[go: up one dir, main page]

 
 
Sign in to use this feature.

Years

Between: -

Subjects

remove_circle_outline
remove_circle_outline
remove_circle_outline
remove_circle_outline
remove_circle_outline
remove_circle_outline
remove_circle_outline
remove_circle_outline
remove_circle_outline

Journals

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

Article Types

Countries / Regions

remove_circle_outline
remove_circle_outline
remove_circle_outline
remove_circle_outline
remove_circle_outline
remove_circle_outline
remove_circle_outline
remove_circle_outline

Search Results (6,206)

Search Parameters:
Keywords = angiogenesis

Order results
Result details
Results per page
Select all
Export citation of selected articles as:
17 pages, 4188 KiB  
Article
Three in One with Dual-Functional Hydrogel of Lactoferrin/NZ2114/LMSH Promoting Staphylococcus aureus-Infected Wound Healing
by Kun Zhang, Xuanxuan Ma, Da Teng, Ruoyu Mao, Na Yang, Ya Hao and Jianhua Wang
Antibiotics 2024, 13(9), 889; https://doi.org/10.3390/antibiotics13090889 (registering DOI) - 15 Sep 2024
Abstract
Wound infections caused by Staphylococcus aureus often result in localized suppurative lesions that severely impede the healing process, so it is urgent to develop a dress with efficient antimicrobial and pro-healing functions. In this study, the bifunctional injectable hydrogel lactoferrin (Lf)/NZ2114/lithium magnesium silicate [...] Read more.
Wound infections caused by Staphylococcus aureus often result in localized suppurative lesions that severely impede the healing process, so it is urgent to develop a dress with efficient antimicrobial and pro-healing functions. In this study, the bifunctional injectable hydrogel lactoferrin (Lf)/NZ2114/lithium magnesium silicate hydrogel (LMSH) was first successfully prepared through the electrostatic interaction method. The physical, biological, and efficacy properties are systematically analyzed with good shear-thinning capacity and biocompatibility. More importantly, it inhibits infection and promotes wound healing in a mouse wound infection model after 14 d treatment, and the bactericidal rate and healing rate were over 99.92% and nearly 100%, respectively. Meanwhile, the massive reduction of inflammatory cells, restoration of tissue structure, and angiogenesis in mice showed the anti-inflammatory and pro-healing properties of the hydrogel. The healed wounds showed thickening with more hair follicles and glands, suggesting that the hydrogel Lf/NZ2114/LMSH (Three in One) could be a better dressing candidate for the treatment of S. aureus-induced wound infections. Full article
(This article belongs to the Special Issue Anti-microbial Coating Innovations to Prevent Infectious Diseases)
Show Figures

Graphical abstract

Graphical abstract
Full article ">Figure 1
<p>The photograph, antimicrobial activity and FT–IR spectrum of the injectable hydrogel Lf/NZ2114/LMSH. (<b>a</b>) The photograph of the synthetic injectable hydrogel; (<b>b</b>) Inhibition zone of NZ2114, Lf, 1% LMSH/Lf/NZ2114, 3% HACC/Lf/NZ2114, 3% SA, 1% LMSH; (<b>c</b>) The FT–IR spectra of 1% LMSH, 1% LMSH + Lf, 1% LMSH + NZ2114, 1% LMSH + Lf + NZ2114, 1.5% LMSH + Lf + NZ2114, 3% HACC, 3% HACC + Lf, 3% HACC + NZ2114, 3% HACC + Lf + NZ2114.</p>
Full article ">Figure 2
<p>The morphology of the synthetic injectable hydrogel. The SEM image of 0.5% LMSH + NZ2114, 1% LMSH + Lf + NZ2114, 1% LMSH + Lf + NZ2114, 3% HACC + Lf + NZ2114, 1% LMSH, 3% HACC, Lf + NZ2114.</p>
Full article ">Figure 3
<p>The viscosity, modulus, and bactericidal properties of different hydrogel samples. (<b>a</b>) The viscosity of 0.5% Lf/NZ2114/LMSH and 1% Lf/NZ2114/LMSH during sheer increase from 0.01 to 100 s<sup>−1</sup>; (<b>b</b>) The viscosity of 3% Lf/NZ2114/HACC during sheer increase from 0.01 to 100 s<sup>−1</sup>; (<b>c</b>) The storage modulus (G′) and loss modulus (G″) of 0.5% Lf/NZ2114/LMSH, 1% Lf/NZ2114/LMSH, and 3% Lf/NZ2114/HACC during strain increase from 0.1% to 1000% at the frequency of 1 Hz; (<b>d</b>,<b>e</b>) The bactericidal rate of 1% Lf/NZ2114/LMSH, 3% Lf/NZ2114/HACC, Lf + NZ2114, Lf, NZ2114, 1% LMSH, 3% HACC against <span class="html-italic">S. aureus</span> CVCC 546 (n = 3). These (−2, −3, −4, −5) are the number of dilutions. (* <span class="html-italic">p</span> &lt; 0.05, *** <span class="html-italic">p</span> &lt; 0.001).</p>
Full article ">Figure 4
<p>The biocompatibility of different hydrogel samples. (<b>a</b>,<b>b</b>) The images and hemolysis rate of Lf, Lf + NZ2114, 1% Lf/NZ2114/LMSH, 3% Lf/NZ2114/HACC, 1% LMSH, 3% HACC, 0.1% Trix-100; (<b>c</b>) Cytotoxicity of HACAT cells co-cultured with Lf, Lf + NZ2114, 1% Lf/NZ2114/LMSH, 3% Lf/NZ2114/HACC, 1% LMSH, 3% HACC. Samples at 1, 2, and 4 days, n = 3; (<b>d</b>) The images of calcein–AM/PI double staining of the HACAT cells that were incubated with 1% Lf/NZ2114/LMSH, 3% Lf/NZ2114/HACC, 1% LMSH, 3% HACC for 1, 2 and 4 days. (Scale bar = 100 μm). * <span class="html-italic">p</span> &lt; 0.05, *** <span class="html-italic">p</span> &lt; 0.001.</p>
Full article ">Figure 5
<p>The body weight, skin load, and wound diagram of mice (n = 3). (<b>a</b>) The weight of mice untreated and treated with Lf + NZ2114, 1% Lf/NZ2114/LMSH, 3% Lf/NZ2114/HACC, 1% LMSH, 3% HACC for 0–14 days; (<b>b</b>) The bacteria of skin untreated and treated with Lf + NZ2114, 1% Lf/NZ2114/LMSH, 3% Lf/NZ2114/HACC, 1% LMSH, 3% HACC samples for 4 d; (<b>c</b>) The macroscopic images of wounds untreated and treated with 1% Lf/NZ2114/LMSH, 3% Lf/NZ2114/HACC samples for 4, 8, 12 and 14 d; (<b>d</b>) The wound healing rate of mice untreated and treated with Lf + NZ2114, 1% Lf/NZ2114/LMSH, 3% Lf/NZ2114/HACC, 1% LMSH, 3% HACC samples for 4, 8, 12, and 14 d. * <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>HE staining of the wounds without any treatment or treated with 1% Lf/NZ2114/LMSH and 3% Lf/NZ2114/HACC samples for 14 days.</p>
Full article ">Figure 7
<p>The cytokine secretion without any treatment or treated with 1% Lf/NZ2114/LMSH and 3% Lf/NZ2114/HACC samples for 14 days. (<b>a</b>–<b>c</b>) The expressions of IL-6, VEGF and EGFR were measured by ELISA kit. (<b>d</b>–<b>g</b>) The expressions of IL-6, VEGF, EGFR, and CD31 were measured at RNA level by qPCR.</p>
Full article ">Figure 8
<p>The immunohistochemistry of without any treatment or treated with 1% Lf/NZ2114/LMSH and 3% Lf/NZ2114/HACC samples for 14 days. CD31 staining of the wounds without any treatment or treated with different hydrogel samples for 14 days. IL-6 staining of the wounds for 14 days. (Red arrows are CD31 binding sites).</p>
Full article ">
15 pages, 9533 KiB  
Article
Photo-Crosslinked Pro-Angiogenic Hydrogel Dressing for Wound Healing
by Wang Zhang, Shuyi Qian, Jia Chen, Tianshen Jian, Xuechun Wang, Xianmin Zhu, Yixiao Dong and Guoping Fan
Int. J. Mol. Sci. 2024, 25(18), 9948; https://doi.org/10.3390/ijms25189948 (registering DOI) - 15 Sep 2024
Viewed by 148
Abstract
Severe burns are one of the most devastating injuries, in which sustained inflammation and ischemia often delay the healing process. Pro-angiogenic growth factors such as vascular endothelial growth factor (VEGF) have been widely studied for promoting wound healing. However, the short half-life and [...] Read more.
Severe burns are one of the most devastating injuries, in which sustained inflammation and ischemia often delay the healing process. Pro-angiogenic growth factors such as vascular endothelial growth factor (VEGF) have been widely studied for promoting wound healing. However, the short half-life and instability of VEGF limit its clinical applications. In this study, we develop a photo-crosslinked hydrogel wound dressing from methacrylate hyaluronic acid (MeHA) bonded with a pro-angiogenic prominin-1-binding peptide (PR1P). The materials were extruded in wound bed and in situ formed a wound dressing via exposure to short-time ultraviolet radiation. The study shows that the PR1P-bonded hydrogel significantly improves VEGF recruitment, tubular formation, and cell migration in vitro. Swelling, Scanning Electron Microscope, and mechanical tests indicate the peptide does not affect the overall mechanical and physical properties of the hydrogels. For in vivo studies, the PR1P-bonded hydrogel dressing enhances neovascularization and accelerates wound closure in both deep second-degree burn and full-thickness excisional wound models. The Western blot assay shows such benefits can be related to the activation of the VEGF–Akt signaling pathway. These results suggest this photo-crosslinked hydrogel dressing efficiently promotes VEGF recruitment and angiogenesis in skin regeneration, indicating its potential for clinical applications in wound healing. Full article
(This article belongs to the Special Issue Advanced Research on Wound Healing 2.0)
Show Figures

Figure 1

Figure 1
<p>Fabrication of an injectable HA-P hydrogel wound dressing. (<b>A</b>) Schematic illustration of the synthesis of MeHA and in situ crosslinking with cysteine-modified PR1P to form a hydrogel wound dressing. (<b>B</b>) Real-time crosslinking rheological measurements of HA and HA-P hydrogels (0.5% <span class="html-italic">w</span>/<span class="html-italic">v</span>) with 30 s exposure to UV radiation. (<b>C</b>) Compressive modulus of HA hydrogels with different material concentrations (gelation with 30 s exposure to UV radiation). (<b>D</b>) Compressive modulus of HA and HA-P hydrogels (0.5% <span class="html-italic">w</span>/<span class="html-italic">v</span>, with 30 s exposure to UV radiation) (mean ± SD, n = 6, <span class="html-italic">** p &lt;</span> 0.01, <span class="html-italic">**** p &lt;</span> 0.0001, ns, not statistically significant).</p>
Full article ">Figure 2
<p>Characterization of HA and HA-P hydrogels. (<b>A</b>,<b>B</b>) SEM micrographs and quantification of the average pore size of freeze-dried HA hydrogels (0.5% <span class="html-italic">w</span>/<span class="html-italic">v</span>) with 30 s, 60 s, and 90 s of UV exposure. (<b>C</b>,<b>D</b>) SEM micrographs and quantification of the average pore size of HA and HA-P hydrogels (0.5% <span class="html-italic">w</span>/<span class="html-italic">v</span>, with 30 s of UV exposure). (<b>E</b>,<b>F</b>) Swelling ratios of HA hydrogels with various UV exposure times and material concentrations. (<b>G</b>) Swelling ratios of HA and HA-P hydrogels (0.5% <span class="html-italic">w</span>/<span class="html-italic">v</span>, with 30 s of UV exposure) (mean ± SD, n = 3, *** <span class="html-italic">p</span> &lt; 0.001, <span class="html-italic">**** p</span> &lt; 0.0001, ns, not statistically significant, scale bar, 100 μm).</p>
Full article ">Figure 3
<p>VEGF recruitment and in vitro angiogenic effect of HA-P hydrogels. (<b>A</b>) Schematic illustration of VEGF recruitment assay. (<b>B</b>) Quantitative analysis of the maintained VEGF within hydrogels shows the HA-P hydrogel binds more VEGF than HA hydrogel does (n = 8). (<b>C</b>) Representative images of cell migration in a scratch wound healing assay after 0, 6, 12, and 24 h. (<b>D</b>) Quantitative analysis of the migration ratio shows HA-P hydrogel loaded with VEGF significantly promotes cell migration compared with the other groups. (<b>E</b>) Representative images of the tube formation of HUVECs. (<b>F</b>,<b>G</b>) Quantitative analysis of capillary length and the number of branch points of the tubule network. The capillary length and branch points in HA-P hydrogels are significantly higher than in the other groups (mean ± SD, n = 3, <span class="html-italic">** p</span> &lt; 0.01, *** <span class="html-italic">p</span> &lt; 0.001, **** <span class="html-italic">p</span> &lt; 0.0001, scale bar, 200 μm).</p>
Full article ">Figure 4
<p>HA-P hydrogel dressing promotes wound regeneration in burns. (<b>A</b>) Representative photos exhibit the wound healing process. (<b>B</b>) Quantitative analysis of residual wound area (%) up to 14 days. HA-P hydrogel treatment shows significant acceleration of healing compared to the control group after day 6. (<b>C</b>) Representative images of H&amp;E staining and (<b>D</b>) Masson’s trichrome staining of the wounds at 14 days post-wounding (scale bar, 500 μm). (<b>E</b>) Quantitative analysis of epithelium thickness and (<b>F</b>) collagen density indicates less epidermis hyperplasia and increased collagen deposition in the HA-P hydrogel treatment group (mean ± SD, n = 8–10, <span class="html-italic">* p</span> &lt; 0.05, <span class="html-italic">** p</span> &lt; 0.01, ns, not statistically significant).</p>
Full article ">Figure 5
<p>HA-P hydrogel dressing enhances angiogenesis and reduces myofibroblasts in burns. (<b>A</b>) Representative images of the CD31<sup>+</sup> staining (green) of different groups at day 14 post-wounding. The nucleus was stained with DAPI (blue). (<b>B</b>,<b>C</b>) Stereological quantification of the surface area and length density of vasculature demonstrates a significant enhancement in angiogenesis for HA-P hydrogel compared with the control group. (<b>D</b>) Representative images of α-SMA<sup>+</sup> staining (red) at day 14 post-wounding. The nucleus was stained with DAPI (blue). (<b>E</b>) Quantitative analysis of the positive area of α-SMA shows the HA-P hydrogel treatment significantly reduces myofibroblasts’ regeneration (mean ± SD, n = 8–10, <span class="html-italic">** p</span> &lt; 0.01, **** <span class="html-italic">p</span> &lt; 0.0001, ns, not statistically significant, scale bar, 100 μm).</p>
Full article ">Figure 6
<p>HA-P hydrogel dressing promoted angiogenesis via activation of the VEGF–Akt signaling pathway. (<b>A</b>) Schematic illustration of the molecular mechanism for HA-P hydrogel dressing which activates the VEGF–Akt signaling pathway in wound healing. (<b>B</b>) Representative images of Western blotting of Akt, p-Akt, and VEGFA in wounds at day 14 post-wounding. (<b>C</b>,<b>D</b>) Quantitative results of Western blotting show that the HA-P hydrogel treatment significantly increases the relative protein expression level of VEGFA and the relative expression ratio of p-Akt/Akt (mean ± SD, n = 8–10, <span class="html-italic">* p</span> &lt; 0.05, ns, not statistically significant).</p>
Full article ">Figure 7
<p>HA-P hydrogel dressing promotes wound healing in a full-thickness excisional wound model. (<b>A</b>) Representative images of the healing process up to 14 days post-wounding. (<b>B</b>) Wound closure curves of different groups show a significant acceleration of healing with the HA-P hydrogel treatment compared to the HA hydrogel and control group from day 4. (<b>C</b>) Representative images of CD31<sup>+</sup> (green) and α-SMA<sup>+</sup> (red) staining at day 14 post-wounding. The nucleus was stained with DAPI (blue). (<b>D</b>,<b>E</b>) Quantitative analysis indicates the HA-P hydrogel treatment significantly improves the angiogenesis and (<b>F</b>) reduces myofibroblasts’ regeneration (mean ± SD, n = 6, ** <span class="html-italic">p</span> &lt; 0.01, **** <span class="html-italic">p</span> &lt; 0.0001, ns, not statistically significant, scale bar, 100 μm).</p>
Full article ">
10 pages, 3157 KiB  
Article
One Copy Number Variation within the Angiopoietin-1 Gene Is Associated with Leizhou Black Goat Meat Quality
by Qun Wu, Xiaotao Han, Yuelang Zhang, Hu Liu, Hanlin Zhou, Ke Wang and Jiancheng Han
Animals 2024, 14(18), 2682; https://doi.org/10.3390/ani14182682 (registering DOI) - 14 Sep 2024
Viewed by 194
Abstract
The ANGPT1 gene plays a crucial role in the regulation of angiogenesis and muscle growth, with previous studies identifying copy number variations (CNVs) within this gene among Leizhou black goats. In this study, we investigated three ANGPT1 CNVs in 417 individuals of LZBG [...] Read more.
The ANGPT1 gene plays a crucial role in the regulation of angiogenesis and muscle growth, with previous studies identifying copy number variations (CNVs) within this gene among Leizhou black goats. In this study, we investigated three ANGPT1 CNVs in 417 individuals of LZBG using quantitative PCR (qPCR), examining the impact of different CNV types on the ANGPT1 gene expression and their associations with growth and meat quality traits. Notably, the ANGPT1 CNV-1 (ARS1_chr14:24950001-24953600) overlaps with protein-coding regions and conserved domains; its gain-of-copies genotype (copies ≥ 3) was significantly correlated with ANGPT1 mRNA expression in muscle tissue (p < 0.01). Furthermore, the gain-of-copies genotype of CNV-1 demonstrated significant correlations with various phenotypic traits, including carcass weight, body weight, shear stress, chest circumference, and cross-sectional area of longissimus dorsi muscle. These findings indicate that the CNV-1 gain-of-copies genotype in the ANGPT1 gene may serve as a valuable marker for selecting Leizhou black goats exhibiting enhanced growth and muscular development characteristics, thereby holding potential applications in targeted breeding programs aimed at improving meat quality. Full article
20 pages, 2620 KiB  
Article
A Comprehensive Genetic and Bioinformatic Analysis Provides Evidence for the Engagement of COVID-19 GWAS-Significant Loci in the Molecular Mechanisms of Coronary Artery Disease and Stroke
by Alexey Loktionov, Ksenia Kobzeva, Anna Dorofeeva, Maryana Babkina, Elizaveta Kolodezhnaya and Olga Bushueva
J. Mol. Pathol. 2024, 5(3), 385-404; https://doi.org/10.3390/jmp5030026 (registering DOI) - 14 Sep 2024
Viewed by 237
Abstract
Cardiovascular diseases (CVDs) significantly exacerbate the severity and mortality of COVID-19. We aimed to investigate whether GWAS-significant SNPs correlate with CVDs in severe COVID-19 patients. DNA samples from 199 patients with severe COVID-19 hospitalized in intensive care units were genotyped using probe-based PCR [...] Read more.
Cardiovascular diseases (CVDs) significantly exacerbate the severity and mortality of COVID-19. We aimed to investigate whether GWAS-significant SNPs correlate with CVDs in severe COVID-19 patients. DNA samples from 199 patients with severe COVID-19 hospitalized in intensive care units were genotyped using probe-based PCR for 10 GWAS SNPs previously implicated in severe COVID-19 outcomes. SNPs rs17713054 SLC6A20-LZTFL1 (risk allele A, OR = 2.14, 95% CI 1.06–4.36, p = 0.03), rs12610495 DPP9 (risk allele G, OR = 1.69, 95% CI 1.02–2.81, p = 0.04), and rs7949972 ELF5 (risk allele T, OR = 2.57, 95% CI 1.43–4.61, p = 0.0009) were associated with increased risk of coronary artery disease (CAD). SNPs rs7949972 ELF5 (OR = 2.67, 95% CI 1.38–5.19, p = 0.003) and rs61882275 ELF5 (risk allele A, OR = 1.98, 95% CI 1.14–3.45, p = 0.01) were linked to a higher risk of cerebral stroke (CS). No associations were observed with AH. Bioinformatics analysis revealed the involvement of GWAS-significant loci in atherosclerosis, inflammation, oxidative stress, angiogenesis, and apoptosis, which provides evidence of their role in the molecular mechanisms of CVDs. This study provides novel insights into the associations between GWAS-identified SNPs and the risk of CAD and CS. Full article
Show Figures

Figure 1

Figure 1
<p>The outline of the study.</p>
Full article ">Figure 2
<p>Graph reflecting the structure and power of the most significant G × G interactions of GWAS loci associated with CAD in severe COVID-19 patients. Notes: the color of the lines reflects the nature of the interaction: orange lines mean moderate synergism, brown means additive (independent) effects; blue and green represent strong and moderate antagonism, respectively; % reflects the strength and direction of the phenotypic effect of gene–gene interaction (% of entropy).</p>
Full article ">Figure 3
<p>Graph reflecting the structure and power of the most significant G × E interactions of GWAS loci associated with CAD in severe COVID-19 patients. Notes: the color of the lines reflects the nature of the interaction: orange and red mean moderate and strong synergism, green —moderate antagonism; brown means additive (independent) effects; % reflects the strength and direction of the phenotypic effect of gene–environmental interaction (% of entropy).</p>
Full article ">Figure 4
<p>Graph reflecting the structure and power of the most significant G × G interactions of GWAS loci associated with CS in severe COVID-19 patients. Notes: the color of the lines reflects the nature of the interaction: red mean strong synergism, blue—pronounced antagonism; brown means additive (independent) effects; % reflects the strength and direction of the phenotypic effect of gene–gene interaction (% of entropy).</p>
Full article ">Figure 5
<p>Graph reflecting the structure and power of the most significant G×E interactions of GWAS loci associated with CS in severe COVID-19 patients. Notes: the color of the lines reflects the nature of the interaction: orange mean moderate synergism, green and blue—moderate and pronounced antagonism; % reflects the strength and direction of the phenotypic effect of gene–gene interaction (% of entropy).</p>
Full article ">Figure 6
<p>cis-eQTL effects and TFs-associated biological processes of GWAS SNPs linked to CVDs.</p>
Full article ">
33 pages, 1865 KiB  
Review
Oxidative Stress and Age-Related Tumors
by Emma Di Carlo and Carlo Sorrentino
Antioxidants 2024, 13(9), 1109; https://doi.org/10.3390/antiox13091109 - 13 Sep 2024
Viewed by 207
Abstract
Oxidative stress is the result of the imbalance between reactive oxygen and nitrogen species (RONS), which are produced by several endogenous and exogenous processes, and antioxidant defenses consisting of exogenous and endogenous molecules that protect biological systems from free radical toxicity. Oxidative stress [...] Read more.
Oxidative stress is the result of the imbalance between reactive oxygen and nitrogen species (RONS), which are produced by several endogenous and exogenous processes, and antioxidant defenses consisting of exogenous and endogenous molecules that protect biological systems from free radical toxicity. Oxidative stress is a major factor in the aging process, contributing to the accumulation of cellular damage over time. Oxidative damage to cellular biomolecules, leads to DNA alterations, lipid peroxidation, protein oxidation, and mitochondrial dysfunction resulting in cellular senescence, immune system and tissue dysfunctions, and increased susceptibility to age-related pathologies, such as inflammatory disorders, cardiovascular and neurodegenerative diseases, diabetes, and cancer. Oxidative stress-driven DNA damage and mutations, or methylation and histone modification, which alter gene expression, are key determinants of tumor initiation, angiogenesis, metastasis, and therapy resistance. Accumulation of genetic and epigenetic damage, to which oxidative stress contributes, eventually leads to unrestrained cell proliferation, the inhibition of cell differentiation, and the evasion of cell death, providing favorable conditions for tumorigenesis. Colorectal, breast, lung, prostate, and skin cancers are the most frequent aging-associated malignancies, and oxidative stress is implicated in their pathogenesis and biological behavior. Our aim is to shed light on the molecular and cellular mechanisms that link oxidative stress, aging, and cancers, highlighting the impact of both RONS and antioxidants, provided by diet and exercise, on cellular senescence, immunity, and development of an antitumor response. The dual role of ROS as physiological regulators of cell signaling responsible for cell damage and diseases, as well as its use for anti-tumor therapeutic purposes, will also be discussed. Managing oxidative stress is crucial for promoting healthy aging and reducing the risk of age-related tumors. Full article
(This article belongs to the Special Issue Reactive Nitrogen Species (RNS) and Redox Signaling in Tumors)
23 pages, 2974 KiB  
Article
Evaluation of Biotechnological Active Peptides Secreted by Saccharomyces cerevisiae with Potential Skin Benefits
by Elisabete Muchagato Maurício, Patrícia Branco, Ana Luiza Barros Araújo, Catarina Roma-Rodrigues, Katelene Lima, Maria Paula Duarte, Alexandra R. Fernandes and Helena Albergaria
Antibiotics 2024, 13(9), 881; https://doi.org/10.3390/antibiotics13090881 - 13 Sep 2024
Viewed by 191
Abstract
Biotechnological active peptides are gaining interest in the cosmetics industry due to their antimicrobial, anti-inflammatory, antioxidant, and anti-collagenase (ACE) effects, as well as wound healing properties, making them suitable for cosmetic formulations. The antimicrobial activity of peptides (2–10 kDa) secreted by Saccharomyces cerevisiae [...] Read more.
Biotechnological active peptides are gaining interest in the cosmetics industry due to their antimicrobial, anti-inflammatory, antioxidant, and anti-collagenase (ACE) effects, as well as wound healing properties, making them suitable for cosmetic formulations. The antimicrobial activity of peptides (2–10 kDa) secreted by Saccharomyces cerevisiae Ethanol-Red was evaluated against dermal pathogens using broth microdilution and challenge tests. ACE was assessed using a collagenase activity colorimetric assay, antioxidant activity via spectrophotometric monitoring of nitrotetrazolium blue chloride (NBT) reduction, and anti-inflammatory effects by quantifying TNF-α mRNA in lipopolysaccharides (LPS)-exposed dermal fibroblasts. Wound healing assays involved human fibroblasts, endothelial cells, and dermal keratinocytes. The peptides (2–10 kDa) exhibited antimicrobial activity against 10 dermal pathogens, with the Minimum Inhibitory Concentrations (MICs) ranging from 125 µg/mL for Staphylococcus aureus to 1000 µg/mL for Candida albicans and Streptococcus pyogenes. In the challenge test, peptides at their MICs reduced microbial counts significantly, fulfilling ISO 11930:2019 standards, except against Aspergillus brasiliensis. The peptides combined with Microcare SB showed synergy, particularly against C. albicans and A. brasilensis. In vitro, the peptides inhibited collagenase activity by 41.8% and 94.5% at 250 and 1000 µg/mL, respectively, and demonstrated antioxidant capacity. Pre-incubation with peptides decreased TNF-α expression in fibroblasts, indicating anti-inflammatory effects. The peptides do not show to promote or inhibit the angiogenesis of endothelial cells, but are able to attenuate fibrosis, scar formation, and chronic inflammation during the final phases of the wound healing process. The peptides showed antimicrobial, antioxidant, ACE, and anti-inflammatory properties, highlighting their potential as multifunctional bioactive ingredients in skincare, warranting further optimization and exploration in cosmetic applications. Full article
(This article belongs to the Special Issue Microbial Natural Products as a Source of Novel Antimicrobials)
Show Figures

Figure 1

Figure 1
<p>Growth profiles of bacteria, i.e., <span class="html-italic">E. coli</span> (<b>A</b>), <span class="html-italic">P. aeruginosa</span> (<b>B</b>), <span class="html-italic">S. aureus</span> (<b>C</b>) and fungi, i.e., <span class="html-italic">A. brasiliensis</span> (<b>D</b>) and <span class="html-italic">C. albicans</span> (<b>E</b>), in body milk formulation in the presence of the peptides (2–10 kDa) at the MIC value and 0.3% (<span class="html-italic">v</span>/<span class="html-italic">v</span>) of Microcare<sup>®</sup> BNA (0.3% B), and the association of both (0.3% B+2–10 kDa). A positive control with 0.6% (<span class="html-italic">v</span>/<span class="html-italic">v</span>) of Microcare<sup>®</sup> BNA and 1% (<span class="html-italic">v</span>/<span class="html-italic">v</span>) of Microcare<sup>®</sup> SB, an ethanol control, and the formulation without any preservative (negative control) were also performed.</p>
Full article ">Figure 2
<p>Inhibition of in vitro collagenase activity by 1,10-phenanthroline (positive control) and by peptides (2–10 kDa) at final concentrations of 50, 250, 500, and 1000 µg/mL. Data are presented as means ± SD (error bars) from three independent measurements. Different letters (a–d) indicate significant differences (<span class="html-italic">p</span> &lt; 0.05) between the peptide concentrations tested (50–1000 µg/mL).</p>
Full article ">Figure 3
<p>Cytotoxicity of the peptides (2–10 kDa) in melanoma cell line MNT1 and normal dermal fibroblasts, keratinocytes, and melanocytes. Cells were exposed to increasing concentrations of the peptide for 48 h and viability was evaluated with the MTS assay. Bars represent the average ± standard deviation.</p>
Full article ">Figure 4
<p>Expression levels of TNF-α in normal dermal fibroblasts after incubation with peptides (2–10 kDa). TNF-α expression after 2 h incubation of fibroblasts with 250 µg/mL of the peptides (2–10 kDa), 1% (<span class="html-italic">v</span>/<span class="html-italic">v</span>) ethanol (vehicle) or medium (control), followed by further 2 h incubation with (+LPS, orange bars) or without (−LPS, blue bars). Bars represent the average and standard deviation of at least three experiments. * <span class="html-italic">p</span>-value &lt; 0.05.</p>
Full article ">Figure 5
<p>Capillary-like tube formation by HUVEC. Representative images of cells after 0 h, 2 h and 6 h of cells seeding on top of Matrigel in F12-K medium supplemented with 250 µg/mL peptides (2–10 kDa), or 1% (<span class="html-italic">v</span>/<span class="html-italic">v</span>) ethanol (Control). Scale bar corresponds to 200 µm.</p>
Full article ">Figure 6
<p>Wound healing assay. Representative images of the wound healing assay at 0 h and 24 h after incubation with 1% (<span class="html-italic">v</span>/<span class="html-italic">v</span>) ethanol (control) or 250 µg/mL of peptides (2–10 kDa) in (<b>A</b>) normal dermal fibroblasts, (<b>B</b>) HUVEC, and (<b>C</b>) keratinocytes. Scale bars correspond to 200 µm. Percentage of wound scratch closure (% remission) after 24 h incubation with 1% (<span class="html-italic">v</span>/<span class="html-italic">v</span>) ethanol (control) or 250 µg/mL of peptides (2–10 kDa) in (<b>D</b>) normal dermal fibroblasts, (<b>E</b>) HUVEC and (<b>F</b>) keratinocytes. Bars represent the mean ± SEM of at least two independent experiments. **** <span class="html-italic">p</span> value &lt; 0.0001, ns—statistically not significant.</p>
Full article ">Figure 7
<p>Superoxide scavenging capacity: (<b>A</b>) Plot of absorbance (Abs 560 nm) as a function of time for the different peptides concentrations; (<b>B</b>) Effect of the peptides on the inhibition of the NBT reduction by the PMS/NADH generated superoxide radical.</p>
Full article ">
15 pages, 707 KiB  
Review
Biomarkers Involved in the Pathogenesis of Hemophilic Arthropathy
by Oana Viola Badulescu, Dragos-Viorel Scripcariu, Minerva Codruta Badescu, Manuela Ciocoiu, Maria Cristina Vladeanu, Carmen Elena Plesoianu, Andrei Bojan, Dan Iliescu-Halitchi, Razvan Tudor, Bogdan Huzum, Otilia Elena Frasinariu and Iris Bararu-Bojan
Int. J. Mol. Sci. 2024, 25(18), 9897; https://doi.org/10.3390/ijms25189897 (registering DOI) - 13 Sep 2024
Viewed by 220
Abstract
Hemophilia, which is a rare disease, results from congenital deficiencies of coagulation factors VIII and IX, respectively, leading to spontaneous bleeding into joints, resulting in hemophilic arthropathy (HA). HA involves complex processes, including synovial proliferation, angiogenesis, and tissue remodeling. Despite ongoing research, factors [...] Read more.
Hemophilia, which is a rare disease, results from congenital deficiencies of coagulation factors VIII and IX, respectively, leading to spontaneous bleeding into joints, resulting in hemophilic arthropathy (HA). HA involves complex processes, including synovial proliferation, angiogenesis, and tissue remodeling. Despite ongoing research, factors contributing to HA progression, especially in adults with severe HA experiencing joint pain, remain unclear. Blood markers, particularly collagen-related ones, have been explored to assess joint health in hemophilia. For example, markers like CTX-I and CTX-II reflect bone and cartilage turnover, respectively. Studies indicate elevated levels of certain markers post-bleeding episodes, suggesting joint health changes. However, longitudinal studies on collagen turnover and basement membrane or endothelial cell markers in relation to joint outcomes, particularly during painful episodes, are scarce. Given the role of the CX3CL1/CX3XR1 axis in arthritis, other studies investigate its involvement in HA. The importance of different inflammatory and bone damage biomarkers should be assessed, alongside articular cartilage and synovial membrane morphology, aiming to enhance understanding of hemophilic arthropathy progression. Full article
(This article belongs to the Special Issue Advances in Rare Diseases Biomarkers)
Show Figures

Figure 1

Figure 1
<p>Biomarkers in HA [<a href="#B38-ijms-25-09897" class="html-bibr">38</a>].</p>
Full article ">
36 pages, 1978 KiB  
Review
Prognostic and Predictive Roles of HER2 Status in Non-Breast and Non-Gastroesophageal Carcinomas
by Erica Quaquarini, Federica Grillo, Lorenzo Gervaso, Giovanni Arpa, Nicola Fazio, Alessandro Vanoli and Paola Parente
Cancers 2024, 16(18), 3145; https://doi.org/10.3390/cancers16183145 - 13 Sep 2024
Viewed by 226
Abstract
The oncogene ERBB2, also known as HER2 or c-ERB2, is located on chromosome 17 (q12). It encodes a tyrosine kinase receptor, the human epidermal growth factor receptor 2 (HER2), involved in neoplastic proliferation, tumor angiogenesis, and invasiveness. Over the past years, [...] Read more.
The oncogene ERBB2, also known as HER2 or c-ERB2, is located on chromosome 17 (q12). It encodes a tyrosine kinase receptor, the human epidermal growth factor receptor 2 (HER2), involved in neoplastic proliferation, tumor angiogenesis, and invasiveness. Over the past years, the introduction of various anti-HER2 therapies has significantly improved outcomes for patients with HER2-positive breast and gastroesophageal carcinomas. More recently, the introduction of a new antibody–drug conjugate, that is trastuzumab deruxtecan, expanded the therapeutic options to low-HER2 breast and gastroesophageal tumors. HER2 protein overexpression is investigated using immunohistochemistry, gene amplification using fluorescence in situ hybridization, and gene mutation using next-generation sequencing. This review evaluated the predictive and prognostic role of HER2 status in various types of epithelial malignant cancers beyond breast and gastroesophageal cancers. We critically analyzed the key published studies, focusing on utilized scoring systems and assays used, and analyzed clinical parameters and therapeutic approaches. Although the evidence about prognostic and predictive roles of HER2 in carcinomas other than breast and gastroesophageal has been widely increasing over the last decade, it still remains investigational, revealing a tumor site-related prognostic and predictive value of the different types of HER2 alterations. However, standardized and validated scoring system assays have not been well-established for many organs. Full article
(This article belongs to the Section Cancer Pathophysiology)
Show Figures

Figure 1

Figure 1
<p>Frequency of HER2 alterations across the tumor types described in this review. Figure adapted from Matthew Cole/Vettoriale stock Alamy, 2016, with copyright permission.</p>
Full article ">Figure 2
<p>(<b>A</b>) Histology of a jejunal adenocarcinoma with a HER2 overexpression (IHC 3+). Note the strong membranous expression of HER2 in most tumor cells. (<b>B</b>) The same case of (<b>A</b>) shows a minor component of the score IHC 1+. Note that in the upper-right corner, there is an area of score 3+ expression.</p>
Full article ">
15 pages, 5976 KiB  
Article
Molecular and Functional Cargo of Plasma-Derived Exosomes in Patients with Hereditary Hemorrhagic Telangiectasia
by Yanru Wang, Linda Hofmann, Diana Huber, Robin Lochbaum, Sonja Ludwig, Cornelia Brunner, Thomas K. Hoffmann, René Lehner and Marie-Nicole Theodoraki
J. Clin. Med. 2024, 13(18), 5430; https://doi.org/10.3390/jcm13185430 - 13 Sep 2024
Viewed by 217
Abstract
Background: Hereditary Hemorrhagic Telangiectasia (HHT) is a genetic disorder leading to frequent bleeding in several organs. As HHT diagnosis is demanding and depends on clinical criteria, liquid biopsy would be beneficial. Exosomes from biofluids are nano-sized vesicles for intercellular communication. Their cargo [...] Read more.
Background: Hereditary Hemorrhagic Telangiectasia (HHT) is a genetic disorder leading to frequent bleeding in several organs. As HHT diagnosis is demanding and depends on clinical criteria, liquid biopsy would be beneficial. Exosomes from biofluids are nano-sized vesicles for intercellular communication. Their cargo and characteristics represent biomarkers for many diseases. Here, exosomes of HHT patients were examined regarding their biosignature. Methods: Exosomes were isolated from the plasma of 20 HHT patients and 17 healthy donors (HDs). The total exosomal protein was quantified, and specific proteins were analyzed using Western blot and antibody arrays. Human umbilical vein endothelial cells (HUVECs) co-incubated with exosomes were functionally examined via immunofluorescence, proliferation, and scratch assay. Results: The levels of the angiogenesis-regulating protein Thrombospondin-1 were significantly higher in HHT compared to HD exosomes. Among HHT, but not HD exosomes, a negative correlation between total exosomal protein and soluble Endoglin (sENG) levels was found. Other exosomal proteins (ALK1, ALK5) and the particle concentration significantly correlated with disease severity parameters (total consultations/interventions, epistaxis severity score) in HHT patients. Functionally, HUVECs were able to internalize both HD and HHT exosomes, inducing a similar change in the F-Actin structure and a reduction in migration and proliferation. Conclusions: This study provided first insights into the protein cargo and function of HHT-derived exosomes. The data indicate changes in sENG secretion via exosomes and reveal exosomal Thrombospondin-1 as a potential biomarker for HHT. Several exosomal characteristics were pointed out as potential liquid biomarkers for disease severity, revealing a possible new way of diagnosis and prognosis of HHT. Full article
Show Figures

Figure 1

Figure 1
<p>Exosome characterization. (<b>A</b>) Representative transmission electron microscopy images of plasma-derived exosomes from a healthy donor (HD) and a patient with Hereditary Hemorrhagic Telangiectasia (HHT). Scale bar = 200 nm. (<b>B</b>) Western blot analysis of HD and HHT exosomes for exosomal markers (CD63, CD9, TSG101), the cellular marker Grp94, and lipoprotein ApoA1. A cell lysate (C) and unprocessed plasma (P) served as positive controls. (<b>C</b>) Representative size distributions of plasma-derived exosomes from HD and HHT patient determined via nanoparticle tracking analysis. (<b>D</b>) Quantitative characteristics of plasma-derived exosomes from HD (n = 17) and HHT patients (n = 18). Box-and-whisker plots represent the median, the 25th and 75th quartiles, and the range.</p>
Full article ">Figure 2
<p>Soluble Endoglin levels in HD and HHT exosomes. (<b>A</b>) Representative Western blot of HD and HHT exosomes for the angiogenesis-related protein sENG and the exosomal marker TSG101. Numbers below lanes indicate band intensities of sENG normalized to TSG101 using lane normalization factors. (<b>B</b>) Normalized sENG values of HD and HHT exosomes (n = 17). Age- and gender-matched pairs of HDs and HHT patients whose plasma was used for exosome isolation are connected by a line. (<b>C</b>) Normalized sENG values of HHT exosomes (n = 17) were correlated to total exosomal protein levels, as determined by Bicinchoninic acid (BCA) assay. Spearman’s rank correlation coefficient (r) and correlation significance (p) were calculated. (<b>D</b>) Normalized sENG values of HD exosomes (n = 17) were correlated to the total exosomal protein levels, as determined by BCA assay. Pearson coefficient of correlation (R2) and correlation significance (p) are shown.</p>
Full article ">Figure 3
<p>Angiogenesis-related proteins in HD and HHT exosomes. (<b>A</b>) Representative antibody array analysis of HD and HHT exosomes (n = 2) for angiogenesis-related proteins including Thrombospondin-1 and platelet factor 4 (PF4). The reference spots’ signal density was used for normalization and quantification of protein levels. (<b>B</b>) Normalized pixel density of Thrombospondin-1 and PF4 was compared between HD and HHT exosomes (n = 6). Box-and-whisker plots represent the median, the 25th and 75th quartiles, and the range. Mann–Whitney test was applied for comparison between HD and HHT exosomes with * corresponding to <span class="html-italic">p</span> ≤ 0.05.</p>
Full article ">Figure 4
<p>Structural changes in HUVECs after exosome incubation. (<b>A</b>) Fluorescence microscopy images at 400× magnification, scale bar = 50 µm. HUVECs were incubated with PKH26-labeled exosomes (orange) for 1 h, 4 h, or 16 h. F-Actin filaments are shown in green, and nuclei in blue (DAPI). (<b>B</b>) Representative images of F-actin structure in HUVECs after exosome incubation. HUVECs were incubated with PBS as control, HD exosomes, or HHT exosomes for 24 h. F-Actin filaments are shown in green, and nuclei in blue (DAPI). (<b>C</b>) Tube formation assay on HUVECs after exosome incubation. HUVECs grown in an extracellular matrix were incubated with PBS as control (n = 4), HD, or HHT exosomes (n = 3) for 16 h. Representative fluorescence microscopy images (50× magnification, scale bar = 500 µm) show the results of tube formation. Using ImageJ 1.53k, the total length of tubes in the analyzed area was calculated for the three groups of treated HUVECs.</p>
Full article ">Figure 5
<p>Proliferation and migration of HUVECs after exosome incubation. (<b>A</b>) Carboxy-fluorescein succinimidyl ester (CFSE) proliferation assay as determined via flow cytometry. CFSE-labeled HUVECs were incubated with PBS as control (n = 4), HD (n = 4) or HHT exosomes (n = 5) for 24 h. Then, these control and exosome-primed HUVECs were analyzed via flow cytometry after 0 h, 24 h, and 48 h. Representative flow cytometry histograms are shown. (<b>B</b>) Box-and-whisker blots show the median, the 25th and 75th quartiles, and the range of the percentage of proliferated HUVECs after 24 h and 48 h. Mann–Whitney test was applied for comparison between groups with * corresponding to <span class="html-italic">p</span> ≤ 0.05. (<b>C</b>) Wound healing assay of HUVECs after exosome incubation. Representative light microscopy images (50× magnification, scale bar = 500 µm) of HUVECs incubated with PBS as control (n = 9), HD (n = 11) or HHT exosomes (n = 13) for 24 h. Then, scratches were induced and documented 0 h, 24 h, and 48 h afterwards. (<b>D</b>) Using ImageJ, the gap width of each scratch was calculated (<a href="#app1-jcm-13-05430" class="html-app">Figure S1</a>). Box-and-whisker blots show the median, the 25th and 75th quartiles, and the range of the gap width reduction 24h and 48h after incubation. Mann–Whitney test was applied for comparison between groups with *** and **** corresponding to <span class="html-italic">p</span> ≤ 0.001 and <span class="html-italic">p</span> ≤ 0.0001.</p>
Full article ">Figure 6
<p>Correlations between exosomal and clinical parameters. (<b>A</b>) Normalized ALK1 values of HHT exosomes (n = 17) were correlated to the number of total consultations and total interventions. (<b>B</b>) Normalized ALK5 values of HHT exosomes (n = 16) were correlated to ESSs at the time of inclusion. (<b>C</b>) Particle concentration of HHT exosome samples (n = 17) was correlated to Epistaxis severity scores (ESS) at the time of inclusion. Pearson coefficient of correlation (R2) and correlation significance (p) are shown for all correlations.</p>
Full article ">
12 pages, 3217 KiB  
Communication
Pravastatin Protects Cytotrophoblasts from Hyperglycemia-Induced Preeclampsia Phenotype
by Ahmed F. Pantho, Sara Mohamed, Janhavi V. Govande, Riddhi Rane, Niraj Vora, Kelsey R. Kelso, Thomas J. Kuehl, Steven R. Lindheim and Mohammad N. Uddin
Cells 2024, 13(18), 1534; https://doi.org/10.3390/cells13181534 - 13 Sep 2024
Viewed by 190
Abstract
There are no effective therapies to prevent preeclampsia (PE). Pravastatin shows promise by attenuating processes associated with PE such as decreased cytotrophoblast (CTB) migration, aberrant angiogenesis, and increased oxidative stress. This study assesses the effects of pravastatin on hyperglycemia-induced CTB dysfunction. Methods: Human [...] Read more.
There are no effective therapies to prevent preeclampsia (PE). Pravastatin shows promise by attenuating processes associated with PE such as decreased cytotrophoblast (CTB) migration, aberrant angiogenesis, and increased oxidative stress. This study assesses the effects of pravastatin on hyperglycemia-induced CTB dysfunction. Methods: Human CTB cells were treated with 100, 150, 200, 300, or 400 mg/dL glucose for 48 h. Some cells were pretreated with pravastatin (1 µg/mL), while others were cotreated with pravastatin and glucose. The expression of urokinase plasminogen activator (uPA), plasminogen activator inhibitor 1 (PAI-1) mRNA, vascular endothelial growth factor (VEGF), placenta growth factor (PlGF), soluble fms-like tyrosine kinase-1 (sFlt-1), and soluble endoglin (sEng) were measured. CTB migration was assayed using a CytoSelect migration assay kit. Statistical comparisons were performed using an analysis of variance with Duncan’s post hoc test. Results: The hyperglycemia-induced downregulation of uPA was attenuated in CTB cells pretreated with pravastatin at glucose levels > 200 mg/dL and cotreated at glucose levels > 300 mg/dL (p < 0.05). Hyperglycemia-induced decreases in VEGF and PlGF and increases in sEng and sFlt-1 were attenuated in both the pretreatment and cotreatment samples regardless of glucose dose (p < 0.05). Pravastatin attenuated hyperglycemia-induced dysfunction of CTB migration. Conclusions: Pravastatin mitigates stress signaling responses in hyperglycemic conditions, weakening processes leading to abnormal CTB migration and invasion associated with PE in pregnancy. Full article
(This article belongs to the Special Issue Signaling Pathways in Pregnancy)
Show Figures

Figure 1

Figure 1
<p>Plot of uPA gene expression for cytotrophoblast cells responding in vitro to hyperglycemia and either pre- or cotreatment with pravastatin. uPA gene expression was increased in CTB cells pretreated with pravastatin at glucose levels &gt; 200 mg/dL, whereas this effect was not seen with cotreatment until &gt;300 mg/dL (<span class="html-italic">n</span> = 6, four replicates each; <span class="html-italic">p</span> &lt; 0.05). Both * and † are statistically significant.</p>
Full article ">Figure 2
<p>Plot of PAI-1 gene expression relative to GAPDH for cytotrophoblast cells responding in vitro to hyperglycemia and either pre- or cotreatment with pravastatin. PAI-1 gene expression was increased in both the pretreatment and cotreatment samples regardless of glucose dose (<span class="html-italic">n</span> = 6, four replicates each; <span class="html-italic">p</span> &lt; 0.05). Both * and † are statistically significant.</p>
Full article ">Figure 3
<p>(<b>A</b>) Plot of VEGF concentration relative to concentration of VEGF at 100 mg/dL glucose concentration, with CTB cells responding to increasing glucose concentrations and no treatment, cotreatment, or pretreatment with pravastatin. Significant increases in VEGF concentrations were seen for both cotreatment and pretreatment groups at all supraphysiologic glucose levels. Both pravastatin pretreatment and cotreatment rescued CTB cells from hyperglycemia-induced downregulation of VEGF. Treatment groups differ (<span class="html-italic">p</span> &lt; 0.001 using ANOVA). Means with different letters differ (<span class="html-italic">n</span> = 8, four replicates each; <span class="html-italic">p</span> &lt; 0.05 using Duncan’s post hoc test). (<b>B</b>) Plot of PlGF concentration relative to concentration of PlGF at 100 mg/dL glucose concentration, with CTB cells responding to increasing glucose concentrations and no treatment, cotreatment, or pretreatment with pravastatin. Significant increases in PlGF concentrations were seen for both cotreatment and pretreatment groups at all supraphysiologic glucose levels. Both pravastatin pretreatment and cotreatment rescued CTB cells from hyperglycemia-induced downregulation of PlGF. Treatment groups differ (<span class="html-italic">p</span> &lt; 0.001 using ANOVA). Means with different letters differ (<span class="html-italic">n</span> = 8, four replicates each; <span class="html-italic">p</span> &lt; 0.05 using Duncan’s post hoc test). (<b>C</b>) Plot of sFLT-1 concentration relative to concentration of sFLT-1 at 100 mg/dL glucose concentration, with CTB cells responding to increasing glucose concentrations and no treatment, cotreatment, or pretreatment with pravastatin. Significant decreases in sFLT-1 concentrations were seen for both cotreatment and pretreatment groups at all supraphysiologic glucose levels. Both pravastatin pretreatment and cotreatment rescued CTB cells from hyperglycemia-induced upregulation of sFlt-1. Treatment groups differ (<span class="html-italic">p</span> &lt; 0.001 using ANOVA). Means with different letters differ (<span class="html-italic">n</span> = 8, four replicates each; <span class="html-italic">p</span> &lt; 0.05 using Duncan’s post hoc test). (<b>D</b>) Plot of sENG concentration relative to concentration of sENG at 100 mg/dL glucose concentration, with CTB cells responding to increasing glucose concentrations and no treatment, cotreatment, or pretreatment with pravastatin. Significant decreases in sENG concentrations were seen for both cotreatment and pretreatment groups at all supraphysiologic glucose levels. Both pravastatin pretreatment and cotreatment rescued CTB cells from hyperglycemia-induced upregulation of sENG. Treatment groups differ (<span class="html-italic">p</span> &lt; 0.001 using ANOVA). Means with different letters differ (<span class="html-italic">n</span> = 8, four replicates each; <span class="html-italic">p</span> &lt; 0.05 using Duncan’s post hoc test).</p>
Full article ">Figure 3 Cont.
<p>(<b>A</b>) Plot of VEGF concentration relative to concentration of VEGF at 100 mg/dL glucose concentration, with CTB cells responding to increasing glucose concentrations and no treatment, cotreatment, or pretreatment with pravastatin. Significant increases in VEGF concentrations were seen for both cotreatment and pretreatment groups at all supraphysiologic glucose levels. Both pravastatin pretreatment and cotreatment rescued CTB cells from hyperglycemia-induced downregulation of VEGF. Treatment groups differ (<span class="html-italic">p</span> &lt; 0.001 using ANOVA). Means with different letters differ (<span class="html-italic">n</span> = 8, four replicates each; <span class="html-italic">p</span> &lt; 0.05 using Duncan’s post hoc test). (<b>B</b>) Plot of PlGF concentration relative to concentration of PlGF at 100 mg/dL glucose concentration, with CTB cells responding to increasing glucose concentrations and no treatment, cotreatment, or pretreatment with pravastatin. Significant increases in PlGF concentrations were seen for both cotreatment and pretreatment groups at all supraphysiologic glucose levels. Both pravastatin pretreatment and cotreatment rescued CTB cells from hyperglycemia-induced downregulation of PlGF. Treatment groups differ (<span class="html-italic">p</span> &lt; 0.001 using ANOVA). Means with different letters differ (<span class="html-italic">n</span> = 8, four replicates each; <span class="html-italic">p</span> &lt; 0.05 using Duncan’s post hoc test). (<b>C</b>) Plot of sFLT-1 concentration relative to concentration of sFLT-1 at 100 mg/dL glucose concentration, with CTB cells responding to increasing glucose concentrations and no treatment, cotreatment, or pretreatment with pravastatin. Significant decreases in sFLT-1 concentrations were seen for both cotreatment and pretreatment groups at all supraphysiologic glucose levels. Both pravastatin pretreatment and cotreatment rescued CTB cells from hyperglycemia-induced upregulation of sFlt-1. Treatment groups differ (<span class="html-italic">p</span> &lt; 0.001 using ANOVA). Means with different letters differ (<span class="html-italic">n</span> = 8, four replicates each; <span class="html-italic">p</span> &lt; 0.05 using Duncan’s post hoc test). (<b>D</b>) Plot of sENG concentration relative to concentration of sENG at 100 mg/dL glucose concentration, with CTB cells responding to increasing glucose concentrations and no treatment, cotreatment, or pretreatment with pravastatin. Significant decreases in sENG concentrations were seen for both cotreatment and pretreatment groups at all supraphysiologic glucose levels. Both pravastatin pretreatment and cotreatment rescued CTB cells from hyperglycemia-induced upregulation of sENG. Treatment groups differ (<span class="html-italic">p</span> &lt; 0.001 using ANOVA). Means with different letters differ (<span class="html-italic">n</span> = 8, four replicates each; <span class="html-italic">p</span> &lt; 0.05 using Duncan’s post hoc test).</p>
Full article ">Figure 4
<p>Both pravastatin pretreatment and cotreatment rescued hyperglycemia-induced CTB cell migration: Serum-starved CTB cells were treated with 100, 150, 200, 300, or 400 mg/dL of glucose (Sigma) for 48 h. Some cells were pretreated with 1 µg/mL of pravastatin before treatment with glucose, while others were cotreated with 1 µg/mL of pravastatin and the above glucose levels. All treated cells were subsequently added to transwell inserts that contained 10 ng/mL EGF and/or 100 ng/mL HGF. CTB cell migration was significantly (* <span class="html-italic">p</span> &lt; 0.05) inhibited by ≥150 mg/dL of glucose that was significantly (* <span class="html-italic">p</span> &lt; 0.05) attenuated by both pretreatment and cotreatment with 1.0 µg/mL pravastatin. Results are presented as mean ± SEM (<span class="html-italic">n</span> = 5, four replicates each). Both * and † are statistically significant.</p>
Full article ">Figure 5
<p>A new model summarizing the inhibitory effect of pravastatin on the stress-signaling pathway that leads to abnormal placentation.</p>
Full article ">
12 pages, 2705 KiB  
Article
3D-Cultured MC3T3-E1-Derived Exosomes Promote Endothelial Cell Biological Function under the Effect of LIPUS
by Xiaohan Liu, Rui Cheng, Hongjuan Cao and Lin Wu
Biomolecules 2024, 14(9), 1154; https://doi.org/10.3390/biom14091154 - 13 Sep 2024
Viewed by 238
Abstract
Porous Ti-6Al-4V scaffold materials can be used to heal massive bone defects because they can provide space for vascularisation and bone formation. During new bone tissue development, rapid vascular ingrowth into scaffold materials is very important. Osteoblast-derived exosomes are capable of facilitating angiogenesis–osteogenesis [...] Read more.
Porous Ti-6Al-4V scaffold materials can be used to heal massive bone defects because they can provide space for vascularisation and bone formation. During new bone tissue development, rapid vascular ingrowth into scaffold materials is very important. Osteoblast-derived exosomes are capable of facilitating angiogenesis–osteogenesis coupling. Low-intensity pulsed ultrasound (LIPUS) is a physical therapy modality widely utilised in the field of bone regeneration and has been proven to enhance the production and functionality of exosomes on two-dimensional surfaces. The impact of LIPUS on exosomes derived from osteoblasts cultured in three dimensions remains to be elucidated. In this study, exosomes produced by osteoblasts on porous Ti-6Al-4V scaffold materials under LIPUS and non-ultrasound stimulated conditions were co-cultured with endothelial cells. The findings indicated that the exosomes were consistently and stably taken up by the endothelial cells. Compared to the non-ultrasound group, the LIPUS group facilitated endothelial cell proliferation and angiogenesis. After 24 h of co-culture, the migration ability of endothelial cells in the LIPUS group was 17.30% higher relative to the non-ultrasound group. LIPUS may represent a potentially viable strategy to promote the efficacy of osteoblast-derived exosomes to enhance the angiogenesis of porous Ti-6Al-4V scaffold materials. Full article
Show Figures

Figure 1

Figure 1
<p>Isolation and identification of exosomes. The separated exosomes were characterised by scanning electron microscopy, nanoparticle tracking analysis (NTA), and identification of surface marker proteins. (<b>A</b>) Methodology for ultrasonic loading of cells and extraction and identification of exosomes. (<b>B</b>) SEM images of MC3T3-E1 were grown within the porous Ti-6Al-4V scaffold. (<b>C</b>) Morphology of exosomes under electron microscopy, scale bar = 200 nm. (<b>D</b>) Particle size distribution and concentration in the exosome suspension, with the size of particles in both LIPUS and control groups ranging between 30 and 300 nm, consistent with known exosomal dimensions. (<b>E</b>) Expression of HSP70, TSG101, and calreticulin in exosomes from each group. Western blot original images can be found in <a href="#app1-biomolecules-14-01154" class="html-app">Supplementary Materials</a>. (<b>F</b>) Measurement of exosome concentration in three independent experiments showed no significant difference between the LIPUS and control groups. ns: <span class="html-italic">p</span> ˃ 0.05.</p>
Full article ">Figure 2
<p>Endothelial cells took up Dil-labelled exosomes. After co-culturing with pre-stained exosomes for 24 h, red fluorescence signals could be observed inside both LIPUS and control group HUVEC cells, with no significant difference in fluorescence signal intensity between the two groups.</p>
Full article ">Figure 3
<p>Exosomes extracted from MC3T3-E1 cells stimulated by ultrasound better promote endothelial cell migration. (<b>A</b>) Schematic of functional experiments. (<b>B</b>) CCK8 assay to assess endothelial cell proliferation. (<b>C</b>) Transwell assay to assess exosome-promoted endothelial cell migration ability, scale bar = 100 μm. (<b>D</b>) Statistical analysis results of endothelial cell migration. One-way analysis of variance was used for testing, and pairwise comparisons were conducted using a post hoc LSD test, n = 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 4
<p>Endothelial cell tube formation experiment. (<b>A</b>) Under microscopic observation, scale bar = 100 um. (<b>B</b>) Image J was used for statistical analysis of the tube formation experiment results: (a) total tube length, (b) number of nodes, (c) number of junctions, (d) number of branches. One-way analysis of variance was used for testing, and pairwise comparisons were conducted using a post hoc LSD test, n = 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 ">
16 pages, 9164 KiB  
Article
Exploring the Role of Fibrin Gels in Enhancing Cell Migration for Vasculature Formation
by Joana A. Moura, Hugh J. Barlow, Shareen H. Doak, Karl Hawkins, Iris Muller and Martin J. D. Clift
J. Funct. Biomater. 2024, 15(9), 265; https://doi.org/10.3390/jfb15090265 - 12 Sep 2024
Viewed by 306
Abstract
A hallmark of angiogenesis is the sprouting of endothelial cells. To replicate this event in vitro, biomaterial approaches can play an essential role in promoting cell migration. To study the capacity of a scaffold of fibrin (fibrinogen:thrombin mix) to support the movement of [...] Read more.
A hallmark of angiogenesis is the sprouting of endothelial cells. To replicate this event in vitro, biomaterial approaches can play an essential role in promoting cell migration. To study the capacity of a scaffold of fibrin (fibrinogen:thrombin mix) to support the movement of the endothelial cells, the migration area of spheroids formed with the HULEC cell line was measured. The cells were first allowed to form a spheroid using the hanging drop technique before being encapsulated in the fibrin gel. The cells’ migration area was then measured after two days of embedding in the fibrin gel. Various conditions affecting fibrin gel polymerization, such as different concentrations of fibrinogen and thrombin, were evaluated alongside rheology, porosity, and fiber thickness analysis to understand how these factors influenced cell behavior within the composite biomaterial. Data point toward thrombin’s role in governing fibrin gel polymerization; higher concentrations result in less rigid gels (loss tangent between 0.07 and 0.034) and increased cell migration (maximum concentration tested: 5 U/mL). The herein presented method allows for a more precise determination of the crosslinking conditions of fibrin gel that can be used to stimulate angiogenic sprouting. Full article
(This article belongs to the Topic Advanced Functional Materials for Regenerative Medicine)
Show Figures

Figure 1

Figure 1
<p>Endothelial cell 2D characterization. (<b>A</b>) Live cell quantification with trypan blue over 7 days. The cell growth fits the following exponential curve: y = 5381e0.2741x. (<b>B</b>) Live and dead cell ratio using trypan blue assay. (<b>C</b>) Picogreen dsDNA quantification over 7 days on the left y-axis. The increase in dsDNA is statistically significant from day 1 to 7 with a <span class="html-italic">p</span>-value of 0.0070. The right y-axis corresponds to the MTS absorbance over the number of live cells. The cell’s metabolism increases from day 1 to 7 and is also statistically significant, with a <span class="html-italic">p</span>-value of 0.0040. (<b>D</b>) Expression of the CD31 marker on days 5 and 7. The increase in CD31 surface marker from day 5 to 7 is statistically significant, with a <span class="html-italic">p</span>-value of 0.044. This is associated also with a decrease in CD31, with a <span class="html-italic">p</span>-value of 0.044. (<b>E</b>) Endothelial cells on day 7 of the cell culture stained with DAPI, Phalloidin, and CD31 (scale bar represents 50 µm). (<b>F</b>) VE-Cadherin (scale bar represents 50 µm). The significance values were taken when <span class="html-italic">p</span>-No value &lt; 0.05, graphically denoted as * <span class="html-italic">p</span> &lt; 0.05 and ** <span class="html-italic">p</span> &lt; 0.01. The error bars correspond to the Standard Deviation. <span class="html-italic">n</span> = 3.</p>
Full article ">Figure 2
<p>Fibrin gel rheological characterization. (<b>A</b>) Storage (G′) and (<b>B</b>) Loss modulus (G″) of fibrin gel in Pascal (Pa). The data are grouped by the fibrinogen concentration. (<b>C</b>) Storage (G′) and (<b>D</b>) Loss modulus (G″) of fibrin gel in Pascal (Pa). The data are grouped by the thrombin concentration. (<b>E</b>) The table below shows the loss tangent values. The color scheme highlights values in the same range. The fibrinogen concentrations tested were 5, 2.5, and 1.25 mg/mL, and for thrombin, they were 5, 1, and 0.1 U/mL. The error bars correspond to the Standard Deviation. The data are between <span class="html-italic">n</span> = 3 and 6. The significance values were taken when <span class="html-italic">p</span> &lt; 0.05, graphically denoted as * <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 and **** <span class="html-italic">p</span> &lt; 0.0001.</p>
Full article ">Figure 3
<p>Confocal microscope images of fibrin gel polymerized with 5, 2.5, and 1.25 mg/mL of fibrinogen and 5, 1, and 0.1 U/mL of thrombin. The fibrinogen solution was mixed with 4% of fibrinogen from human plasma Alexa Fluor 488 conjugated to allow for fiber visualization when excited with a 488 laser. The scale represents 20 µm. <span class="html-italic">n</span> = 6 to 9.</p>
Full article ">Figure 4
<p>Scanning electron microscope images of fibrin gel polymerized with 5, 2.5, and 1.25 mg/mL of fibrinogen and 5, 1, and 0.1 U/mL of thrombin. The scale bar represents 5 µm, except for condition 1.25 mg/mL with 0.1 U/mL, which represents 4 µm. <span class="html-italic">n</span> = 3.</p>
Full article ">Figure 5
<p>Fibrin gel porosity and fiber diameter. (<b>A</b>) Percentage of the area corresponding to pores in the fibrin gel in function of fibrinogen and thrombin concentration (<span class="html-italic">n</span> = 6 to 9). (<b>B</b>) Fiber diameter in micrometers (µm) in function of fibrinogen and thrombin concentration (<span class="html-italic">n</span> = 3). The significance values were taken when <span class="html-italic">p</span> &lt; 0.05, graphically denoted as * <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. The error bars represent the Standard Deviation.</p>
Full article ">Figure 6
<p>Endothelial cell migration area in function of the fibrinogen and thrombin polymerization conditions. The asterisk symbol in the plot represents * <span class="html-italic">p</span> &lt; 0.05. The error bars represent the Standard Deviation. <span class="html-italic">n</span> = 5.</p>
Full article ">Figure 7
<p>Spheroid bright and fluorescence staining. Spheroids on day 2 after fibrin embedding. The column on the left represents hematoxylin stain, and on the right, Hoechst and PI. The scale bar represents 200 µm. <span class="html-italic">n</span> = 5.</p>
Full article ">
21 pages, 1224 KiB  
Review
Towards Targeting Endothelial Rap1B to Overcome Vascular Immunosuppression in Cancer
by Behshid Ghadrdoost Nakhchi, Ramoji Kosuru and Magdalena Chrzanowska
Int. J. Mol. Sci. 2024, 25(18), 9853; https://doi.org/10.3390/ijms25189853 - 12 Sep 2024
Viewed by 404
Abstract
The vascular endothelium, a specialized monolayer of endothelial cells (ECs), is crucial for maintaining vascular homeostasis by controlling the passage of substances and cells. In the tumor microenvironment, Vascular Endothelial Growth Factor A (VEGF-A) drives tumor angiogenesis, leading to endothelial anergy and vascular [...] Read more.
The vascular endothelium, a specialized monolayer of endothelial cells (ECs), is crucial for maintaining vascular homeostasis by controlling the passage of substances and cells. In the tumor microenvironment, Vascular Endothelial Growth Factor A (VEGF-A) drives tumor angiogenesis, leading to endothelial anergy and vascular immunosuppression—a state where ECs resist cytotoxic CD8+ T cell infiltration, hindering immune surveillance. Immunotherapies have shown clinical promise. However, their effectiveness is significantly reduced by tumor EC anergy. Anti-angiogenic treatments aim to normalize tumor vessels and improve immune cell infiltration. Despite their potential, these therapies often cause significant systemic toxicities, necessitating new treatments. The small GTPase Rap1B emerges as a critical regulator of Vascular Endothelial Growth Factor Receptor 2 (VEGFR2) signaling in ECs. Our studies using EC-specific Rap1B knockout mice show that the absence of Rap1B impairs tumor growth, alters vessel morphology, and increases CD8+ T cell infiltration and activation. This indicates that Rap1B mediates VEGF-A’s immunosuppressive effects, making it a promising target for overcoming vascular immunosuppression in cancer. Rap1B shares structural and functional similarities with RAS oncogenes. We propose that targeting Rap1B could enhance therapies’ efficacy while minimizing adverse effects by reversing endothelial anergy. We briefly discuss strategies successfully developed for targeting RAS as a model for developing anti-Rap1 therapies. Full article
Show Figures

Figure 1

Figure 1
<p>Targeting EC Rap1B to overcome VEGF-A-induced vascular immuno-suppression—a model. (<b>A</b>) Rap1B-GTP promotes Vascular Endothelial Growth Factor (VEGF) Receptor 2 (VEGFR2) signaling, inhibits proinflammatory signaling, and is a potential anti-cancer target. (<b>B</b>) Deletion of Rap1B in ECs inhibits tumor growth and promotes leukocyte infiltration (Rap1B<sup>iΔEC</sup> mice). (<b>C</b>) Rap1B mediates VEGF-A-induced suppression of proinflammatory nuclear factor κ-light chain enhancer of activated B cells (NF-κB) signaling, including cell adhesion molecule (CAM) expression, limiting T cell adhesion and recruitment. ICAM—intracellular adhesion molecule; IκB—inhibitor of κB; LFA-1—lymphocyte function associated antigen 1 (integrin αLβ2); TNFR1—tumor necrosis factor receptor 1; VCAM—vascular CAM; VLA-4—very late antigen 4 (integrin α4β1). Signal transduction is indicated by arrows.</p>
Full article ">Figure 2
<p>Strategies for targeting WT and mutant RAS. All structures were visualized using PyMOL, with surface models highlighting the binding interfaces; adapted from [<a href="#B87-ijms-25-09853" class="html-bibr">87</a>]. (<b>A</b>) Direct targeting of GDP-bound KRAS-G12C with covalent inhibitor AMG-510 (sotorasib), binding to the switch II pocket, in orange (PDB: 6OIM). (<b>B</b>) Direct targeting of wild-type (WT) and G12C KRAS with RMC-7977 (PDB: 4OBE). RMC-7977 binds to the switch II groove (SIIG) of RAS; (<b>C</b>) GDP-bound KRAS with the SOS1-mediated nucleotide exchange inhibitor DCAI (PDB: 4DST). The surface model highlights the DCAI pocket in yellow. (<b>D</b>,<b>E</b>). Indirect targeting of RAS: surfaces targeted by inhibitors of SOS (a RAS guanine nucleotide exchange factor, GEF, (<b>D</b>)) or effector protein binding (RAS-binding domain, RBD, (<b>E</b>)) (PDB: 6GJ8). (<b>F</b>) Targeting post-translational modification of RAS with tipifarnib (PDB: 4JV6). This structure shows KRAS in complex with farnesyltransferase and the inhibitor tipifarnib, preventing farnesylation of the HVR within CAAX motif. (<b>G</b>) Allosteric inhibition of Ras by the NS1 monobody (PDB: 5E95). The NS1 binding site is highlighted in purple.</p>
Full article ">
11 pages, 528 KiB  
Review
Copper Serum Levels in the Hemodialysis Patient Population
by Guido Gembillo, Luigi Peritore, Vincenzo Labbozzetta, Alfio Edoardo Giuffrida, Antonella Lipari, Eugenia Spallino, Vincenzo Calabrese, Luca Visconti and Domenico Santoro
Medicina 2024, 60(9), 1484; https://doi.org/10.3390/medicina60091484 - 11 Sep 2024
Viewed by 346
Abstract
Copper is an essential element in the diet of mammals, including humans. It plays an important role in the physiological regulation of various enzymes and is consequently involved in several biological processes such as angiogenesis, oxidative stress regulation, neuromodulation, and erythropoiesis. Copper is [...] Read more.
Copper is an essential element in the diet of mammals, including humans. It plays an important role in the physiological regulation of various enzymes and is consequently involved in several biological processes such as angiogenesis, oxidative stress regulation, neuromodulation, and erythropoiesis. Copper is essential for facilitating the transfer of iron from cells to the bloodstream, which is necessary for proper absorption of dietary iron and the distribution of iron throughout the body. In particular, patients with end-stage renal failure who require renal replacement therapy are at increased risk for disorders of copper metabolism. Many studies on hemodialysis, peritoneal dialysis, and kidney transplant patients have focused on serum copper levels. Some reported mild deficiency, while others reported elevated levels or even toxicity. In some cases, it has been reported that alterations in copper metabolism lead to an increased risk of cardiovascular disease, malnutrition, anemia, or mielopathy. The aim of this review is to evaluate the role of copper in patients undergoing hemodialysis and its potential clinical implications. Full article
(This article belongs to the Section Urology & Nephrology)
Show Figures

Figure 1

Figure 1
<p>The link between copper, hemodialysis, and continuous renal replacement therapy.</p>
Full article ">
25 pages, 9989 KiB  
Article
Characterization of Tissue Immunity Defense Factors of the Lip in Primary Dentition Children with Bilateral Cleft Lip Palate
by Laura Ozola and Mara Pilmane
J. Pers. Med. 2024, 14(9), 965; https://doi.org/10.3390/jpm14090965 - 11 Sep 2024
Viewed by 245
Abstract
Background: Bilateral cleft lip palate is a severe congenital birth defect of the mouth and face. Immunity factors modulate immune response, inflammation, and healing; therefore, they are vital in the assessment of the immunological status of the patient. The aim of this study [...] Read more.
Background: Bilateral cleft lip palate is a severe congenital birth defect of the mouth and face. Immunity factors modulate immune response, inflammation, and healing; therefore, they are vital in the assessment of the immunological status of the patient. The aim of this study is to assess the distribution of Gal-10, CD-163, IL-4, IL-6, IL-10, HBD-2, HBD-3, and HBD-4 in tissue of the bilateral cleft lip palate in primary dentition children. Methods: Five patients underwent cheiloplasty surgery, where five tissue samples of lip were obtained. Immunohistochemical staining, semi-quantitative evaluation, and non-parametric statistical analysis were used. Results: A statistically significant increase in HBD-2, HBD-3, and HBD-4 was found in skin and mucosal epithelium, hair follicles, and blood vessels. A notable increase was also noted in IL-4, IL-6, and IL-10 in the mucosal epithelium and CD163 in blood vessels. The connective tissue of patients presented with a statistically significant decrease in Gal-10, IL-10, and HBD-3. Spearman’s rank correlation revealed multiple significant positive and negative correlations between the factors. Conclusions: Upregulation of CD163 points to increased angiogenesis but the increase in IL-4 and IL-10 as well as the decrease in Gal-10 points to suppression of excessive inflammatory damage. Decreased connective tissue healing and excessive scarring are suggested by the decrease in HBD-3 and IL-10 and the increase in IL-6. Full article
(This article belongs to the Special Issue New Updates in Oral and Maxillofacial Surgery)
Show Figures

Figure 1

Figure 1
<p>Workflow of patient tissue sample selection, processing, and research.</p>
Full article ">Figure 2
<p>Hematoxylin and eosin routine staining of the patient tissue samples (<b>a,b</b>): note vacuolization (<b>a</b>) and inflammatory cell infiltration (<b>b</b>) in skin type epithelium of the lip (arrows). Magnification 200×.</p>
Full article ">Figure 3
<p>Immunohistochemistry of the Gal-10-positive structures in the control and patient tissue samples: (<b>a</b>) control sample with moderate to numerous Gal-10-positive structures in skin epithelium, and moderate in mucosal epithelium, sweat gland ducts, and connective tissue (arrows), 200×; (<b>b</b>) patient sample with few to moderate Gal-10-positive structures in blood vessels, moderate to numerous in skin epithelium, and numerous in mucosal epithelium (arrows), 200×.</p>
Full article ">Figure 4
<p>Immunohistochemistry of the CD-163-positive structures in the control and patient tissue samples: (<b>a</b>) control sample with rare occurrence of CD-163-positive structures in mucosal epithelium and sweat gland ducts, few in the blood vessels, and moderate in connective tissue (arrows), 200×; (<b>b</b>) patient sample with few to moderate CD-163-positive structures in mucosal epithelium and moderate in the skin epithelium (arrows), 200×.</p>
Full article ">Figure 5
<p>Immunohistochemistry of the IL-4-positive structures in the control and patient tissue samples: (<b>a</b>) control sample with no IL-4-positive structures in skin and mucosal epithelium and adipose glands, with rare occurrence in blood vessels, hair follicles, and connective tissue (arrows), 200×; (<b>b</b>) patient sample with few IL-4-positive structures in skin epithelium and connective tissue (arrows), 200×.</p>
Full article ">Figure 6
<p>Immunohistochemistry of the IL-6-positive structures in the control and patient tissue samples: (<b>a</b>) control sample with few IL-6-positive structures in the skin epithelium and moderate in the blood vessels, sweat gland ducts, and connective tissue (arrows), 200×; (<b>b</b>) patient sample with few to moderate IL-6-positive structures in blood vessels, with moderate in skin epithelium and sweat gland ducts, and moderate to numerous in mucosal epithelium and connective tissue (arrows), 200×.</p>
Full article ">Figure 7
<p>Immunohistochemistry of the HBD-2-positive structures in the control and patient tissue samples: (<b>a</b>) control sample with the absence of HBD-2-positive structures in skin epithelium, blood vessels, and connective tissue (arrows), 200×; (<b>b</b>) patient sample with rare occurrence of HBD-2-positive structures in connective tissue, moderate to numerous in skin epithelium and adipose glands, and numerous to abundant in mucosal epithelium (arrows), 200×.</p>
Full article ">Figure 8
<p>Immunohistochemistry of the HBD-3-positive structures in the control and patient tissue samples: (<b>a</b>) control sample with few HBD-3-positive structures in skin and mucosal epithelium, few to moderate in blood vessels, numerous in adipose glands, moderate to numerous in hair follicles, numerous in connective tissue (arrows), 200×; (<b>b</b>) patient sample with rare HBD-3-positive structures in connective tissue, few to moderate in blood vessels and skin epithelium, moderate in mucosal epithelium, numerous in adipose glands and hair follicles, numerous to abundant in sweat gland ducts (arrows), 200×.</p>
Full article ">Figure 9
<p>Immunohistochemistry of the HBD-4-positive structures in the control and patient tissue samples: (<b>a</b>) control sample with no HBD-4-positive structures in skin and mucosal epithelium, adipose glands, hair follicles, and connective tissue, rare in blood vessels (arrows), 200×; (<b>b</b>) patient sample with rare HBD-4-positive structures in blood vessels and connective tissue, moderate in skin epithelium, moderate to numerous in adipose glands and hair follicles, numerous in mucosal epithelium and sweat gland ducts (arrows), 200×.</p>
Full article ">Figure 10
<p>Immunohistochemistry of the IL-10-positive structures in the control and patient tissue samples: (<b>a</b>) control sample with rare occurrence of IL-10-positive structures in adipose glands, few in mucosal epithelium and blood vessels, moderate in skin epithelium, hair follicles, and connective tissue (arrows), 200×; (<b>b</b>) patient sample with few IL-10-positive structures in skin epithelium, few to moderate in connective tissue, moderate in blood vessels and hair follicles, numerous in mucosal epithelium and sweat gland ducts (arrows), 200×.</p>
Full article ">Figure 11
<p>Comparison of immunity defense factor median distribution in patient and control group tissues.</p>
Full article ">Figure 12
<p>Heat-map of correlations between the factors (part one).</p>
Full article ">Figure 13
<p>Heat-map of correlations between the factors (part two).</p>
Full article ">
Back to TopTop