[go: up one dir, main page]

 
 
Sign in to use this feature.

Years

Between: -

Subjects

remove_circle_outline
remove_circle_outline
remove_circle_outline
remove_circle_outline
remove_circle_outline
remove_circle_outline
remove_circle_outline
remove_circle_outline
remove_circle_outline

Journals

remove_circle_outline
remove_circle_outline
remove_circle_outline
remove_circle_outline
remove_circle_outline
remove_circle_outline

Article Types

Countries / Regions

remove_circle_outline
remove_circle_outline
remove_circle_outline
remove_circle_outline
remove_circle_outline

Search Results (477)

Search Parameters:
Keywords = elastin

Order results
Result details
Results per page
Select all
Export citation of selected articles as:
10 pages, 7771 KiB  
Article
A Re-Examination of a Previous Study Relating to Topical Body Formulations: Validating Gene Expression Transcription at Multiple Time Points, and Protein Expression and Translation in an Ex Vivo Model
by Alan D. Widgerow, Mary E. Ziegler and Faiza Shafiq
Cosmetics 2024, 11(5), 159; https://doi.org/10.3390/cosmetics11050159 - 13 Sep 2024
Viewed by 290
Abstract
Introduction: This study was conducted to question the findings of a prior study published in Journal of Drugs in Dermatology (JDD) in September 2023, which reported that a topical firming and toning body lotion (FTB—SkinMedica®, Allergan Aesthetics, an AbbVie Company, Irvine, [...] Read more.
Introduction: This study was conducted to question the findings of a prior study published in Journal of Drugs in Dermatology (JDD) in September 2023, which reported that a topical firming and toning body lotion (FTB—SkinMedica®, Allergan Aesthetics, an AbbVie Company, Irvine, CA, USA) upregulated several genes in a UV-irradiated 3D full-thickness human skin model, outperforming other products, including TransFORM Body Treatment with TriHex Technology® (ATF—Alastin Skincare®, a Galderma company, Fort Worth, TX, USA). Given the unique response reported for FTB, we conducted this study to assess the reproducibility of these results and explore gene expression at multiple time points, along with validating protein expression in an ex vivo model. Materials and Methods: Experiments were conducted using an ex vivo model with photodamaged skin from facelift patients, under an Institutional Review Board-approved study. Skin samples were processed, cultured in transwells with Skin Media, and treated daily with either TransFORM or FTB for 7 days. A control group was left untreated. Gene expression was assessed using RT-PCR on days 1 and 3 and using immunofluorescence after 3 and 7 days of treatment. Skin samples were fixed, paraffin-embedded, sectioned, and stained with an anti-tropoelastin antibody. Fluorescence detection and imaging were conducted to assess protein expression changes. Results: Gene expression data from our study and the initial study showed a few similarities but multiple discrepancies. As opposed to results previously reported at only the 24 h time point, our study was completed at multiple time points and showed a complete reversal of many of these results. For example, COL1A1 expression at 24 h was similar for FTB in both studies but differed for TransFORM, which showed higher levels at 24 h in our study. At day 3, COL1A1 expression decreased markedly for FTB and was sustained for TransFORM. Other genes, such as COL3A1, COL5, ELN, VEGFC, ATG7, ATG12, BECN1, POMP, PSMB5, and PSMB6, exhibited varying expression patterns between the two studies and across different time points. From a translational perspective, histological analysis showed that TransFORM enhanced elastin fiber presence in the dermal–epidermal junction (DEJ) more effectively than FTB at both days 3 and 7. FTB-treated samples maintained a gap in the DEJ, while TransFORM-treated samples exhibited increased cellular proliferation and DEJ undulation, indicative of a healthier regenerative response. Conclusion: This study highlights the problems of examining data and drawing conclusions using a single point of examination. In addition, when a study reports positive results for only one product among a range of eight competitive products, further questioning is essential to exclude the possibility of the experimental model favoring that product. The additional 3-day time point and further translational examination of histological changes paint a completely different picture to that reported in the prior publication. TransFORM outperformed FTB in most gene expressions and histological parameters when assessed over multiple time points in a physiologically relevant ex vivo model. Full article
Show Figures

Figure 1

Figure 1
<p>ECM gene expression. The ex vivo skin model was established, and the skin specimen was left either untreated or treated with TransFORM or FTB for 24 h and 3 days. Gene expression was analyzed using RT-PCR to assess (<b>A</b>) COL1A1 (collagen I), (<b>B</b>) COL3A1 (collage III), (<b>C</b>) COL5 (collagen V), and (<b>D</b>) ELN (elastin) to evaluate the ECM-related genes. The data are presented as the fold-change relative to the untreated sample. The dashed lines represent the approximate expression values presented by Makino et al. [<a href="#B1-cosmetics-11-00159" class="html-bibr">1</a>] as a comparison.</p>
Full article ">Figure 2
<p>Lymphatic vessel and autophagy gene expression. The ex vivo skin model was established, and the skin was left untreated or treated with TransFORM or FTB for 24 h and 3 days. Gene expression was analyzed using RT-PCR to assess (<b>A</b>) VEGFC (vascular endothelial growth factor C) and to evaluate a lymphatic vessel gene; (<b>B</b>) ATG7 (autophagy related 7), (<b>C</b>) ATG12 (autophagy related 12), and (<b>D</b>) BECN1 (Beclin1) were used to evaluate autophagy-related genes. The data are presented as the fold-change relative to the untreated sample. The dashed lines represent the approximate expression values presented by Makino et al. [<a href="#B1-cosmetics-11-00159" class="html-bibr">1</a>] as a comparison.</p>
Full article ">Figure 3
<p>Proteasome gene expression. The ex vivo skin model was established, and the skin was left untreated or treated with TransFORM or FTB for 24 h and 3 days. Gene expression was analyzed using RT-PCR to assess (<b>A</b>) POMP (proteasome maturation protein), (<b>B</b>) PSMB5 (proteasome 20S subunit beta 5), and (<b>C</b>) PSMB6 (proteasome 20S subunit beta 6) in order to evaluate proteasome-related genes. The data are presented as the fold-change relative to the untreated sample. The dashed lines represent the approximate expression values presented by Makino et al. [<a href="#B1-cosmetics-11-00159" class="html-bibr">1</a>] as a comparison.</p>
Full article ">Figure 4
<p>Day 3 tropoelastin expression in an ex vivo model. The ex vivo skin model was established, and the skin was left untreated or treated with TransFORM or FTB for 3 days. The tissue was processed for immunostaining to assess tropoelastin expression (red). The tissue was counter-stained with DAPI (blue) to detect the nuclei.</p>
Full article ">Figure 5
<p>Day 7 tropoelastin expression in an ex vivo model. The ex vivo skin model was established, and the skin was left untreated or treated with TransFORM or FTB for 7 days. The tissue was processed for immunostaining to assess tropoelastin expression (red). The tissue was counter-stained with DAPI (blue) to detect the nuclei.</p>
Full article ">
14 pages, 4138 KiB  
Article
Comparison of Biomechanical and Microstructural Properties of Aortic Graft Materials in Aortic Repair Surgeries
by Haoliang Sun, Zirui Cheng, Xiaoya Guo, Hongcheng Gu, Dalin Tang and Liang Wang
J. Funct. Biomater. 2024, 15(9), 248; https://doi.org/10.3390/jfb15090248 - 28 Aug 2024
Viewed by 453
Abstract
Mechanical mismatch between native aortas and aortic grafts can induce graft failure. This study aims to compare the mechanical and microstructural properties of different graft materials used in aortic repair surgeries with those of normal and dissected human ascending aortas. Five types of [...] Read more.
Mechanical mismatch between native aortas and aortic grafts can induce graft failure. This study aims to compare the mechanical and microstructural properties of different graft materials used in aortic repair surgeries with those of normal and dissected human ascending aortas. Five types of materials including normal aorta (n = 10), dissected aorta (n = 6), human pericardium (n = 8), bovine pericardium (n = 8) and Dacron graft (n = 5) were collected to perform uniaxial tensile testing to determine their material stiffness, and ultimate strength/stretch. The elastin and collagen contents in four tissue groups except for Dacron were quantified by histological examinations, while the material ultrastructure of five material groups was visualized by scanning electron microscope. Statistical results showed that three graft materials including Dacron, human pericardium and bovine pericardium had significantly higher ultimate strength and stiffness than both normal and dissected aortas. Human and bovine pericardia had significantly lower ultimate stretch than native aortas. Histological examinations revealed that normal and diseased aortic tissues had a significantly higher content of elastic fiber than two pericardial tissues, but less collagen fiber content. All four tissue groups exhibited lamellar fiber ultrastructure, with aortic tissues possessing thinner lamella. Dacron was composed of densely coalesced polyethylene terephthalate fibers in thick bundles. Aortic graft materials with denser fiber ultrastructure and/or higher content of collagen fiber than native aortic tissues, exhibited higher ultimate strength and stiffness. This information provides a basis to understand the mechanical failure of aortic grafts, and inspire the design of biomimetic aortic grafts. Full article
(This article belongs to the Special Issue Functional Composite Biomaterials for Tissue Repair)
Show Figures

Figure 1

Figure 1
<p>Sample preparation for uniaxial tensile testing. (<b>a</b>–<b>e</b>) Specimens of normal aorta (<b>a</b>), dissected aorta (<b>b</b>), human pericardium (<b>c</b>), bovine pericardium (<b>d</b>) and Dacron graft (<b>e</b>); (<b>f</b>) Mechanical testing system for uniaxial tensile testing; (<b>g</b>) tissue preparation in dog-bone shape; (<b>h</b>) Recorded images showing the testing process of a sample to material failure. Sample directions (Circ for circumferential direction; Long for longitudinal direction) were indicated in (<b>a</b>,<b>b</b>,<b>e</b>).</p>
Full article ">Figure 2
<p>Representative HE, EVG and Masson images of four sample types including normal aortic tissue (<b>a1</b>–<b>a3</b>), diseased aortic tissue (<b>b1</b>–<b>b3</b>), human pericardial (<b>c1</b>–<b>c3</b>) and bovine pericardial tissues (<b>d1</b>–<b>d3</b>). All scale bars are 50 µm.</p>
Full article ">Figure 3
<p>Comparison of mechanical properties among five material groups. Representative material curves (<b>a</b>), material stiffness (<b>b</b>), ultimate stress (<b>c</b>), and ultimate stretch (<b>d</b>). The red arrow in (<b>a</b>) marks the end of the first segment of material curve of Dacron. The data from the sample in the circumferential direction (or fresh human pericardial sample) were presented as dot markers while data from the longitudinal direction (or fixed human pericardial sample) as a cross marker (* means <span class="html-italic">p</span> &lt; 0.05, ** means <span class="html-italic">p</span> &lt; 0.01, *** means <span class="html-italic">p</span> &lt; 0.001).</p>
Full article ">Figure 4
<p>Comparison of elastic (<b>a</b>) and collagen (<b>b</b>) fiber contents among four tissue groups. The data from each specimen were presented as one dot (* means <span class="html-italic">p</span> &lt; 0.05, *** means <span class="html-italic">p</span> &lt; 0.001).</p>
Full article ">Figure 5
<p>Tissue ultrastructure of five materials using SEM. (<b>a</b>) Normal aortic tissue; (<b>b</b>) Diseases aortic tissue; (<b>c</b>) Human pericardial tissue; (<b>d</b>) Bovine Pericardial tissue; (<b>e</b>,<b>f</b>) Dacron material. Magnification ×1000 for (<b>a</b>–<b>e</b>), ×3000 for (<b>f</b>). All scale bars are 50 µm.</p>
Full article ">
26 pages, 2040 KiB  
Review
Macrofungal Extracts as a Source of Bioactive Compounds for Cosmetical Anti-Aging Therapy: A Comprehensive Review
by Maja Paterska, Bogusław Czerny and Judyta Cielecka-Piontek
Nutrients 2024, 16(16), 2810; https://doi.org/10.3390/nu16162810 - 22 Aug 2024
Viewed by 967
Abstract
For centuries, mushrooms have been used as a component of skincare formulations. Environmental stresses and a modern lifestyle expose the skin to accelerated aging. To slow down this process, natural anti-aging skincare ingredients are being sought. In this review, 52 scientific publications about [...] Read more.
For centuries, mushrooms have been used as a component of skincare formulations. Environmental stresses and a modern lifestyle expose the skin to accelerated aging. To slow down this process, natural anti-aging skincare ingredients are being sought. In this review, 52 scientific publications about the effects of chemical compounds extracted from the fruiting bodies of macrofungi on skin cells were selected. The effects of extracts from nine species that are tested for anti-aging effects have been described. According to available literature data, macrofungi contain many polysaccharides, phenolic compounds, polysaccharide peptides, free amino acids, sterols, proteins, glycosides, triterpenes, alkaloids, which can have an anti-aging effect on the skin by acting as antioxidants, photoprotective, skin whitening, moisturizing, anti-inflammatory and stabilizing collagen, elastin and hyaluronic acid levels in the skin. Full article
(This article belongs to the Section Phytochemicals and Human Health)
Show Figures

Graphical abstract

Graphical abstract
Full article ">Figure 1
<p>Diagram for literature reviews.</p>
Full article ">Figure 2
<p>Factors contributing to skin aging.</p>
Full article ">Figure 3
<p>Schematic presentation of the skin morphology.</p>
Full article ">Figure 4
<p>Anti-aging mushroom bioactive compounds for skin.</p>
Full article ">
14 pages, 4039 KiB  
Article
Oral Yak Whey Protein Can Alleviate UV-Induced Skin Photoaging and Modulate Gut Microbiota Composition
by Diandian Wang, Yaxi Zhou, Jian Zhao, Chao Ren and Wenjie Yan
Foods 2024, 13(16), 2621; https://doi.org/10.3390/foods13162621 - 21 Aug 2024
Viewed by 517
Abstract
Excessive UV exposure can lead to skin roughness, wrinkles, pigmentation, and reduced elasticity, with severe cases potentially causing skin cancer. Nowadays, various anti-photoaging strategies have been developed to maintain skin health. Among them, dietary supplements with anti-photoaging properties are gaining increasing attention. Yak [...] Read more.
Excessive UV exposure can lead to skin roughness, wrinkles, pigmentation, and reduced elasticity, with severe cases potentially causing skin cancer. Nowadays, various anti-photoaging strategies have been developed to maintain skin health. Among them, dietary supplements with anti-photoaging properties are gaining increasing attention. Yak whey protein (YWP) possesses multiple benefits, including anti-inflammatory, antioxidant, and immune-boosting properties, effectively protecting the skin. This study used a mixed UVA and UVB light source to irradiate a nude mouse model, exploring the advantages of YWP in anti-photoaging and regulating gut microbiota. The results indicated that YWP alleviated UV-induced skin damage, wrinkles, dryness, and reduced elasticity by inhibiting oxidative stress, inflammatory factors (IL-1α, IL-6, and TNF-α), and matrix metalloproteinases (MMP-1, MMP-3, and MMP-12), thereby increasing the levels of elastin, type I collagen, and type III collagen in the extracellular matrix (ECM). Additionally, YWP significantly improved the abundance of Firmicutes and Bacteroidota in the gut microbiota of mice, promoting the growth of beneficial bacteria such as Lachnospiraceae_NK4A136_group, Ruminococcus_torques_group, and Clostridia_UCG_014, mitigating the dysbiosis caused by photoaging. These findings underscore the potential of YWP in anti-photoaging and gut microbiota improvement, highlighting it as a promising functional food for enhancing skin and gut health. Full article
Show Figures

Figure 1

Figure 1
<p>Experimental protocol of ultraviolet irradiation and gavage in mice.</p>
Full article ">Figure 2
<p>Macroscopic and histologic effects of YWP on mouse skin. (<b>A</b>) Representative photographs of mouse dorsal skin. (<b>B</b>) HE staining results.</p>
Full article ">Figure 3
<p>Effect of YWP on oxidative stress and inflammation levels in mouse skin. (<b>A</b>) SOD and MDA content in mouse skin (<span class="html-italic">n</span> = 6). (<b>B</b>) mRNA expression of TNF-α, IL-1α, and IL-6d in mouse skin tissue (<span class="html-italic">n</span> = 6). Statistical significance is indicated by distinct lowercase letters at the <span class="html-italic">p</span> &lt; 0.05 level.</p>
Full article ">Figure 4
<p>Effect of YWP on MMPs, elastin, and collagen in mouse skin. (<b>A</b>) Protein expression of MMP-1, MMP-3, MMP-12, and elastin in mouse skin tissues measured by WB (<span class="html-italic">n</span> = 3). (<b>B</b>) RT-qPCR of <span class="html-italic">COL1A1</span> and <span class="html-italic">COL3A1</span> mRNA expression (<span class="html-italic">n</span> = 6). Statistical significance is indicated by distinct lowercase letters at the <span class="html-italic">p</span> &lt; 0.05 level.</p>
Full article ">Figure 5
<p>Effect of YWP on the structure of intestinal flora in mice. (<b>A</b>) Venn diagram showing the number of ASVs in intestinal microbiota across different groups of mice. (<b>B</b>) Beta diversity analysis of intestinal microbiota across different groups of mice. (<b>C</b>–<b>F</b>) Analysis of the number of ASVs in intestinal microbiota across different groups of mice, including the Shannon index, Simpson index, Chao1 index, and Pielou-e index. Statistical significance is indicated by distinct lowercase letters at the <span class="html-italic">p</span> &lt; 0.05 level.</p>
Full article ">Figure 6
<p>Structure of mouse intestinal flora and results of LEfSe analysis. (<b>A</b>) Top 10 relative abundances at the level of mouse intestinal flora phylum in each group. (<b>B</b>) Top 30 relative abundances at the level of mouse intestinal flora genus in each group. (<b>C</b>) Evolutionary branching diagram of LEfSe analysis. (<b>D</b>) Distribution histogram (LDA &gt; 3.5).</p>
Full article ">
15 pages, 29814 KiB  
Case Report
Assessment of Extracellular Matrix Fibrous Elements in Male Dermal Aging: A Ten-Year Follow-Up Preliminary Case Study
by Bogusław Machaliński, Dorota Oszutowska-Mazurek, Przemyslaw Mazurek, Mirosław Parafiniuk, Paweł Szumilas, Alicja Zawiślak, Małgorzata Zaremba, Iwona Stecewicz, Piotr Zawodny and Barbara Wiszniewska
Biology 2024, 13(8), 636; https://doi.org/10.3390/biology13080636 - 20 Aug 2024
Viewed by 419
Abstract
Skin aging is a complex phenomenon influenced by multiple internal and external factors that can lead to significant changes in skin structure, particularly the degradation of key extracellular matrix (ECM) components such as collagen and elastic fibers in the dermis. In this study, [...] Read more.
Skin aging is a complex phenomenon influenced by multiple internal and external factors that can lead to significant changes in skin structure, particularly the degradation of key extracellular matrix (ECM) components such as collagen and elastic fibers in the dermis. In this study, we aimed to meticulously assess the morphological changes within these critical fibrous ECM elements in the dermis of the same volunteer at age 47 and 10 years later (2012 to 2022). Using advanced histological staining techniques, we examined the distribution and characteristics of ECM components, including type I collagen, type III collagen, and elastic fibers. Morphological analysis, facilitated by hematoxylin and eosin staining, allowed for an accurate assessment of fiber bundle thickness and a quantification of collagen and elastic fiber areas. In addition, we used the generalized Pareto distribution for histogram modeling to refine our statistical analyses. This research represents a pioneering effort to examine changes in ECM fiber material, specifically within the male dermis over a decade-long period. Our findings reveal substantial changes in the organization of type I collagen within the ECM, providing insight into the dynamic processes underlying skin aging. Full article
(This article belongs to the Section Cell Biology)
Show Figures

Figure 1

Figure 1
<p>The original H&amp;E, binary, and segmented images using color threshold from the volunteer in 2012 (<b>upper row</b>), and the original H&amp;E, binary, and segmented images using color threshold from 2022 (<b>bottom row</b>). In the segmented images, fibers are marked in red, nuclei in blue, and artifacts or structures not considered in the analysis of red areas are additionally marked in green. The resolution of the images is 2572 × 1928 pixels with an objective magnification ×40 and scale bar 100 <math display="inline"><semantics> <mi mathvariant="sans-serif">μ</mi> </semantics></math>m.</p>
Full article ">Figure 2
<p>Skin dermis of the same volunteer at ages 47 years (2012) and 57 years (2022), stained with H&amp;E. The upper panel images have an objective magnification of ×40, with a scale bar of 100 <math display="inline"><semantics> <mi mathvariant="sans-serif">μ</mi> </semantics></math>m, while the lower panel images have an objective magnification of ×100, with a scale bar of 20 <math display="inline"><semantics> <mi mathvariant="sans-serif">μ</mi> </semantics></math>m. Arrows indicate the dermal fibroblasts in the images.</p>
Full article ">Figure 3
<p>Skin dermis of the same volunteer at ages 47 years (2012) and 57 years (2022), stained with Mallory’s trichrome, silver impregnation, and Weigert’s/orcein. Objective magnification is ×40, and the scale bar represents 100 <math display="inline"><semantics> <mi mathvariant="sans-serif">μ</mi> </semantics></math>m.</p>
Full article ">Figure 3 Cont.
<p>Skin dermis of the same volunteer at ages 47 years (2012) and 57 years (2022), stained with Mallory’s trichrome, silver impregnation, and Weigert’s/orcein. Objective magnification is ×40, and the scale bar represents 100 <math display="inline"><semantics> <mi mathvariant="sans-serif">μ</mi> </semantics></math>m.</p>
Full article ">Figure 4
<p>Histogram of the width of collagen fiber bundles observed in H&amp;E images.</p>
Full article ">Figure 5
<p>Histograms of segmented images stained with Mallory’s trichrome, H&amp;E, and Weigert’s/orcein, representing collagen type I (H&amp;E, Mallory), elastic fibers (Weigert’s/orcein), and collagen type III (silver). Green and blue—2012, red and black—2022.</p>
Full article ">Figure 6
<p>Estimated <span class="html-italic">k</span> and <math display="inline"><semantics> <mi>σ</mi> </semantics></math> values for two-parameter generalized Pareto distribution (GPD) with 95% confidence intervals (rectangle) for Mallory’s trichrome, H&amp;E, and Weigert’s/orcein-stained images. Green and blue—2012, red and black—2022.</p>
Full article ">Figure 7
<p>Estimated <span class="html-italic">k</span> and <math display="inline"><semantics> <mi>σ</mi> </semantics></math> values for the two-parameter generalized Pareto distribution (GPD) with 95% confidence intervals (rectangle) for silver impregnation. Green—2012, red—2022.</p>
Full article ">
11 pages, 2846 KiB  
Article
Efficacy of Vitamin B12 and Adenosine Triphosphate in Enhancing Skin Radiance: Unveiled with a Drug–Target Interaction Deep Learning-Based Model
by Hyeyeon Chun, Hyejin Lee, Jongwook Kim, Hyerin Yeo, Kyongeun Hyung, Dayoung Song, Moonju Kim, Seung-Hyun Jun and Nae-Gyu Kang
Curr. Issues Mol. Biol. 2024, 46(8), 9082-9092; https://doi.org/10.3390/cimb46080537 - 20 Aug 2024
Viewed by 625
Abstract
Skin radiance is crucial for enhancing facial attractiveness and is negatively affected by factors like hyperpigmentation and aging-related changes. Current treatments often lack comprehensive solutions for improving skin radiance. This study aimed to develop a cosmetic formula that enhances skin radiance by reducing [...] Read more.
Skin radiance is crucial for enhancing facial attractiveness and is negatively affected by factors like hyperpigmentation and aging-related changes. Current treatments often lack comprehensive solutions for improving skin radiance. This study aimed to develop a cosmetic formula that enhances skin radiance by reducing hyperpigmentation and improving skin regeneration by targeting specific receptors—the endothelin receptor type B (EDNRB) for hyperpigmentation and the adiponectin receptor 1 (ADIPOR1) for sagging and wrinkles. To achieve this, we used artificial intelligence technologies to screen and select ingredients with an affinity for EDNRB and ADIPOR1. Vitamin B12 (VitB12) was identified as a molecule that targets EDNRB, which is involved in melanogenesis. Adenosine triphosphate (ATP) targets ADIPOR1, which is associated with skin regeneration. VitB12 successfully inhibited intracellular calcium elevation and melanogenesis induced by endothelin-1. In contrast, ATP increased the mRNA expression of collagen and elastin and promoted wound healing. Moreover, the VitB12 and ATP complex significantly increased the expression of hyaluronan synthases, which are crucial for skin hydration. Furthermore, in human participants, the application of the VitB12 and ATP complex to one-half of the face significantly improved skin radiance, elasticity, and texture. Our findings provide valuable insights for the development of skincare formulations. Full article
(This article belongs to the Section Molecular Pharmacology)
Show Figures

Figure 1

Figure 1
<p>Inhibition of ET-1-evoked Ca<sup>2+</sup> signals by VitB12. (<b>a</b>) Representative fluorescent images of MNT-1 cells. Images are acquired before (F<sub>0</sub>) and after the reagent treatment for 1 min. From each experiment, microscopy images with the maximum fluorescence signal after the reagent treatment are shown (F<sub>max</sub>). Scale bar = 100 μm (<b>b</b>) Bar graph showing the decrease in fluorescent intensity after VitB12 or BQ788 treatment. The fluorescence intensities were quantified using Image J. The data are shown as mean ± s.e.m. *** <span class="html-italic">p</span> &lt; 0.001 by Tukey’s HSD test after the one-way ANOVA method.</p>
Full article ">Figure 2
<p>Inhibition of ET-1-induced melanogenesis by VitB12. (<b>a</b>) Melanin levels were measured after 48 h of incubation with ET-1. (<b>b</b>) Relative expression of MC1R mRNA. (<b>c</b>) Relative expression of MITF mRNA. Significance was determined by unpaired Student’s <span class="html-italic">t</span>-test (# <span class="html-italic">p</span> &lt; 0.05) or one-way ANOVA (* <span class="html-italic">p</span> &lt; 0.05, ** <span class="html-italic">p</span> &lt; 0.01, *** <span class="html-italic">p</span> &lt; 0.001).</p>
Full article ">Figure 3
<p>Effects of ATP on dermal fibrous component and wound healing (<b>a</b>) Relative expression of COL1 mRNA. (<b>b</b>) Relative expression of ELN mRNA. (<b>c</b>) Representative images of cell migration assay. Scale bar: 200 μm. (<b>d</b>) Bar graph showing the rate of cell migration (%) after 24 h. Data represent the mean ± s.e.m. Significance was determined by unpaired Student’s <span class="html-italic">t</span>-test (* <span class="html-italic">p</span> &lt; 0.05, *** <span class="html-italic">p</span> &lt; 0.001).</p>
Full article ">Figure 4
<p>Increase in the expression of hyaluronan synthases by the VitB12 and ATP complex. (<b>a</b>) Relative expression of HAS-2 mRNA. (<b>b</b>) Relative expression of HAS-3 mRNA. (<b>c</b>) Representative images of HaCaT cells stained with anti-HAS-2 antibody (green) and co-stained with nuclear stain (DAPI, blue). Cells were treated with the VitB12 and ATP complex for 48 h. Scale bar: 10 μm (<b>d</b>) Bar graph showing the fluorescent intensity of HAS-2 after treatment with the VitB12 and ATP complex. The images were analyzed using Image J. The fluorescence intensity is expressed in arbitrary units (A.U.). Data represent mean ± s.e.m., with n = 7 samples. Significance was determined by unpaired Student’s <span class="html-italic">t</span>-test (** <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 the active toner on skin radiance, elasticity, and texture. (<b>a</b>) Facial images of Derma-View before (0 weeks) and after (4 weeks) treatment with placebo toner (right side) and active toner (left side). (<b>b</b>) Facial images using the texture small filter from the Antera3D. (<b>c</b>) Change in radiance (gloss unit, %) measured by Glossmeter (<b>d</b>) Changes in elasticity (R2, %) measured by Cutometer. (<b>e</b>) Changes in texture (Ra, %) analyzed by Antera3D. The measured values, gloss unit (G.U.), R2 (%), and Ra (A.U.), are normalized to the average value at 0 weeks. Data represent means ± s.e.m. Statistical analysis was conducted using repeated measures ANOVA (* <span class="html-italic">p</span> &lt; 0.05, ** <span class="html-italic">p</span> &lt; 0.01, *** <span class="html-italic">p</span> &lt; 0.001).</p>
Full article ">
17 pages, 2384 KiB  
Article
Exploring the Potential of Saphenous Vein Grafts Ex Vivo: A Model for Intimal Hyperplasia and Re-Endothelialization
by Nur A’tiqah Haron, Mohamad Fikeri Ishak, Muhammad Dain Yazid, Ubashini Vijakumaran, Roszita Ibrahim, Raja Zahratul Azma Raja Sabudin, Hafiza Alauddin, Nur Ayub Md Ali, Hairulfaizi Haron, Muhammad Ishamuddin Ismail, Mohd Ramzisham Abdul Rahman and Nadiah Sulaiman
J. Clin. Med. 2024, 13(16), 4774; https://doi.org/10.3390/jcm13164774 - 14 Aug 2024
Viewed by 464
Abstract
Coronary artery bypass grafting (CABG) utilizing saphenous vein grafts (SVGs) stands as a fundamental approach to surgically treating coronary artery disease. However, the long-term success of CABG is often compromised by the development of intimal hyperplasia (IH) and subsequent graft failure. Understanding the [...] Read more.
Coronary artery bypass grafting (CABG) utilizing saphenous vein grafts (SVGs) stands as a fundamental approach to surgically treating coronary artery disease. However, the long-term success of CABG is often compromised by the development of intimal hyperplasia (IH) and subsequent graft failure. Understanding the mechanisms underlying this pathophysiology is crucial for improving graft patency and patient outcomes. Objectives: This study aims to explore the potential of an ex vivo model utilizing SVG to investigate IH and re-endothelialization. Methods: A thorough histological examination of 15 surplus SVG procured from CABG procedures at Hospital Canselor Tuanku Muhriz, Malaysia, was conducted to establish their baseline characteristics. Results: SVGs exhibited a mean diameter of 2.65 ± 0.93 mm with pre-existing IH averaging 0.42 ± 0.13 mm in thickness, alongside an observable lack of luminal endothelial cell lining. Analysis of extracellular matrix components, including collagen, elastin, and glycosaminoglycans, at baseline and after 7 days of ex vivo culture revealed no significant changes in collagen but demonstrated increased percentages of elastin and glycosaminoglycans. Despite unsuccessful attempts at re-endothelialization with blood outgrowth endothelial cells, the established ex vivo SVG IH model underscores the multifaceted nature of graft functionality and patency, characterized by IH presence, endothelial impairment, and extracellular matrix alterations post-CABG. Conclusions: The optimized ex vivo IH model provides a valuable platform for delving into the underlying mechanisms of IH formation and re-endothelialization of SVG. Further refinements are warranted, yet this model holds promise for future research aimed at enhancing graft durability and outcomes for CAD patients undergoing CABG. Full article
(This article belongs to the Section Vascular Medicine)
Show Figures

Graphical abstract

Graphical abstract
Full article ">Figure 1
<p>Patients’ demographic and clinical characteristics. (<b>A</b>) Gender distribution, (<b>B</b>) race distribution, (<b>C</b>) patients’ age group distribution, (<b>D</b>) number of CVD associated risk factors per patient, and (<b>E</b>) number of grafts needed per patient. Number of subjects, <span class="html-italic">n</span> = 35.</p>
Full article ">Figure 2
<p>Vein graft histological analysis. (<b>A</b>) Representative image of intimal hyperplasia. Arrows denote the medial thickness (MT) and intimal hyperplasia (IH) at 200× magnification. (<b>B</b>) Measurement of vessel diameter. (<b>C</b>) Measurement of intimal medial and intimal hyperplasia thickness were performed using Fiji software. (<b>D</b>) The sections of the saphenous vein (top panel) and umbilical cord (bottom panel) as a positive control were stained with (<b>i</b>,<b>iv</b>) CD31 + αSMA, (<b>ii</b>,<b>v</b>) CD309 + CD133, and (<b>iii</b>,<b>vi</b>) CD146 + CD45. Blue arrow shows the endothelial layer of SV and UA, respectively. (<b>E</b>) Percentage of extracellular matrix (%) using special staining to visualize thin and thick collagen, elastin, and glycosaminoglycans. Scale bar represents 200 µm and 2 mm for 40× and 200× magnification, respectively. All analysis was performed with <span class="html-italic">n</span> = 6.</p>
Full article ">Figure 3
<p>Analysis of intimal hyperplasia of surplus saphenous vein grafts ex vivo model. (<b>A</b>) Intimal hyperplasia (IH) thickness on day 1 and day 7 for direct method (DM) and scrapping method (SM). The triangle shows there is significant difference observed for group SM, * <span class="html-italic">p</span> &lt; 0.05. (<b>B</b>) Representative image of IH thickness for both DM and SM on day 1 and day 7. (<b>C</b>) Percentage of thin and thick collagen in day 1 and day 7 of DM and SM. (<b>D</b>) Percentage of elastin and glycosaminoglycans on day 1 and day 7 of DM and SM. Quantification was performed by using Fiji software. The scale bar represents 100 µm. Analysis was performed with <span class="html-italic">n</span> = 6. • representing each data point or replicates.</p>
Full article ">Figure 4
<p>Re-endothelialization of saphenous vein grafts intimal hyperplasia ex vivo model. (<b>A</b>) Intimal hyperplasia (IH) thickness on day 1 and day 7 for scrapping method (SM) and blood outgrowth endothelial cell (BOEC). The triangle shows there is a significant difference was observed for group BOEC, * <span class="html-italic">p</span> &lt; 0.05. (<b>B</b>) Representative image of IH thickness for both SM and BOEC on day 1 and day 7. (<b>C</b>) Percentage of thin and thick collagen on day 1 and day 7 for group SM and BOEC. (<b>D</b>) Percentage of elastin and glycosaminoglycans on day 1 and day 7 of SM and BOEC groups. Quantification was performed using Fiji software. The scale bar represents 100 µm. Analysis was performed with <span class="html-italic">n</span> = 6. • representing each data point or replicates.</p>
Full article ">
17 pages, 4178 KiB  
Article
Elastin-Derived Peptide-Based Hydrogels as a Potential Drug Delivery System
by Othman Al Musaimi, Keng Wooi Ng, Varshitha Gavva, Oscar M. Mercado-Valenzo, Hajira Banu Haroon and Daryl R. Williams
Gels 2024, 10(8), 531; https://doi.org/10.3390/gels10080531 - 12 Aug 2024
Viewed by 838
Abstract
A peptide-based hydrogel sequence was computationally predicted from the Ala-rich cross-linked domains of elastin. Three candidate peptides were subsequently synthesised and characterised as potential drug delivery vehicles. The elastin-derived peptides are Fmoc-FFAAAAKAA-NH2, Fmoc-FFAAAKAA-NH2, and Fmoc-FFAAAKAAA-NH2. All three [...] Read more.
A peptide-based hydrogel sequence was computationally predicted from the Ala-rich cross-linked domains of elastin. Three candidate peptides were subsequently synthesised and characterised as potential drug delivery vehicles. The elastin-derived peptides are Fmoc-FFAAAAKAA-NH2, Fmoc-FFAAAKAA-NH2, and Fmoc-FFAAAKAAA-NH2. All three peptide sequences were able to self-assemble into nanofibers. However, only the first two could form hydrogels, which are preferred as delivery systems compared to solutions. Both of these peptides also exhibited favourable nanofiber lengths of at least 1.86 and 4.57 µm, respectively, which are beneficial for the successful delivery and stability of drugs. The shorter fibre lengths of the third peptide (maximum 0.649 µm) could have inhibited their self-assembly into the three-dimensional networks crucial to hydrogel formation. Full article
(This article belongs to the Special Issue Recent Advances in Gels Engineering for Drug Delivery (2nd Edition))
Show Figures

Figure 1

Figure 1
<p>Elastin protein sequence. Colours and underlining represent the abundance of repeated sequences of the Ala-rich cross-linking domain.</p>
Full article ">Figure 2
<p>Flory–Huggins interaction parameter Chi (χ) calculated computationally using Materials Studio software (initial screening).</p>
Full article ">Figure 3
<p>Flory–Huggins interaction parameter Chi (χ) calculated computationally using Materials Studio software versus VPGVG.</p>
Full article ">Figure 4
<p>Predicted molecular structures of EDP-1, EDP-2, and EDP-3 (along with H-bond distance 0.400 Å) using I-TASSER [<a href="#B59-gels-10-00531" class="html-bibr">59</a>,<a href="#B60-gels-10-00531" class="html-bibr">60</a>,<a href="#B61-gels-10-00531" class="html-bibr">61</a>].</p>
Full article ">Figure 5
<p>Chemical structure of the three peptide sequences investigated in this work.</p>
Full article ">Figure 6
<p>Critical aggregation concentration (CAC) of the selected peptides (<b>left</b>). Image of the three peptides to show the hydrogel formation (<b>right</b>). (<b>A</b>) EDP-1, (<b>B</b>) EDP-2, (<b>C</b>) EDP-3. (<b>D</b>) Digital images of the formed EDP-1 and EDP-2 hydrogels. Sigmoidal function fitted within the yellow region, where the red tangent shows the CAC.</p>
Full article ">Figure 7
<p>TEM morphology of the three peptides (scale bar= 100 nm).</p>
Full article ">Figure 8
<p>CD spectra for the three peptides investigated in this work.</p>
Full article ">Figure 9
<p>Shear modulus versus time plots for peptides EDP-1 and EDP-2 showing the storage modulus (G′) and loss modules (G″). Peptide solution concentration: 1% (wt/v).</p>
Full article ">
26 pages, 17613 KiB  
Article
RiboScreenTM Technology Delivers a Ribosomal Target and a Small-Molecule Ligand for Ribosome Editing to Boost the Production Levels of Tropoelastin, the Monomeric Unit of Elastin
by Bjoern Wimmer, Jan Schernthaner, Genevieve Edobor, Andreas Friedrich, Katharina Poeltner, Gazmend Temaj, Marlies Wimmer, Elli Kronsteiner, Mara Pichler, Hanna Gercke, Ronald Huber, Niklas Kaefer, Mark Rinnerthaler, Thomas Karl, Jan Krauß, Thomas Mohr, Christopher Gerner, Helmut Hintner, Michael Breitenbach, Johann W. Bauer, Christin Rakers, Daniel Kuhn, Joerg von Hagen, Norbert Müller, Adriana Rathner and Hannelore Breitenbach-Kolleradd Show full author list remove Hide full author list
Int. J. Mol. Sci. 2024, 25(15), 8430; https://doi.org/10.3390/ijms25158430 - 1 Aug 2024
Viewed by 896
Abstract
Elastin, a key structural protein essential for the elasticity of the skin and elastogenic tissues, degrades with age. Replenishing elastin holds promise for anti-aging cosmetics and the supplementation of elastic activities of the cardiovascular system. We employed RiboScreenTM, a technology for [...] Read more.
Elastin, a key structural protein essential for the elasticity of the skin and elastogenic tissues, degrades with age. Replenishing elastin holds promise for anti-aging cosmetics and the supplementation of elastic activities of the cardiovascular system. We employed RiboScreenTM, a technology for identifying molecules that enhance the production of specific proteins, to target the production of tropoelastin. We make use of RiboScreenTM in two crucial steps: first, to pinpoint a target ribosomal protein (TRP), which acts as a switch to increase the production of the protein of interest (POI), and second, to identify small molecules that activate this ribosomal protein switch. Using RiboScreenTM, we identified ribosomal protein L40, henceforth eL40, as a TRP switch to boost tropoelastin production. Drug discovery identified a small-molecule hit that binds to eL40. In-cell treatment demonstrated activity of the eL40 ligand and delivered increased tropoelastin production levels in a dose-dependent manner. Thus, we demonstrate that RiboScreenTM can successfully identify a small-molecule hit capable of selectively enhancing tropoelastin production. This compound has the potential to be developed for topical or systemic applications to promote skin rejuvenation and to supplement elastic functionality within the cardiovascular system. Full article
Show Figures

Figure 1

Figure 1
<p>Pictographic representation of RiboScreen<sup>TM</sup> Technology. On the left-hand side, the first step of the RiboScreen<sup>TM</sup> technology is shown in Panel (1). Two tools are used. First is a screening library of yeast vehicles, each of which is depleted for one of the eighty eukaryotic ribosomal proteins (RPs), the ribosomal variant strain (RVS) screening library. In cyan, a depleted ribosome for a ribosomal protein (white oval) is shown. In grey, a wild-type ribosome is shown in a naïve vehicle. The second tool is a dual luciferase assay to monitor protein expression levels of the protein of interest (POI, magenta), the cellular target. The reporter of the POI carries a C-terminally tagged Firefly luciferase (FF, green), and a Renilla (REN, yellow) luciferase reporter serves as internal control. In the middle section, panel (1) presents the identification of the actual drug target, the target ribosomal protein (TRP) (red arrow), demarcated by its altered functional availability (white oval), which leads to an increase in the production level of the POI, (panel (1), right). Panel (2) shows orthologous yeast and human TRPs (grey ovals) on the left, which serve as protein baits for structure-based virtual screening to identify small-molecule binders. The compounds identified in this way from the screening of a library of small molecules are then computationally docked into the binding sites of the TRP and scored for their binding affinities to identify potential ligands. The virtual screening results are post-processed to select the representative hits (small colored circles) for further analysis (panel (2), left). Panel (3) shows the experimental validation of the hit molecules using RiboScreen<sup>TM</sup> technology. Naïve vehicles, equipped with identical protein reporters as employed in the initial screening (POI-FF and REN), are used (Panel (3), left), but here treated with TRP ligands (red dotted arrow). Hit compounds are identified based on their demonstrated activity to boost POI production levels. In Panel (4), further steps of drug development are listed for completeness.</p>
Full article ">Figure 2
<p>Exon structure of ELN mRNA encoding Tropoelastin. On top, the ELN exon structure is depicted, with the individual exons numbered as present in the human mRNA (modified from [<a href="#B47-ijms-25-08430" class="html-bibr">47</a>]). The bottom shows isoform 6 splice variant, lacking several exons, among them exon 26A.</p>
Full article ">Figure 3
<p>Reporter protein production levels in the ribosomal variant strain (RVS) screening library. This plot correlates, in a single data point for each ribosomal variant strain vehicle, the mean protein production level of the internal control REN reporter, normalized to the mean wild-type signal on the <span class="html-italic">y</span>-axis and the mean protein production level of the TE-FF reporter normalized to the mean wild type signal on the <span class="html-italic">x</span>-axis. A circle centered at a 100% expression level (i.e., 1) of both reporter proteins in the wild-type vehicle (marked in yellow) and with a radius of 50% difference in expression level, corresponding to the 2-fold standard deviation of the overall mean of measurements in all RVS contains the TE-FF and REN reporter expression profiles from the majority of screening vehicles. The data point obtained for the RVS, carrying a depletion for eL40, as encoded by RPL40A (marked in green), but not as encoded by RPL40B (marked in grey), signals that the modification of this functional availability of eL40 boosts TE-FF production 2.8-fold while leaving REN expression levels unaltered. This identifies eL40 as a potential target ribosomal protein (TRP) for customized boost of protein production levels of tropoelastin. Minor TRP species, where REN expression levels are also affected, are represented by the reporter protein expression levels observed in RVS depleted for ribosomal proteins uS15 and uL2 as encoded by RPS13 and by RPL2B (marked in grey). Note that the gene nomenclature and protein nomenclature are different for ribosomal proteins in yeast [<a href="#B40-ijms-25-08430" class="html-bibr">40</a>].</p>
Full article ">Figure 4
<p>Ribosomal protein eL40 is conserved between yeast and humans in sequence, structure, and topological position on the ribosome. (<b>A</b>) Sequence comparison between the two yeast paralogous eL40 proteins, as well as that to their human orthologue, is shown, with the N-terminal ubiquitin tag included. (<b>B</b>) eL40 resides in the 60S subunit of eukaryotic ribosomes, yeast shown to the left (PDB code 7B7D) and human shown to the right (PDB code 6QZP), and ribbon models of yeast and human eL40 show their integration into the 60S subunit, with a close-up in (<b>C</b>). In (<b>D</b>), a visualization of the domain architecture of eL40 depicts the non-globular extension of the protein, which anchors the protein within the rRNA scaffold of the ribosome (grey box). The central domain, as well as the N-terminus and C-terminus, remains accessible for other intermolecular interactions.</p>
Full article ">Figure 5
<p>The spatial–functional conservation of eL40 on yeast and human ribosome. (<b>A</b>) During the generation of the translation-competent ribosomal subunits, eL40 (blue) is the penultimate ribosomal protein to arrive at the 60S subunit, as shown from the inter-subunit side for yeast on the left (PDB code 7B7D) and for human on the right (PDB code 6QZB). eL40 is positioned atop the Sarcin–Ricin loop (yellow). Upon the arrival of the last ribosomal protein to be incorporated into the 60S subunit, uL16 (magenta), the 60 S subunit, attains its final, translation-competent configuration, which is ready to form the translation-competent ribosome by joining the mRNA-associated 40S subunit. (<b>B</b>) On the translation-competent ribosome, yeast to the left (PDB code 5JUU) and human to the right (PDB code 6Z6N) eL40, in close proximity to the small subunit protein eS31 and the Sarcin–Ricin loop of the 60S subunit, respectively, form the landing platform (factor binding site), for elongation factors (EF), which drive the rate of protein synthesis. eEF2, in exemplary form, is shown binding to the factor binding site.</p>
Full article ">Figure 6
<p>Ligand binding sites of eL40 and the visualization of binding pockets. (<b>A</b>) Coulombic surface representation of yeast (left, PDB code 3J77) and human (right, PDB code 6XA1) eL40 proteins, overlaid on their secondary structures. (<b>B</b>) Potential small-molecule binding pockets on yeast eL40, to the left, (yellow, near Ser94, PDB code 3J77) and human eL40, to the right (yellow, near Asp92, PDB code 6EK0), were identified by druggability analysis (Sitemap [<a href="#B72-ijms-25-08430" class="html-bibr">72</a>]) using the amino acid sequence of the eL40 proteins from the ribosome X-ray structures (minus rRNA). These pockets are adjacent (yellow patches) and, depending on protein dynamics, could potentially form a larger, common binding cavity.</p>
Full article ">Figure 7
<p>Validation of small-molecule eL40 ligands for customized boost of tropoelastin production levels. Candidate small-molecule ligands were tested in a range of concentrations from 1 nM to 100 µM. Firefly tagged tropoelastin luciferase signals (blue) and Renilla signals (orange) obtained upon treatments are shown normalized to the untreated control. The compounds are listed per increasing number, and C17 showed a dose-dependent and significant response at 100 μM, with a 1.7-fold boost in tropoelastin reporter expression, and Renilla expression was unaltered (middle panel, yellow rectangle). Examples of minor candidate eL40 activators are C7 and C25 (mute yellow). A representative of non-active eL40 ligands is C22 (lower panel, white).</p>
Full article ">Figure 8
<p>Molecular structures of rpL40 ligands. UPAC nomenclature, molecular structure and space filling models of compounds C17, C7, C25, and C22 are shown.</p>
Full article ">Figure 9
<p>Molecular interactions of C17 with yeast and human rpL40. (<b>A</b>) Predicted docking poses for C17 with yeast (PDB code 3J77) and human (PDB code 6EK0) eL40 structures. This figure illustrates the predicted binding modes of compound C17 to both yeast and human eL40 orthologues. On the left, the complex structure with the yeast eL40 is shown, where the carbon atoms of C17 are highlighted in green. On the right, the complex structure with the human eL40 is displayed, with C17’s carbon atoms in cyan. The molecular surfaces of both structures are depicted in panel (<b>A</b>). Panel (<b>B</b>) highlights the key interactions: the charged interactions between the piperidine group of C17 and the aspartate 92 residue, and the hydrogen bond between the phenolic group of C17 and the backbone of alanine 107.</p>
Full article ">
15 pages, 16991 KiB  
Article
ADAR1 Is Essential for Smooth Muscle Homeostasis and Vascular Integrity
by Dunpeng Cai and Shi-You Chen
Cells 2024, 13(15), 1257; https://doi.org/10.3390/cells13151257 - 26 Jul 2024
Viewed by 757
Abstract
Vascular smooth muscle cells (VSMCs) play a critical role in maintaining vascular integrity. VSMC dysfunction leads to numerous vascular diseases. Adenosine deaminases acting on RNA 1 (ADAR1), an RNA editing enzyme, has shown both RNA editing and non-editing functions. Global deletion of ADAR1 [...] Read more.
Vascular smooth muscle cells (VSMCs) play a critical role in maintaining vascular integrity. VSMC dysfunction leads to numerous vascular diseases. Adenosine deaminases acting on RNA 1 (ADAR1), an RNA editing enzyme, has shown both RNA editing and non-editing functions. Global deletion of ADAR1 causes embryonic lethality, but the phenotype of homozygous ADAR1 deletion specifically in SMCs (ADAR1sm-/-) remains to be determined. By crossing ADAR1fl/fl mice with Myh11-CreERT2 mice followed by Tamoxifen induction, we found that ADAR1sm-/- leads to lethality in adult mice 14 days after the induction. Gross examination revealed extensive hemorrhage and detrimental vascular damage in different organs. Histological analyses revealed destruction of artery structural integrity with detachment of elastin laminae from VSMCs in ADAR1sm-/- aortas. Furthermore, ADAR1sm-/- resulted in severe VSMC apoptosis and mitochondrial dysfunction. RNA sequencing analyses of ADAR1sm-/- aorta segments demonstrated profound transcriptional alteration of genes impacting vascular health including a decrease in fibrillin-1 expression. More importantly, ADAR1sm-/- disrupts the elastin and fibrillin-1 interaction, a molecular event essential for artery structure. Our results indicate that ADAR1 plays a critical role in maintaining SMC survival and vascular stability and resilience. Full article
(This article belongs to the Special Issue Role of Vascular Smooth Muscle Cells in Cardiovascular Disease)
Show Figures

Graphical abstract

Graphical abstract
Full article ">Figure 1
<p>ADAR1 deletion in smooth muscle cells leads to lethality due to vascular defects in mice. Myh11-CreERT2 (WT) and ADAR1sm-/- mice were injected with tamoxifen (1 mg/day, i.p. for 5 days). (<b>A</b>) Mouse survival rates post-tamoxifen administration. (<b>B</b>) Incidences of hemorrhage in various organs of ADAR1sm-/- mice at 13 days post-injection; n = 6. (<b>C</b>) Brain gross images from control and ADAR1sm-/- mice 13 days post-injection. Enlarged panel 1 on the right displays a significant blood clot beneath the brain, indicative of brain herniation. Enlarged panel 2 indicates hemorrhage at the circle of Willis. (<b>D</b>) Mouse brain cross-sections show ventricular hemorrhage in ADAR1sm-/- mice. H&amp;E staining reveals small artery dissection causing hemorrhage in ADAR1sm-/- mouse brain hippocampus areas. Scale bar: 50 μm. Lower panels showing enlarged images in the rectangle boxes in the upper panels. Arrows: normal artery in WT mouse brain and diseased artery in brain of ADAR1sm-/- mice.</p>
Full article ">Figure 2
<p>ADAR1 deletion in smooth muscle cells damages the integrity of elastin fibers. Myh11-CreERT2 (WT) and ADAR1sm-/- mice were injected with tamoxifen (1 mg/day, i.p. for 5 days). (<b>A</b>) H&amp;E and EVG staining of thoracic aorta sections of WT and ADAR1sm-/- mice 13 days post-tamoxifen injection. * Breaks in elastin fibers. Scale bar: 50 μm. (<b>B</b>) Transmission electron microscopy (TEM) images of WT and ADAR1sm-/- mouse aorta segments 13 days post-tamoxifen injection. Green arrows indicate the thickness of elastin lamina, and red arrows the thickness of SMC layer. Scale Bar: 4 μm. (<b>C</b>) Quantification of the elastin degradation index in WT and ADAR1sm-/- mouse aortas. <sup>&amp;</sup> <span class="html-italic">p</span> &lt; 0.001 vs. WT, n = 6. (<b>D</b>) Quantification of the SMC layer and elastin laminae thicknesses in WT and ADAR1sm-/- mouse aortas. <sup>&amp;</sup> <span class="html-italic">p</span> &lt; 0.01 vs. WT, n = 6. (<b>E</b>) Quantification of hemorrhage events in aortas of WT and ADAR1sm-/- mice. <sup>&amp;</sup> <span class="html-italic">p</span> &lt; 0.001 vs. WT, n = 6.</p>
Full article ">Figure 3
<p>ADAR1 deletion causes detachment of smooth muscle layers from elastic fibers. Myh11-CreERT2 (WT) and ADAR1sm-/- mice were injected with tamoxifen (1 mg/day, i.p. for 5 days). Thirteen days later, descending thoracic aortas were analyzed by electron microscopy. (<b>A</b>) Scanning electron microscopy images of WT and ADAR1sm-/- mouse aorta media layers, * gaps between elastin lamina and SMC layers (upper panels). The microstructure of the extracellular matrix (ECM) in the descending aorta was observed by transmission electron microscopy. ‘E’ indicates elastin lamina; ‘M’ indicates SMCs; asterisks (*) highlight the gaps between elastin lamina and SMCs (lower panels). (<b>B</b>) Quantification of the percentage of medial layers with gaps between elastin lamina and SMCs. <sup>&amp;</sup> <span class="html-italic">p</span> &lt; 0.001 vs. WT, n = 6. (<b>C</b>) Scanning electron microscopy (SEM) of the descending aorta segments after formic acid digestion and freeze-drying, which preserved only the elastin. Red arrows indicate elastin fibers in WT tissues, which is absent in ADAR1SM-/- aortas, shown in longitudinal view. The transverse view shows the degradation of elastin lamina in ADAR1sm-/- aortas. (<b>D</b>) Quantification of the degradation of elastin fibers connecting different layers of elastin laminae. <sup>&amp;</sup> <span class="html-italic">p</span> &lt; 0.001 vs. WT, n = 6.</p>
Full article ">Figure 4
<p>RNA-seq analyses reveal critical signaling pathways and genes altered by ADAR1sm-/- in aortic SMCs. Myh11-CreERT2 (WT) and ADAR1sm-/- mice were injected with tamoxifen (1 mg/day, i.p. for 5 days). Thirteen days later, SMCs from thoracic aortas were isolated, total RNAs were extracted, and bulk RNA sequencing was conducted. (<b>A</b>) Volcano plot shows that numerous genes were up- (red dots) or downregulated (green dots) by ADAR1sm-/- (adjusted <span class="html-italic">p</span>-value &lt; 0.01, log2 fold change &gt; 1). Six independent samples from each group were analyzed. (<b>B</b>) KEGG pathway enrichment analyses of differentially expressed mRNAs in ADAR1SM-/- mice compared with WT mice, based on transcriptomic data from 6 mouse aortas per group. (<b>C</b>) Editing site distribution in aortic SMCs based on sequencing data. Most editing sites are within gene regions, particularly the introns and 3′ untranslated regions, with ADAR1sm-/- significantly reducing editing site counts in these regions. * <span class="html-italic">p</span> &lt; 0.01 vs. WT in each group, n = 3. (<b>D</b>) Approximately half of the editing sites are in the repetitive regions, primarily within the short interspersed nuclear elements, long terminal repeats, and long interspersed nuclear elements. ADAR1sm-/- substantially abolishes the editing sites in all these regions. * <span class="html-italic">p</span> &lt; 0.01 vs. WT in each group, n = 3.</p>
Full article ">Figure 5
<p>ADAR1sm-/- causes SMC apoptosis in aortas via mitochondrial pathway. Myh11-CreERT2 (WT) and ADAR1sm-/- mice were injected with tamoxifen (1 mg/day, i.p. for 5 days). Thirteen days post-tamoxifen injection, descending thoracic aortas were isolated and processed. (<b>A</b>) TUNEL staining of aorta sections from WT and ADAR1sm-/- mice. Scale bar: 40 μm. (<b>B</b>) Quantification of TUNEL-positive cells. * <span class="html-italic">p</span> &lt; 0.001 vs WT, n = 6. (<b>C</b>) Cleaved Caspase-3 (Cl-caspase3), cleaved PARP (Cl-PARP), BCL-1, Bax, cytochrome C, and GAPDH levels in WT and ADAR1sm-/- mouse aortic media (adventitia removed) analyzed by Western blot. (<b>D</b>) Flow cytometry assessment of PI and Annexin V staining of the single-cell suspensions from aorta media layers. (<b>E</b>) Quantification of PI- and Annexin V-positive cells (percentages of total cells) for each group. * <span class="html-italic">p</span> &lt; 0.001 vs WT, n = 6. (<b>F</b>) Flow cytometry assessment of JC-1 staining of single-cell suspensions from aorta media layers. Red dots indicate high mitochondrial membrane potential in WT SMCs, and blue dots show low membrane potential in ADAR1sm-/- SMCs. (<b>G</b>) Quantification of JC-1 staining, shown as fold changes. * <span class="html-italic">p</span> &lt; 0.001 vs WT, n = 6. (<b>H</b>) TEM images of aorta media layers of WT and ADAR1sm-/- mice 14 days post-tamoxifen injection. Red arrows indicate mitochondria. (<b>I</b>) Quantification of mitochondria numbers per square micron, * <span class="html-italic">p</span> &lt; 0.001 vs WT, n = 6.</p>
Full article ">Figure 6
<p>ADAR1 deletion in smooth muscle disrupts the fibrillin–1-elastin interaction in aorta media layers. Myh11-CreERT2 (WT) and ADAR1sm-/- mice were injected with tamoxifen (1 mg/day, i.p. for 5 days). Thirteen days post-tamoxifen injection, descending thoracic aortas were isolated and processed. (<b>A</b>,<b>B</b>) In situ proximity ligation assays (PLAs) were conducted to assess fibrillin-1–elastin interactions in WT and ADAR1sm-/- mouse aortas. DAPI stains the nuclei. Scale bar: 20 μm. (<b>B</b>) Quantification of PLA signals (percentages of aorta areas). * <span class="html-italic">p</span> &lt; 0.001 vs WT, n = 6. (<b>C</b>) Coimmunoprecipitation assays detecting fibrillin-1–elastin interactions in mouse aorta media layers. Normal IgG (control), fibrillin-1, or elastin antibodies were used for immunoprecipitation (IP), followed by immunoblotting (IB) with antibodies against fibrillin-1 and elastin. (<b>D</b>) High-magnification TEM images of the aorta media layers of WT and ADAR1sm-/- mice. Asterisks (*) indicate fibrillin microfibrils, which were quantified as percentages of aorta areas. * <span class="html-italic">p</span> &lt; 0.001 vs WT, n = 6. (<b>E</b>) In situ proximity ligation assays (PLAs) confirmed fibrillin-1–elastin interactions in healthy human (Ctrl) and Marfan Syndrome (MFS) patients’ aorta media layers. DAPI stains the nuclei. Scale bar: 20 μm. (<b>F</b>) Quantification of PLA signal (percentages of aorta areas) shown in (<b>E</b>). * <span class="html-italic">p</span> &lt; 0.001 vs Ctrl, n = 6.</p>
Full article ">
16 pages, 6398 KiB  
Article
Comparative Analysis of Decellularization Methods for the Production of Decellularized Umbilical Cord Matrix
by Yang Li, Yang Zhang and Guifeng Zhang
Curr. Issues Mol. Biol. 2024, 46(7), 7686-7701; https://doi.org/10.3390/cimb46070455 - 19 Jul 2024
Viewed by 677
Abstract
The importance of decellularized extracellular matrix (dECM) as a natural biomaterial in tissue engineering and regenerative medicine is rapidly growing. The core objective of the decellularization process is to eliminate cellular components while maximizing the preservation of the ECM’s primary structure and components. [...] Read more.
The importance of decellularized extracellular matrix (dECM) as a natural biomaterial in tissue engineering and regenerative medicine is rapidly growing. The core objective of the decellularization process is to eliminate cellular components while maximizing the preservation of the ECM’s primary structure and components. Establishing a rapid, effective, and minimally destructive decellularization technique is essential for obtaining high-quality dECM to construct regenerative organs. This study focused on human umbilical cord tissue, designing different reagent combinations for decellularization protocols while maintaining a consistent processing time. The impact of these protocols on the decellularization efficiency of human umbilical cord tissue was evaluated. The results suggested that the composite decellularization strategy utilizing trypsin/EDTA + Triton X-100 + sodium deoxycholate was the optimal approach in this study for preparing decellularized human umbilical cord dECM. After 5 h of decellularization treatment, most cellular components were eliminated, confirmed through dsDNA quantitative detection, hematoxylin and eosin (HE) staining, and DAPI staining. Meanwhile, Masson staining, periodic acid-silver methenamine (PASM) staining, periodic acid-Schiff (PAS) staining, and immunofluorescent tissue section staining results revealed that the decellularized scaffold retained extracellular matrix components, including collagen and glycosaminoglycans (GAGs). Compared to native umbilical cord tissue, electron microscopy results demonstrated that the microstructure of the extracellular matrix was well preserved after decellularization. Furthermore, Fourier-transform infrared spectroscopy (FTIR) findings indicated that the decellularization process successfully retained the main functional group structures of extracellular matrix (ECM) components. The quantitative analysis of collagen, elastin, and GAG content validated the advantages of this decellularization process in preserving and purifying ECM components. Additionally, it was confirmed that this decellularized matrix exhibited no cytotoxicity in vitro. This study achieved short-term decellularization preparation for umbilical cord tissue through a combined decellularization strategy. Full article
Show Figures

Graphical abstract

Graphical abstract
Full article ">Figure 1
<p>The dsDNA contents for each decellularization protocol. Data are 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.</p>
Full article ">Figure 2
<p>Immunohistochemical staining for each decellularization protocol. Scale bar: 50 μm. Data are mean ± SD; <span class="html-italic">n</span> = 3; ns: not significant; *: <span class="html-italic">p</span> &lt; 0.05; **: <span class="html-italic">p</span> &lt; 0.01; ***: <span class="html-italic">p</span> &lt; 0.001.</p>
Full article ">Figure 3
<p>Immunofluorescence staining for each decellularization protocol. Scale bar: 100 μm. Data are mean ± SD; <span class="html-italic">n</span> = 3; **: <span class="html-italic">p</span> &lt; 0.01; ***: <span class="html-italic">p</span> &lt; 0.001.</p>
Full article ">Figure 4
<p>SEM analysis for each decellularization protocol. Scale bar: 50 μm. Orange arrows: cell structure.</p>
Full article ">Figure 5
<p>FTIR spectra for each decellularization protocol. (<b>A</b>) FTIR spectra. (<b>B</b>) Feature peak area of collagen spectra. (<b>C</b>) Feature peak area of GAG spectra. Data are 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.</p>
Full article ">Figure 6
<p>Quantitative analysis of collagen, elastin, and GAGs for each decellularization protocol. Data are mean ± SD; <span class="html-italic">n</span> = 3; ns: not significant; *: <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 7
<p>Cytotoxicity (in vitro) of each decellularization protocol. (<b>A</b>) Cell morphology observation. Scale bar: 100 μm. (<b>B</b>) Cell survival rate. Data are mean ± SD; <span class="html-italic">n</span> = 3; ns: not significant.</p>
Full article ">
14 pages, 3212 KiB  
Article
A Single Injection of ADRCs Does Not Prevent AAA Formation in Rats in a Randomized Blinded Design
by Egle Kavaliunaite, Pratibha Dhumale, Charlotte Harken Jensen, Søren P. Sheikh, Jes S. Lindholt and Jane Stubbe
Int. J. Mol. Sci. 2024, 25(14), 7591; https://doi.org/10.3390/ijms25147591 - 10 Jul 2024
Viewed by 669
Abstract
There is a pressing need for alternative medical treatments for abdominal aortic aneurysms (AAAs). Mesenchymal regenerative cells derived from adipose tissue (ADRCs) have shown potential in modulating the inflammation and immune responses that drive AAA progression. We hypothesized that ADRCs could reduce inflammation [...] Read more.
There is a pressing need for alternative medical treatments for abdominal aortic aneurysms (AAAs). Mesenchymal regenerative cells derived from adipose tissue (ADRCs) have shown potential in modulating the inflammation and immune responses that drive AAA progression. We hypothesized that ADRCs could reduce inflammation and preserve vascular integrity, potentially slowing the progression of AAA. In our study, subcutaneous adipose tissue was harvested from male Sprague Dawley rats, from which ADRCs were isolated. AAA was induced in these rats using intraluminal porcine pancreatic elastase, followed by intravenous administration of either ADRCs (106 cells) or saline (0.1 mL). We monitored the progression of AAA through weekly ultrasound, and the rats were sacrificed on day 28 for histological analysis. Our results showed no significant difference in the inner abdominal aortic diameter at day 28 between the control group (172% ± 73%, n = 17) and the ADRC-treated group (181% ± 75%, n = 15). Histological analyses of AAA cross-sections also revealed no significant difference in the infiltration of neutrophils or macrophages between the two groups. Furthermore, the integrity and content of elastin in the tunica media were similar between groups. These findings indicate that a single injection of ADRCs does not inhibit the development of AAA in rats in a randomized blinded study. Full article
Show Figures

Figure 1

Figure 1
<p>The treatment effect of ADRCs on abdominal aortic aneurysm progression. (<b>a</b>) Representative examples of ultrasound scans of AAA measurements in the control group and ADRC treatment group prior to surgery and 7 days, 14 days, 21 days, and 28 days post-surgery. C: Circumference (green circle), mm; A: area of green circle mm<sup>2</sup>; d1: inner to inner diameter (yellow 1+: mm); d2: Diameter (mm). (<b>b</b>) Percentage increase of AAA after initiating AAA till 28 days after treatment in the control group (n = 17) and the ADRC treatment group (n = 15), as well as percentage increase of suprarenal aorta in the control group (n = 17) and the suprarenal aorta in the ADRC treatment group (n = 15). Values are presented as mean ± standard deviation.</p>
Full article ">Figure 2
<p>ADRC effect on elastin in the aneurysm wall. (<b>a</b>) Representative micrograph of aneurysmal cross-sections in the control group and ADRC-treated group. (<b>b</b>) The percentage of elastin in tunica media on day 28 in the control group and the treatment group (n = 17/15). (<b>c</b>) Example image of a cross-sectional abdominal aortic aneurysm divided into eight areas. (<b>d</b>) Below, micrographs representing the score of elastin degradation (1 = preserved elastin architecture; 4 = total disruption of complete concentric elastin lamellae). (<b>e</b>) The average mean score of elastin degradation in tunica media in the control group and the treatment group (n = 17/15). Values are presented as mean ± standard deviation. Each dot represents the quantification of the individual rats.</p>
Full article ">Figure 2 Cont.
<p>ADRC effect on elastin in the aneurysm wall. (<b>a</b>) Representative micrograph of aneurysmal cross-sections in the control group and ADRC-treated group. (<b>b</b>) The percentage of elastin in tunica media on day 28 in the control group and the treatment group (n = 17/15). (<b>c</b>) Example image of a cross-sectional abdominal aortic aneurysm divided into eight areas. (<b>d</b>) Below, micrographs representing the score of elastin degradation (1 = preserved elastin architecture; 4 = total disruption of complete concentric elastin lamellae). (<b>e</b>) The average mean score of elastin degradation in tunica media in the control group and the treatment group (n = 17/15). Values are presented as mean ± standard deviation. Each dot represents the quantification of the individual rats.</p>
Full article ">Figure 3
<p>Infiltration of neutrophil cells in AAA wall. (<b>a</b>) Representative micrographs of MPO staining from aneurysmal tissue in the control and the ADRC treatment groups; left micrographs 20× magnification, right 40×. (<b>b</b>) Number of MPO-positive cells per mm<sup>2</sup> in the two groups. Each dot represents the quantification of the individual rats. Values are presented as mean ± standard deviation (n = 17/15).</p>
Full article ">Figure 4
<p>Infiltrating macrophages into the aneurysmal wall. (<b>a</b>) A representative micrograph of CD68 staining from aneurysmal tissue in the control and the ADRC-treated group. (<b>b</b>) Percentage of CD68 cells’ positive stained area (n = 16/13). Each dot represents the quantification of the individual rats. Values are presented as mean ± standard deviation.</p>
Full article ">
35 pages, 10426 KiB  
Review
Bridging Nature and Engineering: Protein-Derived Materials for Bio-Inspired Applications
by Taufiq Nawaz, Liping Gu, Jaimie Gibbons, Zhong Hu and Ruanbao Zhou
Biomimetics 2024, 9(6), 373; https://doi.org/10.3390/biomimetics9060373 - 20 Jun 2024
Cited by 2 | Viewed by 1147
Abstract
The sophisticated, elegant protein-polymers designed by nature can serve as inspiration to redesign and biomanufacture protein-based materials using synthetic biology. Historically, petro-based polymeric materials have dominated industrial activities, consequently transforming our way of living. While this benefits humans, the fabrication and disposal of [...] Read more.
The sophisticated, elegant protein-polymers designed by nature can serve as inspiration to redesign and biomanufacture protein-based materials using synthetic biology. Historically, petro-based polymeric materials have dominated industrial activities, consequently transforming our way of living. While this benefits humans, the fabrication and disposal of these materials causes environmental sustainability challenges. Fortunately, protein-based biopolymers can compete with and potentially surpass the performance of petro-based polymers because they can be biologically produced and degraded in an environmentally friendly fashion. This paper reviews four groups of protein-based polymers, including fibrous proteins (collagen, silk fibroin, fibrillin, and keratin), elastomeric proteins (elastin, resilin, and wheat glutenin), adhesive/matrix proteins (spongin and conchiolin), and cyanophycin. We discuss the connection between protein sequence, structure, function, and biomimetic applications. Protein engineering techniques, such as directed evolution and rational design, can be used to improve the functionality of natural protein-based materials. For example, the inclusion of specific protein domains, particularly those observed in structural proteins, such as silk and collagen, enables the creation of novel biomimetic materials with exceptional mechanical properties and adaptability. This review also discusses recent advancements in the production and application of new protein-based materials through the approach of synthetic biology combined biomimetics, providing insight for future research and development of cutting-edge bio-inspired products. Protein-based polymers that utilize nature’s designs as a base, then modified by advancements at the intersection of biology and engineering, may provide mankind with more sustainable products. Full article
(This article belongs to the Special Issue Bio-Inspired Design for Structure Applications)
Show Figures

Graphical abstract

Graphical abstract
Full article ">Figure 1
<p>Multiple protein sequence alignment of Collagen type IV: Alpha 1–6 subunits shown the conserved G-X-Y repeated motifs.</p>
Full article ">Figure 2
<p>Collagen type IV Alpha 1–6 subunits’ structures predicted by AlphaFold Monomer v2.0 pipelne. The predicted local distance difference test (pLDDT) is a per-residue measure of local confidence. It is scaled from 0 to 100, with higher scores indicating higher confidence. Color legend: <span style="background:#0052DB"> </span> Very high (pLDDT &gt; 90); <span style="background:#00CCF5"> </span> High (90 &gt; pLDDT &gt; 70); <span style="background:#FFDC00"> </span> Low (70 &gt; pLDDT &gt; 50); <span style="background:#FF7D3F"> </span> Very low (pLDDT &lt; 50).</p>
Full article ">Figure 3
<p>Schematic illustration of the naturally occurring protein-based biomaterial structures and their application, using silk fibroin as an example. (<b>Left</b>) Silk fibroin produced by the silkworm <span class="html-italic">Bombyx mori</span> consists of a heavy chain, a light chain, and a glycoprotein, P25. (<b>Right</b>) Examples of utilizing silk fibrin for various applications, including the silk fibroin hydrogel.</p>
Full article ">Figure 4
<p>AlphaFold v2.0-predicted tertiary structure of type I keratin KRT 31 (<b>A</b>) and type II keratin IKRT 82 (<b>B</b>) from human hair. Gene ID: Q15323 (KRT31); Q9NSB4 (KRT82).The predicted local distance difference test (pLDDT) is a per-residue measure of local confidence. It is scaled from 0 to 100, with higher scores indicating higher confidence. Color legend: <span style="background:#0052DB"> </span> Very high (pLDDT &gt; 90); <span style="background:#00CCF5"> </span> High (90 &gt; pLDDT &gt; 70); <span style="background:#FFDC00"> </span> Low (70 &gt; pLDDT &gt; 50); <span style="background:#FF7D3F"> </span> Very low (pLDDT &lt; 50).</p>
Full article ">Figure 5
<p>Sequence alignment for 13 human elastin isoforms (a-m).</p>
Full article ">Figure 6
<p>Human elastin K (E7EN65) structure predicted by the AlphaFold v2.0. The predicted local distance difference test (pLDDT) is a per-residue measure of local confidence. It is scaled from 0 to 100, with higher scores indicating higher confidence. Color legend <span style="background:#00CCF5"> </span> High (90 &gt; pLDDT &gt; 70); <span style="background:#FFDC00"> </span> Low (70 &gt; pLDDT &gt; 50); <span style="background:#FF7D3F"> </span> Very low (pLDDT &lt; 50).</p>
Full article ">Figure 7
<p>Sequence alignment for resilin isoform A (NP_611157.1, 620 aa) and isoform B (NP_995860.1, 575 aa) missing 45 aa (aa341–385) from <span class="html-italic">D. melanogaster</span>.</p>
Full article ">Figure 8
<p><span class="html-italic">D. melanogaster</span> resilin isoform A (NP_611157.1, 620 aa) structure predicted by AlphaFold v2.0. The predicted local distance difference test (pLDDT) is a per-residue measure of local confidence. It is scaled from 0 to 100, with higher scores indicating higher confidence. Color legend: <span style="background:#0052DB"> </span> Very high (pLDDT &gt; 90); <span style="background:#00CCF5"> </span> High (90 &gt; pLDDT &gt; 70); <span style="background:#FFDC00"> </span> Low (70 &gt; pLDDT &gt; 50); <span style="background:#FF7D3F"> </span> Very low (pLDDT &lt; 50).</p>
Full article ">Figure 9
<p>Wheat HMW and LMW glutenin structures predicted by AlphaFold v2.0. The predicted local distance difference test (pLDDT) is a per-residue measure of local confidence. It is scaled from 0 to 100, with higher scores indicating higher confidence. Color legend: <span style="background:#0052DB"> </span> Very high (pLDDT &gt; 90); <span style="background:#00CCF5"> </span> High (90 &gt; pLDDT &gt; 70); <span style="background:#FFDC00"> </span> Low (70 &gt; pLDDT &gt; 50); <span style="background:#FF7D3F"> </span> Very low (pLDDT &lt; 50).</p>
Full article ">
16 pages, 4506 KiB  
Article
Gasdermin D Inhibitor Necrosulfonamide Alleviates Angiotensin II-Induced Abdominal Aortic Aneurysms in Apolipoprotein E-Deficient Mice
by Jia Guo, Qing Zhang, Zhidong Li, Min Qin, Jinyun Shi, Yan Wang, Wenjia Ai, Junjie Ju, Makoto Samura, Philip S Tsao and Baohui Xu
Biomolecules 2024, 14(6), 726; https://doi.org/10.3390/biom14060726 - 19 Jun 2024
Viewed by 991
Abstract
Abdominal aortic aneurysm (AAA) is a chronic aortic disease that lacks effective pharmacological therapies. This study was performed to determine the influence of treatment with the gasdermin D inhibitor necrosulfonamide on experimental AAAs. AAAs were induced in male apolipoprotein E-deficient mice by subcutaneous [...] Read more.
Abdominal aortic aneurysm (AAA) is a chronic aortic disease that lacks effective pharmacological therapies. This study was performed to determine the influence of treatment with the gasdermin D inhibitor necrosulfonamide on experimental AAAs. AAAs were induced in male apolipoprotein E-deficient mice by subcutaneous angiotensin II infusion (1000 ng/kg body weight/min), with daily administration of necrosulfonamide (5 mg/kg body weight) or vehicle starting 3 days prior to angiotensin II infusion for 30 days. Necrosulfonamide treatment remarkably suppressed AAA enlargement, as indicated by reduced suprarenal maximal external diameter and surface area, and lowered the incidence and reduced the severity of experimental AAAs. Histologically, necrosulfonamide treatment attenuated medial elastin breaks, smooth muscle cell depletion, and aortic wall collagen deposition. Macrophages, CD4+ T cells, CD8+ T cells, and neovessels were reduced in the aneurysmal aortas of necrosulfonamide- as compared to vehicle-treated angiotensin II-infused mice. Atherosclerosis and intimal macrophages were also substantially reduced in suprarenal aortas from angiotensin II-infused mice following necrosulfonamide treatment. Additionally, the levels of serum interleukin-1β and interleukin-18 were significantly lower in necrosulfonamide- than in vehicle-treated mice without affecting body weight gain, lipid levels, or blood pressure. Our findings indicate that necrosulfonamide reduced experimental AAAs by preserving aortic structural integrity as well as reducing mural leukocyte accumulation, neovessel formation, and systemic levels of interleukin-1β and interleukin-18. Thus, pharmacologically inhibiting gasdermin D activity may lead to the establishment of nonsurgical therapies for clinical AAA disease. Full article
(This article belongs to the Section Molecular Medicine)
Show Figures

Figure 1

Figure 1
<p>Study design and experimental approaches. Eight-week-old male apolipoprotein E-deficient mice received continuous subcutaneous infusion of angiotensin II (1000 ng/min/kg body weight) for 28 days to induce abdominal aortic aneurysms. Necrosulfonamide (NSA) (5 mg/kg body weight) or vehicle was given by daily oral gavage, beginning 3 days prior to angiotensin II infusion, for 30 days. Blood pressure measurements were performed prior to (day 0) and 28 days following angiotensin II infusion. Influences on abdominal aortic aneurysms were evaluated via aortic morphological measurements and histological analyses at sacrifice. Blood specimens were collected 28 days after angiotensin II infusion for the measurements of serum lipid and inflammatory cytokine levels. CD: cluster of differentiation. ELISA: enzyme-linked immunosorbent assay.</p>
Full article ">Figure 2
<p>Necrosulfonamide (NSA) treatment attenuates experimental abdominal aortic aneurysm enlargement. Male apolipoprotein E-deficient mice were subcutaneously infused with angiotensin II (1000 ng/min/kg body weight) for abdominal aortic aneurysm induction. Mice were treated with vehicle or NSA (5 mg/kg body weight) for 30 days, initiating 3 days prior to angiotensin II infusion. (<b>A</b>) Scheme for morphological evaluation of angiotensin II-induced abdominal aortic aneurysms. (<b>B</b>) Mean ± standard deviation of adjacent infrarenal external aortic diameter. (<b>C</b>–<b>H</b>) Median and interquartile range (25% and 75%) of maximal suprarenal external aortic diameter (<b>C</b>), ratio of maximal suprarenal to adjacent infrarenal external aortic diameter (<b>D</b>), suprarenal aortic arch length (<b>E</b>), suprarenal aortic chord length (<b>F</b>), ratio of suprarenal arch length to chord length (<b>G</b>) and maximal suprarenal aortic area (<b>H</b>). Non-parametric Mann–Whitney, 0.05 &lt; # <span class="html-italic">p</span> &lt; 0.1 and * <span class="html-italic">p</span> &lt; 0.05 compared to vehicle treatment. Dotted lines: the average value for each morphological parameter measurement from age-matched saline-infused apolipoprotein E-deficient mice.</p>
Full article ">Figure 3
<p>Necrosulfonamide (NSA) treatment attenuates the formation and severity of experimental abdominal aortic aneurysms (AAAs). Male apolipoprotein E-deficient mice were infused with angiotensin (Ang) II (1000 ng/min/kg body weight) for 28 days to induce AAAs. (<b>A</b>) Representative abdominal aortic aneurysm images in the vehicle and NSA treatment groups. (<b>B</b>) AAA incidence. AAA was defined by at least a 50% increase in suprarenal aortic diameter over that in the age-matched saline-infused mice, the presence of aortic dissection, or death caused by aneurysm rupture. Fisher’s exact test, * <span class="html-italic">p</span> &lt; 0.05 compared to vehicle treatment. (<b>C</b>) Survival rate. Log-rank test, 0.05 &lt; # <span class="html-italic">p</span> &lt; 0.1 compared to vehicle treatment. (<b>D</b>) Distribution of AAA severity score. AAAs were graded as grade I to V based on the presence of intramural thrombus as well as the shape and number of aneurysms. (<b>E</b>) Quantification of AAA severity score (median and interquartile range (25% and 75%)) in two treatment groups. Non-parametric Mann–Whitney, ** <span class="html-italic">p</span> &lt; 0.01 compared to vehicle treatment.</p>
Full article ">Figure 4
<p>Necrosulfonamide (NSA) treatment reduces medial elastin breaks, smooth muscle cell (SMC) depletion, and collagen deposition in experimental abdominal aortic aneurysms. Apolipoprotein E-deficient mice were sacrificed 28 days following angiotensin II infusion. Aortas were harvested, sectioned (8 μm), and stained via the elastic Verhoeff’s Van Gieson staining for medial elastin, SMC α-actin antibody for SMCs, and Masson trichome staining for collagen deposition. (<b>A</b>) Representative histological staining images for elastin (black to blue/black), SMCs (red), and collagens (blue). (<b>B</b>,<b>C</b>) Quantification of medial elastin breaks. (<b>D</b>,<b>E</b>) Quantification of SMC α-actin-positive area in aortic cross-sections (ACSs). (<b>F</b>–<b>H</b>) Quantification of collagen-positive area in ACSs. All data in (<b>C</b>,<b>E</b>,<b>G</b>,<b>H</b>) are normalized by aortic internal perimeter in ACSs. Student’s <span class="html-italic">t</span> test ((<b>D</b>,<b>F</b>): mean ± standard deviation) and non-parametric Mann–Whitney ((<b>B</b>,<b>C</b>,<b>G</b>,<b>H</b>): median and interquartile range (25% and 75%)), * <span class="html-italic">p</span> &lt;0.05, ** <span class="html-italic">p</span> &lt; 0.01 and *** <span class="html-italic">p</span> &lt; 0.001 compared to vehicle treatment.</p>
Full article ">Figure 5
<p>Necrosulfonamide (NSA) treatment reduces aortic leukocyte accumulation and angiogenesis in experimental abdominal aortic aneurysms. Frozen aortic sections from vehicle- and NSA (5 mg/kg body weight)-treated, angiotensin II-infused mice were fixed with acetone and stained with antibodies against CD68 for macrophages, CD4 for CD4<sup>+</sup> T cells, CD8 for CD8<sup>+</sup> T cells, CD31 for neovessels. (<b>A</b>) Representative immunohistochemical staining images for macrophages, CD4<sup>+</sup> T cells, CD8<sup>+</sup> T cells, and neovessels in two treatment groups. (<b>B</b>) Quantification (mean ± standard deviation) of macrophage accumulation. (<b>C</b>–<b>E</b>) Quantification (median and interquartile range) of CD4<sup>+</sup> T cells (<b>C</b>), CD8<sup>+</sup> T cells (<b>D</b>), and CD31<sup>+</sup> neovessels (<b>E</b>) per aortic cross-section (ACS). Student’s <span class="html-italic">t</span> test (<b>B</b>) and non-parametric Mann–Whitney (<b>C</b>–<b>E</b>), * <span class="html-italic">p</span> &lt; 0.05 compared to vehicle treatment. CD: Cluster of differentiation.</p>
Full article ">Figure 6
<p>Necrosulfonamide (NSA) treatment reduces atherosclerotic lesion size and alleviates macrophage accumulation in male apolipoprotein E-deficient mice following angiotensin II infusion. Frozen aortic sections were prepared from differentially treated angiotensin II-infused apolipoprotein E-deficient mice, fixed with 10% formalin or acetone, and stained with hematoxylin and eosin (acetone-fixed sections) for total atherosclerotic lesions, Oil Red O for lipid lesions (10% formalin-fixed sections), and CD68 for macrophage accumulation (acetone-fixed sections). (<b>A</b>) Representative histological images of total lesions (hematoxylin and eosin stain), lipid lesions (Oil Red O stain), and macrophage accumulation (CD68 monoclonal antibody immunostaining). (<b>B</b>,<b>C</b>) Quantification (median and interquartile range) of total atherosclerotic lesions (<b>B</b>) and total atherosclerotic lesions normalized by aortic internal perimeter (<b>C</b>). (<b>D</b>,<b>E</b>) Quantification (mean ± standard deviation) of lipid deposition (<b>D</b>) and lipid deposition normalized by aortic internal perimeter (<b>E</b>). (<b>F</b>–<b>H</b>) Quantification (median and interquartile range) of macrophage accumulation (<b>F</b>), macrophage accumulation normalized by total lesion size (<b>G</b>), and macrophage accumulation normalized by aortic internal perimeter (<b>H</b>). Student’s <span class="html-italic">t</span> test (<b>D</b>,<b>E</b>) and non-parametric Mann–Whitney (<b>B</b>,<b>C</b>,<b>F</b>–<b>H</b>), * <span class="html-italic">p</span> &lt; 0.05 and ** <span class="html-italic">p</span> &lt; 0.01 compared to vehicle treatment.</p>
Full article ">Figure 7
<p>Necrosulfonamide (NSA) treatment reduces the serum levels of interleukin (IL)-1β and IL-18 in experimental abdominal aortic aneurysms. Twenty-eight days after angiotensin II infusion, sera were prepared from differentially treated apolipoprotein E-deficient mice. Serum levels of IL-1β (<b>A</b>) and IL-18 (<b>B</b>) were assessed via enzyme-linked immunosorbent assay. All data are mean ± standard deviation. Student’s <span class="html-italic">t</span> test, * <span class="html-italic">p</span> &lt; 0.05 compared to vehicle treatment.</p>
Full article ">Figure 8
<p>Necrosulfonamide (NSA) treatment does not affect body weight gain, lipid levels, or blood pressure in experimental abdominal aortic aneurysms. (<b>A</b>) Body weight in vehicle- and NSA-treated apolipoprotein E-deficient mice at the baseline (day 0) and 28 days after angiotensin (Ang) II infusion. (<b>B</b>,<b>C</b>) Serum levels of total cholesterol (<b>B</b>) and triglycerides (<b>C</b>) from vehicle- and NSA-treated apolipoprotein E-deficient mice 28 days after Ang II infusion. (<b>D</b>,<b>E</b>) Systolic (<b>D</b>) and diastolic (<b>E</b>) blood pressure of vehicle- and NSA-treated apolipoprotein E-deficient mice at the baseline and 28 days after Ang II infusion. All data are presented as the mean ± standard deviation. Two-way analysis of variance test (<b>A</b>,<b>D</b>,<b>E</b>) or Student’s <span class="html-italic">t</span> test (<b>B</b>,<b>C</b>) showed no significant difference between the two treatment groups at the same timepoint.</p>
Full article ">
12 pages, 7894 KiB  
Article
The Anti-Atherosclerotic Effects of Endothelin Receptor Antagonist, Bosentan, in Combination with Atorvastatin—An Experimental Study
by Marianna Stasinopoulou, Nikolaos Kostomitsopoulos and Nikolaos P. E. Kadoglou
Int. J. Mol. Sci. 2024, 25(12), 6614; https://doi.org/10.3390/ijms25126614 - 16 Jun 2024
Viewed by 3383
Abstract
Bosentan, an endothelin receptor antagonist (ERA), has potential anti-atherosclerotic properties. We investigated the complementary effects of bosentan and atorvastatin on the progression and composition of the atherosclerotic lesions in diabetic mice. Forty-eight male ApoE/ mice were fed high-fat diet (HFD) [...] Read more.
Bosentan, an endothelin receptor antagonist (ERA), has potential anti-atherosclerotic properties. We investigated the complementary effects of bosentan and atorvastatin on the progression and composition of the atherosclerotic lesions in diabetic mice. Forty-eight male ApoE/ mice were fed high-fat diet (HFD) for 14 weeks. At week 8, diabetes was induced with streptozotocin, and mice were randomized into four groups: (1) control/COG: no intervention; (2) ΒOG: bosentan 100 mg/kg/day per os; (3) ATG: atorvastatin 20 mg/kg/day per os; and (4) BO + ATG: combined administration of bosentan and atorvastatin. The intra-plaque contents of collagen, elastin, monocyte chemoattractant protein-1 (MCP-1), tumor necrosis factor-a (TNF-a), matrix metalloproteinases (MMP-2, -3, -9), and TIMP-1 were determined. The percentage of lumen stenosis was significantly lower across all treated groups: BOG: 19.5 ± 2.2%, ATG: 12.8 ± 4.8%, and BO + ATG: 9.1 ± 2.7% compared to controls (24.6 ± 4.8%, p < 0.001). The administration of both atorvastatin and bosentan resulted in significantly higher collagen content and thicker fibrous cap versus COG (p < 0.01). All intervention groups showed lower relative intra-plaque concentrations of MCP-1, MMP-3, and MMP-9 and a higher TIMP-1concentration compared to COG (p < 0.001). Importantly, latter parameters presented lower levels when bosentan was combined with atorvastatin compared to COG (p < 0.05). Bosentan treatment in diabetic, atherosclerotic ApoE/ mice delayed the atherosclerosis progression and enhanced plaques’ stability, showing modest but additive effects with atorvastatin, which are promising in atherosclerotic cardiovascular diseases. Full article
Show Figures

Figure 1

Figure 1
<p>All active groups (BOG, ATG, and BO + ATG) had significantly smaller plaques compared to controls in <span class="html-italic">ApoE<sup>−</sup>/<sup>−</sup></span> mice. Representative images and quantifications of aortic valve sections stained with hematoxylin/eosin across all groups.</p>
Full article ">Figure 2
<p>Quantification of immunohistochemical staining with antibodies against MMP-2, MMP-3, MMP-9, TIMP-1, TNF-a, and MCP-1. * <span class="html-italic">p</span> &lt; 0.05 vs. COG, # <span class="html-italic">p</span> &lt; 0.05 vs. BOG, || <span class="html-italic">p</span> &lt; 0.05 vs. AΤG.</p>
Full article ">Figure 3
<p>Representative images of immunohistochemical staining across all groups with antibodies against (<b>a</b>) MMP-2 (upper panel), (<b>b</b>) MMP-3 (middle panel), and (<b>c</b>) TIMP-1 (lower panel). Sections were counterstained with H&amp;E.</p>
Full article ">
Back to TopTop