[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

Journals

Article Types

Countries / Regions

remove_circle_outline
remove_circle_outline
remove_circle_outline
remove_circle_outline
remove_circle_outline

Search Results (188)

Search Parameters:
Keywords = Bradykinin

Order results
Result details
Results per page
Select all
Export citation of selected articles as:
19 pages, 2746 KiB  
Article
LDL-c/HDL-c Ratio and NADPH-Oxidase-2-Derived Oxidative Stress as Main Determinants of Microvascular Endothelial Function in Morbidly Obese Subjects
by Jorge Santos, José M. La Fuente, Argentina Fernández, Paula Ruano and Javier Angulo
Antioxidants 2024, 13(9), 1139; https://doi.org/10.3390/antiox13091139 - 20 Sep 2024
Viewed by 537
Abstract
The identification of obese subjects at higher risk for cardiovascular disease (CVD) is required. We aimed to characterize determinants of endothelial dysfunction, the initial step to CVD, in small omental arteries of visceral fat from obese subjects. The influences of analytical parameters and [...] Read more.
The identification of obese subjects at higher risk for cardiovascular disease (CVD) is required. We aimed to characterize determinants of endothelial dysfunction, the initial step to CVD, in small omental arteries of visceral fat from obese subjects. The influences of analytical parameters and vascular oxidative stress mediated by NADPH-oxidase-2 (NOX2) on endothelial function were determined. Specimens were obtained from 51 obese subjects undergoing bariatric surgery and 14 non-obese subjects undergoing abdominal surgery. Obese subjects displayed reduced endothelial vasodilation to bradykinin (BK). Endothelial vasodilation (pEC50 for BK) among obese subjects was significantly and negatively associated with low-density lipoprotein cholesterol (LDL-c)/high-density lipoprotein cholesterol (HDL-c) ratio (r = -0.510, p = 0.0001) in both women and men, while other metabolic parameters and comorbidities failed to predict endothelial function. The vascular expression of NOX2 was upregulated in obese subjects and was related to decreased endothelial vasodilation (r = −0.529, p = 0.0006, n = 38) and increased oxidative stress (r = 0.783, p = 0.0044, n = 11) in arterial segments. High LDL-c/HDL-c (>2) and high NOX2 (above median) were independently associated with reduced endothelial function, but the presence of both conditions was related to a further impairment. Concomitant elevated LDL-c/HDL-c ratio and high vascular expression of NOX2 would exacerbate endothelial impairment in obesity and could reveal a deleterious profile for cardiovascular outcomes among obese subjects. Full article
(This article belongs to the Special Issue NADPH Oxidases (NOXs))
Show Figures

Figure 1

Figure 1
<p><b>Obesity is related to the impairment of endothelium-dependent vasodilation through endothelial hyperpolarization and nitric oxide-mediated relaxation in human mesenteric small arteries.</b> Endothelium-dependent vasodilations induced by bradykinin (BK, 1 nM to 10 µM) in human mesenteric arteries (HMAs) contracted with the thromboxane analogue, U46619 (10–30 nM) (<b>A</b>,<b>B</b>), or 25–35 mM K<sup>+</sup> © obtained from non-obese subjects (control, BMI &lt; 30) and from subjects with morbid obesity (obese, BMI &gt; 35). Panel (<b>A</b>) shows BK-induced responses in HMAs without treatment, while HMAs were treated with indomethacin (INDO, 10 µM) plus N<sup>G</sup>-nitro-L-arginine methyl ester (L-NAME, 100 µM) in panel (<b>B</b>) and with INDO (10 µM) in panel (<b>C</b>). Data are expressed as the mean ± S.E.M. of the percentage of relaxation. <span class="html-italic">n</span> indicates the number of patients from whom the vessels were collected. *** indicates <span class="html-italic">p</span> &lt; 0.001 vs. control by a two-factor ANOVA test.</p>
Full article ">Figure 2
<p><b>Low-density lipoprotein cholesterol (LDL-c)/high-density lipoprotein cholesterol (HDL-c) ratio determines endothelial vasodilation in obese subjects.</b> Upper panels show linear regressions of LDL-c/HDL-c ratio versus pEC<sub>50</sub> for bradykinin (BK) in subjects with morbid obesity (<b>A</b>–<b>C</b>). pEC<sub>50</sub> values for BK for each subject were obtained in human mesenteric arteries (HMAs) contracted with the thromboxane analogue, U46619 (10–30 nM) (<b>A</b>,<b>B</b>), or 25–35 mM K<sup>+</sup> (<b>C</b>). Lower panels show complete BK-induced vasodilations in HMAs contracted with the thromboxane analogue, U46619 (10–30 nM) (<b>D</b>,<b>E</b>), or 25–35 mM K<sup>+</sup> (<b>F</b>) obtained from obese subjects displaying LDL-c/HDL-c ratios equal to or below 2 (LDL-c/HDL-c ≤ 2) versus those with LDL-c/HDL-c above 2 (LDL-c/HDL-c &gt; 2). Data in panels (<b>A</b>,<b>D</b>) were obtained in HMAs without treatment, while HMAs were treated with indomethacin (INDO, 10 µM) plus N<sup>G</sup>-nitro-L-arginine methyl ester (L-NAME, 100 µM) in panels (<b>B</b>,<b>E</b>) and with INDO (10 µM) in panels (<b>C</b>,<b>F</b>). <span class="html-italic">n</span> indicates the number of patients from whom the determinations were obtained. Coefficients of determination and probability (<span class="html-italic">p</span>) values are indicated in upper panels. Significant associations are highlighted in bold plus italic. Data in lower panels are expressed as the mean ± S.E.M. of the percentage of relaxation. * indicates <span class="html-italic">p</span> &lt; 0.05; *** <span class="html-italic">p</span> &lt; 0.001 vs. control by a two-factor ANOVA test.</p>
Full article ">Figure 3
<p><b>A negative association of low-density lipoprotein cholesterol (LDL-c)/high-density lipoprotein cholesterol (HDL-c) ratio with endothelial vasodilation is observed in both female and male obese subjects.</b> Upper panels show vasodilations induced by bradykinin (BK, 1 nM to 10 µM) in HMAs contracted with the thromboxane analogue, U46619 (10–30 nM) (<b>A</b>,<b>B</b>), or 25–35 mM K+ (<b>C</b>) obtained from female (women) and male (men) obese subjects. Data in panel (<b>A</b>) were obtained in HMAs without treatment, while HMAs were treated with indomethacin (INDO, 10 µM) plus N<sup>G</sup>-nitro-L-arginine methyl ester (L-NAME, 100 µM) in panel (<b>B</b>) and with INDO (10 µM) in panel (<b>C</b>). Data are expressed as the mean ± S.E.M. of the percentage of relaxation. Lower panels show linear regressions of LDL-c/HDL-c ratio versus pEC<sub>50</sub> for BK in female (<b>D</b>) and male (<b>E</b>) subjects with morbid obesity. pEC<sub>50</sub> values for BK for each subject were obtained in HMAs contracted with the thromboxane analogue, U46619. Coefficients of determination and probability (<span class="html-italic">p</span>) values are indicated. Significant associations are highlighted in bold plus italic.</p>
Full article ">Figure 4
<p><b>The vascular expression of NADPH-oxidase-2 (NOX2) is related to increased oxidative stress and reduced endothelial vasodilation in obese subjects.</b> Panels (<b>A</b>,<b>B</b>) show representative immunofluorescence detection of NOX2 protein in sections of small mesenteric arteries from a control and an obese subject, respectively. Magnification ×200. High immunoreactivity (green fluorescence) is detected in the arterial wall of the obese subject. Nuclei are stained with DAPI (blue). Panel (<b>C</b>) shows representative immunoblots for the detection of NOX2 and corresponding β-actin in mesenteric artery homogenates from non-obese (control, <b>C</b>) and obese (Ob) subjects as well as the quantification of the expression assays. Data are expressed as the mean ± S.E.M. of NOX2 band intensities normalized by respective β-actin band intensities. * indicates <span class="html-italic">p</span> &lt; 0.05 vs. control by the Mann–Whitney U test. Panels (<b>D</b>–<b>F</b>) show linear regressions of NOX2 expression (NOX2/β-actin ratio) versus pEC<sub>50</sub> (<b>D</b>,<b>E</b>) or E<sub>max</sub> for BK (<b>F</b>) in subjects with morbid obesity. BK-induced responses for each subject were obtained in HMAs contracted with the thromboxane analogue, U46619 (10–30 nM) (<b>D</b>,<b>E</b>), or 25–35 mM K<sup>+</sup> (<b>F</b>). Data in panel (<b>D</b>) were obtained in HMAs without treatment, while HMAs were treated with indomethacin (INDO, 10 µM) plus N<sup>G</sup>-nitro-L-arginine methyl ester (L-NAME, 100 µM) in panel (<b>E</b>) and with INDO (10 µM) in panel (<b>F</b>). Panels (<b>G</b>,<b>H</b>) are representative images showing the detection of dihydroethidium (DHE) fluorescence (red) and DAPI staining (blue) in HMA sections from an obese subject with low (<b>G</b>) and high (H) DHE/DAPI fluorescence ratios. Magnification ×400. Panel (<b>I</b>) shows the linear regression of NOX2 expression in HMA homogenates versus superoxide generation determined by DHE/DAPI fluorescence in HMA sections from patients with morbid obesity, while panel (<b>J</b>) shows the linear regression of DHE/DAPI fluorescence versus pEC<sub>50</sub> for BK in subjects with morbid obesity. pEC<sub>50</sub> values for BK for each subject were obtained in HMAs contracted with U46619 and without further treatment. Coefficients of determination (r<sup>2</sup>) and probability (<span class="html-italic">p</span>) values are indicated in panels (<b>D</b>–<b>F</b>,<b>I</b>,<b>J</b>). Significant associations are highlighted in bold plus italic. <span class="html-italic">n</span> always indicates the number of subjects from whom the tissues were collected for determinations.</p>
Full article ">Figure 5
<p><b>The vascular expression of NADPH-oxidase-4 (NOX4) is not related to endothelial vasodilation in obese subjects.</b> Panel (<b>A</b>) shows representative immunoblots for the detection of NOX4 and corresponding β-actin in mesenteric artery homogenates from non-obese (control, C) and obese (Ob) subjects as well as the quantification of the expression assays. Data are expressed as the mean ± S.E.M. of NOX4 band intensities normalized by respective β-actin band intensities. <span class="html-italic">p</span> by the Mann–Whitney U test is indicated. Panels (<b>B</b>–<b>D</b>) show linear regressions of NOX4 expression (NOX4/β-actin ratio) versus pEC<sub>50</sub> (<b>B</b>,<b>C</b>) or E<sub>max</sub> for BK (<b>D</b>) in subjects with morbid obesity. BK-induced responses for each subject were obtained in HMAs contracted with the thromboxane analogue, U46619 (10–30 nM) (<b>B</b>,<b>C</b>), or 25–35 mM K<sup>+</sup> (<b>D</b>). Data in panel (<b>B</b>) were obtained in HMAs without treatment, while HMAs were treated with indomethacin (INDO, 10 µM) plus N<sup>G</sup>-nitro-L-arginine methyl ester (L-NAME, 100 µM) in panel (<b>C</b>) and with INDO (10 µM) in panel (<b>D</b>). Coefficients of determination (r<sup>2</sup>) and probability (<span class="html-italic">p</span>) values are indicated in panels (<b>B</b>–<b>D</b>). <span class="html-italic">n</span> always indicates the number of subjects from whom the tissues were collected for determinations.</p>
Full article ">Figure 6
<p><b>The endothelial vasodilation of human mesenteric arteries (HMAs) from obese subjects is additively impaired by high LDL-c/HDL-c ratio and high vascular expression of NADPH-oxidase-2 (NOX2).</b> Panel (<b>A</b>) shows vasodilations induced by bradykinin (BK, 1 nM to 10 µM) in HMAs contracted with the thromboxane analogue, U46619 (10–30 nM), obtained from obese subjects with low LDL-c/HDL-c ratio (≤2) and low vascular expression of NOX2 (below median) (low/low), from obese subjects with high LDL-c/HDL-c ratio (above 2) but low NOX2 expression (high LDL-c/HDL-c), from obese subjects with high vascular expression of NOX2 (above median) but low LDL-c/HDL-c ratio (high NOX2) and from obese subjects with both high LDL-c/HDL-c and NOX2 expression (high/high). Data are expressed as the mean ± S.E.M. of the percentage of relaxation. <span class="html-italic">n</span> indicates the number of subjects from whom the tissues were collected for determinations. ** indicates <span class="html-italic">p</span> &lt; 0.01, *** <span class="html-italic">p</span> &lt; 0.001 vs. low/low, ††† <span class="html-italic">p</span> &lt; 0.001 vs. high/high by two-factor ANOVA test corrected by Bonferroni’s test. Right panels show mean ± S.E.M. of pEC<sub>50</sub> (<b>B</b>) and E<sub>max</sub> (<b>C</b>) values for BK corresponding to each subject group. § <span class="html-italic">p</span> &lt; 0.05, §§ <span class="html-italic">p</span> &lt; 0.01 vs. low/low by Kruskal–Wallis followed by Dunn’s test.</p>
Full article ">
15 pages, 4932 KiB  
Article
The Impact of the Angiotensin-Converting Enzyme Inhibitor Lisinopril on Metabolic Rate in Drosophila melanogaster
by Denise Vecchie’, Julia M. Wolter, Jesse Perry, Patricia Jumbo-Lucioni and Maria De Luca
Int. J. Mol. Sci. 2024, 25(18), 10103; https://doi.org/10.3390/ijms251810103 - 20 Sep 2024
Viewed by 486
Abstract
Evidence suggests that angiotensin-converting enzyme inhibitors (ACEIs) may increase metabolic rate by promoting thermogenesis, potentially through enhanced fat oxidation and improved insulin. More research is, however, needed to understand this intricate process. In this study, we used 22 lines from the Drosophila Genetic [...] Read more.
Evidence suggests that angiotensin-converting enzyme inhibitors (ACEIs) may increase metabolic rate by promoting thermogenesis, potentially through enhanced fat oxidation and improved insulin. More research is, however, needed to understand this intricate process. In this study, we used 22 lines from the Drosophila Genetic Reference Panel to assess the metabolic rate of virgin female and male flies that were either fed a standard medium or received lisinopril for one week or five weeks. We demonstrated that lisinopril affects the whole-body metabolic rate in Drosophila melanogaster in a genotype-dependent manner. However, the effects of genotypes are highly context-dependent, being influenced by sex and age. Our findings also suggest that lisinopril may increase the Drosophila metabolic rate via the accumulation of a bradykinin-like peptide, which, in turn, enhances cold tolerance by upregulating Ucp4b and Ucp4c genes. Finally, we showed that knocking down Ance, the ortholog of mammalian ACE in Malpighian/renal tubules and the nervous system, leads to opposite changes in metabolic rate, and that the effect of lisinopril depends on Ance in these systems, but in a sex- and age-specific manner. In conclusion, our results regarding D. melanogaster support existing evidence of a connection between ACEI drugs and metabolic rate while offering new insights into this relationship. Full article
(This article belongs to the Special Issue Drosophila: A Versatile Model in Biology and Medicine)
Show Figures

Figure 1

Figure 1
<p>Dual role of angiotensin-converting enzyme (ACE). ACE converts Angiotensin I into Angiotensin II and degrades bradykinin. The drug lisinopril inhibits ACE, whereas losartan antagonizes the binding of Angiotensin II to its type-1 receptor.</p>
Full article ">Figure 2
<p>Changes in metabolic rate induced by lisinopril treatment in 22 DGRP lines. The response to lisinopril is influenced by genotype, sex, and age. Data report the means ± standard error of whole-body oxygen consumption rate adjusted for live body weight (<span class="html-italic">n</span> = 8–10 individual flies).</p>
Full article ">Figure 3
<p>Sensitivity of metabolic rate to lisinopril. (<b>a</b>,<b>b</b>) Values represent sensitivity index in DGRP young (panel (<b>a</b>)) and old (panel (<b>b</b>)) flies, averaged over sex. Sensitivity indices were calculated by taking the difference in the average oxygen consumption rate between lisinopril-treated and untreated flies of each genotype and dividing it by the average difference in treated and untreated flies across all genotypes. Lines are ranked by their sensitivity, with positive values indicating a higher metabolic rate with lisinopril, relative to untreated and negative values indicating a lower metabolic rate.</p>
Full article ">Figure 4
<p>Lisinopril does not alter the metabolic rate through changes in locomotor activity. (<b>a</b>) There is no significant effect of lisinopril on the climbing ability of DGRP lines displaying the greatest change in metabolic rate in response to lisinopril at one week of age. Data report climbing indices for untreated (white dots) and lisinopril-treated (red dots) flies averaged over both sexes and ages. (<b>b</b>) The effect of lisinopril on the climbing ability of DGRP lines displaying the greatest change in metabolic rate in response to lisinopril at five weeks of age depends on genotype and sex. **** <span class="html-italic">p</span> &lt; 0.0001 after post hoc Tukey’s tests for multiple comparisons. In both panels, error bars represent standard errors.</p>
Full article ">Figure 5
<p>Losartan affects metabolic rate in <span class="html-italic">D. melanogaster</span>. (<b>a</b>,<b>b</b>) DGRP lines exhibiting the greatest change in metabolic rate in response to lisinopril in young or old flies received losartan for one week (panel (<b>a</b>)) and five weeks (panel (<b>b</b>)), respectively. In both panels, data are averaged over both sexes and error bars represent standard errors. Significant comparisons within each DGRP line were determined by post hoc Tukey’s tests at <span class="html-italic">p</span> &lt; 0.05 and are indicated by different letters.</p>
Full article ">Figure 6
<p>Effects of lisinopril on acute cold tolerance and UCP genes. (<b>a</b>,<b>b</b>) Acute cold tolerance in lisinopril-treated and untreated DGRP lines exhibiting the greatest change in metabolic rate in response to lisinopril at one week (panel (<b>a</b>)) and five weeks (panel (<b>b</b>)) of age. Lisinopril reduces time to recover from chill-induced coma only in DGRP_367 young males (panel (<b>a</b>)) and DGRP_808 old males (panel (<b>b</b>)). (<b>c</b>) The expression of <span class="html-italic">Ucp4b</span> and <span class="html-italic">Ucp4c</span> genes is significantly increased in the head of DGRP_367 young male flies following lisinopril treatment. Transcript levels of each target gene were normalized to <span class="html-italic">rp49</span> and <span class="html-italic">α-tubulin</span>. In all panels, ** <span class="html-italic">p</span> &lt; 0.01 and **** <span class="html-italic">p</span> &lt; 0.0001 after post hoc Tukey’s tests for multiple comparisons. Error bars represent standard errors.</p>
Full article ">Figure 7
<p>The effect of lisinopril on whole-body metabolic rate depends on <span class="html-italic">Ance</span> in an organ system-, sex-, and age-specific manner. (<b>a</b>) Metabolic rate of flies with <span class="html-italic">Ance</span> knocked down in Malpighian tubules (<span class="html-italic">c42</span>-<span class="html-italic">Gal4</span> &gt; <span class="html-italic">Ance</span>-RNAi) and of their corresponding <span class="html-italic">w</span><sup>1118</sup> isogenic line (control). A four-way ANOVA revealed a significant genotype-by-treatment-by-sex interaction effect on oxygen consumption rate (F<sub>1,301</sub> = 47.40, <span class="html-italic">p</span> &lt; 0.0001). (<b>b</b>) Metabolic rate of flies with <span class="html-italic">Ance</span> knocked down in the nervous system (<span class="html-italic">elav</span>-<span class="html-italic">Gal80</span> &gt; <span class="html-italic">Ance</span>-RNAi) and of their corresponding <span class="html-italic">w</span><sup>1118</sup> isogenic line (control). A four-way ANOVA revealed a significant genotype-by-treatment-by-age interaction effect on oxygen consumption rate (F<sub>1,302</sub> = 5.65, <span class="html-italic">p</span> = 0.0180). In both panels, * <span class="html-italic">p</span> &lt; 0.05, ** <span class="html-italic">p</span> &lt; 0.01, *** <span class="html-italic">p</span> &lt; 0.001, and **** <span class="html-italic">p</span> &lt; 0.0001 after post hoc Tukey’s tests for multiple comparisons. Error bars represent standard errors.</p>
Full article ">Figure 8
<p>Lisinopril increases the whole-body metabolic rate of young flies with AD-like symptoms regardless of sex. A four-way ANOVA revealed a statistically significant effect of the genotype-by-treatment-by-age interaction term (F<sub>1,301</sub> = 3.92, <span class="html-italic">p</span> = 0.0486) on metabolic rate. *** <span class="html-italic">p</span> &lt; 0.001 after post hoc Tukey’s tests for multiple comparisons. Error bars represent standard errors.</p>
Full article ">
20 pages, 3630 KiB  
Article
Tissue Kallikrein-1 Suppresses Type I Interferon Responses and Reduces Depressive-Like Behavior in the MRL/lpr Lupus-Prone Mouse Model
by Priyanka S. Bhoj, Cassandra Nocito, Namdev S. Togre, Malika Winfield, Cody Lubinsky, Sabeeya Khan, Nikhita Mogadala, Alecia Seliga, Ellen M. Unterwald, Yuri Persidsky and Uma Sriram
Int. J. Mol. Sci. 2024, 25(18), 10080; https://doi.org/10.3390/ijms251810080 - 19 Sep 2024
Viewed by 723
Abstract
Excessive production and response to Type I interferons (IFNs) is a hallmark of systemic lupus erythematosus (SLE). Neuropsychiatric lupus (NPSLE) is a common manifestation of human SLE, with major depression as the most common presentation. Clinical studies have demonstrated that IFNα can cause [...] Read more.
Excessive production and response to Type I interferons (IFNs) is a hallmark of systemic lupus erythematosus (SLE). Neuropsychiatric lupus (NPSLE) is a common manifestation of human SLE, with major depression as the most common presentation. Clinical studies have demonstrated that IFNα can cause depressive symptoms. We have shown that the kallikrein–kinin system (KKS) [comprised of kallikreins (Klks) and bradykinins] and angiotensin-converting enzyme inhibitors suppressed Type I IFN responses in dendritic cells from lupus-prone mice and human peripheral blood mononuclear cells. Tissue Klk genes are decreased in patients with lupus, and giving exogenous Klk1 ameliorated kidney pathology in mice. We retro-orbitally administered mouse klk1 gene-carrying adenovirus in the Murphy Roths Large lymphoproliferative (MRL/lpr) lupus-prone mice at early disease onset and analyzed immune responses and depressive-like behavior. Klk1 improved depressive-like behavior, suppressed interferon-responsive genes and neuroinflammation, and reduced plasma IFNα levels and proinflammatory cytokines. Klk1 also reduced IFNAR1 and JAK1 protein expression, important upstream molecules in Type I IFN signaling. Klk1 reduced bradykinin B1 receptor expression, which is known to induce proinflammatory response. Together, these findings suggest that Klk1 may be a potential therapeutic candidate to control IFNα production/responses and other inflammatory responses in SLE and NPSLE. Full article
(This article belongs to the Topic Inflammation: The Cause of all Diseases 2.0)
Show Figures

Figure 1

Figure 1
<p>Administration of <span class="html-italic">Ad-klk1</span> increases Klk1 protein expression (<b>A</b>) Western blot of Klk1 protein (MW: 28 kDa) and (<b>B</b>) immunohistochemical results (indicated by pink staining) of Klk1 protein expression in the kidney of the spontaneous cohort. Images were taken at 40× magnification and analyzed using ImageJ (1.54f software). The results are expressed as mean ±SEM. An unpaired <span class="html-italic">t</span>-test was used; * <span class="html-italic">p</span> ≤ 0.05 and *** <span class="html-italic">p</span> ≤ 0.001 compared to the Ad-GFP group; <span class="html-italic">n =</span> 7–8 mice/group.</p>
Full article ">Figure 2
<p>Tissue Klk1 decreases IRGs and <span class="html-italic">hmgb1</span> gene expression in the brain and periphery: (<b>A</b>) Spontaneous MRL/lpr mice and (<b>B</b>) IFNα-induced MRL/lpr mice. The data, obtained by qPCR, were analyzed using an unpaired <span class="html-italic">t</span>-test and are expressed as the mean ±SEM of the fold change in gene expression in the Ad-Klk1 group compared to the Ad-GFP group (marked as a solid black line at 1.0 on the Y axis). * <span class="html-italic">p</span> ≤ 0.05 and ** <span class="html-italic">p</span> ≤ 0.01 compared to the Ad-GFP group; <span class="html-italic">n =</span> 7–8 mice/group in the spontaneous cohort, and <span class="html-italic">n =</span> 5 mice/group in the IFNα-induced cohort.</p>
Full article ">Figure 3
<p>Tissue Klk1 reduces inflammatory cytokine levels in the plasma: (<b>A</b>) Spontaneous MRL/lpr mice and (<b>B</b>) IFNα-induced MRL/lpr mice. Multiplex cytokine analysis was performed using the MSD ELISA kit. The data were analyzed using an unpaired <span class="html-italic">t</span>-test and are expressed as the mean ±SEM. * <span class="html-italic">p</span> ≤ 0.05 compared to the Ad-GFP group; <span class="html-italic">n =</span> 7–8 mice/group in the spontaneous cohort, and <span class="html-italic">n =</span> 5 mice/group in the IFNα-induced cohort.</p>
Full article ">Figure 4
<p>Tissue Klk1 decreases plasma IFNα levels and proteins in Type I IFN signaling: (<b>A</b>) Plasma IFNα levels in the spontaneous MRL/lpr mice were measured using the high sensitivity Verikine ELISA kit and are expressed as the mean ±SEM of IFNα levels (pg/mL). Western blot analysis of IFNAR1 (MW: 64 kDa) and JAK1 (MW: 80 kDa) protein levels in the (<b>B</b>) kidney and (<b>C</b>) spleen of spontaneous MRL/lpr mice. Data were analyzed using an unpaired <span class="html-italic">t</span>-test and are expressed as the mean ±SEM of the fold change in protein levels in the Ad-Klk1 group compared to the Ad-GFP group (marked as a solid black line at 1.0 on the Y axis). * <span class="html-italic">p</span> ≤ 0.05 and ** <span class="html-italic">p</span> ≤ 0.01 compared to the Ad-GFP group; <span class="html-italic">n =</span> 7–8 mice/group.</p>
Full article ">Figure 5
<p>Tissue Klk1 decreases depressive-like behavior and depression-related gene expression in the spontaneous MRL/lpr mice: (<b>A</b>) Tail suspension test (TST). (<b>B</b>) Plasma serotonin levels were estimated by using the ELISA kit. (<b>C</b>) Gene expression analysis of serotonin regulators in the brain; the results are expressed as the mean ±SEM of the fold change in gene expression in the Ad-Klk1 group compared to the Ad-GFP group (marked as a solid black line at 1.0 on the Y axis). (<b>D</b>) Correlation of the immobility time in the TST with plasma IFNα levels and serotonin-related markers (brain <span class="html-italic">isg15</span>, <span class="html-italic">5htt</span>, and <span class="html-italic">tph-2</span>). The data were analyzed using an unpaired <span class="html-italic">t</span>-test and * <span class="html-italic">p</span> ≤ 0.05 compared to the Ad-GFP group; <span class="html-italic">n =</span> 7–8 mice/group.</p>
Full article ">Figure 6
<p>Tissue Klk1 reduced microglial inflammatory state and number. Immunohistochemical staining of IBA-1 in microglia. Data were analyzed using an unpaired <span class="html-italic">t</span>-test and are expressed as the mean ±SEM of the percentage area of IBA-1 staining in the cortex region of the brain at 40× magnification (brown staining, indicated by yellow arrows) using ImageJ (1.54f software). * <span class="html-italic">p</span> ≤ 0.05 compared to Ad-GFP group; <span class="html-italic">n =</span> 7–8 mice/group.</p>
Full article ">Figure 7
<p>Tissue Klk1 decreases C3 and IgG protein levels in the kidney: Western blot analysis of C3 (MW: 50 kDa) and IgG (MW: 50 kDa) protein levels in the kidney of spontaneous MRL/lpr mice. Data were analyzed using an unpaired <span class="html-italic">t</span>-test and are expressed as the mean ±SEM of the fold change in protein levels in the Ad-Klk1 group compared to the Ad-GFP group (marked as a solid black line at 1.0 on the Y axis). * <span class="html-italic">p</span> ≤ 0.05 compared to the Ad-GFP group; <span class="html-italic">n =</span> 7–8 mice/group.</p>
Full article ">Figure 8
<p>Tissue Klk1 alters BK receptor expression in the spontaneous MRL/lpr mice: (<b>A</b>) Fold change in <span class="html-italic">b1r</span> and <span class="html-italic">b2r</span> gene expression in the brain of the Ad-Klk1 group compared to those of the Ad-GFP group (marked as a solid black line at 1.0 on the Y axis), as measured by qPCR. Fold change in B1R (MW: 40 kDa) and B2R (MW: 50 kDa) protein levels in the (<b>B</b>) kidneys and (<b>C</b>) spleen of the Ad-Klk1 group compared to those of the Ad-GFP group (marked as a solid black line at 1.0 on the Y axis), as measured by western blot. The results are expressed as mean ±SEM. An unpaired <span class="html-italic">t</span>-test was used; * <span class="html-italic">p</span> ≤ 0.05 and <span class="html-italic">p</span> ≤ 0.001 compared to the Ad-GFP group; <span class="html-italic">n =</span> 7–8 mice/group.</p>
Full article ">Figure 9
<p>Scheme of the experiment: One cohort of nine-week-old MRL/lpr mice was administered with IFNα- or saline-carrying osmotic pump (IFNα-induced response) on day −2, followed by the retro-orbital administration of <span class="html-italic">Ad-klk1</span> or <span class="html-italic">Ad-gfp</span> vector on day 0. The spontaneous cohort received <span class="html-italic">Ad-klk1</span> or control (<span class="html-italic">Ad-gfp</span>) vector only. Behavioral studies were performed on day 15, and the mice were sacrificed on day 17 to collect the brain, spleen, kidneys, and blood.</p>
Full article ">Figure 10
<p>Tissue Klk1 suppresses Type I IFN responses through IFNAR1, BRs, and PAR2 and reduces depressive-like behavior in the MRL/lpr lupus-prone mice: The present study demonstrates that tissue Klk1 administration reduces IFNα, IFNAR1, JAK1, and IRF7, consequently suppressing IRG expression. Our previous study [<a href="#B11-ijms-25-10080" class="html-bibr">11</a>] has shown a reduction in signal transducer and activator of transcription (STAT) phosphorylation (shown in blue color) following bradykinin (BK) administration, which, along with other pathway proteins (shown in grey color), will be studied in detail in the future. We observed a consequential reduction in the levels of tryptophan hydroxylase 2 (TPH2, a rate-limiting enzyme in serotonin biosynthesis), serotonin (5-HT), serotonin transporter (5-HTT), and 5-hydroxytryptamine receptor 2A (HTR2A, a serotonin receptor). We also observed depressive-like behaviors in MRL/lpr mice. Alleviation of inflammatory and IFN responses may be mediated via the alteration in BK receptors (B1R and B2R) and protease-activated receptor-2 (PAR2) expression; the mechanisms need to be studied in detail (indicated with dashed arrows). The therapeutic effects of Klk1 are shown in green notation, and SLE and NPSLE pathogenesis are represented in red notation.</p>
Full article ">
12 pages, 1103 KiB  
Review
Escin’s Action on Bradykinin Pathway: Advantageous Clinical Properties for an Unknown Mechanism?
by Gianmarco Marcianò, Cristina Vocca, Demirhan Dıraçoğlu, Rotinda Özdaş Sevgin and Luca Gallelli
Antioxidants 2024, 13(9), 1130; https://doi.org/10.3390/antiox13091130 - 19 Sep 2024
Viewed by 612
Abstract
Escin, extracted from horse chestnut (Aesculus hippocastanum) has anti-edema and anti-inflammatory effects. It is used to treat several clinical conditions, including venous insufficiency, pain, inflammation, and edema. Considering escin’s pharmacodynamic, the inhibition of the bradykinin pathway represents a particular effect, decreasing [...] Read more.
Escin, extracted from horse chestnut (Aesculus hippocastanum) has anti-edema and anti-inflammatory effects. It is used to treat several clinical conditions, including venous insufficiency, pain, inflammation, and edema. Considering escin’s pharmacodynamic, the inhibition of the bradykinin pathway represents a particular effect, decreasing the local edema and conferring an advantage in comparison to other compounds. In this narrative review, we described the effects of escin considering its effects on bradykinin pathway. Full article
Show Figures

Figure 1

Figure 1
<p>Escin antioxidant, anti-inflammatory, anti-edematous, and venotonic properties. Escin exerts its antioxidant effect, inhibiting several mediators or processes (minus sign), but promoting other molecules like PECAM-1 (plus sign) and general endothelial function. The increase in venous tone has been observed in different models and (PG)F2α is a possible mediator of this process. Inflammation is reversed by acting on the vascular level, but also on other mediators, cytokines, and transcriptional factors. Interestingly, escin displayed a glucocorticoid-like activity. AP-1, activator protein 1; NF-κB, nuclear factor kappa-light-chain-enhancer of activated B cells; NO<sup>•</sup>, nitric oxide; PECAM-1, platelet endothelial cell adhesion molecule-1; PG, prostaglandin; PL, phospholipase; TNF, Tumor Necrosis Factor.</p>
Full article ">Figure 2
<p>Escin’s action on the bradykinin pathway. Poor evidence exists about the precise mechanism of escin’s action on the bradykinin pathway. Three possible hypotheses may be related to (1) the action on glucocorticoid receptor, that in some models has shown to reduce the activity of bradykinin; (2) NO<sup>•</sup> inhibition, an indirect mechanism, since bradykinin may raise NO<sup>•</sup> concentration; (3) a possible action on coagulation cascade, since <span class="html-italic">Aesculus Ippocastanum</span> has shown anti-coagulant properties, but involving mainly the bark and not the seeds or seed shell. In the figure, we observe the conversion of prekallikrein in kallikrein by XIIa and then the production of bradykinin by kallikrein through its catalytic action on HK. GR, glucocorticoid receptor; HK, high-molecular-weight kininogen; NO<sup>•</sup>, nitric oxide; XII, activated twelve factor.</p>
Full article ">
19 pages, 5369 KiB  
Article
What Are the Key Factors for the Detection of Peptides Using Mass Spectrometry on Boron-Doped Diamond Surfaces?
by Juvissan Aguedo, Marian Vojs, Martin Vrška, Marek Nemcovic, Zuzana Pakanova, Katerina Aubrechtova Dragounova, Oleksandr Romanyuk, Alexander Kromka, Marian Varga, Michal Hatala, Marian Marton and Jan Tkac
Nanomaterials 2024, 14(15), 1241; https://doi.org/10.3390/nano14151241 - 24 Jul 2024
Viewed by 1064
Abstract
We investigated the use of boron-doped diamond (BDD) with different surface morphologies for the enhanced detection of nine different peptides by matrix-assisted laser desorption/ionisation mass spectrometry (MALDI-MS). For the first time, we compared three different nanostructured BDD film morphologies (Continuous, Nanograss, and Nanotips) [...] Read more.
We investigated the use of boron-doped diamond (BDD) with different surface morphologies for the enhanced detection of nine different peptides by matrix-assisted laser desorption/ionisation mass spectrometry (MALDI-MS). For the first time, we compared three different nanostructured BDD film morphologies (Continuous, Nanograss, and Nanotips) with differently terminated surfaces (-H, -O, and -F) to commercially available Ground Steel plates. All these surfaces were evaluated for their effectiveness in detecting the nine different peptides by MALDI-MS. Our results demonstrated that certain nanostructured BDD surfaces exhibited superior performance for the detection of especially hydrophobic peptides (e.g., bradykinin 1–7, substance P, and the renin substrate), with a limit of detection of down to 2.3 pM. Further investigation showed that hydrophobic peptides (e.g., bradykinin 1–7, substance P, and the renin substrate) were effectively detected on hydrogen-terminated BDD surfaces. On the other hand, the highly acidic negatively charged peptide adrenocorticotropic hormone fragment 18–39 was effectively identified on oxygen-/fluorine-terminated BDD surfaces. Furthermore, BDD surfaces reduced sodium adduct contamination significantly. Full article
(This article belongs to the Special Issue Carbon-Based Nanomaterials for Biomedicine Applications)
Show Figures

Figure 1

Figure 1
<p>SEM images (top, angular, and cross-sectional views) of BDD morphologies for Continuous film and structured Nanograss or Nanotips.</p>
Full article ">Figure 2
<p>Raman spectra of the as-grown (Continuous) and structured (Nanograss and Nanotip) BDD films.</p>
Full article ">Figure 3
<p>Wetting properties of as-grown (Continuous) and structured (Nanotip) BDD samples after different surface termination treatments.</p>
Full article ">Figure 4
<p>XPS survey spectra of (<b>a</b>) Continuous BDD, (<b>b</b>) Nanograss BDD, and (<b>c</b>) Nanotip BDD samples as-prepared (black) and with -H (red), -O (green), and -F (blue) terminations.</p>
Full article ">Figure 5
<p>Photograph of prepared MALDI-MS chip with BDD spots deposited.</p>
Full article ">Figure 6
<p>MALDI-MS workflow and MS spectra of the standard peptide mixture solution (0.2 pmol μL<sup>−1</sup>) performed in positive mode using the DHB matrix in all the cases: (<b>A</b>) Ground Steel, (<b>B</b>) Continuous BDD, and (<b>C</b>) Nanograss BDD.</p>
Full article ">Figure 7
<p>Classes of amino acids present in the peptides expressed in percentages (%). Brad: Bradykinin 1–7; Ang II: Angiotensin II; Ang I: Angiotensin I; Sub P: Substance P; Bomb: Bombesin; Ren S: Renin Substrate; AC 1: adrenocorticotropic hormone fragment 1–17; AC 18: adrenocorticotropic hormone fragment 18–39; Som: Somatostatin. Peptides highlighted in green showed significantly better LODs when detected at BDD interfaces in comparison to the commercially available Ground Steel chip. This figure was made based on data taken from <a href="#app1-nanomaterials-14-01241" class="html-app">Table S1</a>. The dashed green line represents the average percentages of hydrophobic amino acids present in Brad, Sub P, and Ren S.</p>
Full article ">Figure 8
<p>Three-dimensional (3D) plots showing the dependences of the LOD<sub>ratio</sub> value on the <span class="html-italic">m</span>/<span class="html-italic">z</span> and charge.</p>
Full article ">Figure 9
<p>The influence of work function (WF) on LOD<sub>ratio</sub> values for the Sub P peptide. The data point marked by the red circle was not a part of the linear plot.</p>
Full article ">Scheme 1
<p>Preparation of BDD interfaces.</p>
Full article ">
14 pages, 1787 KiB  
Article
Itch and Pain Behaviors in Irritant Contact Dermatitis Produced by Sodium Lauryl Sulfate in Mice
by Nathalie M. Malewicz-Oeck, Zhe Zhang, Steven G. Shimada and Robert H. LaMotte
Int. J. Mol. Sci. 2024, 25(14), 7718; https://doi.org/10.3390/ijms25147718 - 14 Jul 2024
Viewed by 891
Abstract
Irritant contact dermatitis (ICD) is a nonspecific skin inflammation caused by irritants, leading to itch and pain. We tested whether differential responses to histamine-dependent and -independent pruritogens can be evoked in ICD induced by sodium lauryl sulfate (SLS). An ICD mouse model was [...] Read more.
Irritant contact dermatitis (ICD) is a nonspecific skin inflammation caused by irritants, leading to itch and pain. We tested whether differential responses to histamine-dependent and -independent pruritogens can be evoked in ICD induced by sodium lauryl sulfate (SLS). An ICD mouse model was established with 5% SLS in acetone versus a vehicle topically applied for 24 h to the cheek. Site-directed itch- and pain-like behaviors, occurring spontaneously and in response to mechanical, thermal, and chemical stimuli (histamine, ß-alanine, BAM8-22, and bradykinin) applied to the cheek, were recorded before (day 0) and after irritant removal (days 1, 2, 3, and 4). Skin inflammation was assessed through visual scoring, ultrasound, and measurements of skin thickness. SLS-treated mice exhibited hyperalgesia-like behavior in response to mechanical and heat stimuli on day 1 compared to the controls. SLS mice exhibited more spontaneous wipes (pain) but not scratching bouts (itch) on day 1. Pruritogen injections caused more scratching but not wiping in SLS-treated mice compared to the controls. Only bradykinin increased wiping behavior compared to saline. SLS-treated mice developed noticeable erythema, scaling, and increased skin thickness on days 1 and 2. SLS induced cutaneous inflammation and behavioral signs of spontaneous pain and itching, hyperalgesia to mechanical and heat stimuli and a chemical algogen, and enhanced itch response to pruritogens. These sensory reactions preceded the inflammation peak and lasted up to two days. Full article
(This article belongs to the Section Molecular Immunology)
Show Figures

Graphical abstract

Graphical abstract
Full article ">Figure 1
<p>Effects of SLS-induced ICD on DS derived in response to vFFs and a heated contact thermode placed to the cheek. Stimuli were applied before (day 0) and 1 h (day 1), 24 h (day 2), 48 h (day 3), and 72 h (day 4) after SLS or vehicle application. The mean DSs are shown in black for SLS and grey for the control group, respectively. (<b>a</b>) Mean DSs ± SDs for an innocuous vFF of 0.23 mN and mean DSs for a noxious vFF with a 100 µm tip diameter applied with three different bending forces (2, 10, and 20 mN). (<b>b</b>) Mean DSs ± SDs for innocuous (38 °C) and noxious t (52 °C) temperature stimuli. Significant differences before vs. after application for 24 h within groups are marked as follows: * <span class="html-italic">p</span> &lt; 0.05, ** <span class="html-italic">p</span> &lt; 0.01. Significant differences between the groups are marked as follows: † <span class="html-italic">p</span> &lt; 0.05, †† <span class="html-italic">p</span> &lt; 0.01. Data: means ± SDs. N = 12 mice/group.</p>
Full article ">Figure 2
<p>Effects of SLS-induced ICD on spontaneous scratching and wiping behaviors. Spontaneous scratching bouts (<b>a</b>) and wipes (<b>b</b>) were counted 1 h (day 1) and 24 h (day 2) after removal of the distilled water (control group) or 5% SLS (SLS group). * <span class="html-italic">p</span> &lt; 0.05 indicates significance between groups. Data: means ± SEMs. N = 12 mice/group.</p>
Full article ">Figure 3
<p>Effects of SLS on scratching and wiping evoked by intradermal injection of saline, bradykinin, and each pruritogen. Chemicals were injected into cheek skin after a 24 h treatment with either distilled water (control group) or 5% SLS (SLS group). (<b>a</b>,<b>b</b>) Mean numbers of scratching bouts (<b>a</b>) and wipes (<b>b</b>) evoked by the intradermal injection of each chemical into skin of the cheek previously treated with distilled water. (<b>c</b>,<b>d</b>) Mean numbers of scratching bouts (<b>c</b>) and wipes (<b>d</b>) evoked by an injection (intradermally) of each chemical into the cheek skin area previously treated with 5% SLS. (<b>e</b>,<b>f</b>) Comparison of the effects of SLS vs. distilled water on the mean numbers of scratching bouts (<b>e</b>) and wipes (<b>f</b>) evoked by single chemical injection. Significant differences for comparison with saline are marked as follows: * <span class="html-italic">p</span> &lt; 0.05, ** <span class="html-italic">p</span> &lt; 0.01. Significant differences in panels (<b>e</b>,<b>f</b>) for comparison between the control group and the SLS group are marked as follows: ** <span class="html-italic">p</span> &lt; 0.01. Error bars: SEMs. N = 12 mice/group.</p>
Full article ">Figure 4
<p>Effects of SLS-induced ICD on the severity of inflammation of cheek skin. (<b>a</b>) Exemplary photographs of the skin of the cheek for each experimental condition and group (SLS group vs. control group) before and each day after removing the PEEK cup showing a noticeable change in erythema and scaling. Erythema (<b>b</b>) and scaling (<b>c</b>) were separately evaluated on a scale from 0 to 4, corresponding to “none” to “very marked”. The thickness of a fold of cheek skin (<b>d</b>) was obtained with a micrometer. Significant differences for comparison between days are marked as follows: * <span class="html-italic">p</span> &lt; 0.05, ** <span class="html-italic">p</span> &lt; 0.01. Significant differences between the groups (SLS versus control) are marked as follows: † <span class="html-italic">p</span> &lt; 0.05, †† <span class="html-italic">p</span> &lt; 0.01. Error bars: SEMs. N = 12 mice/group.</p>
Full article ">Figure 5
<p>Effects of SLS-induced ICD on cheek skin assessed with ultrasound. (<b>a</b>) Exemplary ultrasound skin images (<b>a</b>) were used to measure the changes in different skin layers in SLS and control groups over time (<b>b</b>). Skin thickness measurements obtained from ultrasound images (<b>c</b>) and thickness measurements of different layers of the epidermis (<b>d</b>), dermis (<b>e</b>), and hypodermis (<b>f</b>). Significant differences are marked as follows: * <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; 0.0001. N = 10–12 mice/group.</p>
Full article ">Figure 6
<p>The schematic experimental schedule for the establishment and characterization of the ICD model. For acquisition of ultrasound images, a different group of mice were briefly anesthetized. Arrows symbolize days of testing.</p>
Full article ">
14 pages, 1964 KiB  
Article
Assessment of Radiolabelled Derivatives of R954 for Detection of Bradykinin B1 Receptor in Cancer Cells: Studies on Glioblastoma Xenografts in Mice
by Miho Shukuri, Satoru Onoe, Tsubasa Karube, Risa Mokudai, Hayate Wakui, Haruka Asano, Shin Murai and Hiromichi Akizawa
Pharmaceuticals 2024, 17(7), 902; https://doi.org/10.3390/ph17070902 - 7 Jul 2024
Viewed by 726
Abstract
Bradykinin B1 receptor (B1R) has garnered attention as a cancer therapeutic and diagnostic target. Several reports on radiolabelled derivatives of B1R antagonists have shown favourable properties as imaging agents in cells highly expressing hB1R following transfection. In the present study, we assessed whether [...] Read more.
Bradykinin B1 receptor (B1R) has garnered attention as a cancer therapeutic and diagnostic target. Several reports on radiolabelled derivatives of B1R antagonists have shown favourable properties as imaging agents in cells highly expressing hB1R following transfection. In the present study, we assessed whether radiolabelled probes can detect B1R endogenously expressed in cancer cells. To this end, we evaluated 111In-labelled derivatives of a B1R antagonist ([111In]In-DOTA-Ahx-R954) using glioblastoma cell lines (U87MG and U251MG) with different B1R expression levels. Cellular uptake studies showed that the specific accumulation of [111In]In-DOTA-Ahx-R954 in U87MG was higher than that in U251MG, which correlated with B1R expression levels. Tissue distribution in U87MG-bearing mice revealed approximately 2-fold higher radioactivity in tumours than in the muscle in the contralateral leg. The specific accumulation of [111In]In-DOTA-Ahx-R954 in the tumour was demonstrated by the reduction in the tumour-to-plasma ratios in nonlabelled R954-treated mice. Moreover, ex vivo autoradiographic images revealed that the intratumoural distribution of [111In]In-DOTA-Ahx-R954 correlated with the localisation of B1R-expressing glioblastoma cells. In conclusion, we demonstrated that [111In]In-DOTA-Ahx-R954 radioactivity correlated with B1R expression in glioblastoma cells, indicating that radiolabelled derivatives of the B1R antagonist could serve as promising tools for elucidating the involvement of B1R in cancer. Full article
Show Figures

Figure 1

Figure 1
<p>Chemical structure of [<sup>111</sup>In]In-DOTA-Ahx-R954.</p>
Full article ">Figure 2
<p>Specific accumulation of [<sup>111</sup>In]In-DOTA-Ahx-R954 in cultured U87MG and U251MG cells. (<b>A</b>) Immunoblotting of B1R; (<b>B</b>) Time-dependent changes in cell accumulated radioactivity following incubation with [<sup>111</sup>In]In-DOTA-Ahx-R954. Data are expressed as the mean ± SD (30 min in U87MG, <span class="html-italic">n</span> = 3, and in others, <span class="html-italic">n</span> = 4). <span class="html-italic">p</span> &lt; 0.001 at all time points except 0 min, U87MG vs. U251MG; (<b>C</b>) Radioactivity of [<sup>111</sup>In]In-DOTA-Ahx-R954 in cells in the absence (control) or presence of 100 µM R954 (+R954) or [des-Arg<sup>10</sup>]-kallidin (+[des-Arg<sup>10</sup>]-kallidin). Data are expressed as mean ± SD (<span class="html-italic">n</span> = 4). * <span class="html-italic">p</span> &lt; 0.01, ** <span class="html-italic">p</span> &lt; 0.001, vs. control group (assay performed in the absence of antagonists and agonists).</p>
Full article ">Figure 3
<p>Binding of increasing concentrations of [<sup>111</sup>In]In-DOTA-Ahx-R954 to cultured U87MG and U251MG cells. Binding of [<sup>111</sup>In]In-DOTA-Ahx-R954 to U87MG (<b>A</b>) and U251MG (<b>B</b>) cells in the absence (total binding) or presence (nonspecific binding; NSB) of an excess (100 μM) of R954. Specific binding was obtained by subtracting NSB from total binding. Data are expressed as the mean ± SD (<span class="html-italic">n</span> = 4).</p>
Full article ">Figure 4
<p>Intratumoural distribution of [<sup>111</sup>In]In-DOTA-Ahx-R954 and [<sup>14</sup>C]iodoantipyrine. (<b>A</b>) Representative ex vivo autoradiographs of [<sup>111</sup>In]In-DOTA-Ahx-R954 (60 min post-injection) and [<sup>14</sup>C]iodoantipyrine (1 min post-injection) obtained from the same slide; (<b>B</b>) Radioactivity data (tumour-to-muscle ratio) in the ex vivo autoradiographs are expressed as mean ± SD (in [<sup>111</sup>In]In-DOTA-Ahx-R954, <span class="html-italic">n</span> = 5, and in [<sup>14</sup>C]iodoantipyrine, <span class="html-italic">n</span> = 4).</p>
Full article ">Figure 5
<p>B1R expression and intratumoural distribution of GFAP-positive glioblastoma cells. (<b>A</b>) Representative photomicrographs of double immunofluorescent labelling for B1R (green) and GFAP (red) in U87MG tumour tissue; (<b>B</b>) Representative macro image of GFAP immunoreactivity detected using an Amersham™ Typhoon™ scanner; (<b>C</b>) Correlation between the radioactivity of [<sup>111</sup>In]In-DOTA-Ahx-R954 and the fluorescence intensity of GFAP. The analysis was evaluated using the value of the tumour-to-muscle ratio. The correlation coefficient was 0.96.</p>
Full article ">
11 pages, 936 KiB  
Opinion
Reducing Brain Edema Using Berotralstat, an Inhibitor of Bradykinin, Repurposed as Treatment Adjunct in Glioblastoma
by Richard E. Kast
Neuroglia 2024, 5(3), 223-233; https://doi.org/10.3390/neuroglia5030016 - 2 Jul 2024
Viewed by 935
Abstract
Glioblastomas synthesize, bear receptors for, and respond to bradykinin, triggering migration and proliferation. Since centrifugal migration into uninvolved surrounding brain tissue occurs early in the course of glioblastoma, this attribute defeats local treatment attempts and is the primary reason current treatments almost always [...] Read more.
Glioblastomas synthesize, bear receptors for, and respond to bradykinin, triggering migration and proliferation. Since centrifugal migration into uninvolved surrounding brain tissue occurs early in the course of glioblastoma, this attribute defeats local treatment attempts and is the primary reason current treatments almost always fail. Stopping bradykinin-triggered migration would be a step closer to control of this disease. The recent approval and marketing of an oral plasma kallikrein inhibitor, berotralstat (Orladeyo™), and pending FDA approval of a similar drug, sebetralstat, now offers a potential method for reducing local bradykinin production at sites of bradykinin-mediated glioblastoma migration. Both drugs are approved for treating hereditary angioedema. They are ideal for repurposing as a treatment adjunct in glioblastoma. Furthermore, it has been established that peritumoral edema, a common problem during the clinical course of glioblastoma, is generated in large part by locally produced bradykinin via kallikrein action. Both brain edema and the consequent use of corticosteroids both shorten survival in glioblastoma. Therefore, by (i) migration inhibition, (ii) growth inhibition, (iii) edema reduction, and (iv) the potential for less use of corticosteroids, berotralstat may be of service in treatment of glioblastoma, slowing disease progression. This paper recounts the details and past research on bradykinin in glioblastoma and the rationale of treating it with berotralstat. Full article
Show Figures

Figure 1

Figure 1
<p>A simplified schematic of the kallikrein–bradykinin system taken from Lima et al. [<a href="#B10-neuroglia-05-00016" class="html-bibr">10</a>] showing locus of action of four FDA/EMA-approved drugs to inhibit bradykinin functions, berotralstat, ecallantide, icatibant, and sebetralstat, all FDA approved to treat HAE. Concentrated human plasma and recombinant C-1 esterase are also FDA approved and marketed for this. The amplifying feedback cycle between coagulation factor XII and prekallikrein is shown in red rectangle. Blue oval represents a glioma cell. Not shown in this schematic is the blood–brain barrier-opening effect of excess bradykinin signaling associated with and common in GB.</p>
Full article ">Figure 2
<p>Schematic showing simplified locus of action of how adding celecoxib and/or dapsone has potential to augment berotralstat, lowering edema further. The bradykinin-amplifying feedback loop is depicted, where bradykinin agonism at B2R increases B2R expression, forming an amplifying loop as long as enough bradykinin is present. Bradykinin’s T1/2 is &lt;40 s [<a href="#B33-neuroglia-05-00016" class="html-bibr">33</a>]. This means that this amplification cycle is rapidly shut off by berotralstat or any plasma kallikrein inhibitor that would stop bradykinin release. B2R, the bradykinin receptor 2; HMWK, high molecular weight kininogen; the dashed red arrow indicates what would happen if plasma kallikrein were not inhibited.</p>
Full article ">
19 pages, 2297 KiB  
Article
Increased Prolylcarboxypeptidase Expression Can Serve as a Biomarker of Senescence in Culture
by Nicholas Glen Boullard, Jason J. Paris, Zia Shariat-Madar and Fakhri Mahdi
Molecules 2024, 29(10), 2219; https://doi.org/10.3390/molecules29102219 - 9 May 2024
Viewed by 1022
Abstract
Prolylcarboxypeptidase (PRCP, PCP, Lysosomal Pro-X-carboxypeptidase, Angiotensinase C) controls angiotensin- and kinin-induced cell signaling. Elevation of PRCP appears to be activated in chronic inflammatory diseases [cardiovascular disease (CVD), diabetes] in proportion to severity. Vascular endothelial cell senescence and mitochondrial dysfunction have consistently been shown [...] Read more.
Prolylcarboxypeptidase (PRCP, PCP, Lysosomal Pro-X-carboxypeptidase, Angiotensinase C) controls angiotensin- and kinin-induced cell signaling. Elevation of PRCP appears to be activated in chronic inflammatory diseases [cardiovascular disease (CVD), diabetes] in proportion to severity. Vascular endothelial cell senescence and mitochondrial dysfunction have consistently been shown in models of CVD in aging. Cellular senescence, a driver of age-related dysfunction, can differentially alter the expression of lysosomal enzymes due to lysosomal membrane permeability. There is a lack of data demonstrating the effect of age-related dysfunction on the expression and function of PRCP. To explore the changes in PRCP, the PRCP-dependent prekallikrein (PK) pathway was characterized in early- and late-passage human pulmonary artery endothelial cells (HPAECs). Detailed kinetic analysis of cells treated with high molecular weight kininogen (HK), a precursor of bradykinin (BK), and PK revealed a mechanism by which senescent HPAECs activate the generation of kallikrein upon the assembly of the HK–PK complex on HPAECs in parallel with an upregulation of PRCP and endothelial nitric oxide (NO) synthase (eNOS) and NO formation. The NO production and expression of both PRCP and eNOS increased in early-passage HPAECs and decreased in late-passage HPAECs. Low activity of PRCP in late-passage HPAECs was associated with rapid decreased telomerase reverse transcriptase mRNA levels. We also found that, with an increase in the passage number of HPAECs, reduced PRCP altered the respiration rate. These results indicated that aging dysregulates PRCP protein expression, and further studies will shed light into the complexity of the PRCP-dependent signaling pathway in aging. Full article
Show Figures

Graphical abstract

Graphical abstract
Full article ">Figure 1
<p>Plasma kallikrein activity in various HPAE cell passages. A total of 40,000 HPAE cells/well were cultured in a 96-well plate for a total of 17 to 18 h. An amount of 20 nM PK was incubated with 20 nM HK bound to HPAE cells at 37 °C. The liberation of paranitroaniline (pNA) from chromogenic substrate (S2302) by kallikrein was measured at 405 nm. Inset, endothelial (HPACs) PRCP activates PK (a zymogen) to kallikrein, which leads to the release of paranitroaniline (pNA) from S2302. All values are expressed as mean + SEM of triplicate points of and represent at least three independent experiments.</p>
Full article ">Figure 2
<p>PRCP expression is age-dependent. (<b>A</b>), Representative Western blots of PRCP protein expression of whole-cell lysates from both working passages and late passages (P2–P30). (<b>B</b>), Densitometric analysis of relative density from three independent Western blot experiments. Bars display mean + SEM. (<b>C</b>), Densitometric analysis of RT-PCR of PRCP in cultured endothelial cells from HPAECs of both working and late passages of three independent total RNA preparations. Total RNA was isolated from HPAECs and then amplified by RT-PCR. Amplified DNA (111 bp fragment) was resolved on 1.5% agarose gel. Values are expressed as mean + SEM of triplicate points of 3 independent experiments.</p>
Full article ">Figure 3
<p>eNOS plays a role in maintaining endothelial integrity during aging. (<b>A</b>), NO production by HPAECs. The kallikrein–kinin generating pathway induces nitric oxide (NO) generation in a time-dependent manner. Both working passages (P9, P10, P17, P18)- and late passages (P25, P26, P29, P30, P40)- HPAE cells were incubated with 300 nM HK and 300 nM PK for 17 h at 37 °C. The solution was collected to measure the amount of nitrate+nitrite (the final products of NO metabolism) in each sample using a fluorometric assay. Data are presented as mean + SEM (<span class="html-italic">n</span> = 3). (<b>B</b>), Densitometric analysis of RT-PCR of eNOS mRNA in cultured endothelial cells from HPAECs of both working and late passages of three independent total RNA preparations. Values are expressed as mean ± SEM of three independent experiments.</p>
Full article ">Figure 4
<p>Senescence and telomere length. (<b>A</b>), Representative photographs of β-gal-positive staining cells. The cells were examined on a light microscope with an attached camera using a 10× objective. The β-gal-positive cells were blue. Representative images were used for figures. Positive staining cells were determined (<span class="html-italic">n</span> = 3). (<b>B</b>), The relative levels of β-gal-positive staining in various HPAEC passages. The fraction of β-gal-positive cells was determined by counting the number of blue cells in 5 non-overlapping images from various fields of view within each well. The β-gal-positive (blue) cell count was divided by the number of total cell count to calculate the percentage senescent cells. Graph shows mean data ± SEM from three experiments, with at least 500 cells counted per condition per experiment. (<b>C</b>), Reduced transcription of telomerase (hTERT) mRNA. HPAECs were continuously passaged, and hTERT mRNA was determined. The decrease in telomerase was accompanied by an increase in PRCP at working passages (P5–P20). (<b>D</b>), The effect of aging on the level of FGF-2 expression. HPAECs were passaged, and FGF-2 expression was determined. Each bar displays mean data + SEM from three experiments.</p>
Full article ">Figure 5
<p>UM8190 treatment decreases oxygen consumption. (<b>A</b>), UM8190 structure. (<b>B</b>), Oxygen consumption of HPAECs as a result of mitochondrial activity was plotted in the absence of UM8190. Representative trace served as control. (<b>C</b>), After a basal measurement of ORC, measurement of OCR for HPAECs (5 × 10<sup>5</sup> cells/mL) followed by the sequential addition of UM8190 (10–150 μM) with a measurement of OCR as indicated by arrows. OCRs were measured three times and plotted as a function of time. (<b>D</b>), The resulting UM8190 effects on OCR are plotted as a percentage of control. Data shown are the mean ± SEM. N = 3 of three independent experiments.</p>
Full article ">Figure 6
<p>Effect of UM8190 on the respiration of digitonin-permeabilized HPAE cells. (<b>A</b>), Effect of UM8190 on Complex I. (<b>B</b>), Oxygen consumption rate as a percentage of control. (<b>C</b>), The inhibitory effect of UM8190 on Complex II. (<b>D</b>), Effect of UM8190 on cytochrome C oxidase. (<b>E</b>), A schematic diagram summarizes the effects of UM8190 on complexes of mitochondrial respiratory chain. Data are representative traces from three independent experiments.</p>
Full article ">Figure 7
<p>UM8190 compound inhibits mitochondrial Complex I ROS production. All the values are expressed as mean standard error of the mean of at least three independent experiments. Inset, a schematic diagram showing the proposed mechanism by which UM8190 exerts its effects on the HPAECs including the kallikrein–kinin forming pathway and mitochondrial complex I.</p>
Full article ">
11 pages, 2138 KiB  
Article
Dabsylated Bradykinin Is Cleaved by Snake Venom Proteases from Echis ocellatus
by Julius Abiola, Anna Maria Berg, Olapeju Aiyelaagbe, Akindele Adeyi and Simone König
Biomedicines 2024, 12(5), 1027; https://doi.org/10.3390/biomedicines12051027 - 7 May 2024
Viewed by 1018
Abstract
The vasoactive peptide bradykinin (BK) is an important member of the renin–angiotensin system. Its discovery is tightly interwoven with snake venom research, because it was first detected in plasma following the addition of viper venom. While the fact that venoms liberate BK from [...] Read more.
The vasoactive peptide bradykinin (BK) is an important member of the renin–angiotensin system. Its discovery is tightly interwoven with snake venom research, because it was first detected in plasma following the addition of viper venom. While the fact that venoms liberate BK from a serum globulin fraction is well described, its destruction by the venom has largely gone unnoticed. Here, BK was found to be cleaved by snake venom metalloproteinases in the venom of Echis ocellatus, one of the deadliest snakes, which degraded its dabsylated form (DBK) in a few minutes after Pro7 (RPPGFSP↓FR). This is a common cleavage site for several mammalian proteases such as ACE, but is not typical for matrix metalloproteinases. Residual protease activity < 5% after addition of EDTA indicated that DBK is also cleaved by serine proteases to a minor extent. Mass spectrometry-based protein analysis provided spectral proof for several peptides of zinc metalloproteinase-disintegrin-like Eoc1, disintegrin EO4A, and three serine proteases in the venom. Full article
(This article belongs to the Special Issue Renin-Angiotensin System in Cardiovascular Biology)
Show Figures

Figure 1

Figure 1
<p>Scan of a TLC plate showing the results of DBK digestion by snake venom at different time points. The reaction is already complete after a few minutes. The serum control visualizes the location of DBK and its fragment DBK1-8 (fragment DBK1-5 is only visible with contrast enhancement and thus not seen here [<a href="#B14-biomedicines-12-01027" class="html-bibr">14</a>]). The product of venom digestion was shown by MS analysis to be fragment DBK1-7 (<a href="#biomedicines-12-01027-f002" class="html-fig">Figure 2</a>).</p>
Full article ">Figure 2
<p>Peptide fragment ion spectrum for DBK1-7 detected after DBK digest by <span class="html-italic">E. ocellatus</span> venom. For analysis, the TLC spot was scraped off the plate, and the peptide was extracted and subjected to target MS/MS. Ions were labeled according to the b- and y-ion series for amino acid residue losses from either end of the peptide (for theoretical fragment ion masses and original spectrum, see <a href="#app1-biomedicines-12-01027" class="html-app">Supplementary Figure S1</a>). The star indicates an intense ion derived from the dabsyl label [<a href="#B14-biomedicines-12-01027" class="html-bibr">14</a>].</p>
Full article ">Figure 3
<p>Fragment ion spectra for peptides measured in <span class="html-italic">E. ocellatus</span> venom digest using target MS/MS on the doubly charged precursor. Matches from: (<b>A</b>) zinc metalloproteinase-disintegrin-like protein H3, <span class="html-italic">Vipera ammodytes ammodytes</span>, R4NNL0, note zoom ranges; (<b>B</b>) metalloproteinase (fragment), <span class="html-italic">E. coloratus</span>, E9JG63; (<b>C</b>) disintegrin EO4A, <span class="html-italic">E. ocellatus</span>, Q3BER3. Ions were labeled according to the b- and the y-ion series for amino acid residue losses from either end of the peptide (for theoretical fragment ion masses and original spectra, see <a href="#app1-biomedicines-12-01027" class="html-app">Supplementary Figures S2–S4</a>).</p>
Full article ">Figure 3 Cont.
<p>Fragment ion spectra for peptides measured in <span class="html-italic">E. ocellatus</span> venom digest using target MS/MS on the doubly charged precursor. Matches from: (<b>A</b>) zinc metalloproteinase-disintegrin-like protein H3, <span class="html-italic">Vipera ammodytes ammodytes</span>, R4NNL0, note zoom ranges; (<b>B</b>) metalloproteinase (fragment), <span class="html-italic">E. coloratus</span>, E9JG63; (<b>C</b>) disintegrin EO4A, <span class="html-italic">E. ocellatus</span>, Q3BER3. Ions were labeled according to the b- and the y-ion series for amino acid residue losses from either end of the peptide (for theoretical fragment ion masses and original spectra, see <a href="#app1-biomedicines-12-01027" class="html-app">Supplementary Figures S2–S4</a>).</p>
Full article ">Figure 4
<p>Fragment ion spectra for peptides measured in <span class="html-italic">E. ocellatus</span> venom digest using target MS/MS on the doubly charged precursor. Matches from: (<b>A</b>,<b>B</b>) serine protease (fragment, <span class="html-italic">E. ocellatus</span>, D5KRX9; (<b>C</b>) serine protease (fragment, <span class="html-italic">E. ocellatus</span>, D5KRY1). Ions were labeled according to the b- and the y-ion series for amino acid residue losses from either end of the peptide (for theoretical fragment ion masses and original spectra, see <a href="#app1-biomedicines-12-01027" class="html-app">Supplementary Figures S7–S9</a>).</p>
Full article ">
28 pages, 7682 KiB  
Article
Evaluation of Urtica dioica Phytochemicals against Therapeutic Targets of Allergic Rhinitis Using Computational Studies
by Erick Bahena Culhuac and Martiniano Bello
Molecules 2024, 29(8), 1765; https://doi.org/10.3390/molecules29081765 - 12 Apr 2024
Viewed by 1258
Abstract
Allergic rhinitis (AR) is a prevalent inflammatory condition affecting millions globally, with current treatments often associated with significant side effects. To seek safer and more effective alternatives, natural sources like Urtica dioica (UD) are being explored. However, UD’s mechanism of action remains unknown. [...] Read more.
Allergic rhinitis (AR) is a prevalent inflammatory condition affecting millions globally, with current treatments often associated with significant side effects. To seek safer and more effective alternatives, natural sources like Urtica dioica (UD) are being explored. However, UD’s mechanism of action remains unknown. Therefore, to elucidate it, we conducted an in silico evaluation of UD phytochemicals’ effects on known therapeutic targets of allergic rhinitis: histamine receptor 1 (HR1), neurokinin 1 receptor (NK1R), cysteinyl leukotriene receptor 1 (CLR1), chemoattractant receptor-homologous molecule expressed on type 2 helper T cells (CRTH2), and bradykinin receptor type 2 (BK2R). The docking analysis identified amentoflavone, alpha-tocotrienol, neoxanthin, and isorhamnetin 3-O-rutinoside as possessing a high affinity for all the receptors. Subsequently, molecular dynamics (MD) simulations were used to analyze the key interactions; the free energy of binding was calculated through Generalized Born and Surface Area Solvation (MMGBSA), and the conformational changes were evaluated. Alpha-tocotrienol exhibited a high affinity while also inducing positive conformational changes across all targets. Amentoflavone primarily affected CRTH2, neoxanthin targeted NK1R, CRTH2, and BK2R, and isorhamnetin-3-O-rutinoside acted on NK1R. These findings suggest UD’s potential to treat AR symptoms by inhibiting these targets. Notably, alpha-tocotrienol emerges as a promising multi-target inhibitor. Further in vivo and in vitro studies are needed for validation. Full article
(This article belongs to the Topic Plant Extracts and Their Therapeutic Effects)
Show Figures

Figure 1

Figure 1
<p>Graphical representation of the interactions between HR1 and (<b>A</b>) alpha-tocotrienol, (<b>B</b>) amentoflavone, and (<b>C</b>) isorhamnetin 3-O-rutinoside.</p>
Full article ">Figure 2
<p>Graphical representation of the interactions of NKR1 in complex with (<b>A</b>) aprepitant, (<b>B</b>) alpha-tocotrienol, (<b>C</b>) amentoflavone, (<b>D</b>) isorhamnetin-3-O-rutinoside, and (<b>E</b>) neoxanthin.</p>
Full article ">Figure 3
<p>Graphical representation of the interactions of CLR1 in complex with (<b>A</b>) zafirlukast, (<b>B</b>) alpha-tocotrienol, (<b>C</b>) amentoflavone, (<b>D</b>) isorhamnetin-3-O-rutinoside, and (<b>E</b>) neoxanthin.</p>
Full article ">Figure 4
<p>Graphical representation of the interactions of CRTH2 in complex with (<b>A</b>) fevipiprant, (<b>B</b>) alpha-tocotrienol, (<b>C</b>) amentoflavone, (<b>D</b>) isorhamnetin-3-O-rutinoside, and (<b>E</b>) neoxanthin.</p>
Full article ">Figure 5
<p>Graphical representation of the interactions of BK2R in complex with (<b>A</b>) JSM-10292, (<b>B</b>) alpha-tocotrienol, (<b>C</b>) amentoflavone, (<b>D</b>) isorhamnetin-3-O-rutinoside, and (<b>E</b>) neoxanthin.</p>
Full article ">Figure 6
<p>Structural comparison of HR1 (intracellular views): (<b>A</b>) alpha-tocotrienol complex (red) dynamics at 0, 50, and 100 ns compared to histamine experimental structure (cyan). (<b>B</b>) Alpha-tocotrienol complex at 0 ns (red) and 100 ns (orange). (<b>C</b>) Alpha-tocotrienol complex at 0 ns (red) compared to doxepin experimental structure (black). (<b>D</b>) Alpha-tocotrienol complex at 100 ns (red) compared to doxepin experimental structure (black). White arrows illustrate the measurements between TMs. Measurements were performed with VMD based on the alpha carbon of TM3’s residue 130, TM6’s residue 408, and TM7’s residue 472.</p>
Full article ">Figure 7
<p>Structural comparison of NKR1 (intracellular view). (<b>A</b>) Alpha-tocotrienol complex dynamics (red) at 0, 50, and 100 ns compared to experimental structure of NKR1/aprepitant (black). (<b>B</b>) Alpha-tocotrienol complex (red) compared to experimental structure of NKR1/substance P (cyan). (<b>C</b>) MD aprepitant complex dynamics (silver) at 0, 50, and 100 ns compared to experimental structure of NKR1/aprepitant (black). (<b>D</b>) MD aprepitant complex (silver) compared to experimental structure of NKR1/substance P (cyan). (<b>E</b>) Isorhamnetin-3-O-rutinoside complex dynamics (purple) at 0, 50, and 100 ns compared to experimental structure of NKR1/aprepitant (black). (<b>F</b>) Isorhamnetin-3-O-rutinoside complex compared to experimental structure of NKR1/substance P (cyan). (<b>G</b>) Neoxanthin Complex dynamics (blue) at 0, 50, and 100 ns compared to experimental structure of NKR1/aprepitant (black). (<b>H</b>) Neoxanthin complex compared to experimental structure of NKR1/substance P (Cyan). Measurements were performed with VMD based on the alpha carbon of TM3’s residue 135, TM4’s residue 144, TM5’s residue 224, and TM6’s residue 242. White arrows illustrate the measurements between TMs.</p>
Full article ">Figure 8
<p>Structural comparison of CLR1 (intracellular view). (<b>A</b>) Alpha-tocotrienol complex dynamics (red) at 0, 50, and 100 ns compared to experimental structure of CLR1/zafirlukast (black). (<b>B</b>) Alpha-tocotrienol complex at 0 ns (red) and 100 ns (orange). (<b>C</b>) MD zafirlukast complex dynamics (silver) at 0, 50, and 100 ns compared to experimental structure of CLR1/zafirlukast (black). (<b>D</b>) Zafirlukast complex at 0 ns (silver) and 100 ns (grey). (<b>E</b>) Alpha-tocotrienol complex (red) compared to MD zafirlukast at 100 ns (silver). Measurements were performed with VMD based on the alpha carbon of TM5’s residue 221 and TM6’s residue 225. White arrows illustrate the measurements between TMs.</p>
Full article ">Figure 9
<p>Structural comparison of CRTH2 (intracellular view). (<b>A</b>) Alpha-tocotrienol complex dynamics (red) at 0, 50, and 100 ns compared to experimental structure of CRTH2/15R-methyl-PGD2 (cyan). (<b>B</b>) Fevipiprant complex dynamics (silver) at 0, 50, and 100 ns compared to experimental structure of CRTH2/15R-methyl-PGD2 (cyan). (<b>C</b>) Amentoflavone complex dynamics (green) at 0, 50, and 100 ns compared to experimental structure of CRTH2/15R-methyl-PGD2 (cyan). (<b>D</b>) neoxanthin complex dynamics (blue) at 0, 50, and 100 ns compared to experimental structure of CRTH2/15R-methyl-PGD2 (cyan).</p>
Full article ">Figure 10
<p>Structural comparison of BK2R (intracellular view). (<b>A</b>) Alpha-tocotrienol complex dynamics (red) at 0, 50, and 100 ns compared to experimental structure of BK2R/bradykinin (cyan) (<b>B</b>) Alpha-tocotrienol complex at 0 ns (red) and 100 ns (orange). (<b>C</b>) JSM-10292 complex dynamics (silver) at 0, 50, and 100 ns compared to experimental structure of BK2R/bradykinin (cyan). (<b>D</b>) JSM-10292 complex at 0 ns (silver) and 100 ns (grey). (<b>E</b>) Alpha-tocotrienol complex (red) compared to JSM-10292 at 100 ns (silver). Measurements were performed with VMD based on the alpha carbon of TM5’s residue 253 and TM6’s residue 266. White arrows illustrate the measurements between TMs.</p>
Full article ">Figure 11
<p>Pathways associated with AR targets. The graph was made using Shinny GO 8.0 Analysis [<a href="#B74-molecules-29-01765" class="html-bibr">74</a>].</p>
Full article ">
14 pages, 2596 KiB  
Review
Sebetralstat: A Rapidly Acting Oral Plasma Kallikrein Inhibitor for the On-Demand Treatment of Hereditary Angioedema
by Edward P. Feener, Rebecca L. Davie, Nivetha Murugesan, Stephen J. Pethen, Sally L. Hampton, Michael D. Smith, Paul K. Audhya and Chris M. Yea
Drugs Drug Candidates 2024, 3(2), 328-341; https://doi.org/10.3390/ddc3020019 - 7 Apr 2024
Viewed by 2081
Abstract
Sebetralstat is a novel, potent, and selective oral plasma kallikrein inhibitor drug candidate in clinical development for the on-demand treatment of hereditary angioedema (HAE). Upon binding, sebetralstat induces a conformational change in the active site of plasma kallikrein, which contributes to its high [...] Read more.
Sebetralstat is a novel, potent, and selective oral plasma kallikrein inhibitor drug candidate in clinical development for the on-demand treatment of hereditary angioedema (HAE). Upon binding, sebetralstat induces a conformational change in the active site of plasma kallikrein, which contributes to its high potency (Ki 3 nM) and selectivity (>1500 fold) against other serine proteases. Its physiochemical properties promote both rapid dissolution in the stomach and rapid absorption in the upper intestine that contribute to its fast and efficient absorption. A single oral administration of sebetralstat rapidly provides near-complete inhibition of plasma kallikrein and blockade of high-molecular-weight kininogen cleavage as early as 15 min, which drives its clinical efficacy. In a phase 2 clinical trial, sebetralstat significantly reduced the time to beginning of symptom relief (p < 0.0001) with median times of 1.6 h (95% CI: 1.5–3.0) with sebetralstat versus 9.0 h (4.0–17.2) with placebo. KONFIDENT (NCT05259917) is a phase 3 clinical trial assessing the on-demand use of sebetralstat for HAE. If successful, this trial could support the approval of sebetralstat as the first noninvasive, on-demand treatment option to rapidly halt HAE attacks and provide fast symptom relief. Full article
(This article belongs to the Special Issue Drugs of the Kallikrein-Kinin System)
Show Figures

Graphical abstract

Graphical abstract
Full article ">Figure 1
<p>Structure of sebetralstat (KVD900).</p>
Full article ">Figure 2
<p>Evolution of plasma kallikrein (PKa) inhibitors with reduced basicity in the P1 group. The human PKa IC<sub>50</sub> and ligand efficiency (LE) for each compound are indicated.</p>
Full article ">Figure 3
<p>(<b>A</b>). Crystal structure of sebetralstat complex with PKa (PDB code 8A3Q). (<b>B</b>). Key interactions of sebetralstat with residues in the active site of PKa. Network of π-interactions observed: TYR 174/terminal pyridone (face-to-face); TRP 215/phenyl linker (face-to-face) and pyrazole core (edge-to-face); HIS 57/pyrazole core (face-to-face). Hydrogen bonds are observed: GLY 99 backbone N–H/pyridone carbonyl, SER 214 backbone carbonyl/amide N–H, and LYS 192 side chain N–H/amide carbonyl. The pyridine P1 group occupies the S1 subpocket without forming specific polar interactions with any of the amino acids, including ASP 189. (<b>C</b>). Overlays of crystal structures 8A3Q (protein shown in pink, sebetralstat in green) and 2ANW (protein in gray, benzamidine in yellow) to highlight the movement of key residues in the S4 region to enable an extended network of π-stacking interactions with sebetralstat. Images created using BioSolvIT SeeSAR v13.0.1; BioSolveIT GmbH, Sankt Augustin, Germany, 2023, <a href="http://www.biosolveit.de/SeeSAR" target="_blank">www.biosolveit.de/SeeSAR</a> (accessed on 11 December 2023).</p>
Full article ">Figure 4
<p>Dose–response effects of sebetralstat on PKa activity and HK cleavage in human plasma. (<b>A</b>) Representative graphs showing dose–response effects of sebetralstat in DXS-stimulated PKa activity in plasma from a healthy volunteer and a patient with HAE-C1INH. (<b>B</b>) The percentage of HK after DXS-stimulation of the plasma was determined in the presence or absence of sebetralstat. Bar graph showing the percentage of HK after DXS stimulation compared to HK in unstimulated healthy volunteer plasma (expressed as % mean ± SEM). cHK, cleaved HK; DXS, dextran sulfate; HK, high-molecular-weight kininogen; PKa, plasma kallikrein; SEM, standard error of the mean. <span class="html-italic">p</span>-values: **** &lt; 0.0001, * &lt; 0.05. Figure panels (<b>A</b>,<b>B</b>) were reprinted/adapted with permission from Ref. [<a href="#B44-ddc-03-00019" class="html-bibr">44</a>]. 2022 KalVista Pharmaceuticals, Inc. <span class="html-italic">Clin. Exp. Allergy</span> 2022, 52, 1059–1070 Duckworth et al. Pharmacological Suppression of the Kallikrein Kinin System with KVD900: An Orally Available Plasma Kallikrein Inhibitor for the On-demand Treatment of Hereditary Angioedema.</p>
Full article ">Figure 5
<p>Effects of sebetralstat on PKa activity and HK cleavage in healthy volunteers. Sebetralstat was administered as a single dose (600-mg tablet) to 12 healthy male volunteers under fasted conditions. Plasma was collected pre- and postdose at the times indicated. (<b>A</b>) Sebetralstat concentration in plasma (gray). PKa activity (black) in DXS-stimulated plasma postdose compared with plasma obtained before sebetralstat administration. (<b>B</b>) The relative concentration of HK after DXS stimulation as a percentage of total HK and cHK (combined) compared to reference control plasma (arithmetic mean ± SD). (<b>C</b>) Population PD data in healthy volunteers administered a single dose of sebetralstat at 160, 300, or 600 mg, and patients with HAE administered a single 600-mg dose. Data show arithmetic mean percent DXS-stimulated PKa activity in plasma samples collected. The indicated timepoints are postdose compared to predose plasma. DXS, dextran sulfate; HK, high-molecular-weight kininogen; cHK, cleaved high-molecular-weight kininogen; SD, standard deviation of the mean. Adapted with permission from Ref. [<a href="#B43-ddc-03-00019" class="html-bibr">43</a>]. 2021, Andreas Maetzel, Michael D. Smith, Edward J. Duckworth, Sally L. Hampton, Gian Marco De Donatis, Nivetha Murugesan, Louise J. Rushbrooke, Lily Li, Danielle Francombe, Edward P. Feener, and Christopher M. Yea.</p>
Full article ">Figure 6
<p>Pharmacokinetic and pharmacodynamic effects of a single oral administration of 600 mg of sebetralstat in a tablet formulation on plasma in participants with HAE-C1INH. Plasma was obtained predose and at the indicated times after sebetralstat administration up to 4 h, the final timepoint measured. (<b>A</b>) Mean plasma concentrations of sebetralstat for participants with HAE-C1INH (<span class="html-italic">n</span> = 68 dosed/enrolled, <span class="html-italic">n</span> = 42 with plasma samples) are shown in the gray line (ng/mL). PKa activity in DXS-stimulated plasma from a representation cohort of 26 participants with HAE-C1INH is shown in the black line. (<b>B</b>) HK and PK were quantified using a capillary-based immunoassay in DXS-stimulated whole plasma from a representative cohort of 6 individuals with HAE-C1INH. Bar graphs show percentage compared with plasma PK and HK levels in the absence of DXS stimulation (arithmetic mean ± SEM). DXS, dextran sulfate; HK, high-molecular-weight kininogen; PK, plasma prekallikrein; SD, standard deviation of the mean (modified from [<a href="#B44-ddc-03-00019" class="html-bibr">44</a>,<a href="#B54-ddc-03-00019" class="html-bibr">54</a>]). Figure panel A reprinted/adapted with permission from Ref. [<a href="#B44-ddc-03-00019" class="html-bibr">44</a>]. 2022 KalVista Pharmaceuticals, Inc. Figure panel B was published in <span class="html-italic">Lancet</span>, 401(10375), Aygören-Pürsün E, et al., An investigational oral plasma kallikrein inhibitor for on-demand treatment of hereditary angioedema: a two-part, randomised, double-blind, placebo-controlled, crossover phase 2 trial, pages 458-469, Copyright Elsevier 2023.</p>
Full article ">Figure 7
<p>Time to symptom relief and proportion of attacks rated at least “a little better” on the Patient Global Impression of Change (PGI-C) for two consecutive timepoints within 12 h of study drug administration. Number at risk is the count of individuals who have not yet experienced the event and have not been censored (modified from [<a href="#B54-ddc-03-00019" class="html-bibr">54</a>]). This figure was published in <span class="html-italic">Lancet</span>, 401(10375), Aygören-Pürsün E, et al., An investigational oral plasma kallikrein inhibitor for on-demand treatment of hereditary angioedema: a two-part, randomised, double-blind, placebo-controlled, crossover phase 2 trial, pages 458–469, Copyright Elsevier 2023.</p>
Full article ">Figure 8
<p>Schema of the effects of sebetralstat on the kallikrein kinin system (KKS). Sebetralstat inhibits plasma kallikrein (PKa)-mediated cleavage of high-molecular-weight kininogen (HK) that generates bradykinin and cleaved HK (cHK). Sebetralstat also inhibits the activation of Factor XII and the generation of PKa by the positive feedback amplification of the KKS. HAE, hereditary angioedema.</p>
Full article ">
25 pages, 5740 KiB  
Review
New Light on Plants and Their Chemical Compounds Used in Polish Folk Medicine to Treat Urinary Diseases
by Beata Olas, Waldemar Różański, Karina Urbańska, Natalia Sławińska and Magdalena Bryś
Pharmaceuticals 2024, 17(4), 435; https://doi.org/10.3390/ph17040435 - 28 Mar 2024
Viewed by 1695
Abstract
This review contains the results of Polish (Central Europe) ethnomedical studies that describe the treatment of urinary tract diseases with wild and cultivated plants. The study includes only the plants that are used to treat the urinary tract, excluding prostate diseases. A review [...] Read more.
This review contains the results of Polish (Central Europe) ethnomedical studies that describe the treatment of urinary tract diseases with wild and cultivated plants. The study includes only the plants that are used to treat the urinary tract, excluding prostate diseases. A review of the literature was carried out to verify the pharmacological use of the plants mentioned in the interviews. Based on this, the study reviews the pharmacological activities of all the recorded species and indicates their most important chemical compounds. Fifty-three species (belonging to 30 families) were selected for the study. The Compositae (eight species), Rosaceae (six species), and Apiaceae (six species) are the most common families used in the treatment of urinary diseases in Polish folk medicine. Both in vitro and in vivo studies have confirmed that many of these plant species have beneficial properties, such as diuretic, antihyperuricemic, antimicrobial, and anti-inflammatory activity, or the prevention of urinary stone formation. These effects are exerted through different mechanisms, for example, through the activation of bradykinin B2 receptors, inhibition of xanthine oxidase, or inhibition of Na+-K+ pump. Many plants used in folk medicine are rich in phytochemicals with proven effectiveness against urinary tract diseases, such as rutin, arbutin, or triterpene saponins. Full article
(This article belongs to the Section Natural Products)
Show Figures

Figure 1

Figure 1
<p>Summary of activities of plants used in Polish folk medicine to treat urinary disorders (confirmed by scientific studies). The arrows indicate an increase (upward-facing arrow) or decrease (downward-facing arrow) in the excretion or blood concentration of compounds and markers. Compilation of data from <a href="#sec2-pharmaceuticals-17-00435" class="html-sec">Section 2</a>.</p>
Full article ">Figure 2
<p>The mechanisms of action of plants toward urinary diseases and disorders. Upward-facing green arrow indicates increase, while downward-facing red arrow indicates decrease. Compilation of data from <a href="#sec1-pharmaceuticals-17-00435" class="html-sec">Section 1</a> and <a href="#sec2-pharmaceuticals-17-00435" class="html-sec">Section 2</a>, as well as [<a href="#B119-pharmaceuticals-17-00435" class="html-bibr">119</a>,<a href="#B120-pharmaceuticals-17-00435" class="html-bibr">120</a>,<a href="#B121-pharmaceuticals-17-00435" class="html-bibr">121</a>]. UTIs—urinary diseases.</p>
Full article ">Figure 3
<p>Main chemical compounds of selected plants used in Polish folk medicine to treat urinary diseases and their potential molecular targets of action. Upward-facing green arrows indicate an increase, while downward-facing red arrows indicate a decrease. COX, cyclooxygenase; KIM1, kidney injury molecule 1; TNF-α, tumor necrosis factor alpha; TGF-β, transforming growth factor β; MMP 9, matrix metalloproteinase 9; XO, xanthine oxidase.</p>
Full article ">
28 pages, 8226 KiB  
Article
Modular Hub Genes in DNA Microarray Suggest Potential Signaling Pathway Interconnectivity in Various Glioma Grades
by Marco A. Orda, Peter Matthew Paul T. Fowler and Lemmuel L. Tayo
Biology 2024, 13(4), 206; https://doi.org/10.3390/biology13040206 - 23 Mar 2024
Cited by 5 | Viewed by 1984
Abstract
Gliomas have displayed significant challenges in oncology due to their high degree of invasiveness, recurrence, and resistance to treatment strategies. In this work, the key hub genes mainly associated with different grades of glioma, which were represented by pilocytic astrocytoma (PA), oligodendroglioma (OG), [...] Read more.
Gliomas have displayed significant challenges in oncology due to their high degree of invasiveness, recurrence, and resistance to treatment strategies. In this work, the key hub genes mainly associated with different grades of glioma, which were represented by pilocytic astrocytoma (PA), oligodendroglioma (OG), anaplastic astrocytoma (AA), and glioblastoma multiforme (GBM), were identified through weighted gene co-expression network analysis (WGCNA) of microarray datasets retrieved from the Gene Expression Omnibus (GEO) database. Through this, four highly correlated modules were observed to be present across the PA (GSE50161), OG (GSE4290), AA (GSE43378), and GBM (GSE36245) datasets. The functional annotation and pathway enrichment analysis done through the Database for Annotation, Visualization, and Integrated Discovery (DAVID) showed that the modules and hub genes identified were mainly involved in signal transduction, transcription regulation, and protein binding, which collectively deregulate several signaling pathways, mainly PI3K/Akt and metabolic pathways. The involvement of several hub genes primarily linked to other signaling pathways, including the cAMP, MAPK/ERK, Wnt/β-catenin, and calcium signaling pathways, indicates potential interconnectivity and influence on the PI3K/Akt pathway and, subsequently, glioma severity. The Drug Repurposing Encyclopedia (DRE) was used to screen for potential drugs based on the up- and downregulated hub genes, wherein the synthetic progestin hormones norgestimate and ethisterone were the top drug candidates. This shows the potential neuroprotective effect of progesterone against glioma due to its influence on EGFR expression and other signaling pathways. Aside from these, several experimental and approved drug candidates were also identified, which include an adrenergic receptor antagonist, a PPAR-γ receptor agonist, a CDK inhibitor, a sodium channel blocker, a bradykinin receptor antagonist, and a dopamine receptor agonist, which further highlights the gene network as a potential therapeutic avenue for glioma. Full article
(This article belongs to the Section Biochemistry and Molecular Biology)
Show Figures

Graphical abstract

Graphical abstract
Full article ">Figure 1
<p>Network indices to approximate scale-free topology through average number of connections per gene in the networks. Numbers represent the scale free topology model fit per dataset at given soft-thresholding power. The green box indicates the point at which the index across the datasets has stabilized (β = 10).</p>
Full article ">Figure 2
<p>Approximate linear relationship for the GBM dataset at β = 10 presented by log-log plot.</p>
Full article ">Figure 3
<p>Dendrogram of gene clustering on TOM-based dissimilarity and module split sensitivities of GBM dataset. The assorted colors represent identified gene co-expression modules corresponding to the portion of the dendrogram directly above.</p>
Full article ">Figure 4
<p>Module preservation analysis performed on the gene co-expression modules from the GBM network in (<b>a</b>) pilocytic astrocytoma, (<b>b</b>) oligodendroglioma, and (<b>c</b>) anaplastic astrocytoma datasets. The modules in the yellow partition pertain to highly preserved modules (Z &gt; 10); green, yellow, brown, and gray exhibited high preservation in all datasets.</p>
Full article ">Figure 5
<p>Top-enriched terms for the green, yellow, brown, and gray modules in terms of (<b>a</b>) biological processes, (<b>b</b>) cellular component, (<b>c</b>) molecular functions, and (<b>d</b>) KEGG pathways.</p>
Full article ">Figure 6
<p>The identified top 10 hub gene networks based on the PPI networks of the (<b>a</b>) green, (<b>b</b>) yellow, (<b>c</b>) brown, and (<b>d</b>) gray modules. Nodes represent the hub genes, and edges represent interactions between connected nodes. Colors represent ranking with each network, from red (high) to yellow (low).</p>
Full article ">Figure 7
<p>KEGG pathways of modules with increasing preservation per glioma grade (I, II, III, and IV) based on module preservation and pathway enrichment analysis results. Signaling pathways within the range of glioma grade are highly preserved (z &gt; 10).</p>
Full article ">Figure 8
<p>Cellular processes interconnectivity based on hub gene functions.</p>
Full article ">Figure A1
<p>Boxplot of microarray datasets (<b>a</b>) PA, (<b>b</b>) OG, and (<b>c</b>) AA and (<b>d</b>) GBM datasets after data preparation and filtering.</p>
Full article ">Figure A2
<p>Sample clustering dendrogram of (<b>a</b>) PA, (<b>b</b>) OG, (<b>c</b>) AA, and (<b>d</b>) GBM datasets. The ends of each branch represent the individual samples in the datasets.</p>
Full article ">Figure A3
<p>Ranked expression plots and ranked connectivity plots for dataset comparison: (<b>a</b>,<b>b</b>) PA vs. OG; (<b>c</b>,<b>d</b>) PA vs. AA; (<b>e</b>,<b>f</b>) OG vs. AA; (<b>g</b>,<b>h</b>) GBM vs. PA; (<b>i</b>,<b>j</b>) GBM vs. AA; and (<b>k</b>,<b>l</b>) GBM vs. OG.</p>
Full article ">
14 pages, 611 KiB  
Review
Five-Membered Nitrogen Heterocycles Angiotensin-Converting Enzyme (ACE) Inhibitors Induced Angioedema: An Underdiagnosed Condition
by Niki Papapostolou, Stamatios Gregoriou, Alexander Katoulis and Michael Makris
Pharmaceuticals 2024, 17(3), 360; https://doi.org/10.3390/ph17030360 - 10 Mar 2024
Viewed by 1928
Abstract
Angiotensin-converting enzyme (ACE) inhibitors are used primarily in the treatment of hypertension, heart failure, and in the acute phase of myocardial infarction. Lisinopril [N2-[(1S)-1-car-boxy-3-phenylpropyl]-L-lysyl-L-proline], enalapril [(S)-1-[N-[1-(ethoxycarbonyl)-3-phenylpropyl]-L-alanyl]-L-proline] and ramipril [2-aza-bicyclo-[3.3.0]-octane-3-carboxylic acid] are all five-membered heterocycles and three of the most prevalent ACE [...] Read more.
Angiotensin-converting enzyme (ACE) inhibitors are used primarily in the treatment of hypertension, heart failure, and in the acute phase of myocardial infarction. Lisinopril [N2-[(1S)-1-car-boxy-3-phenylpropyl]-L-lysyl-L-proline], enalapril [(S)-1-[N-[1-(ethoxycarbonyl)-3-phenylpropyl]-L-alanyl]-L-proline] and ramipril [2-aza-bicyclo-[3.3.0]-octane-3-carboxylic acid] are all five-membered heterocycles and three of the most prevalent ACE inhibitors in clinical use worldwide. ACE inhibitor-induced angioedema (AE) is clinically characterized by self-limited edema of the dermis and subcutaneous lipid tissue, localized on face skin, oral mucosa and tongue in most cases. However, severe episodes of intestinal AE misdiagnosed as acute appendicitis and laryngeal AE requiring incubation have been reported. The pathophysiology of ACE inhibitor-induced angioedema is attributed to the accumulation of bradykinin, which is a potent vasodilator with proinflammatory activity that is normally degraded by angiotensin-converting enzyme (ACE) and aminopeptidase P; however, a small proportion of treated patients is affected. Given that patients do not respond to anti-H1 antihistamines and steroids, early clinical recognition and discontinuation of the ACE inhibitors are the treatments of choice for the long-term management of ACE inhibitor- induced angioedema. The search period of the present review was set up until November 2023, and its aim is to shed light on the broader context of ACE inhibitor-induced angioedema, exploring aspects such as clinical presentation, pathophysiology, and therapeutic considerations in this potentially life-threatening condition. The exploration of alternative drug options such as angiotensin II receptor blockers, the potential association of coadministration of DPP-4 inhibitors with ACE inhibitors, the presentation of angioedema and the significant clinical importance of this condition are also discussed. By focusing on the chemical structure of ACE inhibitors, specifically their nitrogen-based heterocycles—an attribute shared by over 880 drugs approved by the FDA within the pharmaceutical industry—this review emphasizes the pivotal role of nitrogen scaffolds in drug design and underscores their relevance in ACE inhibitor pharmacology. Full article
(This article belongs to the Special Issue Nitrogen Containing Scaffolds in Medicinal Chemistry 2023)
Show Figures

Figure 1

Figure 1
<p>Pathways involved in bradykinin mediated angioedema.</p>
Full article ">
Back to TopTop