold design on stem cell behaviors such as quiescence, activation, proliferation, differentiation, and self-renewal. The heatmap provides insights into how scaffold properties can be strategically modulated to optimize muscle regeneration and stem cell function, summarizing findings based on references listed in <a href="#bioengineering-11-01245-t004" class="html-table">Table 4</a>.</p>
Full article ">Figure 8
<p>Network mapping of scaffold properties and muscle stem cell markers. This figure illustrates the estimated relationships between various scaffold properties (e.g., stiffness, alignment, porosity) and muscle stem cell markers, including Pax7, Myf5, Myod1, Myogenin, and others. The network displays interactions based on literature-derived estimates linking scaffold features to different stem cell fates such as quiescence, activation, proliferation, differentiation, and self-renewal. Scaffold properties, represented as nodes, interact through specific pathways with muscle stem cell markers and their regulatory networks. Color-coded nodes indicate stem cell fates influenced by scaffold design: blue represents quiescence, green denotes activation, yellow signifies proliferation, red indicates differentiation, and purple corresponds to self-renewal. The pathways connecting these nodes highlight the mechanistic role of scaffold properties in modulating muscle stem cell behavior and fate, reflecting how physical cues from scaffolds translate into cellular responses. This diagram emphasizes the complexity and interconnectedness of scaffold-induced regulation, illustrating their critical influence on stem cell signaling and tissue regeneration dynamics. The network is informed by selective literature reviews, as referenced in <a href="#bioengineering-11-01245-t004" class="html-table">Table 4</a>.</p>
Full article ">
11 pages, 2963 KiB  
Article
Microstructural and Surface Texture Evaluation of Orthodontic Microimplants Covered with Bioactive Layers Enriched with Silver Nanoparticles
by Magdalena Sycińska-Dziarnowska, Magdalena Ziąbka, Katarzyna Cholewa-Kowalska, Gianrico Spagnuolo, Hyo-Sang Park, Steven J. Lindauer and Krzysztof Woźniak
J. Funct. Biomater. 2024, 15(12), 371; https://doi.org/10.3390/jfb15120371 (registering DOI) - 9 Dec 2024
Viewed by 103
Abstract
Bacterial infections are a common cause of clinical complications associated with the use of orthodontic microimplants. Biofilm formation on their surfaces and subsequent infection of peri-implant tissues can result in either exfoliation or surgical removal of these medical devices. In order to improve [...] Read more.
Bacterial infections are a common cause of clinical complications associated with the use of orthodontic microimplants. Biofilm formation on their surfaces and subsequent infection of peri-implant tissues can result in either exfoliation or surgical removal of these medical devices. In order to improve the properties of microimplants, hybrid coatings enriched with silver nanoparticles, calcium, and phosphorus were investigated. The present study aimed to assess the microstructure of commercially available microimplants composed of a medical TiAlV (Ti6Al4V) alloy covered with organic–inorganic layers obtained by the sol–gel method using the dip-coating technique. The microstructures and elemental surface compositions of the sterile, etched, and layer-modified microimplants were characterized by scanning electron microscopy with X-ray energy-dispersive spectroscopy (SEM-EDS). Elements such as silver (Ag), calcium (Ca), phosphorus (P), silicon (Si), oxygen (O), and carbon (C) were detected on the microimplant’s surface layer. The SEM observations revealed that control microimplants (unetched) had smooth surfaces with only manufacturing-related embossing, while etching in hydrofluoric acid increased the surface roughness and introduced fluoride onto the microimplants. Layers with only silver nanoparticles reduced the roughness of the implant surface, and no extrusion was observed, while increased roughness and emerging porosity were observed when the layers were enriched with calcium and phosphorus. The highest roughness was observed in the microimplants etched with AgNPs and CaP, while the AgNPs-only layer showed a reduction in the roughness average parameter due to lower porosity. Enhancing the effectiveness of microimplants can be achieved by applying selective surface treatments to different parts. By keeping the outer tissue contact area smooth while making the bone contact area rough to promote stronger integration with bone tissue, the overall performance of the implants can be significantly improved. Full article
(This article belongs to the Special Issue Feature Papers in Dental Biomaterials (2nd Edition))
Show Figures

Figure 1

Figure 1
<p>SEM images of microimplant surface and EDS analysis of Ti (<b>a</b>–<b>d</b>), Ti-etched (<b>e</b>–<b>h</b>), Ti- etched with layer with AgNPs (<b>i</b>–<b>l</b>), Ti-etched with layer with AgNPs and enriched with CaP (<b>m</b>–<b>p</b>), Ti with layer with AgNPs (<b>r</b>–<b>u</b>), and Ti with layer with AgNPs and enriched with CaP (<b>w</b>–<b>z</b>). Magnification: 130×, 500×, and 1000×.</p>
Full article ">Figure 2
<p>SEM images of plate samples composed of Ti6Al4V alloy: layer with AgNPs (<b>a</b>) and layer with AgNPs and CaP (<b>b</b>). Magnification: 500×.</p>
Full article ">Figure 3
<p>Confocal laser scanning microscopy images of microimplant surfaces showing the roughness (Ra) of Ti (<b>a</b>), Ti-etched (<b>b</b>), Ti-etched with layer with AgNPs (<b>c</b>), Ti-etched with layer with AgNPs and enriched with CaP (<b>d</b>), Ti with layer with AgNPs (<b>e</b>), and Ti with layer with AgNPs and enriched with CaP (<b>f</b>).</p>
Full article ">
16 pages, 2198 KiB  
Article
Inhibitory Effects of Gliadin Hydrolysates on BACE1 Expression and APP Processing to Prevent Aβ Aggregation
by Chin-Yu Lin, Cheng-Hong Hsieh, Pei-Yu Lai, Ching-Wei Huang, Yung-Hui Chung, Shang-Ming Huang and Kuo-Chiang Hsu
Int. J. Mol. Sci. 2024, 25(23), 13212; https://doi.org/10.3390/ijms252313212 (registering DOI) - 9 Dec 2024
Viewed by 174
Abstract
Alzheimer’s disease (AD), a leading neurodegenerative disorder, is closely associated with the accumulation of amyloid-beta (Aβ) peptides in the brain. The enzyme β-secretase (BACE1), pivotal in Aβ production, represents a promising therapeutic target for AD. While bioactive peptides derived from food protein hydrolysates [...] Read more.
Alzheimer’s disease (AD), a leading neurodegenerative disorder, is closely associated with the accumulation of amyloid-beta (Aβ) peptides in the brain. The enzyme β-secretase (BACE1), pivotal in Aβ production, represents a promising therapeutic target for AD. While bioactive peptides derived from food protein hydrolysates have neuroprotective properties, their inhibitory effects on BACE1 remain largely unexplored. In this study, we evaluated the inhibitory potential of protein hydrolysates from gliadin, whey, and casein proteins prepared using bromelain, papain, and thermolysin. Through in vitro and cellular assays, bromelain-hydrolyzed gliadin (G-Bro) emerged as the most potent BACE1 inhibitor, with an IC50 of 0.408 mg/mL. G-Bro significantly reduced BACE1 expression and amyloid precursor protein (APP) processing in N2a/PS/APP cell cultures, suggesting its potential to attenuate Aβ aggregation. The unique peptide profile of G-Bro likely contributes to its inhibitory effect, with proline residues disrupting β-sheets, lysine residues introducing positive charges that hinder aggregation, hydrophobic residues stabilizing binding interactions, and glutamine residues enhancing solubility and stability. These findings highlight gliadin hydrolysates, particularly G-Bro, as potential natural BACE1 inhibitors with applications in dietary interventions for AD prevention. However, further studies are warranted to elucidate specific peptide interactions and their bioactivity in neural pathways to better understand their therapeutic potential. Full article
Show Figures

Figure 1

Figure 1
<p>Degree of hydrolysis of proteins hydrolyzed by (<b>a</b>) papain, (<b>b</b>) bromelain, and (<b>c</b>) thermolysin.</p>
Full article ">Figure 2
<p>Effect of G-Bro on the viability of N2a/PS/APP cells. Data are expressed as mean ± SD of three independent experiments. Statistical analyses were performed using one-way ANOVA test. Non-treated cells were considered as control. ** <span class="html-italic">p</span> &lt; 0.01 vs. control.</p>
Full article ">Figure 3
<p>Effect of G-Bro on BACE1 expression in N2a/PS/APP cells. Data are expressed as mean ± SD of three independent experiments. Statistical analyses were performed using one-way ANOVA test. Band intensity was evaluated with Image J Version 1.54. Non-treated cells were considered as control. * <span class="html-italic">p</span> &lt; 0.05 and ** <span class="html-italic">p</span> &lt; 0.01 vs. control.</p>
Full article ">Figure 4
<p>Effect of G-Bro on sAPP production in N2a/PS/APP cells. Data are expressed as mean ± SD of three independent experiments. Statistical analyses were performed using one-way ANOVA test. Non-treated cells were considered as control. * <span class="html-italic">p</span> &lt; 0.05 vs. control.</p>
Full article ">Figure 5
<p>Effect of G-Bro on Aβ aggregate formation in N2a/PS/APP cells. Data are expressed as mean ± SD of three independent experiments. Statistical analyses were performed using one-way ANOVA test. Non-treated cells were considered as control. * <span class="html-italic">p</span> &lt; 0.05 and ** <span class="html-italic">p</span> &lt; 0.01 vs. control.</p>
Full article ">Figure 6
<p>The overall framework of the study.</p>
Full article ">
12 pages, 3478 KiB  
Article
The Effect of Conditioned Medium from Angiopoietin-1 Gene-Modified Mesenchymal Stem Cells on Wound Healing in a Diabetic Mouse Model
by Qiong Deng, Shenzhen Pan, Fangzhou Du, Hongfei Sang, Zhixin Cai, Xiaoyu Xu, Qian Wei, Shuang Yu, Jingzhong Zhang and Chenglong Li
Bioengineering 2024, 11(12), 1244; https://doi.org/10.3390/bioengineering11121244 (registering DOI) - 9 Dec 2024
Viewed by 168
Abstract
Introduction: Mesenchymal stem cells (MSCs) have been introduced as a promising treatment for diabetic wounds. The effects of stem cell therapy are thought to be caused by bioactive molecules secreted by stem cells. Stem cell-based gene therapies can target bioactive molecules. Therefore, treatment [...] Read more.
Introduction: Mesenchymal stem cells (MSCs) have been introduced as a promising treatment for diabetic wounds. The effects of stem cell therapy are thought to be caused by bioactive molecules secreted by stem cells. Stem cell-based gene therapies can target bioactive molecules. Therefore, treatment using conditioned medium (CM) derived from genetically engineered stem cells has been proposed as an alternative option for diabetic ulcer care. Methods: MSCs derived from human umbilical cords were obtained and engineered to overexpress the angiogenin-1 gene (MSCsAng1) through plasmid transfection. This study extracted conditioned medium from MSCs (MSC-CM) or MSCsAng1(MSCAng1-CM) for wound treatment applications. Via in vitro experiments, the proangiogenic effects of MSCAng1-CM were assessed via the migration and tube formation of human umbilical vein endothelial cells (HUVECs). Furthermore, the efficacy of MSCAng1-CM in promoting wound healing, re-epithelialization, hair follicle, and angiogenesis was evaluated via a diabetic mouse skin defect model. Results: In vitro assays demonstrated that MSCAng1-CM significantly enhanced HUVECs’ functions, including migration and tube formation. In vivo assays revealed that MSCAng1-CM exhibited notable advancements in healing speed, re-epithelialization, hair follicle, and angiogenesis. Conclusion: These results indicate that MSCAng1-CM can promote wound healing in diabetic mice and make the vascular structure in regenerated tissues more stable without inducing tissue fibrosis, providing a new therapeutic strategy for treating diabetic skin wounds. This provides a valuable theoretical basis for further research on regenerative medicine and cell therapy. Full article
Show Figures

Figure 1

Figure 1
<p>Construction of MSCs<sup>Ang1</sup> and preparation and characterization of CM. (<b>A</b>) Schematic representation of the lentiviral plasmid carrying Ang1 and GFP linked via a T2A peptide. (<b>B</b>) Immunofluorescence staining of GFP in MSCs<sup>Ang1</sup>. (<b>C</b>) Quantitative analysis of GFP fluorescence intensity in MSCs<sup>Ang1</sup> showed a high GFP-positive expression rate. (<b>D</b>) qPCR analysis revealed that the mRNA expression level of Ang1 in MSCs<sup>Ang1</sup> was significantly elevated compared to MSCs. (<b>E</b>) Ang1 protein levels in MSC-CM and MSC<sup>Ang1</sup>-CM were assessed, demonstrating a significant increase in Ang1 expression in MSC<sup>Ang1</sup>-CM relative to MSC-CM. Total protein content in MSC-CM and MSC<sup>Ang1</sup>-CM was quantified using Coomassie Brilliant Blue staining. Scale bar: 100 μm. All data are presented as mean ± SD. *** <span class="html-italic">p</span> &lt; 0.001.</p>
Full article ">Figure 2
<p>Effects of different treatments on HUVECs’ migration and tube formation. (<b>A</b>) Images of HUVECs’ migration changes after 24 h and 48 h of different treatments. (<b>B</b>) Quantitative statistical analysis of cell migration rate showed that compared to Control and MSC-CM, MSC<sup>Ang1</sup>-CM significantly promoted HUVECs’ migration. (<b>C</b>) Number of HUVEC migrations in the transwell chamber after 24 h of different treatments. (<b>D</b>) Quantitative statistical analysis of the number of cell migrations shows that MSC<sup>Ang1</sup>-CM significantly promoted the number of HUVEC migrations compared to Control and MSC-CM. (<b>E</b>) Tubulogenic images of HUVECs after treatment with Control, MSC-CM, and MSC<sup>Ang1</sup>-CM for 24 h. (<b>F</b>) Quantitative statistical analysis of tubulogenesis experiments showed that compared to MSC-CM and Control, the total length of tubes in the MSC<sup>Ang1</sup>-CM group was significantly increased, with statistically significant differences observed among the three groups. Control: Replacement of 50% of the original medium of HUVECs with fresh D/F medium. Scale bar: 100 μm. All data are presented as mean ± SD. ** <span class="html-italic">p</span> &lt; 0.01, *** <span class="html-italic">p</span> &lt; 0.001.</p>
Full article ">Figure 3
<p>Effect of MSC<sup>Ang1</sup>-CM on wound healing in diabetic mice. (<b>A</b>) Dynamic changes and (<b>B</b>) healing rates of diabetic mice wounds under different treatments. (<b>C</b>) Quantitative statistical analysis of dynamic changes in wound healing showed that compared with PBS and MSC-CM groups, the MSC<sup>Ang1</sup>-CM treatment group significantly accelerated the healing of skin wounds in diabetic mice. (<b>D</b>) H&amp;E staining of diabetic mice wound tissues after 14 days of different treatments, orange arrows indicate hair follicles, and green arrows indicate blood vessels. (<b>E</b>) Quantitative statistical analysis of the Epithelial Gap (EG) in diabetic mice wounds after 14 days of treatments with PBS, MSC-CM, and MSC<sup>Ang1</sup>-CM showed that compared with PBS and MSC-CM EGs, the EG in the MSC<sup>Ang1</sup>-CM treatment group was significantly reduced and the re-epithelialization of skin wounds was significantly accelerated. (<b>F</b>) Quantitative statistical analysis of wound thickness in diabetic mice on day 14 showed that the skin thickness in the MSC-CM and MSC<sup>Ang1</sup>-CM groups was increased compared to the PBS group, while there was no significant difference between the MSC-CM and MSC<sup>Ang1</sup>-CM groups. All data are presented as mean ± SD. ns <span class="html-italic">p</span> &gt; 0.05, ** <span class="html-italic">p</span> &lt; 0.01, *** <span class="html-italic">p</span> &lt; 0.001.</p>
Full article ">Figure 4
<p>MSC<sup>Ang1</sup>-CM promotes regeneration of hair follicles and formation of blood vessels in diabetic mice wound migration repair area. (<b>A</b>) Dual immunofluorescence staining images of K14 and K17 in diabetic mice wounds after 14 days of different treatments. (<b>B</b>) Quantitative statistical analysis of hair follicle numbers in diabetic mice wounds on day 14 showed that compared with PBS and MSC-CM treatment, the number of K17-positive hair follicles in the migration repair area of the diabetic wound in the MSC<sup>Ang1</sup>-CM treatment group was significantly higher. (<b>C</b>) Dual immunofluorescence staining images of CD34 and αSMA in the wound migration repair area. (<b>D</b>,<b>E</b>) Quantitative statistical analysis of CD34+ vessels and CD34+αSMA+ double-positive mature vessel structures; (<b>D</b>) results showed that, compared to the PBS and MSC-CM groups, the MSC<sup>Ang1</sup>-CM treatment group had a significant increase in CD34+ vascular-like structures; (<b>E</b>) results showed that, compared to the PBS and MSC-CM groups, the MSC<sup>Ang1</sup>-CM treatment group was significantly increased in CD34+α-SMA+ double-positive mature vascular-like structures in diabetic mice wounds. (<b>F</b>) Western blot results showed that, after 14 days of treatment with PBS, MSC-CM, and MSC<sup>Ang1</sup>-CM, the MSC<sup>Ang1</sup>-CM group significantly upregulated the expression level of CD34 in the diabetic skin wound tissues, with β-Actin as the internal reference. (<b>G</b>) Semi-quantitative statistical analysis of CD34 showed that compared to the PBS and MSC-CM groups, the expression level of CD34 in the diabetic skin wound tissues treated with MSC<sup>Ang1</sup>-CM for 14 days was significantly increased, with a statistically significant difference between them. Scale bars: 50 μm, 100 μm. All data are presented as mean ± SD. * <span class="html-italic">p</span> &lt; 0.05, ** <span class="html-italic">p</span> &lt; 0.01, *** <span class="html-italic">p</span> &lt; 0.001.</p>
Full article ">Figure 5
<p>Impact of Ang1 secreted by engineered MSCs<sup>Ang1</sup> on fibrosis. (<b>A</b>–<b>C</b>) Expression levels of Col1a1, Col3a1, and the Col1a1/Col3a1 ratio in HUVECs after 24 h of different treatments. The results showed that both Col1a1 and Col3a1 expression levels were elevated in the MSC<sup>Ang1</sup>-CM compared to Control and MSC-CM; however, there were no significant differences in the Col1a1/Col3a1 ratio. (<b>D</b>–<b>F</b>) Expression levels of Col1a1, Col3a1, and the Col1a1/Col3a1 ratio in diabetic skin wound tissues after 14 days of different treatments. The results showed that both Col1a1 and Col3a1 expression levels were elevated in the MSC<sup>Ang1</sup>-CM compared to PBS and MSC-CM; however, there were no significant differences in the Col1a1/Col3a1 ratio among different treatments. Control: Replacement of 50% of the original medium of HUVECs with fresh D/F medium ns <span class="html-italic">p</span> &gt; 0.05, * <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 ">
11 pages, 3982 KiB  
Communication
Bioactive Agrocomposite for Tissue Engineering and Bone Regeneration
by Miguel Suffo, Celia Pérez-Muñoz, Daniel Goma-Jiménez, Carlos Revenga, Pablo Andrés-Cano and Miguel Ángel Cauqui-López
Inventions 2024, 9(6), 123; https://doi.org/10.3390/inventions9060123 (registering DOI) - 9 Dec 2024
Viewed by 177
Abstract
Background: This study describes a novel biomaterial consisting of a mixture of biphasic bioceramic obtained from waste generated by the sugar industry (Carbocal) and a medical-grade epoxy resin adhesive called LOCTITE® M31 CLTM. The objective was to demonstrate the possibility of coating [...] Read more.
Background: This study describes a novel biomaterial consisting of a mixture of biphasic bioceramic obtained from waste generated by the sugar industry (Carbocal) and a medical-grade epoxy resin adhesive called LOCTITE® M31 CLTM. The objective was to demonstrate the possibility of coating non-bioactive and non-biodegradable metallic surfaces on implantable elements. Methods: After preparation, the mixture was applied to the surfaces of hip prostheses composed of two distinct materials: polyetherimide and grade 5 titanium. In both cases, adhesion tests produced favourable results. Additionally, cell cultures were conducted using human foetal osteoblastic cell lines (hFOB 1.19). Results: It was observed that the mixture did not affect the proliferation of bone cells. Conclusions: This composite material was found to promote the growth of bone cells, suggesting its potential for fostering bone tissue development. Full article
(This article belongs to the Section Inventions and Innovation in Biotechnology and Materials)
Show Figures

Figure 1

Figure 1
<p>Sample preparation: (<b>a</b>) composite materials formed before coating the samples. Nomenclature I–VII corresponds to the compositions indicated in <a href="#inventions-09-00123-t002" class="html-table">Table 2</a>; (<b>b</b>) samples after coating 3 implantable surfaces (PEI-ULTEM1010<sup>®</sup> from the manufacturer Sabic (Riyadh, Saudi Arabia), Vitalium<sup>®</sup> from the manufacturer Dentsply Sirona (Charlotte, NC, USA), and Ti-6Al-4V); (<b>c</b>) implantable surfaces with adhesion test coatings: 1a—on femoral head on Vitalium material; 1b—different concentrations on small samples of Vitalium; 2a—epoxy only on flat surface of the intermediate part of the femoral stem of a hip prosthesis, manufactured in fused deposition (FDM) in material U1010; 2b—mixture VII on flat surface of the intermediate part of the femoral stem of a hip prosthesis, manufactured in FDM in material U1010; 2c—mixture VII on flat surface of the distal part of the femoral stem of a hip prosthesis, on loan from Stryker Iberia S.L. (Alcobendas, Madrid, Spain) in Ti-6Al-4V material.</p>
Full article ">Figure 2
<p>Instrument with V-shaped blades at 30° used for the grating in the adhesion test.</p>
Full article ">Figure 3
<p>Viability of human osteoblasts after 24 h, 48 h, 72 h, and 7 days of incubation with the different biomaterials tested. Culture medium in the absence of any biomaterial was used as a positive control, while 70% methanol was used as a negative control. X: sample tested; Y: viability (%).</p>
Full article ">Figure 4
<p>Growth of cells adhered to the surface of the biomaterial VI; (<b>a</b>) cells fixed with 70% methanol; (<b>b</b>) live cells.</p>
Full article ">Figure 5
<p>Growth of cells adhered to the surface of the biomaterial VII; (<b>a</b>) cells fixed with 70% methanol; (<b>b</b>) live cells.</p>
Full article ">Figure 6
<p>Viability of human osteoblasts after 24 h, 48 h, 72 h, and 7 days of incubation with samples VI and VII. Culture medium in the absence of any biomaterial was used as a positive control, while 70% methanol was used as a negative control. X: measure of time (hours/days); Y: viability (%).</p>
Full article ">Figure 7
<p>Results of adhesion tests of the biomaterial to implantable surfaces; (<b>a</b>) coating applied on the distal part of the Ti5 hip prosthesis; (<b>b</b>) coating applied to the femoral head of the hip prosthesis made of Vitalium; (<b>c</b>) coating applied on the intermediate area of the femoral stem of the hip prosthesis, manufactured in FDM 3D printing in U1010 material. In the enlarged view, two coatings can be seen, type I on the left and type VII on the right.</p>
Full article ">
9 pages, 1782 KiB  
Proceeding Paper
Advancing Functional Food Innovation: A Patent Landscape Analysis of Lamiaceae Bioactives Through Cooperative Patent Classification Systems
by Reda El Boukhari and Ahmed Fatimi
Biol. Life Sci. Forum 2024, 38(1), 1; https://doi.org/10.3390/blsf2024038001 (registering DOI) - 9 Dec 2024
Viewed by 48
Abstract
Medicinal plants from the Lamiaceae family hold significant promise as functional food ingredients due to their high content of essential dietary fiber and bioactive compounds. Lamiaceae plants are rich in phenolic acids, flavonoids, and alkaloids, contributing to their antioxidant and anti-inflammatory properties. This [...] Read more.
Medicinal plants from the Lamiaceae family hold significant promise as functional food ingredients due to their high content of essential dietary fiber and bioactive compounds. Lamiaceae plants are rich in phenolic acids, flavonoids, and alkaloids, contributing to their antioxidant and anti-inflammatory properties. This study utilizes a comprehensive patent analysis to explore recent trends in functional foods developed from Lamiaceae plants. We examined patents from databases using Cooperative Patent Classification (CPC) codes relevant to dietetic products and food compositions. Findings indicate a surge in patents related to Lamiaceae-based dietary supplements, particularly those targeting metabolic health, anti-aging, cognitive function, and bone and liver health. Mentha, Scutellaria, Salvia, and Perilla are the most represented genera, with dietary supplements showing potential in chronic disease prevention. This analysis highlights the growing commercial and therapeutic interest in Lamiaceae-derived functional foods, particularly for preventive health applications. Full article
(This article belongs to the Proceedings of The 4th International Electronic Conference on Nutrients)
Show Figures

Figure 1

Figure 1
<p>The number of <span class="html-italic">Lamiaceae</span>-based functional food patents granted per year between 2001 and 2024 (up to July 16).</p>
Full article ">Figure 2
<p>The top 10 jurisdictions identified by their number of patents related to <span class="html-italic">Lamiaceae</span>-based functional food.</p>
Full article ">Figure 3
<p>The main <span class="html-italic">Lamiaceae</span> genera concerned by the collected patents.</p>
Full article ">Figure 4
<p>The top 10 plant families associated with <span class="html-italic">Lamiaceae</span> from the collected patent documents.</p>
Full article ">Figure 5
<p>A summary of the most cited dietary supplement targets from the studied patent documents.</p>
Full article ">
10 pages, 826 KiB  
Article
Data Mining for the Characterization of a Paper Prototype Obtained with Bacterial Cellulose Derived from Banana and Pineapple By-Products
by Juan Diego Valenzuela-Cobos, Simón Pérez-Martínez, Manuel Fiallos-Cárdenas and Fabricio Guevara-Viejó
Appl. Sci. 2024, 14(23), 11426; https://doi.org/10.3390/app142311426 (registering DOI) - 9 Dec 2024
Viewed by 250
Abstract
The primary objective of this research is to evaluate the feasibility of two of the most prevalent agricultural residues in Ecuador, banana peels and pineapple peels, as a carbon source in the culture medium of Komagataeibacter hansenii for the production of bacterial cellulose [...] Read more.
The primary objective of this research is to evaluate the feasibility of two of the most prevalent agricultural residues in Ecuador, banana peels and pineapple peels, as a carbon source in the culture medium of Komagataeibacter hansenii for the production of bacterial cellulose (BC) and BC-based paper. This analysis includes an assessment of the productivity parameters of the obtained BC and the quality parameters of the BC-based paper, employing multivariate statistical methodologies. The experimental design consisted of fifteen treatments: T1 served as the control using the standard HS medium, while treatments T2–T8 used banana peel extracts (BPE), and treatments T9–T15 used pineapple peel extracts (PPE) at concentrations from 10% to 40% (v/v). Extracts were prepared with tailored pretreatments for each type of peel to optimize bioactive compound recovery. Standardized fermentation and purification conditions were applied, and once the cellulose was obtained, additives and coating agents were incorporated to produce paper samples from each treatment. The results indicated that higher BPE concentrations (T5, T6, T7, and T8) correlated significantly with increased Weight and Yield of BC, as well as improved grammage and water content in the BC-based paper. This highlights that efficient paper production is influenced by the quality of the bacterial cellulose used, with BPE-based media yielding optimal results due to their nutrient composition, which promotes bacterial growth and metabolic activity. This approach suggests a pathway for advancing sustainable and economical paper production. Full article
(This article belongs to the Special Issue Innovative Engineering Technologies for the Agri-Food Sector)
Show Figures

Figure 1

Figure 1
<p>Biplot for the productivity parameters of BC obtained from the 15 study treatments (T1–T15). The evaluated variables were weight (W), yield (Y), and substrate conversion ratio (SCR).</p>
Full article ">Figure 2
<p>Biplot for the quality parameters of BC-based paper obtained from the 15 study treatments (T1–T15). The evaluated variables were grammage (GM), brightness (BG), water content (WC), and opacity (OP).</p>
Full article ">
21 pages, 1660 KiB  
Article
Impact of Conventional Pasteurization, High Temperature Short Time, Ultra-High Temperature, and Storage Time on Physicochemical Characteristics, Bioactive Compounds, Antioxidant Activity, and Microbiological Quality of Fruit Nectars
by Natalia Polak, Stanisław Kalisz, Elżbieta Hać-Szymańczuk and Bartosz Kruszewski
Foods 2024, 13(23), 3963; https://doi.org/10.3390/foods13233963 (registering DOI) - 8 Dec 2024
Viewed by 383
Abstract
Berries are a valuable source of numerous bioactive compounds, and they have an interesting organoleptic profile. Unfortunately, their low storage life determines the need for their preservation. Among the various methods used in this regard, it was decided to use the High Temperature [...] Read more.
Berries are a valuable source of numerous bioactive compounds, and they have an interesting organoleptic profile. Unfortunately, their low storage life determines the need for their preservation. Among the various methods used in this regard, it was decided to use the High Temperature Short Time (HTST) (90 °C/15 s) and Ultra-High Temperature (UHT) (130 °C/5 s) methods to preserve the produced fruit nectar blends (strawberry–blackcurrant and strawberry–chokeberry). For comparison, the nectars were also preserved using conventional pasteurization (90 °C/10 min). Physicochemical, chromatographic, and microbiological determinations were carried out in the tested nectars before and immediately after processing, as well as after 1, 2, 3, 4, and 6 months of refrigerated storage. All methods allowed for the significant inactivation of selected microbial groups. Non-significant changes were observed as a result of HTST and UHT processing in the context of pH, TSS, and titratable acidity. Varied major changes occurred in the content of bioactive components (TPC—decrease or increase by 2–4%, TAC—decrease by 3–20%, vitamin C—decrease by 15–78%), antioxidant activity (decrease or increase by 3–9%), and nephelometric turbidity (decrease or increase by 11–65%). Both nectars showed better quality and nutritional value after the HTST and UHT processes compared to treatment with classic pasteurization. Storage affected the degradation of bioactive compounds, reduced antioxidant activity, increased turbidity, and caused the brightening of samples together with reducing redness and yellowness. Considering the results obtained, it is reasonable to recommend the use of the HTST and UHT methods in industrial conditions for the preservation of liquid fruit and vegetable products such as juices, nectars, and beverages. Full article
(This article belongs to the Special Issue Advancing Food Safety through PCR and Modern Detection Techniques)
Show Figures

Figure 1

Figure 1
<p>Total color difference (∆E*) between nectar samples. PT—traditional pasteurization, HTST—high temperature short time, UHT—ultra-high temperature, SB—strawberry–blackcurrant nectar, SC—strawberry–chokeberry nectar.</p>
Full article ">Figure 2
<p>Heatmap of correlations between content of bioactive compounds (TAC, TPC, vitamin C content, individual anthocyanins), color parameters (L*, a*, b*), and antioxidant activity (AA), based on total results obtained for all samples.</p>
Full article ">Figure 3
<p>Chromatograms of anthocyanins present in raw nectars: strawberry–blackcurrant (SB) nectar on the left, and strawberry–chokeberry (SC) nectar on the right. Identified anthocyanins: 1—delphinidin–3–O–glucoside, 2—delphinidin–3–O–rutinoside, 3—cyanidin–3–O–galactoside, 4—cyanidin–3–O–glucoside, 5—cyanidin–3–O–rutinoside, 6—cyanidin–3–O–arabinoside, 7—pelargonidin–3–O–glucoside, 8—pelargonidin–3–O–arabinoside, 9—cyanidin–3–O–xyloside.</p>
Full article ">Figure 4
<p>DPPH radical scavenging activity [%] of strawberry–blackcurrant nectar (<b>A</b>) and straw-berry-chokeberry nectar (<b>B</b>); ABTS radical scavenging activity [%] of strawberry–blackcurrant nectar (<b>C</b>) and strawberry–chokeberry nectar (<b>D</b>). HTST—high temperature short time, PT—traditional pasteurization, UHT—ultra-high temperature. Values marked with different capital letters are significantly different (<span class="html-italic">p</span> &lt; 0.05) and concern changes between preservation methods in SB or SC nectars in specific weeks of storage. Values marked with different small letters are significantly different (<span class="html-italic">p</span> &lt; 0.05) and concern changes during storage in SB or SC nectars for specific preservation methods.</p>
Full article ">
17 pages, 4770 KiB  
Article
Effect of Enterococcus hirae GS22 Fermentation-Assisted Extraction on the Physicochemical and Bioactivities of Sea Cucumber Intestinal Polysaccharides
by Xiqian Tan, Xiaoqing Wang, Fangchao Cui, Ali Zeshan, Dangfeng Wang, Xuepeng Li and Jianrong Li
Molecules 2024, 29(23), 5800; https://doi.org/10.3390/molecules29235800 (registering DOI) - 8 Dec 2024
Viewed by 492
Abstract
The sea cucumber intestine (SI), a secondary product from sea cucumber processing, contains polysaccharides as one of its active ingredients, and fermentation is an effective method for extracting bioactive substances from food by-products. In this study, to explore the effect of Enterococcus hirae [...] Read more.
The sea cucumber intestine (SI), a secondary product from sea cucumber processing, contains polysaccharides as one of its active ingredients, and fermentation is an effective method for extracting bioactive substances from food by-products. In this study, to explore the effect of Enterococcus hirae GS22 fermentation on the extraction of SI polysaccharides, the polysaccharides were extracted through the SI with and without Enterococcus hirae GS22 fermentation, and the obtained polysaccharides were designated as SC-PF and SC-P. The extraction yield, the structural characteristics, and the biological functions of the polysaccharides were then evaluated. The results indicated that Enterococcus hirae GS22 could grow well using SI as the substrate and that fermentation could improve the extraction yield of the polysaccharide from 0.48% to 0.63%, decrease the molecular weight (Mw), and change the monosaccharide composition. The diameter of SC-PF was smaller than SC-P, and the absolute value of the zeta potential of SC-PF was found to be lower than SC-P. Fermentation does not change the functional group or the thermal ability of the polysaccharide. SC-PF had better antioxidant ability than SC-P; the DPPH and superoxide anion scavenging ability were 96.3% and 36.5%, respectively. SC-PF also showed nearly 1.3- and 1.1-fold higher inhibition of α-glucosidase and α-amylase as compared to SC-P. The current results showed that E. hirae GS22 fermentation has the potential to extract SI polysaccharides with better prebiotic abilities. Full article
(This article belongs to the Collection Advances in Glycosciences)
Show Figures

Figure 1

Figure 1
<p>Bacterial growth and pH variation during fermentation: (<b>a</b>) bacterial count; (<b>b</b>) pH value. Different lowercase letters indicate statistically significant differences (<span class="html-italic">p</span> &lt; 0.05).</p>
Full article ">Figure 2
<p>The spectra and morphology analysis results of the polysaccharides: (<b>a</b>) UV spectra; (<b>b</b>) FTIR spectra; (<b>c</b>) SEM image of SC-P (5000×); (<b>d</b>) SEM image of SC-PF (5000×); (<b>e</b>) AFM image.</p>
Full article ">Figure 3
<p>Structural and thermal analysis of the polysaccharides: (<b>a</b>) XRD analysis; (<b>b</b>) Congo Red analysis; (<b>c</b>) DSC analysis.</p>
Full article ">Figure 4
<p>The antioxidant abilities of the polysaccharides: (<b>a</b>) DPPH scavenging ability; (<b>b</b>) superoxide radical scavenging ability. Different lowercase letters indicate statistically significant differences (<span class="html-italic">p</span> &lt; 0.05).</p>
Full article ">Figure 5
<p>Functional properties of the polysaccharides: (<b>a</b>) α-amylase activity inhibition; (<b>b</b>) α-glucosidase activity inhibition; (<b>c</b>) Glc absorption capacity; (<b>d</b>) cholesterol absorption capacity. Different lowercase letters indicate statistically significant differences (<span class="html-italic">p</span> &lt; 0.05).</p>
Full article ">
19 pages, 3637 KiB  
Article
Valorization of Hom Thong Banana Peel (Musa sp., AAA Group) as an Anti-Melanogenic Agent Through Inhibition of Pigmentary Genes and Molecular Docking Study
by Pichchapa Linsaenkart, Wipawadee Yooin, Supat Jiranusornkul, Korawan Sringarm, Chaiwat Arjin, Pornchai Rachtanapun, Kittisak Jantanasakulwong, Juan M. Castagnini and Warintorn Ruksiriwanich
Int. J. Mol. Sci. 2024, 25(23), 13202; https://doi.org/10.3390/ijms252313202 (registering DOI) - 8 Dec 2024
Viewed by 428
Abstract
Prolonged and unprotected exposure to the environment explicitly influences the development of hyperpigmented lesions. The enzyme tyrosinase (TYR) is a key target for regulating melanin synthesis. Several bioactive compounds derived from plant extracts have been found to possess potent anti-melanogenesis properties against TYR. [...] Read more.
Prolonged and unprotected exposure to the environment explicitly influences the development of hyperpigmented lesions. The enzyme tyrosinase (TYR) is a key target for regulating melanin synthesis. Several bioactive compounds derived from plant extracts have been found to possess potent anti-melanogenesis properties against TYR. In particular, the potential of banana peels from various varieties has garnered interest due to their application in skin hyperpigmentation treatment. A molecular docking study demonstrated interactions between rosmarinic acid, which is predominantly found in all Hom Thong peel extracts, and the active site of TYR (PDB ID: 2Y9X) at residues HIS263, VAL283, SER282, and MET280, with the lowest binding energy of −5.05 kcal/mol, showing the strongest interaction. Additionally, Hom Thong banana peels are rich in phenolic compounds that could inhibit melanin content and tyrosinase activity in both human and mouse melanoma cells. These effects may be attributed to the suppression of gene expression related to melanogenesis, including the regulator gene MITF and pigmentary genes TYR, TRP-1, and DCT, indicating effects comparable to those of the standard treatment groups with arbutin and kojic acid. Our findings indicated the potential of Hom Thong peel extracts as anti-melanogenic agents. Full article
Show Figures

Figure 1

Figure 1
<p>The molecular interaction profile of mushroom tyrosinase (PDB ID: 2Y9X) and the ligands after molecular docking studies binding poses of (<b>a</b>) L-tyrosine; (<b>b</b>) L-DOPA; (<b>c</b>) β-arbutin; and (<b>d</b>) kojic acid visualized by the BIOVIA Discovery Studio Visualizer.</p>
Full article ">Figure 2
<p>The molecular interaction profile of (<b>a</b>) 2D and (<b>b</b>) 3D structures of <span class="html-italic">(R)</span>-rosmarinic acid towards mushroom tyrosinase (PDB ID: 2Y9X) visualized by the BIOVIA Discovery Studio Visualizer and PyMOL, respectively.</p>
Full article ">Figure 3
<p>Effects of Hom Thong banana peel extracts on cell viability in (<b>a</b>) G361 human melanoma cells at 48 h; (<b>b</b>) B16 mouse melanoma cells at 48 h; (<b>c</b>) G361 human melanoma cells at 72 h; and (<b>d</b>) B16 mouse melanoma cells at 72 h. Data are expressed as the mean ± SD. Significant differences between samples are indicated by different letters (a, b, c, d, e, and f) with <span class="html-italic">p</span> &lt; 0.05. Ct: control, Water: aqueous extract of Hom Thong banana peel, 50EtOH: 50% ethanolic extract of Hom Thong banana peel, 95EtOH: 95% ethanolic extract of Hom Thong banana peel.</p>
Full article ">Figure 3 Cont.
<p>Effects of Hom Thong banana peel extracts on cell viability in (<b>a</b>) G361 human melanoma cells at 48 h; (<b>b</b>) B16 mouse melanoma cells at 48 h; (<b>c</b>) G361 human melanoma cells at 72 h; and (<b>d</b>) B16 mouse melanoma cells at 72 h. Data are expressed as the mean ± SD. Significant differences between samples are indicated by different letters (a, b, c, d, e, and f) with <span class="html-italic">p</span> &lt; 0.05. Ct: control, Water: aqueous extract of Hom Thong banana peel, 50EtOH: 50% ethanolic extract of Hom Thong banana peel, 95EtOH: 95% ethanolic extract of Hom Thong banana peel.</p>
Full article ">Figure 4
<p>Effects of Hom Thong banana peel extracts and their bioactive compounds on (<b>a</b>) melanin content in G361 human melanoma cells; (<b>b</b>) melanin content in B16 mouse melanoma cells; (<b>c</b>) tyrosinase activity in G361 human melanoma cells; and (<b>d</b>) tyrosinase activity in B16 mouse melanoma cells. Figures of L-dopachrome formation in (<b>e</b>) G361 human melanoma cells and (<b>f</b>) B16 mouse melanoma cells. Data are expressed as the mean ± SD. Significant differences between samples are indicated by different letters (a, b, c, and d) with <span class="html-italic">p</span> &lt; 0.05. Water: aqueous extract of Hom Thong banana peel, 50EtOH: 50% ethanolic extract of Hom Thong banana peel, 95EtOH: 95% ethanolic extract of Hom Thong banana peel.</p>
Full article ">Figure 4 Cont.
<p>Effects of Hom Thong banana peel extracts and their bioactive compounds on (<b>a</b>) melanin content in G361 human melanoma cells; (<b>b</b>) melanin content in B16 mouse melanoma cells; (<b>c</b>) tyrosinase activity in G361 human melanoma cells; and (<b>d</b>) tyrosinase activity in B16 mouse melanoma cells. Figures of L-dopachrome formation in (<b>e</b>) G361 human melanoma cells and (<b>f</b>) B16 mouse melanoma cells. Data are expressed as the mean ± SD. Significant differences between samples are indicated by different letters (a, b, c, and d) with <span class="html-italic">p</span> &lt; 0.05. Water: aqueous extract of Hom Thong banana peel, 50EtOH: 50% ethanolic extract of Hom Thong banana peel, 95EtOH: 95% ethanolic extract of Hom Thong banana peel.</p>
Full article ">Figure 5
<p>Effects of Hom Thong banana peel extracts and their bioactive compounds on relative gene expression of (<b>a</b>) <span class="html-italic">MITF</span> in G361 human melanoma cells; (<b>b</b>) <span class="html-italic">MITF</span> in B16 mouse melanoma cells; (<b>c</b>) <span class="html-italic">TYR</span> in G361 human melanoma cells; (<b>d</b>) <span class="html-italic">TYR</span> in B16 mouse melanoma cells; (<b>e</b>) <span class="html-italic">TRP-1</span> in G361 human melanoma cells; (<b>f</b>) <span class="html-italic">TRP-1</span> in B16 mouse melanoma cells; (<b>g</b>) <span class="html-italic">DCT</span> in G361 human melanoma cells; and (<b>h</b>) <span class="html-italic">DCT</span> in B16 mouse melanoma cells. Data are expressed as the mean ± SD. Significant differences between samples are indicated by different letters (a, b, and c) with <span class="html-italic">p</span> &lt; 0.05. Water: aqueous extract of Hom Thong banana peel, 50EtOH: 50% ethanolic extract of Hom Thong banana peel, 95EtOH: 95% ethanolic extract of Hom Thong banana peel.</p>
Full article ">Figure 5 Cont.
<p>Effects of Hom Thong banana peel extracts and their bioactive compounds on relative gene expression of (<b>a</b>) <span class="html-italic">MITF</span> in G361 human melanoma cells; (<b>b</b>) <span class="html-italic">MITF</span> in B16 mouse melanoma cells; (<b>c</b>) <span class="html-italic">TYR</span> in G361 human melanoma cells; (<b>d</b>) <span class="html-italic">TYR</span> in B16 mouse melanoma cells; (<b>e</b>) <span class="html-italic">TRP-1</span> in G361 human melanoma cells; (<b>f</b>) <span class="html-italic">TRP-1</span> in B16 mouse melanoma cells; (<b>g</b>) <span class="html-italic">DCT</span> in G361 human melanoma cells; and (<b>h</b>) <span class="html-italic">DCT</span> in B16 mouse melanoma cells. Data are expressed as the mean ± SD. Significant differences between samples are indicated by different letters (a, b, and c) with <span class="html-italic">p</span> &lt; 0.05. Water: aqueous extract of Hom Thong banana peel, 50EtOH: 50% ethanolic extract of Hom Thong banana peel, 95EtOH: 95% ethanolic extract of Hom Thong banana peel.</p>
Full article ">Figure 6
<p>The scheme summarizes melanin biosynthesis.</p>
Full article ">
20 pages, 883 KiB  
Article
Exploring Hyaluronidase and Alpha-Glucosidase Inhibition Activities of the Hydrothermal Extract of Coffee Silverskin Obtained from a Central Composite Design
by Thavy Kit, Agita Rachmala Ginting, Punnanee Sumpavapol, Lita Chheang and Sudtida Pliankarom Thanasupsin
Processes 2024, 12(12), 2805; https://doi.org/10.3390/pr12122805 (registering DOI) - 8 Dec 2024
Viewed by 226
Abstract
Coffee silverskin (CS), the main by-product of coffee roasting production, contains various valuable bioactive compounds in its chemical compositions. Hydrothermal water extraction (HDTE) is one of the promising techniques for valorizing the organic fraction of CS into functional bioactive ingredients, which can be [...] Read more.
Coffee silverskin (CS), the main by-product of coffee roasting production, contains various valuable bioactive compounds in its chemical compositions. Hydrothermal water extraction (HDTE) is one of the promising techniques for valorizing the organic fraction of CS into functional bioactive ingredients, which can be further exploited in various applications. This study aimed to evaluate the hyaluronidase and α-glucosidase inhibition activities of the CS extracts obtained under optimized water extraction conditions. Process optimization was performed using central composite design response surface methodology (CCD-RSM) with a broader range of extraction temperatures (25, 137.5, and 250 °C), reaction times (5, 38.5, and 72 min), and solid-to-liquid ratios (1:10, 1:80, and 1:150). The highest yield of 39.62% was obtained at 137.5 °C, with a reaction time of 72 min and an S/L ratio of 1:80. The total caffeoylquinic acid contents (T-CQA) were quantified based on the sum of three major isomers, including 3-CQA, 4-CQA, and 5-CQA. The results revealed that the highest T-CQA (2.76 ± 0.20 mg/g CS) was significantly obtained (p < 0.05) by subcritical water extraction (SWE) at 143.2 °C with an S/L ratio of 1:10 and an extraction time of 10.41 min. At such conditions, the total phenolic content (TPC), antioxidant properties (AP), and caffeine were 96.13 mg gallic acid equivalence per gram (GAE/g) CS, 20.85 ± 0.17 mg Trolox equivalence per gram (TE/g) CS, and 10.84 ± 1.25 mg/g CS, respectively. The 50% inhibition capacity (IC50) of hyaluronidase and α-glucosidase inhibition of the CS extracted were 5.00 mg/mL and 9.00 mg/mL, respectively. Our results supported the potential direct or indirect applications of CS, such as hydrothermal CS extract (HDT-CSE), in functional food or drinks. Repurposing CS residue to manufacture new products can efficiently reduce the amount of organic waste in landfills, thus conserving resources and energy and contributing to a lower overall carbon footprint in coffee production. Full article
(This article belongs to the Section Environmental and Green Processes)
Show Figures

Figure 1

Figure 1
<p>Waste management hierarchy for coffee silverskin.</p>
Full article ">Figure 2
<p>Schematic diagram of the experimental setup.</p>
Full article ">Figure 3
<p>Contour plots for the effects of the extraction temperature (A), reaction time (B), and solid-to-liquid ratio (C): (<b>a</b>–<b>c</b>) on the extraction yield (Y1); (<b>d</b>–<b>f</b>) on the TPC (Y2); (<b>g</b>–<b>i</b>) on the AP (Y3); (<b>j</b>–<b>l</b>) on the T-CQA (Y4).</p>
Full article ">Figure 4
<p>HPLC chromatogram of bioactive compounds in the coffee silverskin extract.</p>
Full article ">Figure 5
<p>Inhibition activities of 3-CQA, 5-CQA, and CSE.</p>
Full article ">Figure 6
<p>Schematic diagram of the experimental setup and key findings.</p>
Full article ">
15 pages, 3170 KiB  
Article
Preparation and Characterization of Small-Size and Strong Antioxidant Nanocarriers to Enhance the Stability and Bioactivity of Curcumin
by Shanshan Tie, Yujin Yang, Jiawei Ding, Yanyan Li, Mengmeng Xue, Jianrui Sun, Fang Li, Qiuxia Fan, Ying Wu and Shaobin Gu
Foods 2024, 13(23), 3958; https://doi.org/10.3390/foods13233958 (registering DOI) - 8 Dec 2024
Viewed by 396
Abstract
The purpose of this study was to design nanocarriers with small-size and antioxidant properties for the effective encapsulation of curcumin. Here, procyanidins, vanillin, and amino acids were used to successfully prepare nanocarriers of a controllable size in the range of 328~953 nm and [...] Read more.
The purpose of this study was to design nanocarriers with small-size and antioxidant properties for the effective encapsulation of curcumin. Here, procyanidins, vanillin, and amino acids were used to successfully prepare nanocarriers of a controllable size in the range of 328~953 nm and to endow antioxidant ability based on a one-step self-assembly method. The reaction involved a Mannich reaction on the phenolic hydroxyl groups of procyanidins, aldehyde groups of vanillin, and amino groups of amino acids. Subsequently, curcumin nanoparticles were prepared by loading curcumin with this nanocarrier, and the encapsulation efficiency of curcumin was 85.97%. Compared with free curcumin, the antioxidant capacity and photothermal stability of the embedded curcumin were significantly improved, and it could be slowly released into simulated digestive fluid. Moreover, using the corticosterone-induced PC12 cell injury model, the cell viability increased by 23.77% after the intervention of curcumin nanoparticles, and the cellular antioxidant capacity was also significantly improved. The nanoparticles prepared in this work can effectively improve the solubility, stability, and bioactivity of curcumin, which provides a reference for the embedding and delivery of other hydrophobic bioactive compounds. Full article
(This article belongs to the Section Food Nutrition)
Show Figures

Figure 1

Figure 1
<p>Preparation and particle size characterization of nanocarriers. (<b>a</b>) A schematic diagram of the preparation of NCs−1~NCs−8 using PCs, vanillin, and amino acids as raw materials. (<b>b</b>) Particle size, polydispersity index (PDI), and (<b>c</b>) particle size distribution of amino acid-dependent nanocarriers NCs−1~NCs−3. (<b>d</b>) Particle size, PDI, and (<b>e</b>) particle size distribution of PC-dependent nanocarriers NCs−1 and NCs−4~NCs−5. (<b>f</b>) Particle size, PDI, and (<b>g</b>) particle size distribution of vanillin-dependent nanocarriers NCs−1 and NCs−6~NCs−8. Note: the lower letters a, b, and c indicate that there are statistically significant differences between the samples.</p>
Full article ">Figure 2
<p>Formation and spectral characterization. (<b>a</b>) Schematic illustration of the reaction pathway of NCs. FTIR spectra of (<b>b</b>) vanillin (Van), Lys, (<b>c</b>) PCs, and NCs. (<b>d</b>) UV-Vis spectra and (<b>e</b>) crystal structure of PCs and NCs.</p>
Full article ">Figure 3
<p>Preparation and characterization of Cur NPs. (<b>a</b>) A schematic diagram of preparation, SEM image, and (<b>b</b>) embedding efficiency (EE) of Cur NPs. (<b>c</b>) FTIR spectra, (<b>d</b>) UV−vis spectra and (<b>e</b>) crystal structure of Cur NPs.</p>
Full article ">Figure 4
<p>Antioxidant capacity experiment. (<b>a</b>) DPPH and (<b>b</b>) ABTS radical scavenging activities for Cur, NCs, and Cur NPs. Note: the lower letters a, b, and c indicate that there are statistically significant differences between the samples.</p>
Full article ">Figure 5
<p>(<b>a</b>) UV irradiation and (<b>b</b>) thermal stability analyses for Cur and Cur NPs. Note: the lower letters a, b, and indicate that there are statistically significant differences between the samples.</p>
Full article ">Figure 6
<p>(<b>a</b>) Schematic diagram of simulated digestion and (<b>b</b>) release profile of Cur NPs.</p>
Full article ">Figure 7
<p>Cell viability analysis. (<b>a</b>) The effect of different concentrations of CORT on the viability of PC12 cells. Effect of (<b>b</b>) NCs, (<b>c</b>) Cur, and (<b>d</b>) Cur NPs on the cell viability of 400 μM CORT-induced PC12 cells. Note: the lower letters a−e indicate that there are statistically significant differences between the samples.</p>
Full article ">Figure 8
<p>Effect of Cur, NCs, and Cur NPs on the levels of (<b>a</b>) T-SOD, (<b>b</b>) CAT, and (<b>c</b>) MDA in PC12 cells induced by CORT. Note: the lower letters a, b, c, and d indicate that there are statistically significant differences between the samples.</p>
Full article ">Figure 9
<p>(<b>a</b>) Optical images of Cur NPs and NCs. (<b>b</b>) Optical images and hemolysis rate (HR) of negative control, positive control, Cur NPs, and NCs after treatment of red blood cells.</p>
Full article ">
15 pages, 1329 KiB  
Article
Antimicrobial Mixture Based on Micronized Kaolinite and Ziziphora Essential Oil as a Promising Formulation for the Management of Infected Wounds
by Aigerim A. Karaubayeva, Tolkyn Bekezhanova, Karlygash Zhaparkulova, Katarzyna Susniak, Jan Sobczynski, Paulina Kazimierczak, Agata Przekora, Krystyna Skalicka-Wozniak, Łukasz Kulinowski, Anna Glowniak-Lipa, Zuryiadda B. Sakipova and Izabela Korona-Głowniak
Int. J. Mol. Sci. 2024, 25(23), 13192; https://doi.org/10.3390/ijms252313192 (registering DOI) - 8 Dec 2024
Viewed by 342
Abstract
Kaolinite stands out as a promising natural geomaterial for developing new therapeutic systems aimed at addressing global health challenges, such as multidrug-resistant infections. In this study, we report on the formulation and biological activity of a therapeutic mixture composed of white micronized kaolinite [...] Read more.
Kaolinite stands out as a promising natural geomaterial for developing new therapeutic systems aimed at addressing global health challenges, such as multidrug-resistant infections. In this study, we report on the formulation and biological activity of a therapeutic mixture composed of white micronized kaolinite (KAO) and Ziziphora essential oil (ZEO), intended for topical application on infected wounds. GC–MS analysis revealed that the primary component of ZEO is pulegone, constituting 72.98% of the oil. ZEO demonstrated good bioactivity against bacterial and fungal strains (MIC 1.25–5 mg/mL). Additionally, ZEO at a concentration of 0.0156% (0.156 mg/mL) was found to significantly stimulate collagen synthesis. The antimicrobial activity of the tested KAO–ZEO mixture formulation (30% KAO/0.25% ZEO in an excipient base) showed the highest effectiveness against Candida spp. (MIC 0.08–25 mg/mL) and Gram-positive bacteria (MIC 0.16–25 mg/mL), with lower activity against Gram-negative bacteria (MIC 25–50 mg/mL). Moreover, the KAO–ZEO mixture was nontoxic (cell viability near 100%) to human skin fibroblasts according to the ISO 10993-5 standard and promoted collagen synthesis by skin cells. This is the first documented formulation combining KAO and ZEO, demonstrating significant antimicrobial properties along with the ability to stimulate collagen production in fibroblasts. These properties highlight KAO–ZEO as a promising novel treatment, which may synergize with current care standards and improve wound healing outcomes. Full article
Show Figures

Figure 1

Figure 1
<p>The GC–MS chromatogram of Ziziphora EO (ZEO).</p>
Full article ">Figure 2
<p>The picture of gel formulation containing micronized kaolinite clay and Ziziphora essential oil.</p>
Full article ">Figure 3
<p>Evaluation of the biological properties of ZEO and KAO–ZEO on normal human skin fibroblasts (BJ): (<b>a</b>) Screening cytotoxicity test on different concentrations of ZEO ranging from 0.0039% to 0.2500% and cytotoxicity assessment of KAO–ZEO extract prepared by soaking 100 mg of the sample in 1 mL of the culture medium for 24 h at 37 °C; (<b>b</b>) Evaluation of cell proliferation after exposure to the highest non-toxic concentration of ZEO (0.0156%) and KAO–ZEO extract (prepared at the ratio 100 mg/mL); (<b>c</b>) Collagen synthesis assessment after exposure to the highest non-toxic concentration of ZEO (0.0156%) and KAO–ZEO extract (prepared at the ratio 100 mg/mL); (control—cells maintained in the culture medium without the ZEO and KAO–ZEO; * statistically significant results considered at <span class="html-italic">p</span> &lt; 0.05 compared to the control cells according to One-way ANOVA with post hoc Dunnett’s test).</p>
Full article ">
14 pages, 5665 KiB  
Article
Sustainable Lipase Immobilization: Chokeberry and Apple Waste as Carriers
by Karina Jasińska, Maksym Nowosad, Aleksander Perzyna, Andrzej Bielacki, Stanisław Dziwiński, Bartłomiej Zieniuk and Agata Fabiszewska
Biomolecules 2024, 14(12), 1564; https://doi.org/10.3390/biom14121564 (registering DOI) - 8 Dec 2024
Viewed by 279
Abstract
In the modern world, the principles of the bioeconomy are becoming increasingly important. Recycling and reusability play a crucial role in sustainable development. Green chemistry is based on enzymes, but immobilized biocatalysts are still often designed with synthetic polymers. Insoluble carriers for immobilized [...] Read more.
In the modern world, the principles of the bioeconomy are becoming increasingly important. Recycling and reusability play a crucial role in sustainable development. Green chemistry is based on enzymes, but immobilized biocatalysts are still often designed with synthetic polymers. Insoluble carriers for immobilized biocatalysts, particularly those derived from agro-industrial waste such as mesoporous lignocellulosic materials, offer a promising alternative. By using waste materials as support for enzymes, we can reduce the environmental impact of waste disposal and contribute to the development of efficient bioprocessing technologies. The current study aimed to assess the possibility of using apple and chokeberry pomace as carriers for the immobilization of Palatase 20000L (lipase from Rhizomucor miehei). The analysis of lignocellulosic materials revealed that chokeberry pomace has a higher neutral detergent fiber (NDF) and lignin contents than apple pomace. Moreover, Scanning Electron Microscopy (SEM) observations indicated similar compact structures in both pomaces. The lipase activity assays demonstrated that immobilization of lipase from R. miehei onto apple and chokeberry pomace improves their properties, especially the synthetic activity. The findings highlight the potential of utilizing fruit pomaces not only as a source of bioactive compounds but also in enhancing enzyme stability for industrial applications. Full article
(This article belongs to the Special Issue Recent Advances in the Enzymatic Synthesis of Bioactive Compounds)
Show Figures

Figure 1

Figure 1
<p>Scanning electron microphotographs (×200 and ×600) of (<b>A</b>,<b>B</b>)—native chokeberry pomace and (<b>C</b>,<b>D</b>)—native apple pomace.</p>
Full article ">Figure 2
<p>The hydrolytic activities of immobilized lipase onto chokeberry pomace—ChoP (native, after hexane and ethanol treatment) and apple pomace—AP (native, after hexane and ethanol treatment, measured initially (dark magenta bars) and after 6 months of storage (violet bars). Means with the same letter a or b for initial activity and A for activity after 6 months of storage did not differ significantly (α = 0.05). Means with * within one immobilized lipase preparation differ significantly (α = 0.05).</p>
Full article ">Figure 3
<p>The synthetic activities of immobilized lipase onto chokeberry pomace—ChoP (native, after hexane and ethanol treatment) and apple pomace—AP (native, after hexane and ethanol treatment, measured initially (dark magenta bars) and after 6 months of storage (violet bars). Means with the same letter (a–e) or (A–D) did not differ significantly. Means with * within one immobilized lipase preparation differ significantly (α = 0.05).</p>
Full article ">Figure 4
<p>The specific hydrolytic activities of Palatase 20000L in the free form (PAL) and immobilized onto different carrier forms. Means with the same letter (a–c) did not differ significantly (α = 0.05). Abbreviations: ChoP—chokeberry pomace, AP—apple pomace.</p>
Full article ">Figure 5
<p>The specific synthetic activities of Palatase 20000L in the free form (PAL) and immobilized onto different carrier forms. Means with the same letter (a–f) did not differ significantly (α = 0.05). Abbreviations: ChoP—chokeberry pomace, AP—apple pomace.</p>
Full article ">Figure 6
<p>Comparison of hydrolytic activity of biocatalysts for the following substrates: <span class="html-italic">p</span>-nitrophenyl butyrate (C4:0), <span class="html-italic">p</span>-nitrophenyl laurate (C12:0), <span class="html-italic">p</span>-nitrophenyl palmitate (C16:0), and <span class="html-italic">p</span>-nitrophenyl oleate (C18:1). The means compared within one enzyme preparation, marked with different lowercase letters, are statistically different (α = 0.05). Abbreviations: ChoP—chokeberry pomace, AP—apple pomace.</p>
Full article ">Figure 7
<p>The effect of pH (ranging from 5 to 8) on the hydrolytic activity of the produced biocatalyst. The averages compared within a single enzyme preparation across the examined pH range, indicated by distinct lowercase letters, are significantly different (α = 0.05). Abbreviations: ChoP—chokeberry pomace, AP—apple pomace.</p>
Full article ">Figure 8
<p>Recovery analysis of lipase immobilized onto chokeberry pomace. The highest hydrolytic activities of all biocatalysts were defined as 100%. Abbreviation: ChoP—chokeberry pomace.</p>
Full article ">Figure 9
<p>Recovery analysis of lipase immobilized onto apple pomace. The highest hydrolytic activities of all biocatalysts were defined as 100%. Abbreviation: AP—apple pomace.</p>
Full article ">
11 pages, 544 KiB  
Article
Protective Effects of a Brassica nigra Sprout Hydroalcoholic Extract on Lipid Homeostasis, Hepatotoxicity, and Nephrotoxicity in Cyclophosphamide-Induced Toxicity in Rats
by Hassan Barakat, Thamer Aljutaily, Raghad I. Alkhurayji, Huda Aljumayi, Khalid S. Alhejji and Sami O. Almutairi
Metabolites 2024, 14(12), 690; https://doi.org/10.3390/metabo14120690 (registering DOI) - 8 Dec 2024
Viewed by 295
Abstract
Background: Brassica nigra possesses a significant concentration of bioactive compounds and has been demonstrated to have a variety of pharmacological properties, although its sprout has not been extensively studied. Thus, the protective effects of Brassica nigra sprout hydroalcoholic extract (BNSE) on lipid [...] Read more.
Background: Brassica nigra possesses a significant concentration of bioactive compounds and has been demonstrated to have a variety of pharmacological properties, although its sprout has not been extensively studied. Thus, the protective effects of Brassica nigra sprout hydroalcoholic extract (BNSE) on lipid homeostasis, hepatotoxicity, and nephrotoxicity in cyclophosphamide (CYP)-induced toxicity in rats were examined in this study. Methods: Four experimental rat groups (n = 8 for each group) were examined as follows: NR, normal rats that received normal saline by oral gavage daily; CYP, injected with a single dose of CYP at 250 mg kg−1 intraperitoneally (i.p.) and did not receive any treatment, receiving only normal saline by oral gavage daily; CYP + BNSE250, injected with a single dose of CYP at 250 mg kg−1 i.p. and treated with BNSE at 250 mg kg−1 by oral gavage daily for three weeks; and CYP + BNSE500, injected with a single dose of CYP at 250 mg kg−1 i.p. and treated with BNSE at 500 mg kg−1 by oral gavage daily for three weeks. Results: The results indicated a significant increase (p < 0.05) in triglyceride (TG), cholesterol (CHO), low-density lipoprotein cholesterol (LDL-c), and very low-density lipoprotein cholesterol (VLDL-c) levels in CYP-induced toxicity rats. The administration of BNSE at 250 and 500 mg kg−1 significantly (p < 0.05) attenuated TG, CHO, LDL-c, and VLDL-c at values comparable with the NR group. The most efficient treatment for improving the lipid profile and atherogenicity complication was BNSE at 500 mg kg−1, performing even better than 250 mg kg−1. Administrating BNSE at 250 or 500 mg kg−1 improved the liver’s function in a dose-dependent manner. Comparing the lower dose of 250 mg kg−1 of BNSE with 500 mg kg−1 showed that administrating 250 mg kg−1 attenuated alanine transaminase (ALT) by 28.92%, against 33.36% when 500 mg kg−1 was given. A similar trend was observed in aspartate aminotransferase (AST), where 19.44% was recorded for BNSE at 250 mg kg−1 and 34.93% for BNSE at 500 mg kg−1. Higher efficiency was noticed for BNSE at 250 and 500 mg kg−1 regarding alkaline phosphatase (ALP). An improvement of 38.73% for BNSE at 500 mg kg−1 was shown. The best treatment was BNSE at 500 mg kg−1, as it markedly improved liver function, such as total bilirubin (T.B.), in a dose-dependent manner. The administration of BNSE attenuated the total protein (T.P.), albumin, and globulin levels to be close to or higher than the typical values in NR rats. Conclusions: BNSE might be used for its promising hypolipidemic, hepatoprotective, and nephroprotective potential and to prevent diseases related to oxidative stress. Further research on its application in humans is highly recommended. Full article
(This article belongs to the Special Issue Plants and Plant-Based Foods for Metabolic Disease Prevention)
Show Figures

Figure 1

Figure 1
<p>Effects of <span class="html-italic">B. nigra</span> sprout extract at different doses on AI in rats with CYP-induced immunosuppression (mean ± SE), <span class="html-italic">n</span> = 8. <sup>a,b,</sup> and <sup>c</sup>: bars not sharing similar letters differed significantly (<span class="html-italic">p</span> &gt; 0.05), for experimental groups, see Materials and Methods.</p>
Full article ">
Back to TopTop