[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
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 (8,140)

Search Parameters:
Keywords = endothelial cell

Order results
Result details
Results per page
Select all
Export citation of selected articles as:
23 pages, 17221 KiB  
Article
Aged Gut Microbiome Induces Metabolic Impairment and Hallmarks of Vascular and Intestinal Aging in Young Mice
by Chak-Kwong Cheng, Lianwei Ye, Yuanyuan Zuo, Yaling Wang, Li Wang, Fuyong Li, Sheng Chen and Yu Huang
Antioxidants 2024, 13(10), 1250; https://doi.org/10.3390/antiox13101250 - 17 Oct 2024
Viewed by 112
Abstract
Aging, an independent risk factor for cardiometabolic diseases, refers to a progressive deterioration in physiological function, characterized by 12 established hallmarks. Vascular aging is driven by endothelial dysfunction, telomere dysfunction, oxidative stress, and vascular inflammation. This study investigated whether aged gut microbiome promotes [...] Read more.
Aging, an independent risk factor for cardiometabolic diseases, refers to a progressive deterioration in physiological function, characterized by 12 established hallmarks. Vascular aging is driven by endothelial dysfunction, telomere dysfunction, oxidative stress, and vascular inflammation. This study investigated whether aged gut microbiome promotes vascular aging and metabolic impairment. Fecal microbiome transfer (FMT) was conducted from aged (>75 weeks old) to young C57BL/6 mice (8 weeks old) for 6 weeks. Wire myography was used to evaluate endothelial function in aortas and mesenteric arteries. ROS levels were measured by dihydroethidium (DHE) staining and lucigenin-enhanced chemiluminescence. Vascular and intestinal telomere function, in terms of relative telomere length, telomerase reverse transcriptase expression and telomerase activity, were measured. Systemic inflammation, endotoxemia and intestinal integrity of mice were assessed. Gut microbiome profiles were studied by 16S rRNA sequencing. Some middle-aged mice (40–42 weeks old) were subjected to chronic metformin treatment and exercise training for 4 weeks to evaluate their anti-aging benefits. Six-week FMT impaired glucose homeostasis and caused vascular dysfunction in aortas and mesenteric arteries in young mice. FMT triggered vascular inflammation and oxidative stress, along with declined telomerase activity and shorter telomere length in aortas. Additionally, FMT impaired intestinal integrity, and triggered AMPK inactivation and telomere dysfunction in intestines, potentially attributed to the altered gut microbial profiles. Metformin treatment and moderate exercise improved integrity, AMPK activation and telomere function in mouse intestines. Our data highlight aged microbiome as a mechanism that accelerates intestinal and vascular aging, suggesting the gut-vascular connection as a potential intervention target against cardiovascular aging and complications. Full article
Show Figures

Figure 1

Figure 1
<p>Effects of aged-to-young FMT on body parameters. (<b>A</b>) Schematic overview on FMT protocol from aged and young donor mice to young recipient mice. (<b>B</b>) Body weights of aged donor mice (Aged), young-transplanted (Young (Control)) and aged-transplanted young mice (Young (FMT)) after 6-week FMT protocol. (<b>C</b>) Body weight changes and (<b>D</b>) percentage changes in body weights of mice in (<b>B</b>) during the 6-week FMT. (<b>E</b>) Weights of indicated organs of mice in (<b>B</b>) postmortem after the 6-week FMT. (<b>F</b>) Weights of inguinal subcutaneous adipose tissue (ingSAT), perigonadal visceral adipose tissue (pgVAT) and brown adipose tissue (BAT) of mice in (<b>B</b>). (<b>G</b>) Gross appearance of adipose tissues of mice in (<b>B</b>). (<b>H</b>) Glucose tolerance test (GTT) on mice in (<b>B</b>) at week 6 of FMT, and (<b>I</b>) corresponding area under curve (AUC) analysis of glucose over time. (<b>J</b>) Insulin tolerance test (ITT) of mice in (<b>B</b>) at week 6 of FMT, and (<b>K</b>) corresponding AUC analysis of glucose over time. <span class="html-italic">N</span> = 10 per group. Data are mean ± SD. * <span class="html-italic">p</span> &lt; 0.05; Brown-Forsythe and Welch ANOVA and Dunnett T3 test.</p>
Full article ">Figure 2
<p>Effects of aged-to-young FMT on endothelial function. Representative traces for endothelium-dependent relaxations (EDRs) in (<b>A</b>) aortas and (<b>B</b>) mesenteric arteries of Aged, young-transplanted (Young (Control)) and aged-transplanted mice (Young (FMT)). Summary statistics of wire myography on EDRs in (<b>C</b>) aortas and (<b>D</b>) mesenteric arteries from different mouse groups. (<b>E</b>) Dihydroethidium (DHE) staining on en face endothelium of different mouse groups, and (<b>F</b>) corresponding quantification of DHE fluorescence. (<b>G</b>) Lucigenin-enhanced chemiluminescence on aortic ROS levels of different mouse groups. (<b>H</b>) Nitrite levels in aortas of different mouse groups. <span class="html-italic">N</span> = 8 per group. (<b>I</b>) Representative Western blots, and (<b>J</b>,<b>K</b>) quantification of Western blotting on expression of AMPK, p-AMPK at Thr172, eNOS and p-eNOS at Ser1177 in aortas of different mouse groups. <span class="html-italic">N</span> = 6 per group. Data are mean ± SD. * <span class="html-italic">p</span> &lt; 0.05; Brown-Forsythe and Welch ANOVA and Dunnett T3 test.</p>
Full article ">Figure 3
<p>Effects of aged-to-young FMT on vascular and systemic inflammation, and vascular telomere function. (<b>A</b>) RT-PCR on mRNA levels of pro-inflammatory genes in aortas of Aged, young-transplanted (Young (Control)) and aged-transplanted mice (Young (FMT)). (<b>B</b>) ELISA on circulating inflammatory markers of different mouse groups. (<b>C</b>) ELISA on circulating GLP-1 levels of different mouse groups. (<b>D</b>) Tert mRNA level in aortas of different mouse groups. (<b>E</b>) Telomerase activities in aortas of different mouse groups. (<b>F</b>) Relative telomere length in aortas of different mouse groups. <span class="html-italic">N</span> = 8 per group. Data are mean ± SD. * <span class="html-italic">p</span> &lt; 0.05; Brown-Forsythe and Welch ANOVA and Dunnett T3 test.</p>
Full article ">Figure 4
<p>Effects of aged-to-young FMT on intestinal inflammation, telomere function and barrier function. (<b>A</b>) RT-PCR on mRNA levels of pro-inflammatory genes in intestines of Aged, young-transplanted (Young (Control)) and aged-transplanted mice (Young (FMT)). (<b>B</b>) Lucigenin-enhanced chemiluminescence on intestinal ROS levels of different mouse groups. (<b>C</b>) Tert mRNA level in intestines of different mouse groups. (<b>D</b>) Telomerase activities in intestines of different mouse groups. (<b>E</b>) Relative telomere length in intestines of different mouse groups. Endotoxin levels in (<b>F</b>) feces and (<b>G</b>) sera of different mouse groups. ELISA on serum levels of (<b>H</b>) LBP and (<b>I</b>) I-FABP of different mouse groups. (<b>J</b>) Proglucagon mRNA level in intestines of different mouse groups. <span class="html-italic">N</span> = 8 per group. (<b>K</b>) Representative Western blots and quantification of Western blotting on expression of AMPK and p-AMPK at Thr172 in intestines of different mouse groups. <span class="html-italic">N</span> = 6 per group. Data are mean ± SD. * <span class="html-italic">p</span> &lt; 0.05; Brown-Forsythe and Welch ANOVA and Dunnett T3 test.</p>
Full article ">Figure 5
<p>Effects of aged-to-young FMT on gut microbial profiles in young host mice. (<b>A</b>) Principal component analysis (PCA) plot revealing distinct clusters for fecal microbiome samples obtained from young (depicted in blue) and aged (in brown) mice before antibiotic treatment and FMT, highlighting the species contributing to this clustering. <span class="html-italic">N</span> = 8 per group. (<b>B</b>) PCA plot showing the clustering of fecal microbiome samples from young-transplanted (Young (Control); depicted in blue) and aged-transplanted mice (Young (FMT); in red). <span class="html-italic">N</span> = 8 per group. (<b>C</b>) Non-Metric Multi-Dimensional Scaling (NMDS) plot displaying the clustering of microbiome across various mouse groups. <span class="html-italic">N</span> = 6–8 per group. (<b>D</b>) Differential abundance analysis on the mean difference in centered log ratio for enriched species in young, aged, Young (Control) and Young (FMT) mice. <span class="html-italic">N</span> = 8 per group.</p>
Full article ">Figure 6
<p>Effects of chronic metformin treatment and moderate exercise training on intestinal homeostasis. (<b>A</b>) Schematic diagram on chronic metformin treatment and moderate exercise training with the presence and absence of compound C (CC) treatment in middle-aged C57BL/6 mice. Representative Western blots and quantification of Western blotting on expression of AMPK and p-AMPK at Thr172 in intestines of (<b>B</b>) metformin-treated mice, and (<b>C</b>) exercise-trained mice. Lucigenin-enhanced chemiluminescence on intestinal ROS levels of (<b>D</b>) metformin-treated mice, and (<b>E</b>) exercise-trained mice. RT-PCR on mRNA levels of pro-inflammatory genes in intestines of (<b>F</b>) metformin-treated mice, and (<b>G</b>) exercise-trained mice. (<b>H</b>) Tert mRNA level in intestines of different mouse groups. (<b>I</b>) Telomerase activities in intestines of different mouse groups. (<b>J</b>) Relative telomere length in intestines of different mouse groups. ELISA on serum levels of (<b>K</b>) LBP and (<b>L</b>) I-FABP of different mouse groups. <span class="html-italic">N</span> = 6 per group. Data are mean ± SD. * <span class="html-italic">p</span> &lt; 0.05; Brown-Forsythe and Welch ANOVA and Dunnett T3 test.</p>
Full article ">Figure 7
<p>Schematic overview of the study. Aged microbiome induces metabolic impairments and vascular dysfunction in young mice. Aged microbiome causes telomere dysfunction, oxidative stress, and inflammation in intestines and vasculature of young mice. Metformin and moderate exercise potentially retard hallmarks of intestinal aging through AMPK activation. The study highlights the network among multiple aging hallmarks, including dysbiosis, deregulated nutrient sensing, chronic inflammation and telomere attrition, in terms of gut-vascular connection.</p>
Full article ">
14 pages, 8177 KiB  
Article
The Protective Role of Intermedin in Contrast-Induced Acute Kidney Injury: Enhancing Peritubular Capillary Endothelial Cell Adhesion and Integrity Through the cAMP/Rac1 Pathway
by Tingting Gao, Ruiyuan Gu, Heng Wang, Lizheng Li, Bojin Zhang, Jie Hu, Qinqin Tian, Runze Chang, Ruijing Zhang, Guoping Zheng and Honglin Dong
Int. J. Mol. Sci. 2024, 25(20), 11110; https://doi.org/10.3390/ijms252011110 - 16 Oct 2024
Viewed by 239
Abstract
Contrast-induced acute kidney injury (CIAKI) is a common complication with limited treatments. Intermedin (IMD), a peptide belonging to the calcitonin gene-related peptide family, promotes vasodilation and endothelial stability, but its role in mitigating CIAKI remains unexplored. This study investigates the protective effects of [...] Read more.
Contrast-induced acute kidney injury (CIAKI) is a common complication with limited treatments. Intermedin (IMD), a peptide belonging to the calcitonin gene-related peptide family, promotes vasodilation and endothelial stability, but its role in mitigating CIAKI remains unexplored. This study investigates the protective effects of IMD in CIAKI, focusing on its mechanisms, particularly the cAMP/Rac1 signaling pathway. Human umbilical vein endothelial cells (HUVECs) were treated with iohexol to simulate kidney injury in vitro. The protective effects of IMD were assessed using CCK8 assay, flow cytometry, ELISA, and Western blotting. A CIAKI rat model was utilized to evaluate renal peritubular capillary endothelial cell injury and renal function through histopathology, immunohistochemistry, immunofluorescence, Western blotting, and transmission electron microscopy. In vitro, IMD significantly enhanced HUVEC viability and mitigated iohexol-induced toxicity by preserving intercellular adhesion junctions and activating the cAMP/Rac1 pathway, with Rac1 inhibition attenuating these protective effects. In vivo, CIAKI caused severe damage to peritubular capillary endothelial cell junctions, impairing renal function. IMD treatment markedly improved renal function, an effect negated by Rac1 inhibition. IMD protects against renal injury in CIAKI by activating the cAMP/Rac1 pathway, preserving peritubular capillary endothelial integrity and alleviating acute renal injury from contrast media. These findings suggest that IMD has therapeutic potential in CIAKI and highlight the cAMP/Rac1 pathway as a promising target for preventing contrast-induced acute kidney injury in at-risk patients, ultimately improving clinical outcomes. Full article
Show Figures

Figure 1

Figure 1
<p>IMD can antagonize damage to HUVEC viability and apoptosis induced by iohexol. (<b>A</b>) HUVECs were preincubated with IMD (0, 1, 10, 100 nmol/L) for 30 min and then treated with iohexol (10, 20, 40, 80 mgI/mL) for 12 h. A CCK-8 kit was used to test cell viability. (<b>B</b>) HUVECs were preincubated with IMD (10 nmol/L) for 30 min, then were treated with iohexol (40 mgI/mL) or iohexol (40 mgI/mL) +NSC23766(50 μM) for 12 h. Apoptosis was detected using flow cytometry. (<b>C</b>) The apoptosis rate in each group was quantified (<span class="html-italic">n</span> = 6). The results were analyzed using ANOVA, followed by Tukey’s multiple comparisons test for subgroup analysis. All data are expressed as mean ± SD, * <span class="html-italic">p</span> &lt; 0.05 compared to the control group, # <span class="html-italic">p</span> &lt; 0.05. For comparison between groups, ns: not significant.</p>
Full article ">Figure 2
<p>IMD can protect the adherens junction of HUVECs by activating the cAMP/Rac1 pathway. (<b>A</b>) Intermedin induces cAMP production in HUVECs. The concentration of cAMP was measured using ELISA, <span class="html-italic">n</span> = 4. (<b>B</b>,<b>C</b>) HUVECs were treated with 10 nmol of IMD and/or 40 mgI/mL iohexol or 50 μM NSC23766 for 12 h. Whole-cell lysates of HUVECs were collected for immunoblotting analysis of Rac1, VE-cadherin, and GAPDH. The results were analyzed using ANOVA, with Tukey’s multiple comparisons test applied for subgroup analysis. All data are expressed as mean ± SD, * <span class="html-italic">p</span> &lt; 0.05 compared to the control group, <sup>#</sup> <span class="html-italic">p</span> &lt; 0.05. For comparison between groups, ns: not significant.</p>
Full article ">Figure 3
<p>IMD attenuates renal injury in rat CIAKI models, and inhibition of Rac1 might abolish this protective effect. (<b>A</b>) Male SD rats were randomly divided into five groups (<span class="html-italic">n</span> = 6), and the rats in each group were treated as shown in the scheme. (<b>B</b>) Renal morphology (HE, magnification ×400) showed vacuolar degeneration of renal tubules (shown by red arrows) and dilatation of renal tubules (shown by green arrows). (<b>C</b>) Magnification of renal PAS staining (×400) showed the absence of brush edges of renal tubules, vacuolar degeneration, and dilatation of renal tubules. (<b>D</b>) Renal tubular injury score. (<b>E</b>) Serum creatinine levels for the indicated treatments. The results were analyzed using ANOVA, with Tukey’s multiple comparisons test applied for subgroup analysis. All data are expressed as mean ± SD, * <span class="html-italic">p</span> &lt; 0.05 compared to the control group, # <span class="html-italic">p</span> &lt; 0.05. For comparison between groups, ns: not significant.</p>
Full article ">Figure 4
<p>IMD activates the cAMP/Rac1 pathway and alleviates renal peritubular capillary injury in CIAKI rats. (<b>A</b>,<b>B</b>): The renal lysates of rats were collected and the expression of Rac1, VE-cadherin, VEGFR2, and GAPDH were detected using Western blotting. (<b>C</b>) Kidney sections were stained with CD34 immunohistochemical staining (magnification, ×400, scale 30 μm). (<b>D</b>) Kidney sections were stained with CD34 and ICAM1 immunofluorescence staining (magnification, ×200, scale 50 μm). (<b>E</b>) The average optical density of CD34 positive staining in each group was quantified (<span class="html-italic">n</span> = 6). (<b>E</b>) Quantification of the average optical density of ICAM1 immunofluorescence (<span class="html-italic">n</span> = 6). The data were analyzed using ANOVA, followed by Tukey’s multiple comparisons test for subgroup analysis. All data are expressed as mean ± SD, * <span class="html-italic">p</span> &lt; 0.05 compared to the control group, # <span class="html-italic">p</span> &lt; 0.05. For comparison between groups, ns: not significant.</p>
Full article ">Figure 5
<p>IMD protects the PTC endothelial barrier: a representative microphotograph of the ultrastructural changes of peritubular capillary endothelial cells, cell membrane discontinuity (shown by green arrows), and basement membrane fracture (shown by blue arrows) (original magnification, ×12,000; scale, 1 μm).</p>
Full article ">
15 pages, 1641 KiB  
Article
Expression of Myeloperoxidase in Patient-Derived Endothelial Colony-Forming Cells—Associations with Coronary Artery Disease and Mitochondrial Function
by Weiqian Eugene Lee, Elijah Genetzakis, Giannie Barsha, Joshua Vescovi, Carmen Mifsud, Stephen T. Vernon, Tung Viet Nguyen, Michael P. Gray, Stuart M. Grieve and Gemma A. Figtree
Biomolecules 2024, 14(10), 1308; https://doi.org/10.3390/biom14101308 (registering DOI) - 16 Oct 2024
Viewed by 294
Abstract
Background and Aims: Myeloperoxidase (MPO) plays a critical role in the innate immune response and has been suggested to be a surrogate marker of oxidative stress and inflammation, with elevated levels implicated in cardiovascular diseases, such as atherosclerosis and heart failure, as well [...] Read more.
Background and Aims: Myeloperoxidase (MPO) plays a critical role in the innate immune response and has been suggested to be a surrogate marker of oxidative stress and inflammation, with elevated levels implicated in cardiovascular diseases, such as atherosclerosis and heart failure, as well as in conditions like rheumatoid arthritis and cancer. While MPO is well-known in leukocytes, its expression and function in human endothelial cells remain unclear. This study investigates MPO expression in patient-derived endothelial colony-forming cells (ECFCs) and its potential association with CAD and mitochondrial function. Methods: ECFCs were cultured from the peripheral blood of 93 BioHEART-CT patients. MPO expression and associated functions were examined using qRT-PCR, immunochemistry, flow cytometry, and MPO activity assays. CAD presence was defined using CT coronary angiography (CACS > 0). Results: We report MPO presence in patient-derived ECFCs for the first time. MPO protein expression occurred in 70.7% of samples (n = 41) which had nuclear co-localisation, an atypical observation given its conventional localisation in the granules of neutrophils and monocytes. This suggests potential alternative roles for MPO in nuclear processes. MPO mRNA expression was detected in 66.23% of samples (n = 77). CAD patients had a lower proportion of MPO-positive ECFCs compared to non-CAD controls (57.45% vs. 80%, p = 0.04), a difference that persisted in the statin-naïve sub-cohort (53.85% vs. 84.62%, p = 0.02). Non-CAD patients with MPO expression showed upregulated mitochondrial-antioxidant genes (AIFM2, TXNRD1, CAT, PRDX3, PRDX6). In contrast, CAD patients with MPO gene expression had heightened mROS production and mitochondrial mass and decreased mitochondrial function compared to that of CAD patients without MPO gene expression. Conclusions: MPO is present in the nucleus of ECFCs. In non-CAD ECFCs, MPO expression is linked to upregulated mitochondrial-antioxidant genes, whereas in CAD ECFCs, it is associated with greater mitochondrial dysfunction. Full article
Show Figures

Figure 1

Figure 1
<p>MPO protein expression in patient-derived ECFCs. (<b>A</b>) Representative Western blot of MPO protein expression in corresponding patient-derived ECFCs using anti-MPO antibody and anti-β-actin for loading control. (<b>B</b>) MPO protein is co-localised to the nuclei of patient-derived ECFCs. Representative immunocytochemistry images of MPO protein expression co-localised to the nucleus of patient-derived ECFCs, showing nuclei (blue) and MPO granules (red) within the cell (20× magnification). The scale bar represents 100 μM. (<b>C</b>) Representative Western blot of subcellular expression of MPO in patient-derived ECFCs at the soluble nuclear and chromatin-bound nuclear subfractions. At least three biological replicates were used. Original images can be found in <a href="#app1-biomolecules-14-01308" class="html-app">Supplementary Materials</a> file.</p>
Full article ">Figure 2
<p>CAD patients were less likely to express MPO gene, as identified by qRT-PCR. Stacked bar plots showing the association between proportion in MPO gene expression and the presence of CAD in (<b>A</b>) all patients (<span class="html-italic">n</span> = 77; No CAD = 30, CAD = 47), (<b>B</b>) male patients (<span class="html-italic">n</span> = 40; No CAD = 14, CAD = 26) and (<b>C</b>) female patients (<span class="html-italic">n</span> = 36; No CAD = 16, CAD = 20). (<b>D</b>) Statin-naïve patients (<span class="html-italic">n</span> = 54; No CAD = 27, CAD = 27), (<b>E</b>) statin-naïve male patients (<span class="html-italic">n</span> = 27; No CAD = 12, CAD = 15) and (<b>F</b>) statin-naïve female patients (<span class="html-italic">n</span> = 26; No CAD = 14, CAD = 12). Statistical association was analysed using Pearson’s chi-square test (categorial variables). Categorical measurements are shown as percentages. * <span class="html-italic">p</span> &lt; 0.05, ** <span class="html-italic">p</span> &lt; 0.01.</p>
Full article ">Figure 2 Cont.
<p>CAD patients were less likely to express MPO gene, as identified by qRT-PCR. Stacked bar plots showing the association between proportion in MPO gene expression and the presence of CAD in (<b>A</b>) all patients (<span class="html-italic">n</span> = 77; No CAD = 30, CAD = 47), (<b>B</b>) male patients (<span class="html-italic">n</span> = 40; No CAD = 14, CAD = 26) and (<b>C</b>) female patients (<span class="html-italic">n</span> = 36; No CAD = 16, CAD = 20). (<b>D</b>) Statin-naïve patients (<span class="html-italic">n</span> = 54; No CAD = 27, CAD = 27), (<b>E</b>) statin-naïve male patients (<span class="html-italic">n</span> = 27; No CAD = 12, CAD = 15) and (<b>F</b>) statin-naïve female patients (<span class="html-italic">n</span> = 26; No CAD = 14, CAD = 12). Statistical association was analysed using Pearson’s chi-square test (categorial variables). Categorical measurements are shown as percentages. * <span class="html-italic">p</span> &lt; 0.05, ** <span class="html-italic">p</span> &lt; 0.01.</p>
Full article ">Figure 3
<p>MPO gene expression is associated with dysregulated mitochondrial function and dynamics. (<b>A</b>–<b>C</b>) CAD ECFCs with MPO gene expression at baseline had increased (<b>A</b>) mROS production and (<b>B</b>) mitochondrial mass and decreased (<b>C</b>) mitochondrial function. N = 8; no MPO gene expression: N = 3, MPO gene expression: N = 5. Data are represented as mean ± S.E.M. Welch’s <span class="html-italic">t</span>-test was performed.</p>
Full article ">Figure 4
<p>Antioxidant genes were upregulated with MPO gene expression in non-CAD patients. Relative fold gene expression was evaluated in ECFCs without CAD at baseline in (<b>A</b>) <span class="html-italic">AIFM2</span>, (<b>B</b>) <span class="html-italic">TXNRD1</span>, (<b>C</b>) <span class="html-italic">CAT</span>, (<b>D</b>) <span class="html-italic">PRDX3</span> and (<b>E</b>) <span class="html-italic">PRDX6</span>. Each sample was performed in triplicate. N = 30; no MPO gene expression: N = 6, MPO gene expression: N = 24. Data are represented as mean ± S.E.M. Student’s <span class="html-italic">t</span>-test was performed.</p>
Full article ">
13 pages, 2764 KiB  
Article
Cadmium Induces Vascular Endothelial Cell Detachment by Downregulating Claudin-5 and ZO-1 Levels
by Takato Hara, Mayuka Asatsu, Tatsuya Yamagishi, Chinami Ohata, Hitomi Funatsu, Yuzuki Takahashi, Misaki Shirai, Chiaki Nakata, Haruka Katayama, Toshiyuki Kaji, Tomoya Fujie and Chika Yamamoto
Int. J. Mol. Sci. 2024, 25(20), 11035; https://doi.org/10.3390/ijms252011035 - 14 Oct 2024
Viewed by 313
Abstract
Cadmium is a contributing factor to cardiovascular diseases and highly toxic to vascular endothelial cells. It has a distinct mode of injury, causing the de-endothelialization of regions in the monolayer structure of endothelial cells in a concentration-dependent manner. However, the specific molecules involved [...] Read more.
Cadmium is a contributing factor to cardiovascular diseases and highly toxic to vascular endothelial cells. It has a distinct mode of injury, causing the de-endothelialization of regions in the monolayer structure of endothelial cells in a concentration-dependent manner. However, the specific molecules involved in the cadmium toxicity of endothelial cells remain unclear. The purpose of this study was to identify the specific molecular mechanisms through which cadmium affects endothelial detachment. Cadmium inhibited the expression of claudin-5 and zonula occludens (ZO)-1, which are components of tight junctions (strongest contributors to intercellular adhesion), in a concentration- and time-dependent manner. Compared to arsenite, zinc, and manganese, only cadmium suppressed the expression of both claudin-5 and ZO-1 molecules. Moreover, the knockdown of claudin-5 and ZO-1 exacerbated cadmium-induced endothelial cell injury and expansion of the detachment area, whereas their overexpression reversed these effects. CRE-binding protein inhibition reduced cadmium toxicity, suggesting that CRE-binding protein activation is involved in the cadmium-induced inhibition of claudin-5 and ZO-1 expression and endothelial detachment. These findings provide new insights into the toxicological mechanisms of cadmium-induced endothelial injury and risk of cardiovascular disease. Full article
Show Figures

Figure 1

Figure 1
<p>Effects of metals on the morphology and mRNA expression levels of tight junction molecules in vascular endothelial cells. (<b>A</b>) Morphology of vascular endothelial cells. (<b>B</b>) mRNA expression levels of claudin-5, claudin-12, occludin, zonula occludens (ZO)-1, and ZO-2 in cells treated with 2 and 5 μM cadmium, arsenite, and manganase or 20 and 50 μM zinc for 24 h. Values are represented as the mean ± standard error (S.E.) of triplicates. * <span class="html-italic">p</span> &lt; 0.05 and ** <span class="html-italic">p</span> &lt; 0.01 vs. control by Dunnett’s test.</p>
Full article ">Figure 2
<p>Effects of metals on the protein expression levels of tight junction molecules in vascular endothelial cells. Protein expression levels of claudin-5, claudin-12, occludin, ZO-1 (both bands are ZO-1), ZO-2, and ZO-3 in vascular endothelial cells treated with 1, 2, 3, 4, and 5 μM cadmium, arsenite, and manganase or 10, 20, 30, 40, and 50 μM of zinc for 24 h.</p>
Full article ">Figure 3
<p>Cadmium time-dependently inhibits claudin-5 and ZO-1 expression in vascular endothelial cells. (<b>A</b>) mRNA and (<b>B</b>) protein expression levels of claudin-5 and ZO-1 in cells treated with 2 μM cadmium for 4, 8, 12, and 24 h. Values are represented as the mean ± S.E. of triplicates. ** <span class="html-italic">p</span> &lt; 0.01 vs. corresponding control by Student’s <span class="html-italic">t</span>-test.</p>
Full article ">Figure 4
<p>Cadmium toxicity is exacerbated in vascular endothelial cells that suppress the expression of claudin-5 and ZO-1. (<b>A</b>) Cell morphology. (<b>B</b>) Claudin-5 and ZO-1 protein expression levels. (<b>C</b>) Lactate dehydrogenase (LDH) leakage from cells. Vascular endothelial cells were treated with 2 μM cadmium for 24 h after transfection with a control, ZO-1, or claudin-5 small interfering RNA (siRNA). Values are represented as the mean ± S.E. of four samples. * <span class="html-italic">p</span> &lt; 0.05 and ** <span class="html-italic">p</span> &lt; 0.01 vs. without cadmium; <sup>#</sup> <span class="html-italic">p</span> &lt; 0.05 and <sup>##</sup> <span class="html-italic">p</span> &lt; 0.01 vs. corresponding control siRNA by Tukey’s test.</p>
Full article ">Figure 5
<p>Cadmium toxicity is reduced in vascular endothelial cells overexpressing claudin-5 and ZO-1. (<b>A</b>) Cell morphology. (<b>B</b>) Claudin-5 and ZO-1 protein expression levels. (<b>C</b>) LDH leakage from cells. Vascular endothelial cells were treated with 2 μM cadmium for 24 h after transfection with the control, ZO-1, claudin-5, or claudin-5 and ZO-1 vector. Values are represented as the mean ± S.E. of four samples. ** <span class="html-italic">p</span> &lt; 0.01 vs. without cadmium; <sup>##</sup> <span class="html-italic">p</span> &lt; 0.01 vs. corresponding control vector by Tukey’s test.</p>
Full article ">Figure 6
<p>Effect of CRE-binding protein (CREB) inhibitor on cadmium toxicity of vascular endothelial cells. (<b>A</b>) CREB activation by cadmium treatment. Vascular endothelial cells were treated with 2 and 5 μM cadmium for 8 h. (<b>B</b>) Morphology of vascular endothelial cells. (<b>C</b>) LDH leakage from cells. Vascular endothelial cells were pretreated with 10 μM CREB inhibitor 666-15 for 2 h and then treated with 2 and 5 μM cadmium for 24 h (left panels) or 48 h (right panels). Values are represented as the mean ± S.E. of four samples. ** <span class="html-italic">p</span> &lt; 0.01 vs. without cadmium; <sup>##</sup> <span class="html-italic">p</span> &lt; 0.01 vs. without inhibitor by Tukey’s test.</p>
Full article ">Figure 7
<p>Effect of CREB knockdown on cadmium toxicity of vascular endothelial cells. (<b>A</b>) Morphology of vascular endothelial cells. (<b>B</b>) LDH leakage from cells. (<b>C</b>) Cadmium accumulation in cells. Vascular endothelial cells were transfected with the control or CREB siRNA for 24 h and then treated with 2 and 5 μM cadmium for 24 h. Values are represented as the mean ± S.E. of four samples. ** <span class="html-italic">p</span> &lt; 0.01 vs. corresponding without cadmium; <sup>##</sup> <span class="html-italic">p</span> &lt; 0.01 vs. corresponding control siRNA by Tukey’s test. (<b>D</b>,<b>E</b>) Protein and mRNA expression levels of claudin-5, ZO-1, and CREB in vascular endothelial cells. Vascular endothelial cells were transfected with the control or CREB siRNA for 24 h and then treated with 2 μM cadmium for (<b>E</b>) 12 h or (<b>D</b>) 24 h. Values are represented as the mean ± S.E. of triplicates. * <span class="html-italic">p</span> &lt; 0.05 and ** <span class="html-italic">p</span> &lt; 0.01 vs. corresponding without cadmium; <sup>##</sup> <span class="html-italic">p</span> &lt; 0.01 vs. corresponding control siRNA by Tukey’s test.</p>
Full article ">
18 pages, 20751 KiB  
Article
Insights into the Various Cellular Antimicrobial Responses, Biocompatibility, Osteogenesis, Wound Healing, and Angiogenesis of Copper-Doped Nano-Hydroxyapatite Composite Calcium Phosphate Bone Cement In Vitro
by Ssu-Meng Huang, Wen-Cheng Chen, Shih-Ming Liu, Chia-Ling Ko, Jian-Chih Chen and Chi-Jen Shih
J. Compos. Sci. 2024, 8(10), 424; https://doi.org/10.3390/jcs8100424 - 14 Oct 2024
Viewed by 348
Abstract
Calcium phosphate bone cement (CPC) is a popular material for bone remodeling, and nanohydroxyapatite (nHA) represents a breakthrough that has a wide range of clinical applications. During the early stages of bone repair, antibacterial and angiogenesis effects are essential to remodel new bone [...] Read more.
Calcium phosphate bone cement (CPC) is a popular material for bone remodeling, and nanohydroxyapatite (nHA) represents a breakthrough that has a wide range of clinical applications. During the early stages of bone repair, antibacterial and angiogenesis effects are essential to remodel new bone tissues. In this study, an antibacterial effect was achieved by incorporating Cu2+-doped nano-hydroxyapatite (Cu–nHA) synthesized through hydrothermal methods into CPC, and the impact of various amounts of Cu–nHA addition on the antibacterial and mechanical properties of CPC hybridization was evaluated. Moreover, the effects of Cu–nHA/CPC composites on the proliferation and mineralization of mouse progenitor osteoblastic cells (D1 cells) were characterized; the cell migration and angiogenesis ability of vascular endothelial cells (HUVECs) were also studied. Results indicated that incorporating 5 wt.% and 10 wt.% Cu–nHA into CPC led to a practical short-term antibacterial effect on S. aureus but not on E. coli. These Cu–nHA/CPC slurries remained injectable, anti-disintegrative, and non-toxic. Furthermore, compared with pure CPC, these Cu–nHA/CPC slurries demonstrated positive effects on D1 cells, resulting in better proliferation and mineralization. In addition, these Cu–nHA/CPC slurries were more effective in promoting the migration and angiogenesis of HUVECs. These findings indicate that 10 wt.% Cu–nHA/CPC has great application potential in bone regeneration. Full article
(This article belongs to the Section Biocomposites)
Show Figures

Graphical abstract

Graphical abstract
Full article ">Figure 1
<p>FTIR spectra (<b>a</b>) and XRD patterns (<b>b</b>) of CPC-only and Cu–nHA nanoparticle composite CPC, added with 5 wt.% and 10 wt.% Cu–nHA nanoparticles, respectively, were obtained after 1 day of further reaction in Tris buffer solution.</p>
Full article ">Figure 2
<p>Antibacterial activities of CPC only, 5 wt.% Cu–nHA/CPC, and 10 wt.% Cu–nHA against (<b>a</b>) <span class="html-italic">S. aureus</span> and (<b>b</b>) <span class="html-italic">E. coli</span> cultured for 4 days. * Indicates that the groups were significantly different (<span class="html-italic">p</span> &lt; 0.05) based on one-way ANOVA (<span class="html-italic">n</span> = 3).</p>
Full article ">Figure 2 Cont.
<p>Antibacterial activities of CPC only, 5 wt.% Cu–nHA/CPC, and 10 wt.% Cu–nHA against (<b>a</b>) <span class="html-italic">S. aureus</span> and (<b>b</b>) <span class="html-italic">E. coli</span> cultured for 4 days. * Indicates that the groups were significantly different (<span class="html-italic">p</span> &lt; 0.05) based on one-way ANOVA (<span class="html-italic">n</span> = 3).</p>
Full article ">Figure 3
<p>Injectability and anti-dispersibility in ddH<sub>2</sub>O were observed in time, 1 h, and 24 h after the addition of 5 wt.% Cu–nHA/CPC and 10 wt.% Cu–nHA/CPC composites compared with CPC only.</p>
Full article ">Figure 4
<p>(<b>a</b>) Cu–nHA microimages and (<b>b</b>) the morphology and Cu element mapping (green dots) of fracture surfaces were compared among groups of 5 wt.% Cu–nHA/CPC, 10 wt.% Cu–nHA/CPC, and CPC only after soaking in Tris buffer solution for 1 day.</p>
Full article ">Figure 5
<p>Cytotoxicity of 5 wt.% Cu–nHA/CPC, 10 wt.% Cu–nHA/CPC, and CPC-only extracts was compared to evaluate their effect on L929 cells after 1 day and 3 days of culture. Quantitative (<b>a</b>) and qualitative (<b>b</b>) measurements were taken. The red line indicates that if the cell viability is less than 70% compared to the control, the substance extract is toxic to the cells.</p>
Full article ">Figure 6
<p>Optical images of cell migration after culturing HUVECs with nHA/CPC composite extract of nHA with or without Cu<sup>2+</sup> doping for 6, 12, and 24 h.</p>
Full article ">Figure 7
<p>After a 6 h culture of HUVECs with nHA/CPC composite extracts, angiogenesis was assessed through optical (<b>a</b>) and fluorescence staining (<b>b</b>), with or without Cu<sup>2+</sup> doping.</p>
Full article ">Figure 8
<p>Evaluations were conducted to compare the long-term proliferation ability (<b>a</b>), absorbance value at OD<sub>405</sub> (<b>b</b>), and ALP semi-quantitative analysis (<b>c</b>) of nHA nanoparticles with or without Cu<sup>2+</sup> doping composite to CPC in contact culture with D1 cells; * indicates one-way ANOVA of groups, with <span class="html-italic">p</span> &lt; 0.05 indicating significant difference (<span class="html-italic">n</span> = 6).</p>
Full article ">Figure 9
<p>Qualitative ALP staining was performed on the surfaces of nHA/CPC composite samples with nHA, with or without Cu<sup>2+</sup> doping, followed by contact culture with D1 cells for 1–14 days.</p>
Full article ">
12 pages, 947 KiB  
Article
Soluble Urokinase-Type Plasminogen Activator Receptor (suPAR), Growth Differentiation Factor-15 (GDF-15), and Soluble C5b-9 (sC5b-9) Levels Are Significantly Associated with Endothelial Injury Indices in CAR-T Cell Recipients
by Eleni Gavriilaki, Christos Demosthenous, Paschalis Evangelidis, Zoi Bousiou, Ioannis Batsis, Anna Vardi, Despina Mallouri, Eudoxia-Evaggelia Koravou, Nikolaos Spyridis, Alkistis Panteliadou, Georgios Karavalakis, Marianna Masmanidou, Tasoula Touloumenidou, Apostolia Papalexandri, Christos Poziopoulos, Evangelia Yannaki, Ioanna Sakellari, Marianna Politou and Ioannis Papassotiriou
Int. J. Mol. Sci. 2024, 25(20), 11028; https://doi.org/10.3390/ijms252011028 - 14 Oct 2024
Viewed by 542
Abstract
Endothelial injury indices, such as Endothelial Activation and Stress Index (EASIX), modified EASIX (m-EASIX), and simplified EASIX (s-EASIX) scores, have been previously associated with chimeric antigen receptor-T (CAR-T) cell immunotherapy complications. Soluble urokinase-type plasminogen activator receptor (suPAR), growth differentiation factor-15 (GDF-15), and soluble [...] Read more.
Endothelial injury indices, such as Endothelial Activation and Stress Index (EASIX), modified EASIX (m-EASIX), and simplified EASIX (s-EASIX) scores, have been previously associated with chimeric antigen receptor-T (CAR-T) cell immunotherapy complications. Soluble urokinase-type plasminogen activator receptor (suPAR), growth differentiation factor-15 (GDF-15), and soluble C5b-9 (sC5b-9) have been described as markers of endothelial injury post-hematopoietic stem cell transplantation. In the current study, we examined whether suPAR, GDF-15, and sC5b-9 levels were associated with endothelial injury indices in adult CAR-T cell recipients. The levels of these markers were measured in patients before CAR-T cell infusion and in healthy individuals with immunoenzymatic methods. We studied 45 CAR-T cell recipients and 20 healthy individuals as the control group. SuPAR, GDF-15, and sC5b-9 levels were significantly higher in the patients’ group compared to the healthy control group (p < 0.001, in all comparisons). SuPAR levels at baseline were associated with the m-EASIX scores calculated at the same time point (p = 0.020), while suPAR and GDF-15 concentrations were correlated with EASIX scores at day 14 post-infusion (p < 0.001 in both comparisons). Moreover, sC5b-9 levels were correlated with the s-EASIX scores at infusion (p = 0.008) and the EASIX scores at day 14 (p = 0.005). In our study, sC5b9, suPAR, and GDF-15 levels were found to reflect endothelial injury in CAR-T cell recipients. Full article
(This article belongs to the Special Issue Novel Insights into Monoclonal Antibodies in Disease)
Show Figures

Figure 1

Figure 1
<p>(<b>A</b>) sC5b9 levels show a significant correlation with sEASIX0 (<span class="html-italic">p</span> = 0.008) and (<b>B</b>) EASIX14 (<span class="html-italic">p</span> = 0.005) scores. sC5b9 = soluble C5b-9 and s-EASIX0 = simplified Endothelial Activation and Stress Index scores were calculated at the day of CAR-T cell infusion; EASIXDAY14 = Endothelial Activation and Stress Index scores were calculated 14 days post-CAR-T cell infusion.</p>
Full article ">Figure 2
<p>(<b>A</b>) s-EASIX0 score over the median value (1.9, range: 0.74–49.3) was associated with poor OS (<span class="html-italic">p</span> = 0.027); (<b>B</b>) s-EASIX14 score over the median value (3, range: 0.6–74.6) was also associated with poor OS (<span class="html-italic">p</span> = 0.004). Green lines: s-EASIX score below the median; blue lines: s-EASIX score over the median value. S-EASIX0 = simplified Endothelial Activation and Stress Index calculated at the day of the infusion; S-EASIX14 = simplified Endothelial Activation and Stress Index calculated 14 days post-infusion; OS = overall survival.</p>
Full article ">Figure 3
<p>Methodology used in our study. Samples for sC5b-9, suPAR, and GDF-15 measurements were obtained from CAR-T cell recipients before their admission to the cellular therapy unit, and at the same time point, EASIX, mEASIX, and s-EASIX scores were calculated. EASIX and s-EASIX scores were also calculated on the day of CAR-T cell product infusion and 14 days post-infusion. During the post-infusion period, patients were closely monitored for therapy-related toxicities. The minimum follow-up period was 1 month. LDC = lymphodepleting chemotherapy; EASIX = Endothelial Activation and Stress Index; m-EASIX = modified Endothelial Activation and Stress Index; s-EASIX = simplified Endothelial Activation and Stress Index; sC5b-9 = soluble C5b-9; suPAR = soluble urokinase-type plasminogen activator receptor; GDF-15 = growth differentiation factor-15; CRS = cytokine release syndrome; ICANS = immune effector cell-associated neurotoxicity syndrome.</p>
Full article ">
43 pages, 4206 KiB  
Review
Advancements in Polymer Biomaterials as Scaffolds for Corneal Endothelium Tissue Engineering
by Kevin Y. Wu, Myriam Belaiche, Ying Wen, Mazen Y. Choulakian and Simon D. Tran
Polymers 2024, 16(20), 2882; https://doi.org/10.3390/polym16202882 (registering DOI) - 12 Oct 2024
Viewed by 629
Abstract
Corneal endothelial dysfunction is a leading cause of vision loss globally, frequently requiring corneal transplantation. However, the limited availability of donor tissues, particularly in developing countries, has spurred on the exploration of tissue engineering strategies, with a focus on polymer biomaterials as scaffolds [...] Read more.
Corneal endothelial dysfunction is a leading cause of vision loss globally, frequently requiring corneal transplantation. However, the limited availability of donor tissues, particularly in developing countries, has spurred on the exploration of tissue engineering strategies, with a focus on polymer biomaterials as scaffolds for corneal endotlhelium regeneration. This review provides a comprehensive overview of the advancements in polymer biomaterials, focusing on their role in supporting the growth, differentiation, and functional maintenance of human corneal endothelial cells (CECs). Key properties of scaffold materials, including optical clarity, biocompatibility, biodegradability, mechanical stability, permeability, and surface wettability, are discussed in detail. The review also explores the latest innovations in micro- and nano-topological morphologies, fabrication techniques such as electrospinning and 3D/4D bioprinting, and the integration of drug delivery systems into scaffolds. Despite significant progress, challenges remain in translating these technologies to clinical applications. Future directions for research are highlighted, including the need for improved biomaterial combinations, a deeper understanding of CEC biology, and the development of scalable manufacturing processes. This review aims to serve as a resource for researchers and clinician–scientists seeking to advance the field of corneal endothelium tissue engineering. Full article
(This article belongs to the Section Biobased and Biodegradable Polymers)
Show Figures

Figure 1

Figure 1
<p>Histology of the human cornea. Created with BioRender.com.</p>
Full article ">Figure 2
<p>Penetrating keratoplasty (PK), Descemet’s stripping automated keratoplasty (DSAEK), and Descemet’s membrane endothelial keratoplasty (DMEK). Created with BioRender.com.</p>
Full article ">Figure 3
<p>Corneal endothelium tissue engineering. Created with BioRender.com.</p>
Full article ">Figure 4
<p>Schematic of the ECM structure of Descemet’s membrane [<a href="#B25-polymers-16-02882" class="html-bibr">25</a>]. Reprinted with permission from ref. [<a href="#B25-polymers-16-02882" class="html-bibr">25</a>]. Copyright 2021 Elsevier.</p>
Full article ">Figure 5
<p>Topological features with different shapes and sizes that can be used for surface patterning [<a href="#B113-polymers-16-02882" class="html-bibr">113</a>]. Reprinted with permission from ref. [<a href="#B113-polymers-16-02882" class="html-bibr">113</a>]. Copyright 2016 Elsevier.</p>
Full article ">Figure 6
<p>Enhanced protein adsorption and cell adhesion on micropatterned hydrogels. Created with <a href="http://BioRender.com" target="_blank">BioRender.com</a>.</p>
Full article ">Figure 7
<p>Length scale bar illustrating featured resolutions of various biofabrication techniques in comparison with geometric sizes of representative cells and tissues [<a href="#B154-polymers-16-02882" class="html-bibr">154</a>]. Reprinted from ref. [<a href="#B154-polymers-16-02882" class="html-bibr">154</a>], licensed under CC BY 4.0 [<a href="https://creativecommons.org/licenses/by/4.0/" target="_blank">https://creativecommons.org/licenses/by/4.0/</a>]. Accessed on 3 August 2024. “ES” stands for electrospinning and “LEP” stands for low-voltage electrospinning patterning.</p>
Full article ">
13 pages, 16899 KiB  
Article
Prognostic Impact of H19/Cell Adhesion Molecules Circuitry on Prostate Cancer Biopsy
by Valeria Pecci, Francesco Pierconti, Angela Carlino, Francesco Pinto, Ugo Gradilone, Sara De Martino, Dante Rotili, Claudio Grassi, Alfredo Pontecorvi, Carlo Gaetano, Lidia Strigari, Antonella Farsetti and Simona Nanni
Biomedicines 2024, 12(10), 2322; https://doi.org/10.3390/biomedicines12102322 - 12 Oct 2024
Viewed by 372
Abstract
Introduction: Metastatic prostate cancer (PCa) presents a significant challenge in oncology due to its high mortality rate and the absence of effective biomarkers for predicting patient outcomes. Building on previous research that highlighted the critical role of the long noncoding RNA (lncRNA) H19 [...] Read more.
Introduction: Metastatic prostate cancer (PCa) presents a significant challenge in oncology due to its high mortality rate and the absence of effective biomarkers for predicting patient outcomes. Building on previous research that highlighted the critical role of the long noncoding RNA (lncRNA) H19 and cell adhesion molecules in promoting tumor progression under hypoxia and estrogen stimulation, this study aimed to assess the potential of these components as prognostic biomarkers for PCa at the biopsy stage. Methods: This research utilized immunohistochemistry and droplet digital PCR to analyze formalin-fixed paraffin-embedded (FFPE) biopsies, focusing on specific markers within the H19/cell adhesion molecules pathway. Results: A novel multivariate analysis led to a “BioScore”, a composite biomarker score to predict disease progression. This score is based on evaluating five key markers: the expression levels of Hypoxia-Inducible Factor 2 Alpha (HIF-2α), endothelial Nitric Oxide Synthase (eNOS), β4 integrin, E-cadherin transcript (CDH1), and lncRNA H19. The criteria for the “BioScore” involve identifying three out of these five markers, combining elevated levels of HIF-2α, eNOS, β4 integrin, and CDH1 with reduced H19 expression. Conclusions: This finding suggests the possibility of identifying, at the time of biopsy, PCa patients at higher risk of metastasis based on dysregulation in the H19/cell adhesion molecules circuitry. This study provides a valuable opportunity for early intervention in managing PCa, potentially contributing to personalized treatment strategies. Full article
(This article belongs to the Section Cancer Biology and Oncology)
Show Figures

Figure 1

Figure 1
<p>IHC on PCa biopsies and score in the normal and tumoral areas. (<b>A</b>) Representative IHC staining of PCa biopsies with a specific antibody to HIF2α and eNOS in normal (left, scale bar = 84 mm) and tumoral (right, scale bar = 210 mm) areas. The scale bar is shown as a black line. (<b>B</b>) The IHC score was evaluated as Intensity × Quantity (IxQ) in normal (NT) and tumoral (T) areas for eNOS and HIF-2α staining. <span class="html-italic">p</span> values are indicated.</p>
Full article ">Figure 2
<p>Analysis of PCa biopsies and identification of prognostic score based on combinatorial expression of biomarkers. (<b>A</b>) Representative IHC staining of PCa biopsies with antibodies to β4 integrin, HIF2α, and eNOS. Scale bar: 100 μm. (<b>B</b>) Representative gene expression analysis by ddPCR of CDH1, H19, and control gene GAPDH on PCa biopsies and negative controls. (<b>C</b>) ROC curve analysis of PFS using eNOS and β4 integrin expression on PCa biopsies. The dashed line represents the random assignment AUC and 95% confidence interval (CI) details in <a href="#biomedicines-12-02322-t002" class="html-table">Table 2</a>. (<b>D</b>) Left: ROC curve analysis of PFS using the novel “BioScore” determined on PCa biopsies (cut-off and AUC details in <a href="#biomedicines-12-02322-t002" class="html-table">Table 2</a>). Right: Kaplan–Meier curve of PFS for patients with low or high “BioScore” levels (dashed and solid line, respectively). The Hazard Ratio with 95% CI was 0.090 (0.031–0.532). <span class="html-italic">p</span> values are indicated.</p>
Full article ">
14 pages, 1995 KiB  
Article
Cardioprotection by Preconditioning with Intralipid Is Sustained in a Model of Endothelial Dysfunction for Isolated-Perfused Hearts
by Martin Stroethoff, Natalie Schneider, Lea Sung, Jan Wübbolt, André Heinen and Annika Raupach
Int. J. Mol. Sci. 2024, 25(20), 10975; https://doi.org/10.3390/ijms252010975 - 12 Oct 2024
Viewed by 383
Abstract
Endothelial dysfunction (ED) is closely associated with most cardiovascular diseases. Experimental models are needed to analyze the potential impact of ED on cardioprotection in constant pressure Langendorff systems (CPLS). One cardioprotective strategy against ischemia/reperfusion injury (I/RI) is conditioning with the lipid emulsion Intralipid [...] Read more.
Endothelial dysfunction (ED) is closely associated with most cardiovascular diseases. Experimental models are needed to analyze the potential impact of ED on cardioprotection in constant pressure Langendorff systems (CPLS). One cardioprotective strategy against ischemia/reperfusion injury (I/RI) is conditioning with the lipid emulsion Intralipid (IL). Whether ED modulates the cardioprotective effect of IL remains unknown. The aim of the study was to transfer a protocol using a constant flow Langendorff system for the induction of ED into a CPLS, without the loss of smooth muscle cell functionality, and to analyze the cardioprotective effect of IL against I/RI under ED. In isolated hearts of male Wistar rats, ED was induced by 10 min perfusion of a Krebs–Henseleit buffer containing 60 mM KCl (K+), and the vasodilatory response to the vasodilators histamine (endothelial-dependent) and sodium–nitroprusside (SNP, endothelial-independent) was measured. A CPLS was employed to determine cardioprotection of pre- or postconditioning with 1% IL against I/RI. The constant flow perfusion of K+ reduced endothelial response to histamine but not to SNP, indicating reduced vasodilatory functionality of endothelial cells but not smooth muscle cells. Preconditioning with IL reduced infarct size and improved cardiac function while postconditioning with IL had no effect. The induction of ED neither influenced infarct size nor affected the cardioprotective effect by preconditioning with IL. This protocol allows for studies of cardioprotective strategies under ED in CLPS. The protection by preconditioning with IL seems to be mediated independently of a functional endothelium. Full article
Show Figures

Figure 1

Figure 1
<p>Difference of endothelial response (ΔER = ER<sub>1</sub> − ER<sub>2</sub>) before (ER<sub>1</sub>) and after (ER<sub>2</sub>) constant flow perfusion (10 min) with Krebs–Henseleit buffer (KHB) alone (white fill) or KHB containing 60 mM KCl (K+, striped green) to 800 nmol histamine (his, orange borders) or 1 µM sodium nitroprusside (SNP, pink borders). As a positive control for ED induction, a bolus of 1 s with 1% triton (grey fill) was used. Data are mean ± SD, <span class="html-italic">n</span> = 4 (SNP, triton), <span class="html-italic">n</span> = 7 (histamine). One-way ANOVA followed by Šidák’s multiple comparison test. *: <span class="html-italic">p</span> &lt; 0.05, ns: not significant (<span class="html-italic">p</span> &gt; 0.05).</p>
Full article ">Figure 2
<p>Infarct sizes of hearts after ischemia/reperfusion with pre- or post-treatment for 10 min with 1% Intralipid (IL) or vehicle (Con). LV: left ventricle. Data are mean ± SD, <span class="html-italic">n</span> = 7. One-way ANOVA, Dunnett’s multiple comparison test, *: <span class="html-italic">p</span> &lt; 0.05, ns: not significant (<span class="html-italic">p</span> &gt; 0.05).</p>
Full article ">Figure 3
<p>Infarct sizes of hearts after ischemia/reperfusion with post-treatment for 20 min with Intralipid (IL) or vehicle (Con). LV: left ventricle. Data are mean ± SD, <span class="html-italic">n</span> = 6. <span class="html-italic">t</span>-test; ns: not significant.</p>
Full article ">Figure 4
<p>Preconditioning with IL under endothelial dysfunction (ED). Hearts were treated for 10 min before ischemia with 1% Intralipid (IL; blue border) or vehicle (Con; black border). ED was induced by 10 min constant flow perfusion of Krebs–Henseleit buffer (KHB) containing 60 mM KCl (K+; green stripes). The other groups received normal KHB under constant flow conditions (white filling). LV: left ventricle. Data are mean ± SD, <span class="html-italic">n</span> = 7. Two-way ANOVA; * <span class="html-italic">p</span> &lt; 0.05 for effect by conditioning; ns = not significant for effect by ED induction and interaction (<span class="html-italic">p</span> &gt; 0.05).</p>
Full article ">Figure 5
<p>Schematic drawing of the utilized Langendorff system. Krebs–Henseleit buffer (KHB) and KHB containing 1% Intralipid (IL) are perfused in a constant pressure mode (80 mmHg) in separate circuits. To induce endothelial dysfunction, KHB containing 60 mM KCl (K+) is perfused in a constant flow mode. To ensure equivalent control conditions for induction of ED, KHB is perfused in a constant flow mode. The individual modules were switched on or off as required by the respective questions during the experimental setups. Created with <a href="http://BioRender.com" target="_blank">BioRender.com</a>.</p>
Full article ">Figure 6
<p>Experimental timeline for characterization of endothelial dysfunction. Dashed blue lines mark measuring points for coronary perfusion pressure (CPP). BL1/2: baseline at time point 1/2, P: after perfusion of vasodilator, His: histamine, T: triton, SNP: sodium–nitroprusside, KHB flow: constant flow perfusion with Krebs–Henseleit buffer, K+ flow: constant flow perfusion with KHB containing 60 mM KCl.</p>
Full article ">Figure 7
<p>Experimental protocol: conditioning with Intralipid (IL) using a constant pressure Langendorff system with the ability to switch into a constant flow mode for induction of endothelial dysfunction. (<b>a</b>) Pre- or postconditioning with IL for 10 min. (<b>b</b>) Postconditioning with IL for 20 min. (<b>c</b>) Preconditioning with IL under endothelial dysfunction induced by constant flow perfusion of Krebs–Henseleit buffer (KHB) containing 60 mM KCl (K+ flow) and under control condition with constant flow perfusion of KHB alone (KHB flow).</p>
Full article ">
18 pages, 1566 KiB  
Review
The Role of Soluble CD163 (sCD163) in Human Physiology and Pathophysiology
by Andriana Plevriti, Margarita Lamprou, Eleni Mourkogianni, Nikolaos Skoulas, Maria Giannakopoulou, Md Sanaullah Sajib, Zhiyong Wang, George Mattheolabakis, Antonios Chatzigeorgiou, Antonia Marazioti and Constantinos M. Mikelis
Cells 2024, 13(20), 1679; https://doi.org/10.3390/cells13201679 - 11 Oct 2024
Viewed by 553
Abstract
Soluble CD163 (sCD163) is a circulating inflammatory mediator, indicative of acute and chronic, systemic and non-systemic inflammatory conditions. It is the cleavage outcome, consisting of almost the entire extracellular domain, of the CD163, a receptor expressed in monocytic lineages. Its expression is proportional [...] Read more.
Soluble CD163 (sCD163) is a circulating inflammatory mediator, indicative of acute and chronic, systemic and non-systemic inflammatory conditions. It is the cleavage outcome, consisting of almost the entire extracellular domain, of the CD163, a receptor expressed in monocytic lineages. Its expression is proportional to the abundance of CD163+ macrophages. Various mechanisms trigger the shedding of the CD163 receptor or the accumulation of CD163-expressing macrophages, inducing the sCD163 concentration in the circulation and bodily fluids. The activities of sCD163 range from hemoglobin (Hb) scavenging, macrophage marker, decoy receptor for cytokines, participation in immune defense mechanisms, and paracrine effects in various tissues, including the endothelium. It is an established marker of macrophage activation and thus participates in many diseases, including chronic inflammatory conditions, such as atherosclerosis, asthma, and rheumatoid arthritis; acute inflammatory conditions, such as sepsis, hepatitis, and malaria; insulin resistance; diabetes; and tumors. The sCD163 levels have been correlated with the severity, stage of the disease, and clinical outcome for many of these conditions. This review article summarizes the expression and role of sCD163 and its precursor protein, CD163, outlines the sCD163 generation mechanisms, the biological activities, and the known underlying molecular mechanisms, with an emphasis on its impact on the endothelium and its contribution in the pathophysiology of human diseases. Full article
(This article belongs to the Special Issue Immune Cell Effect on the Endothelium)
Show Figures

Figure 1

Figure 1
<p>Schematic representation of the three CD163 transmembrane isoforms. They differ in the length of the intracellular domains, with the 49-amino-acid isoform having dominant expression.</p>
Full article ">Figure 2
<p>(<b>A</b>) Uniform manifold approximation and projection (UMAP) representation of all human cell types based on CD163 expression (data from 483,152 cells from the CZ database). (<b>B</b>) Clustering and analysis of the cell types with the highest CD163 expression.</p>
Full article ">Figure 3
<p>Schematic representation of the hemoglobin clearance mechanism by CD163 in inflammatory conditions. The Hb released by the ruptured red cells is bound to the 3rd SRCR domain of the CD163 extracellular domain, either by itself or as a Hp–Hb complex, which eventually leads to CD163 internalization.</p>
Full article ">Figure 4
<p>Schematic representation of sCD163 generation from the proteolytic cleavage of CD163. The list of the proteases known to cleave CD163 is shown on the left and the cleavage requires ATP consumption (black arrow).</p>
Full article ">
22 pages, 6432 KiB  
Article
Priming Mesenchymal Stem Cells with Lipopolysaccharide Boosts the Immunomodulatory and Regenerative Activity of Secreted Extracellular Vesicles
by Aina Areny-Balagueró, Marta Camprubí-Rimblas, Elena Campaña-Duel, Anna Solé-Porta, Adrián Ceccato, Anna Roig, John G. Laffey, Daniel Closa and Antonio Artigas
Pharmaceutics 2024, 16(10), 1316; https://doi.org/10.3390/pharmaceutics16101316 - 10 Oct 2024
Viewed by 432
Abstract
Background: Mesenchymal stem cells (MSCs)-derived extracellular vesicles (EVs) have been proposed as an alternative to live-cell administration for Acute Respiratory Distress Syndrome (ARDS). MSC-EVs can be chiefly influenced by the environment to which the MSCs are exposed. Here, lipopolysaccharide (LPS) priming of MSCs [...] Read more.
Background: Mesenchymal stem cells (MSCs)-derived extracellular vesicles (EVs) have been proposed as an alternative to live-cell administration for Acute Respiratory Distress Syndrome (ARDS). MSC-EVs can be chiefly influenced by the environment to which the MSCs are exposed. Here, lipopolysaccharide (LPS) priming of MSCs was used as a strategy to boost the natural therapeutic potential of the EVs in acute lung injury (ALI). Methods: The regenerative and immunemodulatory effect of LPS-primed MSC-EVs (LPS-EVs) and non-primed MSC-EVs (C-EVs) were evaluated in vitro on alveolar epithelial cells and macrophage-like THP-1 cells. In vivo, ALI was induced in adult male rats by the intrapulmonary instillation of HCl and LPS. Rats (n = 8 to 22/group) were randomized to receive a single bolus (1 × 108 particles) of LPS-EVs, C-EVs, or saline. Lung injury severity was assessed at 72 h in lung tissue and bronchoalveolar lavage. Results: In vitro, LPS-EVs improved wound regeneration and attenuated the inflammatory response triggered by the P. aeruginosa infection, enhancing the M2 macrophage phenotype. In in vivo studies, LPS-EVs, but not C-EVs, significantly decreased the neutrophilic infiltration and myeloperoxidase (MPO) activity in lung tissue. Alveolar macrophages from LPS-EVs-treated animals exhibited a reduced expression of CXCL-1, a key neutrophil chemoattractant. However, both C-EVs and LPS-EVs reduced alveolar epithelial and endothelial permeability, mitigating lung damage. Conclusions: EVs from LPS-primed MSCs resulted in a better resolution of ALI, achieving a greater balance in neutrophil infiltration and activation, while avoiding the complete disruption of the alveolar barrier. This opens new avenues, paving the way for the clinical implementation of cell-based therapies. Full article
Show Figures

Graphical abstract

Graphical abstract
Full article ">Figure 1
<p>Characterization of C-MSCs and LPS-primed MSCs isolated from rat bone marrow. (<b>A</b>) Immunophenotype of C-MSCs and LPS-MSCs, staining CD105 (FITC, green), CD90 (FITC, green) and CD44 (Texas Red, red) markers. Quantification of the percentage of positive cells for each marker (<span class="html-italic">n</span> = 7–11). The nuclei of the cells were stained with Hoechst (UV light, blue), 20× magnification. Scale bar: 100 µm. (<b>B</b>) Representative images of undifferentiated MSCs and MSCs differentiation towards adipogenic, osteogenic and chondrogenic lineages in vitro. Magnification: 10×. Scale bar: 500 µm. (<b>C</b>) Quantification of the EVs-like particles (<b>right</b>) (<span class="html-italic">n</span> = 5) and protein (<b>left</b>) (<span class="html-italic">n</span> = 7) concentration in C-MSCs and LPS-MSCs media. Data are presented as mean ± SEM (<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. Abbreviations: extracellular vesicles, EVs.</p>
Full article ">Figure 2
<p>Characterization of C- and LPS-MSCs-derived EVs. (<b>A</b>) Representative particle size distributions of C-EVs and LPS-EVs’ pools by Nanosight analysis. (<b>B</b>) Representative cryo-TEM images of C-EVs and LPS-EVs, 12kX magnification. Scale bar: 0.5 µm. (<b>C</b>) Detection of Alix, TSG101 and CD81 surface markers in C-EVs and LPS-EVs by Western Blot analysis. Detection of Calnexin only in cell lysate samples as negative control. Abbreviations: cell lysate, C. Lys.; extracellular vesicles, EVs.</p>
Full article ">Figure 3
<p>Effect of MSCs-derived EVs on wound healing and cell proliferation in vitro. (<b>A</b>) Representative optical images of wound healing in HPAEpiC monolayer. Magnification: 20×. Scale bar: 50 µm. Percentage of wound closure in HPAEpiC 24 h after being treated with C-EVs and LPS-EVs. (<b>B</b>) Percentage of cell viability of HPAEpiC 24 h after being treated with C-EVs and LPS-EVs, considering that non-treated cells (control) had 100% cellular viability. Data are presented as mean ± SEM of eight (<b>A</b>) and four (<b>B</b>) independent experiments with two replicates of each condition; * <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; **** <span class="html-italic">p</span> &lt; 0.0001. Abbreviations: extracellular vesicles, EVs.</p>
Full article ">Figure 4
<p>Effect of MSCs-derived EVs on infected THP-1 cells in vitro. Representation of mRNA expression of pro-inflammatory cytokines, chemoattractant mediators, and M1 and M2 macrophage phenotype markers: IL-1β (<b>A</b>), IL-6 (<b>B</b>), IL-8 (<b>C</b>), CD86 (<b>D</b>) and CD206 (<b>E</b>) in THP-1 cells activated with LPS and infected by <span class="html-italic">P. aeruginosa</span> and treated with C-EVs or LPS-EVs. (<b>F</b>) Ratio of CD86/CD206 expression (M1/M2 ratio). The relative expression of target genes was normalized to RPL37a expression. Data are presented as mean ± SEM of six independent experiments with two or three replicates of each condition. * <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; **** <span class="html-italic">p</span> &lt; 0.0001. Abbreviations: interleukin, IL; extracellular vesicles, EVs.</p>
Full article ">Figure 5
<p>Effect of MSCs-derived EVs on animals’ weight and lung permeability. (<b>A</b>) Monitoring of the animals’ body weight every 24 h, considering 100% as the starting body weight for each group (# <span class="html-italic">p</span> &lt; 0.05; ## <span class="html-italic">p</span> &lt; 0.01 control group vs. HCl + LPS group). (<b>B</b>) Ratio of lung weight/body weight measured at the end of the experiment (grams/grams) (<span class="html-italic">n</span> = 12–22). (<b>C</b>) Total protein concentration (µg/mL) and (<b>D</b>) cells’ concentration (cells/µL) in the BAL fluid at the end point (72 h) (<span class="html-italic">n</span> = 8–13); Data are presented as mean ± SEM * <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; **** <span class="html-italic">p</span> &lt; 0.0001. Abbreviations: extracellular vesicles, EVs.</p>
Full article ">Figure 6
<p>Effect of MSCs-derived EVs on neutrophil infiltration in the intra-alveolar space in vivo at 72 h. Percentage of (<b>A</b>) monocytes, (<b>B</b>) neutrophils and (<b>C</b>) lymphocytes in BAL by flow cytometry. (<b>D</b>) Representative images of BAL cells cytospins stained with Diff-Quick. Neutrophils (red arrows), macrophages (green arrows) and lymphocytes (yellow arrows) are indicated in each image. Magnification: 20×. (<b>E</b>) MPO activity quantification in lung tissue homogenates (mU/total tissue protein (g)). Data are presented as mean ± SEM (<span class="html-italic">n</span> = 8–14); * <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>Effect of MSCs-derived EVs on lung tissue inflammation in vivo at 72 h. mRNA expression of pro-inflammatory cytokines, (<b>A</b>) IL-1β and (<b>B</b>) IL-6, and chemoattractant mediators, (<b>C</b>) CCL-2 and (<b>D</b>) CXCL-1 (mRNA expression correlated vs. GAPDH). Data are presented as mean ± SEM (<span class="html-italic">n</span> = 9–14). * <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. Abbreviations: interleukin, IL.</p>
Full article ">Figure 8
<p>Effect of MSCs-derived EVs on alveolar macrophage inflammation in vivo at 72 h. mRNA expression of pro-inflammatory cytokines, (<b>A</b>) IL-1β; chemoattractant mediators, (<b>B</b>) CXCL-1; and M2 phenotype markers, (<b>C</b>) Arg-1 and (<b>D</b>) MR (mRNA expression correlated vs. GAPDH). Data are presented as mean ± SEM (<span class="html-italic">n</span> = 8–14). * <span class="html-italic">p</span> &lt; 0.05; ** <span class="html-italic">p</span> &lt; 0.01. Abbreviations: interleukin, IL; arginase 1, Arg-1; mannose receptor, MR.</p>
Full article ">Figure 9
<p>Effect of MSCs-derived EVs on the lung injury in vivo at 72 h. Representative images of H&amp;E histological lung sections of control and injured animals with (<b>A</b>) 2.5× and (<b>B</b>) 10× magnification. Scale bars: 200 and 100 µm, respectively. (<b>C</b>) Quantification of the lung injury score (LIS), evaluating hemorrhage, peribronchial infiltration, interstitial edema, pneumocyte hyperplasia and intra-alveolar infiltration, as described in <a href="#app1-pharmaceutics-16-01316" class="html-app">Supplementary Table S3</a>. Data are presented as mean ± SEM (<span class="html-italic">n</span> = 9–14). *** <span class="html-italic">p</span> &lt; 0.001.</p>
Full article ">
22 pages, 6409 KiB  
Article
Intracellular Iron Deficiency and Abnormal Metabolism, Not Ferroptosis, Contributes to Homocysteine-Induced Vascular Endothelial Cell Death
by Wenting Shi, Jing Zhang, Wairong Zhao, Meiyan Yue, Jie Ma, Silu Zeng, Jingyi Tang, Yu Wang and Zhongyan Zhou
Biomedicines 2024, 12(10), 2301; https://doi.org/10.3390/biomedicines12102301 - 10 Oct 2024
Viewed by 340
Abstract
Background/Objectives: Homocysteine (Hcy) and iron are factors co-related with the progression of cardiovascular diseases. The vascular endothelium is an important barrier for physiological homeostasis, and its impairment initiates cardiovascular injury. However, the mechanism underlying Hcy-caused vascular endothelial cell injury and the participation of [...] Read more.
Background/Objectives: Homocysteine (Hcy) and iron are factors co-related with the progression of cardiovascular diseases. The vascular endothelium is an important barrier for physiological homeostasis, and its impairment initiates cardiovascular injury. However, the mechanism underlying Hcy-caused vascular endothelial cell injury and the participation of iron are not fully elucidated. This study aims to investigate the Hcy-induced vascular endothelial injury and iron metabolism dysfunction as well as the underlying molecular mechanism. Methods: Human umbilical vein endothelial cells (HUVECs) were employed as the experimental model to examine the Hcy-induced endothelial injury and its underlying mechanism via various biochemical assays. Results: Hcy suppressed the cell viability and proliferation and caused cell death in a concentration-dependent manner. Hcy induced cell cycle arrest, apoptosis, and autophagy as well as impairment of intracellular energy metabolism. Hcy disrupted the intracellular antioxidant system and mitochondrial function by increasing intracellular ROS, MDA and mitochondrial content, and decreasing the SOD activity and mitochondrial membrane potential. Hcy significantly reduced the GSH-Px activity along with the accumulation of intracellular GSH in a concentration-dependent manner. Ferroptosis inhibitors, Ferrostatin-1 (Fer-1), and Deferoxamine (DFO) significantly decreased the Hcy-caused cytotoxicity accompanied by a reduction in dysregulated mitochondria content, but only DFO ameliorated the elevation of intracellular ROS, and neither Fer-1 nor DFO affected the Hcy-caused reduction in intracellular ATP. In addition, Hcy decreased the intracellular concentration of iron, and supplementing Hcy with various concentrations of Fe3+ increased the cell viability and decreased the LDH release in a concentration-dependent manner. Hcy dramatically decreased the mRNA expression level of transferrin receptor while increasing the mRNA expression levels of transferrin, ferritin light chain, ferritin heavy chain, ferroportin, and SLC7A11. Moreover, Hcy suppressed the protein expression of phosphor-Akt, phosphor-mTOR, Beclin-1, LC3A/LC3B, Nrf2, HO-1, phosphor-MEK1/2, phosphor-ERK1/2, and Caspase-3 in concentration- and time-dependent manners. Conclusions: Hcy-induced vascular endothelial injury is likely to be associated with apoptosis and autophagy, but not ferroptosis. The key underlying mechanisms are involved in the disruption of the intracellular antioxidant system and iron metabolism via regulation of PI3K/Akt/mTOR, MAPKs, Nrf2/HO-1, and iron metabolism. Full article
Show Figures

Figure 1

Figure 1
<p>The effect of Hcy on endothelial cell morphology, cytotoxicity, and cell viability in HUVECs<b>.</b> (<b>A</b>) The cell morphology of HUVECs treated with indicated concentrations of Hcy for 24 h, n = 3. (<b>B</b>) The cytotoxicity and cell viability of Hcy were examined by LDH release and MTT assays, respectively. Data are presented as folds or percentages of the control group, n = 4. (<b>C</b>) HUVECs were suspended with various concentrations (2, 4, and 8 mM) of Hcy and cultured in an RTCA system for 24 h, and the anti-proliferation effect of Hcy is presented as the Cell Index. The Cell Index of the Hcy-treated group and the control group were also summarized at 24 h, n = 3. (<b>D</b>) HUVECs were cultured and attached for 24 h and then treated with various concentrations (2, 4, and 8 mM) of Hcy for another 24 h. The toxicity effect of Hcy was recorded in real time by RTCA. The Cell Index was normalized to the folds of the value at the time point of adding Hcy. The normalized Cell Index of the Hcy-treated group and control group were also summarized after treatment with Hcy for 24 h, n = 3. (<b>E</b>,<b>F</b>) Presentive images and analysis of live and dead cell staining. The green (calcein-AM) and red (PI) fluorescence demonstrate the live and dead cells, respectively, n = 5. Data are presented as mean ± S.E.M. # <span class="html-italic">p</span> ˂ 0.05, ## <span class="html-italic">p</span> ˂ 0.01, and ### <span class="html-italic">p</span> ˂ 0.001 vs. the control group.</p>
Full article ">Figure 2
<p>The effects of Hcy on cell cycle, apoptosis, autophagy, and energy metabolism in HUVECs. (<b>A</b>) The cell cycle was analyzed by PI staining, followed by flow cytometry. The cell population percentages of the G0/G1, S, and G2/M phases were summarized in both the control group and the Hcy-treated group, n = 3. (<b>B</b>) The apoptosis cells were detected by annexin V-FITC and PI double-staining using flow cytometry. The percentages of early (LR) and late (UR) apoptotic cells were calculated and summarized, n = 3. (<b>C</b>–<b>G</b>) HUVECs were co-treated with Hcy (8 mM) with various indicated concentrations of Z-VDA-FMK (n = 3), 3-MA (n = 3), Wort (n = 3), LY294002 (n = 3), or Rapa (n = 4) for 24 h. The cytotoxicity was examined by LDH release assay. Results are presented as folds of the control group. (<b>H</b>) The intracellular ATP concentration (µmol/mg protein) was quantified by the commercially available kit, n = 4. Data are presented as mean ± S.E.M. # <span class="html-italic">p</span> ˂ 0.05, ## <span class="html-italic">p</span> ˂ 0.01, and ### <span class="html-italic">p</span> ˂ 0.001 vs. the control group. * <span class="html-italic">p</span> ˂ 0.05, ** <span class="html-italic">p</span> ˂ 0.01, and *** <span class="html-italic">p</span> ˂ 0.001 vs. the Hcy-treated group.</p>
Full article ">Figure 3
<p>The effect of Hcy on intracellular ROS and lipid peroxidation in HUVECs. (<b>A</b>,<b>B</b>) The HUVECs were treated with Hcy (8 mM) for 24 h, and then the intracellular ROS were indicated by DHE staining. The fluorescence intensity was calculated using ImageJ software (1.49 V), n = 3. (<b>C</b>) The HUVECs were treated with various indicated concentrations of Hcy for 24 h and the SOD activity was measured using a commercially available kit, n = 3. (<b>D</b>–<b>F</b>) Co-treatment of Hcy (8 mM) with indicated concentrations of DPI (n = 3), NAC (n = 5), VitE (n = 3), and liproxtatin-1 (n = 3) for 24 h, followed by cytotoxicity detection using LDH release kit. (<b>G</b>) The HUVECs were treated with various indicated concentrations of Hcy for 24 h. The intracellular level of MDA was measured using a commercially available kit, n = 8. (<b>H</b>) Co-treatment of Hcy (8 mM) with various indicated concentrations of Liproxtatin-1 for 24 h followed by LDH release assay, n = 3. The results were normalized to folds of the control group. Data are presented as mean ± S.E.M. # <span class="html-italic">p</span> ˂ 0.05, ## <span class="html-italic">p</span> ˂ 0.01, and ### <span class="html-italic">p</span> ˂ 0.001 vs. the control group. * <span class="html-italic">p</span> ˂ 0.05, and *** <span class="html-italic">p</span> ˂ 0.001 vs. the Hcy-treated group.</p>
Full article ">Figure 4
<p>The role of ferroptosis on Hcy-induced vascular endothelial cell toxicity in HUVECs. (<b>A</b>,<b>B</b>) The HUVECs were treated with various concentrations (2, 4, and 8 mM) of Hcy for 24 h. The intracellular GSH-Px activity and GSH were measured by commercially available kits, n = 5. (<b>C</b>,<b>D</b>) The HUVECs were treated with Hcy (8 mM) for 24 h. The mRNA expression of GPX4 (n = 7) and SLC7A11 (n = 5) genes was examined by real-time PCR. (<b>E</b>,<b>F</b>) The HUVECs were co-treated with Hcy (8 mM), with various indicated concentrations of Fer-1 or DFO for 24 h, followed by the LDH release assay, n = 3. (<b>G</b>,<b>H</b>) HUVECs were suspended with Hcy (8 mM), with or without Fer-1 (80 µM) or DFO (80 µM), for 24 h. The cell proliferation was recorded in real time by RTCA, n = 4. The results were normalized to folds of the control group. Data are presented as mean ± S.E.M. # <span class="html-italic">p</span> ˂ 0.05, ## <span class="html-italic">p</span> ˂ 0.01, and ### <span class="html-italic">p</span> ˂ 0.001 vs. the control group. * <span class="html-italic">p</span> ˂ 0.05, ** <span class="html-italic">p</span> ˂ 0.01, and *** <span class="html-italic">p</span> ˂ 0.001 vs. the Hcy-treated group.</p>
Full article ">Figure 5
<p>The effect of ferroptosis inhibitors on Hcy-induced intracellular ROS production and mitochondrial dysfunction in HUVECs. (<b>A</b>–<b>C</b>) The HUVECs were treated with Hcy (8 mM), with or without Fer-1 (80 µM) or DFO (80 µM), for 24 h. Then, the intracellular ROS, number of mitochondria, and mitochondria membrane potential were indicated by DHE (<b>A</b>), Mito-tracker (<b>B</b>), and JC-1 (<b>C</b>) staining, respectively, n = 3. (<b>D</b>–<b>F</b>) The fluorescence intensity was calculated using ImageJ software. The results were normalized to folds of the control group. (<b>G</b>) The HUVECs were treated with Hcy (8 mM), with or without Fer-1 (80 µM) or DFO (80 µM), for 24 h. Then, the intracellular concentration of ATP was detected by a commercially available kit, n = 4. Data are presented as mean ± S.E.M. # <span class="html-italic">p</span> ˂ 0.05 and ## <span class="html-italic">p</span> ˂ 0.01 vs. the control group. * <span class="html-italic">p</span> ˂ 0.05, and *** <span class="html-italic">p</span> ˂ 0.001 vs. the Hcy-treated group.</p>
Full article ">Figure 6
<p>The role of iron metabolism in Hcy-induced vascular endothelial cell toxicity in HUVECs. (<b>A</b>) HUVECs were treated with Hcy (8 mM) for 24 h. The intracellular level of iron was measured by a commercially available kit according to its manual, n = 3. (<b>B</b>,<b>C</b>) Co-treatment of Hcy (8 mM) with indicated concentrations of Fe<sup>3+</sup> for 24 h, followed by cell viability and cytotoxicity detections using MTT and LDH release assays, respectively, n = 3. (<b>D</b>) The cell morphology was observed by an inverted microscope equipped with 10× and 20× objective lenses. (<b>E</b>) The HUVECs were treated with or without Hcy (8 mM) for 24 h. The mRNA expressions of transferrin receptor (n = 6), transferrin (n = 5), ferritin light chain (n = 5), ferritin heavy chain (n = 6), and ferriportin (n = 6) genes were examined by real-time PCR. Results are presented as folds or percentages of the control group. Data are presented as mean ± S.E.M. # <span class="html-italic">p</span> ˂ 0.05, ## <span class="html-italic">p</span> ˂ 0.01, and ### <span class="html-italic">p</span> ˂ 0.001 vs. the control group. * <span class="html-italic">p</span> ˂ 0.05, ** <span class="html-italic">p</span> ˂ 0.01, and *** <span class="html-italic">p</span> ˂ 0.001 vs. the Hcy-treated group.</p>
Full article ">Figure 7
<p>The effect of Hcy on Akt/mTOR autophagy signaling in HUVECs. (<b>A</b>) The HUVECs were treated with various concentrations (2, 4, and 8 mM) of Hcy for 24 h. The representative bands in Western blotting analysis. (<b>B</b>–<b>E</b>) The quantitative protein expressions of phospho-mTOR (n = 3), mTOR (n = 3), phosphor-Akt (n = 4), Akt (n = 4), Beclin-1 (n = 4), LC3A/B (n = 4), and GAPDH (n = 4) were detected by Western blotting analysis. GAPDH served as the internal control. The results were normalized to folds of the control group. Data are presented as mean ± S.E.M. # <span class="html-italic">p</span> ˂ 0.05, ## <span class="html-italic">p</span> ˂ 0.01, and ### <span class="html-italic">p</span> ˂ 0.001 vs. the control group.</p>
Full article ">Figure 8
<p>The effect of Hcy on MAPKs and Nrf2/HO-1 signaling in HUVECs. (<b>A</b>) The HUVECs were treated with various concentrations (2, 4, and 8 mM) of Hcy for 24 h. The representative bands in Western blotting analysis. (<b>B</b>–<b>F</b>) The quantitative protein expressions of Nrf2 (n = 4), HO-1 (n = 4), phosphor-MEK1/2 (n = 4), MEK1/2 (n = 4), phosphor-ERK1/2 (n = 3), ERK1/2 (n = 3), Caspase-3 (n = 4), and GAPDH (n = 3) were detected by Western blotting analysis. GAPDH served as the internal control. The results were normalized to folds of the control group. Data are presented as mean ± S.E.M. # <span class="html-italic">p</span> ˂ 0.05, ## <span class="html-italic">p</span> ˂ 0.01, and ### <span class="html-italic">p</span> ˂ 0.001 vs. the control group.</p>
Full article ">Figure 9
<p>Schematic overview of the underlying mechanism of Hcy-induced endothelial death.</p>
Full article ">
16 pages, 3593 KiB  
Article
A Versatile Microfluidic Device System that Lacks a Synthetic Extracellular Matrix Recapitulates the Blood–Brain Barrier and Dynamic Tumor Cell Interaction
by Daniel Santillán-Cortez, Andrés Eliú Castell-Rodríguez, Aliesha González-Arenas, Juan Antonio Suárez-Cuenca, Vadim Pérez-Koldenkova, Denisse Añorve-Bailón, Christian Gabriel Toledo-Lozano, Silvia García, Mónica Escamilla-Tilch and Paul Mondragón-Terán
Bioengineering 2024, 11(10), 1008; https://doi.org/10.3390/bioengineering11101008 - 10 Oct 2024
Viewed by 436
Abstract
Microfluidic systems offer controlled microenvironments for cell-to-cell and cell-to-stroma interactions, which have precise physiological, biochemical, and mechanical features. The optimization of their conditions to best resemble tumor microenvironments constitutes an experimental modeling challenge, particularly regarding carcinogenesis in the central nervous system (CNS), given [...] Read more.
Microfluidic systems offer controlled microenvironments for cell-to-cell and cell-to-stroma interactions, which have precise physiological, biochemical, and mechanical features. The optimization of their conditions to best resemble tumor microenvironments constitutes an experimental modeling challenge, particularly regarding carcinogenesis in the central nervous system (CNS), given the specific features of the blood–brain barrier (BBB). Gel-free 3D microfluidic cell culture systems (gel-free 3D-mFCCSs), including features such as self-production of extracellular matrices, provide significant benefits, including promoting cell–cell communication, interaction, and cell polarity. The proposed microfluidic system consisted of a gel-free culture device inoculated with human brain microvascular endothelial cells (HBEC5i), glioblastoma multiforme cells (U87MG), and astrocytes (ScienCell 1800). The gel-free 3D-mFCCS showed a diffusion coefficient of 4.06 × 10−9 m2·s−1, and it reconstructed several features and functional properties that occur at the BBB, such as the vasculogenic ability of HBEC5i and the high duplication rate of U87MG. The optimized conditions of the gel-free 3D-mFCCS allowed for the determination of cellular proliferation, invasion, and migration, with evidence of both physical and biochemical cellular interactions, as well as the production of pro-inflammatory cytokines. In conclusion, the proposed gel-free 3D-mFCCSs represent a versatile and suitable alternative to microfluidic systems, replicating several features that occur within tumor microenvironments in the CNS. This research contributes to the characterization of microfluidic approaches and could lead to a better understanding of tumor biology and the eventual development of personalized therapies. Full article
(This article belongs to the Section Cellular and Molecular Bioengineering)
Show Figures

Graphical abstract

Graphical abstract
Full article ">Figure 1
<p>Cell culture models used to mimic the microenvironmental complexity of brain tumors. Cell culture platforms represent the increasing complexity of cellular models that are used to study brain tumors. This complexity is related to their ability to mimic real biological microenvironments, 3D architectures, and cell-to-cell and micro-vascularity interactions.</p>
Full article ">Figure 2
<p>Design and fabrication of the microfluidic system for cellular interaction research.</p>
Full article ">Figure 3
<p>The device design and permeability of the proposed microfluidic system: (<b>A</b>) comparative size of device; (<b>B</b>) lateral view of device; (<b>C</b>) progressive permeability in the device, demonstrated by the permeation of blue food stain (upper panel) and Texas red dextran (10WD; lower panel). Time is also indicated for comparison. The yellow arrows show the flow to and communication with the other cell culture microchannels. The white scale bar represents 100 µm.</p>
Full article ">Figure 4
<p>The growth kinetics of the HBEC5i (endothelial cells), U87MG (glioblastoma tumoral cells), and astrocyte cell lines. The top central graphs show the quantification of the growth rate with representative micrographs of the time course of proliferation in the culture plate below each graph. The scale bar represents 100 µm in all.</p>
Full article ">Figure 5
<p>The cell migration ability of (<b>A</b>) U87MG endothelial cells and (<b>B</b>) HBEC5i at different starting cell densities. Migrating cells are indicated in yellow.</p>
Full article ">Figure 6
<p>Culture U87MG in microfluidic system culture. The cell culture process at a density of 2 × 10<sup>4</sup> cells·µL<sup>−1</sup>: a heatmap of the cell distribution at 0 h is shown for cell loading, and as time progresses (48–150 h), cell clusters that were initially formed exhibited 3D projections, which have been described as organoids, and were acquired via confocal microscopy at 150 h. The cellular cytoskeleton is shown with the expression of phalloidin (green), and the nuclei were assessed with DAPI (blue). The scale bar represents 100 µm.</p>
Full article ">Figure 7
<p>Co-culture of U87MG-HBEC5i acquired by confocal microscopy: the cell migration process. The cellular cytoskeleton is shown with the expression of phalloidin (green), and the nuclei were assessed with DAPI (blue). Cell–cell interactions are shown with yellow arrows, which occurred through interconnecting channels. The scale bar represents 100 µm.</p>
Full article ">Figure 8
<p>The culture evolution at different time points and the process of cell invasion. Glioblastoma U87MG cells are shown in green, endothelial HBEC5i cells are shown in red, cell nuclei are shown in blue (DAPI), and cell–cell interactions through interconnecting channels are shown by green and red arrows. The expression of cytokines was carried out by combining the triplicate model to obtain the necessary amount of culture medium (cytokine sample = 1). The scale bar represents 200 µm.</p>
Full article ">
25 pages, 6163 KiB  
Article
A Human Brain-Chip for Modeling Brain Pathologies and Screening Blood–Brain Barrier Crossing Therapeutic Strategies
by Shek Man Chim, Kristen Howell, Alexandros Kokkosis, Brian Zambrowicz, Katia Karalis and Elias Pavlopoulos
Pharmaceutics 2024, 16(10), 1314; https://doi.org/10.3390/pharmaceutics16101314 - 10 Oct 2024
Viewed by 1220
Abstract
Background/Objectives: The limited translatability of preclinical experimental findings to patients remains an obstacle for successful treatment of brain diseases. Relevant models to elucidate mechanisms behind brain pathogenesis, including cell-specific contributions and cell-cell interactions, and support successful targeting and prediction of drug responses in [...] Read more.
Background/Objectives: The limited translatability of preclinical experimental findings to patients remains an obstacle for successful treatment of brain diseases. Relevant models to elucidate mechanisms behind brain pathogenesis, including cell-specific contributions and cell-cell interactions, and support successful targeting and prediction of drug responses in humans are urgently needed, given the species differences in brain and blood-brain barrier (BBB) functions. Human microphysiological systems (MPS), such as Organ-Chips, are emerging as a promising approach to address these challenges. Here, we examined and advanced a Brain-Chip that recapitulates aspects of the human cortical parenchyma and the BBB in one model. Methods: We utilized human primary astrocytes and pericytes, human induced pluripotent stem cell (hiPSC)-derived cortical neurons, and hiPSC-derived brain microvascular endothelial-like cells and included for the first time on-chip hiPSC-derived microglia. Results: Using Tumor necrosis factor alpha (TNFα) to emulate neuroinflammation, we demonstrate that our model recapitulates in vivo-relevant responses. Importantly, we show microglia-derived responses, highlighting the Brain-Chip’s sensitivity to capture cell-specific contributions in human disease-associated pathology. We then tested BBB crossing of human transferrin receptor antibodies and conjugated adeno-associated viruses. We demonstrate successful in vitro/in vivo correlation in identifying crossing differences, underscoring the model’s capacity as a screening platform for BBB crossing therapeutic strategies and ability to predict in vivo responses. Conclusions: These findings highlight the potential of the Brain-Chip as a reliable and time-efficient model to support therapeutic development and provide mechanistic insights into brain diseases, adding to the growing evidence supporting the value of MPS in translational research and drug discovery. Full article
Show Figures

Figure 1

Figure 1
<p>Characterization of the human cortical Brain-Chip. (<b>A</b>) Representative confocal images showing the hiPSC-derived microvascular endothelial-like cells (iBMECs) attached to the porous membrane (vascular channel). (<b>i</b>) Immunostaining against the tight junction marker ZO-1. Stack of Z-series for the vascular channel (left) and high magnification optical section of ZO-1 staining (right) are shown. (<b>ii</b>) Immunostaining against the brain microvascular endothelial cell marker GLUT1 (stack of Z-series). (<b>B</b>) Confocal images of astrocytes (GFAP) with pericytes (NG2) (<b>i</b>) and neurons (MAP2) with microglia (Iba1 and CD68) (<b>ii</b>) attached to the porous membrane in the brain channel. Confocal images (stack of Z-series) of the entire brain channel (<b>top</b>) and high-magnification confocal optical sections (<b>bottom</b>). All cell types were present and uniformly distributed along the entire brain channel. (<b>C</b>) (<b>i</b>): Confocal micrograph (stack of z-series) showing immunofluorescence staining against GFAP (astrocytes) and MAP2 (neurons) coupled with phase contrast for visualization of the porous membrane. (<b>ii</b>): Digital 3D reconstruction of z-series image stacks showing the Brain-Chip from the side. The interrupted line indicates the location of the porous membrane separating the brain from the vascular channel. The nuclear staining (Hoechst) on the vascular side indicates the iBMECs. A GFAP signal is detected in the vascular side (arrows). Arrows in both images indicate the astrocytic end-feet passing through the 7 μm pores extending into the vascular channel. (<b>D</b>) Schematic representation of the experimental design and averaged data from quantitative barrier function analysis via apparent permeability (Papp) to 3 kDa fluorescent dextran crossing through the vascular channel to the brain channel on Days 1 through 6 in microfluidics. Chips with and without iBMECs were examined (N = 6 chips/group). Each data point represents an individual chip. Graph: mean ± SEM. Shaded box: range of Papp values shown in animal models. (<b>E</b>) Confocal images showing GABAergic (VGAT) and glutamatergic neurons (VGLUT1) in the brain channel of the chip. All stainings were performed in Brain-Chips after six days in microfluidics. (<b>F</b>) Examination of functional connectivity between GABAergic and glutamatergic neurons using pharmacology and extracellular glutamate measurements. The experimental design (top) and extracellular glutamate quantification (mean ± SEM) for the indicated time points and treatments are shown. N = 3 chips/group, *** <span class="html-italic">p</span> &lt; 0.001, two-way ANOVA and post hoc Tukey’s test.</p>
Full article ">Figure 2
<p>TNFα-induced neuroinflammation and BBB disruption in the Brain-Chip. (<b>A</b>) Outline of the experimental design. Beginning on Day 2 in microfluidics, 100 ng/mL of TNFα were dosed in the brain channel and replenished 24 h later. Chips dosed with PBS were used as the control. Immunocytochemistry and extracellular glutamate measurements were performed on Day 4. Effluents were collected daily from Day 1 to Day 4 for the BBB permeability assay (Days 1–4) and cytokines/chemokines analysis (Days 2–4). (<b>B</b>) Representative confocal images of microglia (CD68) and neurons (MAP2) (<b>i</b>) and astrocytes (<b>ii</b>). (<b>iii</b>) Averaged data (mean ± SEM) for the number of CD68+ cells and MAP2 intensity. TNFα treatment increases the numbers of CD68-positive cells, indicative of microglial reactivity. The signal intensity of the neuronal dendritic marker MAP2 is decreased, suggesting neuronal dysfunction. High-resolution stacks of z-series from brain channel areas (50% coverage of the channel) were analyzed for each chip. N = 3 chips/treatment. Confocal images of all chips used for MAP and CD68 analysis can be found in <a href="#app1-pharmaceutics-16-01314" class="html-app">Figure S2</a>. The morphology of reactive astrocytes upon TNFα exposure changes from a polygonal state to a more elongated state (see <a href="#app1-pharmaceutics-16-01314" class="html-app">Figure S3</a> for additional chips and images). (<b>C</b>) Averaged data (mean ± SEM) of extracellular glutamate measurements in brain effluents collected at the end of the experiment (day 4). N = 3 chips/group. Red asterisks: comparison with TNFα-treated chips without microglia. Black asterisks: comparison with the respective control. (<b>D</b>) Apparent permeability (Papp) of the barrier across days (mean ± SEM). Papp on Day 2 was measured immediately before TNFα perfusion. Chips with microglia, N = 4; chips without microglia, N = 3. (<b>B</b>–<b>D</b>) * <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, **** <span class="html-italic">p</span> &lt; 0.0001, Student’s <span class="html-italic">t</span>-test (<b>B</b>) and one-way (<b>C</b>) or two-way (<b>D</b>) ANOVA with post hoc Tukey’s test (significantly different compared with all other groups).</p>
Full article ">Figure 3
<p>TNFα-induced secretion of cytokines and chemokines and contribution of microglia. Longitudinal analysis of cytokines and chemokines in brain channel effluents collected at the indicated time points. Effluent collection on Day 2 was performed immediately prior to TNFα dosing (baseline levels of cytokines and chemokines). Brain-Chips with and without microglia were examined to determine their contribution to the observed inflammatory responses. All other cell types (astrocytes, pericytes, glutamatergic and GABAergic neurons, brain microvascular endothelial-like cells) were present in the chips. Brain-Chips (with and without microglia) treated with PBS were used as controls. Graphs: averaged data (mean ± SEM) from 4 chips for each group. Microglial-specific responses are indicated by the blue rectangles.</p>
Full article ">Figure 4
<p>Specificity of the Brain-Chip for human transferrin receptor-mediated BBB crossing. (<b>A</b>) Schematic representation of experimental design. On Day 1 in microfluidics, the vascular channel was perfused with a human antibody specific for human TfR1 or mouse TfR1, or isotype control antibody (human IgG1), at a final concentration ~10 μg/mL. PBS was used as a baseline control. On Day 2, effluents from the brain channel were collected for hIgG1 measurements. The permeability of the barrier was examined on Days 1 and 2 by measuring its Papp to a 3 kDa fluorescent dextran perfused to the vascular channel starting at the beginning of incubation in microfluidics. (<b>B</b>) All chips had a tight barrier with similar Papp values, which were within the range of those shown in published works and in rodent models (shaded box). Each data point represents an individual chip. (<b>C</b>) Measurement of hIgG1 levels in culture media of vascular and brain channels prior to perfusion of the TfR1 antibodies (N = 3 and N = 5 chips, respectively, for vascular and brain channel). hIgG1 was detected in the vascular maintenance medium, as it contained 2% human serum. Data points represent three individual measurements. (<b>D</b>) Quantification of hIgG1 in the vascular channel media prior to perfusion (influent) (<b>i</b>) and in brain channel effluents collected 24 h post-dosing (<b>ii</b>). The percentage of vascular hIgG1 detected in the brain channel for each treatment group was also calculated (<b>iii</b>). Averaged data (mean ± SEM) and individual chip values are shown in (<b>i</b>–<b>iii</b>). N = 3 chips per group. ** <span class="html-italic">p</span> &lt; 0.01, *** <span class="html-italic">p</span> &lt; 0.001, one-way ANOVA with post hoc Tukey’s test. Blue asterisks: comparison with mouse TfR1. Black asterisks: comparison with hIgG1.</p>
Full article ">Figure 5
<p>The human Brain-Chip has the resolution to identify BBB crossing differences between human TfR1 antibodies. (<b>A</b>) Outline of the experimental design. The chips were connected to microfluidics for one day prior to antibody administration of human TfR1 antibodies with varying BBB crossing properties in vivo. The vascular channel was dosed with the antibodies in serum-free culture media at 10 μg/mL. PBS and an isotype antibody were used as negative controls. After 8 h, effluents from the brain channel were collected for antibody measurements. (<b>B</b>) Antibody quantification data. Mean ± SEM and individual chip values are shown. Antibody quantification in vascular channel media prior to perfusion (<b>left</b>) and in the brain channel effluent 8 h post-dosing (<b>middle</b>) are shown. The <b>right</b> graph shows the percentage of the dosed antibodies detected in the brain channel. Four human TfR1 antibody clones with different BBB crossing ability were examined (REGN1, REGN5, REGN12, and REGN28). The clones were tested in mice as conjugates to AAV9 expressing GFP, and their BBB crossing ability was determined based on immunohistochemical detection and quantification of the GFP signal in the brain parenchyma. The numeric immunohistochemical scores and their order based on their BBB crossing abilities are shown below the graphs. N = 6 chips/group. * <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, **** <span class="html-italic">p</span> &lt; 0.0001, one-way ANOVA with post hoc Tukey’s test. Black asterisks: comparison with isotype antibody control. Light blue asterisks: comparison with REGN12. No differences in BBB permeability were observed between groups (<a href="#app1-pharmaceutics-16-01314" class="html-app">Figure S5A</a>).</p>
Full article ">Figure 6
<p>The human Brain-Chip detects different levels of permeation hTfR1 antibody-conjugated AAV9. (<b>A</b>) Schematic of the experimental design. AAV9 expressing GFP and decorated with hTfR1 antibodies were perfused in the vascular channel for two days (Days 1–3). On Day 5, the endothelial-like cells from the vascular channel and the brain channel cells were collected for GFP protein analysis. (<b>B</b>) Graphs showing averaged data (mean ± SEM) and individual chip values of GFP protein quantification analysis in endothelial and brain cell lysates, as indicated. The hTfR1 antibody-conjugated viruses were tested in vivo for their BBB crossing abilities, based on GFP signal intensity scoring in the mouse brain parenchyma, as shown below the graphs. We examined AAV9 conjugated with the same hTfR1 clones tested as purified antibodies in the chip (REGN1, REGN5, REGN12, REGN28) plus viruses conjugated with the antibody clone REGN25, which exhibited low BBB crossing in vivo. Unconjugated AAV9 and hASGR1-conjugated viruses were used as negative controls. hTfR1 antibody-conjugated AAV9: N = 6 chips for each group; PBS, AAV9 and hASGR1-AAV9: 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, **** <span class="html-italic">p</span> &lt; 0.0001, one-way ANOVA with post hoc Tukey’s test. Black asterisks: comparison with PBS, AAV9, and hASGR1-AAV9. Red asterisks: comparison with REGN1. Blue asterisks: comparison with REGN5. Light blue asterisks: comparison with REGN12. All chips had a tight barrier with comparable Papp (<a href="#app1-pharmaceutics-16-01314" class="html-app">Figure S5A</a>). The amount of total protein in endothelial and brain channel cell lysates was comparable between groups (<a href="#app1-pharmaceutics-16-01314" class="html-app">Figure S5B</a>).</p>
Full article ">
15 pages, 3576 KiB  
Article
High-Throughput Transcriptomics Identifies Chemoresistance-Associated Gene Expression Signatures in Human Angiosarcoma
by Glenys Mai Shia Khor, Sara Haghani, Tiffany Rui En Tan, Elizabeth Chun Yong Lee, Bavani Kannan, Boon Yee Lim, Jing Yi Lee, Zexi Guo, Tun Kiat Ko and Jason Yongsheng Chan
Int. J. Mol. Sci. 2024, 25(19), 10863; https://doi.org/10.3390/ijms251910863 - 9 Oct 2024
Viewed by 431
Abstract
Angiosarcomas, clinically aggressive cancers of endothelial origin, are a rare subtype of soft-tissue sarcomas characterized by resistance to chemotherapy and dismal prognosis. In this study, we aim to identify the transcriptomic biomarkers of chemoresistance in angiosarcoma. We examined 72 cases of Asian angiosarcomas, [...] Read more.
Angiosarcomas, clinically aggressive cancers of endothelial origin, are a rare subtype of soft-tissue sarcomas characterized by resistance to chemotherapy and dismal prognosis. In this study, we aim to identify the transcriptomic biomarkers of chemoresistance in angiosarcoma. We examined 72 cases of Asian angiosarcomas, including 35 cases treated with palliative chemotherapy, integrating information from NanoString gene expression profiling, whole transcriptome profiling (RNA-seq), immunohistochemistry, cell line assays, and clinicopathological data. In the chemoresistant cohort (defined as stable disease or progression), we observed the significant overexpression of genes, including SPP1 (log2foldchange 3.49, adj. p = 0.0112), CXCL13, CD48, and CLEC5A, accompanied by the significant enrichment of myeloid compartment and cytokine and chemokine signaling pathways, as well as neutrophils and macrophages. RNA-seq data revealed higher SPP1 expression (p = 0.0008) in tumor tissues over adjacent normal compartments. Immunohistochemistry showed a significant moderate positive correlation between SPP1 protein and gene expression (r = 0.7016; p < 0.00110), while higher SPP1 protein expression correlated with lower chemotherapeutic sensitivity in patient-derived angiosarcoma cell lines MOLAS and ISOHAS. In addition, SPP1 mRNA overexpression positively correlated with epithelioid histology (p = 0.007), higher tumor grade (p = 0.0023), non-head and neck location (p = 0.0576), and poorer overall survival outcomes (HR 1.84, 95% CI 1.07–3.18, p = 0.0288). There was no association with tumor mutational burden, tumor inflammation signature, the presence of human herpesvirus-7, ultraviolet exposure signature, and metastatic state at diagnosis. In conclusion, SPP1 overexpression may be a biomarker of chemoresistance and poor prognosis in angiosarcoma. Further investigation is needed to uncover the precise roles and underlying mechanisms of SPP1. Full article
Show Figures

Figure 1

Figure 1
<p>Transcriptomic analysis of chemoresistant versus chemosensitive angiosarcoma. (<b>A</b>) Volcano plot of the differential gene expression in chemoresistant versus chemosensitive angiosarcoma reveals significant <span class="html-italic">SPP1</span> overexpression in chemoresistant angiosarcoma. (<b>B</b>) Significant upregulation of <span class="html-italic">SPP1</span> expression in tumor compared to matched normal tissue reflected in whole transcriptome sequencing data. (<b>C</b>) Pathway analysis showed that expression of myeloid compartment, cytokine and chemokine signaling, and matrix remodeling and metastasis pathways were upregulated in chemoresistant tumors. Conversely, Hedgehog, Notch, and Wnt signaling pathways were upregulated in chemosensitive tumors. (<b>D</b>) Cell-type analysis of chemoresistant tumors suggested enrichment of myeloid cells including neutrophils, macrophages, and NK cells, whereas a greater population of mast cells, dendritic cells (DC), and regulatory T cells (Treg) were observed in chemosensitive tumors.</p>
Full article ">Figure 2
<p>Chemotherapy response of angiosarcoma correlates with immuno-oncogenic pathways and cell types. (<b>A</b>) Significance of association of top three immuno-oncology pathways with chemotherapy non-response versus response in angiosarcoma. (<b>B</b>) Significance of association of top three cell types with chemotherapy non-response versus response in angiosarcoma.</p>
Full article ">Figure 3
<p>Transcriptomic analysis of <span class="html-italic">SPP1</span>-high and <span class="html-italic">SPP1</span>-low angiosarcoma. (<b>A</b>) Volcano plot of differentially expressed genes between <span class="html-italic">SPP1</span>-high and <span class="html-italic">SPP1</span>-low tumors (n = 72) revealed significant overexpression of various genes, including <span class="html-italic">SLC11A1</span>, <span class="html-italic">PLOD2</span>, <span class="html-italic">CXCL3</span>, <span class="html-italic">IL6</span>, <span class="html-italic">FSTL3</span>, <span class="html-italic">CXCL2</span>, <span class="html-italic">TREM1</span> and <span class="html-italic">IL1β</span>. (<b>B</b>–<b>D</b>) NanoString pathway analysis and gene-specific analysis (GSA) revealed significant overexpression of myeloid compartment pathway genes (<span class="html-italic">SLC11A1</span>, <span class="html-italic">CXCL3</span>, <span class="html-italic">CXCL2</span>, <span class="html-italic">TREM1</span>, <span class="html-italic">IL1β</span>), cytokine and chemokine signaling pathway genes (<span class="html-italic">CXCL3</span>, <span class="html-italic">IL6</span>, <span class="html-italic">CXCL2</span>, <span class="html-italic">IL1β</span>), and matrix remodeling and metastasis pathway genes (<span class="html-italic">PLOD2</span>). (<b>C</b>) Cell-type analysis of <span class="html-italic">SPP1</span>-high tumors suggested enrichment of myeloid cells, including neutrophils, macrophages, and NK cells.</p>
Full article ">Figure 4
<p>Clinicopathological characteristics associated with <span class="html-italic">SPP1</span> expression in angiosarcoma. (<b>A</b>) Presence of epithelioid histology was significantly associated with higher <span class="html-italic">SPP1</span> gene expression levels. (<b>B</b>) Higher FNCLCC grading was significantly associated with higher <span class="html-italic">SPP1</span> gene expression levels. (<b>C</b>,<b>D</b>) Tumor mutational burden (TMB) and tumor inflammatory signature (TIS) scores were not associated with <span class="html-italic">SPP1</span> gene expression levels.</p>
Full article ">Figure 5
<p>Spatial distribution of <span class="html-italic">SPP1</span>. (<b>A</b>) Spatial transcriptomics analysis showed that <span class="html-italic">SPP1</span> expression displayed significant heterogeneity in terms of spatial distribution in all samples, with focal clusters of cells expressing higher levels of <span class="html-italic">SPP1</span> scattered throughout the tumor tissues. (<b>B</b>) Spatial distribution of various cell types in the tumor tissues. (<b>C</b>) Violin plots showing <span class="html-italic">SPP1</span> expression across all cell types in angiosarcoma, including stromal, immune, and tumor cells, and it was most enriched in myeloid cells.</p>
Full article ">Figure 6
<p>Prognostic implications of <span class="html-italic">SPP1</span>. (<b>A</b>) IHC images at 400× magnification (scale bar, 100 µM) of tissue sections stained with anti-osteopontin antibody (<span class="html-italic">SPP1</span>). Representative images for high and low <span class="html-italic">SPP1</span> H-scores are shown. (<b>B</b>) Scatterplot showing significant moderate correlation between <span class="html-italic">SPP1</span> expression levels from IHC H-scores and NanoString (Spearman’s r = 0.7016; <span class="html-italic">p</span> &lt; 0.0110). (<b>C</b>) Western blot reflected higher <span class="html-italic">SPP1</span> protein expression in ISOHAS as compared to MOLAS cell line. (<b>D</b>) In the patient-derived MOLAS and ISOHAS cell lines, treatment with paclitaxel and doxorubicin resulted in reduced viability in a dose-dependent manner, with ISOHAS demonstrating higher cell viability than MOLAS upon treatment with either chemotherapeutic drug. *, <span class="html-italic">p</span> &lt; 0.05. (<b>E</b>,<b>F</b>) Kaplan–Meier curves showing survival probability in patients with <span class="html-italic">SPP1</span>-high versus <span class="html-italic">SPP1</span>-low angiosarcoma. Patients with <span class="html-italic">SPP1</span>-high angiosarcoma showed poorer overall survival (CI 1.07 to 3.18, HR 1.84, <span class="html-italic">p</span> = 0.0288), along with a trend toward poorer progression-free survival (CI 0.98 to 3.06, HR 1.74, <span class="html-italic">p</span> = 0.0566) compared to patients with <span class="html-italic">SPP1</span>-low angiosarcoma.</p>
Full article ">
Back to TopTop