[go: up one dir, main page]

 
 
Sign in to use this feature.

Years

Between: -

Subjects

remove_circle_outline
remove_circle_outline
remove_circle_outline
remove_circle_outline
remove_circle_outline
remove_circle_outline
remove_circle_outline
remove_circle_outline
remove_circle_outline

Journals

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

Article Types

Countries / Regions

remove_circle_outline
remove_circle_outline
remove_circle_outline
remove_circle_outline
remove_circle_outline
remove_circle_outline
remove_circle_outline

Search Results (2,922)

Search Parameters:
Keywords = angiotensin

Order results
Result details
Results per page
Select all
Export citation of selected articles as:
12 pages, 407 KiB  
Article
Use and Benefit of Sacubitril/Valsartan in Elderly Patients with Heart Failure with Reduced Ejection Fraction
by Luis Nieto Roca, Marcelino Cortés García, Jorge Balaguer Germán, Antonio José Bollas Becerra, José María Romero Otero, José Antonio Esteban Chapel, Carlos Rodríguez López, Ana María Pello Lázaro, Mikel Taibo Urquía and José Tuñón
J. Clin. Med. 2024, 13(16), 4772; https://doi.org/10.3390/jcm13164772 - 14 Aug 2024
Abstract
Background: Heart failure (HF) is a highly prevalent syndrome in elderly subjects. Currently, multiple drugs have shown clinical benefits in patients with HF and reduced ejection fraction (HFrEF). However, evidence is scarce in elderly patients (beyond 75 years old), even more so for [...] Read more.
Background: Heart failure (HF) is a highly prevalent syndrome in elderly subjects. Currently, multiple drugs have shown clinical benefits in patients with HF and reduced ejection fraction (HFrEF). However, evidence is scarce in elderly patients (beyond 75 years old), even more so for the latest drugs, such as angiotensin receptor-neprilysin inhibitors (ARNIs). This study aims to evaluate the use and benefits of ARNIs in elderly patients with HFrEF. Methods: A prospective observational cohort study was designed. Patients with left ventricular systolic dysfunction (defined by left ventricular ejection fraction [LVEF] < 40%) and age ≥ 75 years from January 2016 to December 2020 were prospectively included. Patients with an indication for ARNIs at inclusion or throughout follow-up were selected. Clinical, electrocardiographic and echocardiographic variables were collected. Results: A total of 616 patients were included, 34.4% of them female, with a mean age of 83.3 years, mean LVEF of 28.5% and ischemic etiology in 53.9% of patients. Only 14.3% of patients were taking ARNIs. After a mean follow-up of 34 months, 50.2% of patients died, and 62.2% had a cardiac event (total mortality or hospital admission due to HF). Multivariate Cox regression analysis showed that the use of ARNIs was independently and significantly associated with lower rates of mortality [HR 0.36 (95% CI 0.21–0.61)], with similar results in relation to all-cause mortality in a propensity-score-matched analysis [HR 0.33 (95% CI 0.19–0.57)]. Conclusions: We observed an important underuse of ARNIs in a cohort of elderly HFrEF patients, in which treatment with ARNIs was associated with a significant reduction in mortality. Greater implementation of clinical practice guidelines in this group of patients could improve their prognosis. Full article
(This article belongs to the Special Issue Clinical Management of Patients with Heart Failure)
Show Figures

Graphical abstract

Graphical abstract
Full article ">Figure 1
<p>Kaplan–Meier curves for mortality, first showing results for the overall population and then after PS (propensity score) matching according to ARNI use. ARNIs: angiotensin receptor-neprilysin inhibitors.</p>
Full article ">
11 pages, 1503 KiB  
Review
Purinergic Receptor Antagonists: A Complementary Treatment for Hypertension
by Rocio Bautista-Pérez and Martha Franco
Pharmaceuticals 2024, 17(8), 1060; https://doi.org/10.3390/ph17081060 - 13 Aug 2024
Viewed by 206
Abstract
The treatment of hypertension has improved in the last century; attention has been directed to restoring several altered pathophysiological mechanisms. However, regardless of the current treatments, it is difficult to control blood pressure. Uncontrolled hypertension is responsible for several cardiovascular complications, such as [...] Read more.
The treatment of hypertension has improved in the last century; attention has been directed to restoring several altered pathophysiological mechanisms. However, regardless of the current treatments, it is difficult to control blood pressure. Uncontrolled hypertension is responsible for several cardiovascular complications, such as chronic renal failure, which is frequently observed in hypertensive patients. Therefore, new approaches that may improve the control of arterial blood pressure should be considered to prevent serious cardiovascular disorders. The contribution of purinergic receptors has been acknowledged in the pathophysiology of hypertension; this review describes the participation of these receptors in the alteration of kidney function in hypertension. Elevated interstitial ATP concentrations are essential for the activation of renal purinergic receptors; this becomes a fundamental pathway that leads to the development and maintenance of hypertension. High ATP levels modify essential mechanisms implicated in the long-term control of blood pressure, such as pressure natriuresis, the autoregulation of the glomerular filtration rate and renal blood flow, and tubuloglomerular feedback responses. Any alteration in these mechanisms decreases sodium excretion. ATP stimulates the release of vasoactive substances, causes renal function to decline, and induces tubulointerstitial damage. At the same time, a deleterious interaction involving angiotensin II and purinergic receptors leads to the deterioration of renal function. Full article
(This article belongs to the Special Issue Pharmacological Advances for Treatment in Hypertension 2.0)
Show Figures

Figure 1

Figure 1
<p>Renal hemodynamics in rats with Ang II-mediated hypertension (14 days, Ang II, 435 ng/kg/min) during an acute infusion of a broad purinergic receptor blocker (PPADS, 30 mg⋅kg<sup>−1</sup>) and specific P2X1 (NF 449, 30 nM⋅kg<sup>−1</sup>⋅h<sup>−1</sup>) and P2X7 (A 438079, 80 μm⋅kg<sup>−1</sup>) receptor antagonists. The purinergic antagonists induced a decrease in the afferent and efferent resistances (* &lt; 0.05 to 0.019; ο &lt; 0.05 vs. Sham) that produced a significant elevation in the glomerular blood flow; as a result of these changes, the single-nephron glomerular filtration rate returned to near-normal levels. These results show that, in Ang II-induced hypertension, the renal vasoconstriction induced by Ang II is associated with important actions of the P2X1 and P2X7 receptors (Modified by Franco et al. [<a href="#B32-pharmaceuticals-17-01060" class="html-bibr">32</a>,<a href="#B34-pharmaceuticals-17-01060" class="html-bibr">34</a>]).</p>
Full article ">Figure 2
<p>The vasoconstrictor effects of Ang II (14 days) and both P2X1 and P2X7 receptors are shown in this figure. Ang II infusion induced systemic hypertension and an elevation of interstitial fluid concentrations of ATP as well as local Ang II. The direct effect of Ang II associated with its regulatory response to hypertension produced renal vasoconstriction. The glomerular hemodynamics were characterized by an increase in the afferent resistance (AR) and efferent resistance (ER), which resulted in a decrease in the glomerular blood flow (GBF) and a diminished ultrafiltration coefficient (Kf). These alterations produced a reduction in the single-nephron glomerular filtration rate. These changes induced renal ischemia, leading to an overexpression of P2X receptors in the smooth muscle of intrarenal arterioles. Simultaneously, tubulointerstitial inflammatory cell infiltration contributed to a further elevation in interstitial ATP and the overexpression of P2X receptors in the intrarenal arterioles and on the surface of inflammatory cells. These changes induced the release of cytokines, growth factors, and chemoattractant factors; these factors exacerbated inflammatory cell infiltration and the intensity of renal vasoconstriction. Under these conditions, oxidative stress, the augmentation of adenosine (ADO), a decrease in nitric oxide (NO), an increase in the local production of Ang II (RAS), and the stimulation of sympathetic tone (SNS) developed. These alterations adjusted the sodium excretion and reduced pressure natriuresis, leading to decreased sodium excretion compared to the expected level for the elevation of blood pressure. This resulted in sodium retention, and salt-sensitive hypertension developed (modified from Graciano et al. 2008 [<a href="#B47-pharmaceuticals-17-01060" class="html-bibr">47</a>] and Franco et al. 2019 [<a href="#B43-pharmaceuticals-17-01060" class="html-bibr">43</a>]).</p>
Full article ">
13 pages, 5142 KiB  
Article
Spike Protein of SARS-CoV-2 Activates Cardiac Fibrogenesis through NLRP3 Inflammasomes and NF-κB Signaling
by Huynh Van Tin, Lekha Rethi, Satoshi Higa, Yu-Hsun Kao and Yi-Jen Chen
Cells 2024, 13(16), 1331; https://doi.org/10.3390/cells13161331 - 11 Aug 2024
Viewed by 2920
Abstract
Background: The spike protein of severe acute respiratory syndrome coronavirus 2 (SARS-CoV-2) is crucial to viral entry and can cause cardiac injuries. Toll-like receptor 4 (TLR4) and NOD-, LPR-, and pyrin-domain-containing 3 (NLRP3) inflammasome are critical immune system components implicated in cardiac fibrosis. [...] Read more.
Background: The spike protein of severe acute respiratory syndrome coronavirus 2 (SARS-CoV-2) is crucial to viral entry and can cause cardiac injuries. Toll-like receptor 4 (TLR4) and NOD-, LPR-, and pyrin-domain-containing 3 (NLRP3) inflammasome are critical immune system components implicated in cardiac fibrosis. The spike proteins activate NLRP3 inflammasome through TLR4 or angiotensin-converting enzyme 2 (ACE2) receptors, damaging various organs. However, the role of spike proteins in cardiac fibrosis in humans and the interactions of spike proteins with NLRP3 inflammasomes and TLR4 remain poorly understood. Methods: We utilized scratch assays, Western blotting, and immunofluorescence to evaluate the migration, fibrosis signaling, mitochondrial calcium levels, reactive oxygen species (ROS) production, and cell morphology of cultured human cardiac fibroblasts (CFs) treated with spike (S1) proteins for 24 h with or without an anti-ACE2 neutralizing antibody, a TLR4 blocker, or an NLRP3 inhibitor. Results: S1 protein enhanced CFs migration and the expressions of collagen 1, α-smooth muscle actin, transforming growth factor β1 (TGF-β1), phosphorylated SMAD2/3, interleukin 1β (IL-1β), and nuclear factor kappa-light-chain-enhancer of activated B cells (NF-κB). S1 increased ROS production but did not affect mitochondrial calcium content and cell morphology. Treatment with an anti-ACE2 neutralizing antibody attenuated the effects of S1 on collagen 1 and TGF-β1 expressions. Moreover, NLRP3 (MCC950) and NF-kB inhibitors, but not the TLR4 inhibitor TAK-242, prevented the S1-enhanced CFs migration and overexpression of collagen 1, TGF-β1, and IL-1β. Conclusion: S1 activates human CFs by priming NLRP3 inflammasomes through NF-κB signaling in an ACE2-dependent manner. Full article
(This article belongs to the Special Issue Insight into Cardiomyopathy)
Show Figures

Graphical abstract

Graphical abstract
Full article ">Figure 1
<p>S1 protein enhanced CFs activation. Treatment with S1 protein (5 nM) for 24 h increased cell migration (<b>A</b>) but not cell proliferation (<b>B</b>) in CFs. (<b>C</b>) Additionally, S1 protein also elevated pro-COL1A1 and α-SMA protein expressions. <span class="html-italic">n</span> = 4 independent experiments.</p>
Full article ">Figure 2
<p>S1 protein increased profibrotic signaling. Treatment with S1 protein (5 nM) for 24 h increased protein expressions of TGF-β1 and pSMAD2/3 measured using Immunoblot ((<b>A</b>) <span class="html-italic">n</span> = 4 independent experiments), secretion of TGF-β1 in cultured medium measured using ELISA assays ((<b>B</b>) <span class="html-italic">n</span> = 5 independent experiments), and TGF-β1 mRNA quantified using RT-qPCR in CFs ((<b>C</b>) <span class="html-italic">n</span> = 4 independent experiments).</p>
Full article ">Figure 3
<p>Impact of NLRP3 and TLR4 signaling on S1 protein-mediated CFs migration. Pretreatment with MCC950 (10 µM (<b>A</b>)) but not TAK-242 (1 µM (<b>B</b>)) blocked the effects of S1 protein (5 nM) treatment for 24 h on cell migration. <span class="html-italic">n</span> = 3 independent experiments.</p>
Full article ">Figure 4
<p>The role of NLRP3 signaling on S1 protein-induced fibrotic markers in CFs. MCC950 effectively blocked the effect of S1 on expressions of fibrotic markers including pro-COL1A1, TGF-β1, and its downstream target, pSMAD2/3. <span class="html-italic">n</span> = 4 independent experiments.</p>
Full article ">Figure 5
<p>Role of TLR4 signaling on S1 protein-induced fibrotic markers in CFs. TAK-242 did not change the effect of S1 on pro-COL1A1 and TGF-β1 expressions in CFs (10 µM, B). <span class="html-italic">n</span> = 4 independent experiments.</p>
Full article ">Figure 6
<p>Role of NF-κB on S1 protein-mediated CFs migration and expressions of fibrotic factors. Pretreatment with BAY 11-7082 (an NF-κB inhibitor, 3 µM) completely blocked the effects of S1 protein (5 nM for 24 h) on CFs migration ((<b>A</b>) <span class="html-italic">n</span> = 5 independent experiments) and the protein expressions of IL1-β cleavage, TGF-β1, and pSMAD2/3 ((<b>B</b>,<b>C</b>) <span class="html-italic">n</span> = 4 independent experiments).</p>
Full article ">Figure 7
<p>S1 activated CFs in an ACE2-dependent manner. ACE2 neutralization antibody (5 µM) effectively blocked the enhanced CFs migration induced by the S1 protein (5 nM for 24 h) (<b>A</b>), along with suppressing the protein expressions of pro-COL1A1, TGF-β1, and phosphorylated NF-κB (p-p65) (<b>B</b>). <span class="html-italic">n</span> = 4 independent experiments.</p>
Full article ">Figure 8
<p>Mechanisms underlying S1 protein-induced activation of CFs and promotion of ECM protein synthesis. Key processes involve activation of the NLRP3 inflammasome, ACE2/NF-κB signaling, and ROS formation. Upon binding to ACE2, the S1 protein initiates a signaling cascade that activates NF-κB, a transcription factor promoting the expression of inflammation-related genes, including those required for NLRP3 inflammasome and pro-interleukin (IL)-1β. ROS formation triggers NLRP3 inflammasome activation, leading to processing of pro-IL-1β into its mature form by caspase-1. Mature IL-1β is released extracellularly, binds to its receptor, and initiates a signaling cascade, enhancing TGF-β1 production, promoting CFs activation, and inducing ECM synthesis, ultimately contributing to cardiac fibrosis. Abbreviations: SARS-CoV-2: severe acute respiratory syndrome coronary virus 2, CFs: cardiac fibroblasts, ECM: extra cellular matrix, EMT: epithelial-mesenchymal transition, NLRP3: NLR family pyrin domain containing 3, ACE2: angiotensin converting enzyme 2, NF-κB: nuclear factor kappa-light-chain-enhancer of activated B cells, ROS: reactive oxygen species, IL-1β: Interleukin 1 beta.</p>
Full article ">
16 pages, 1217 KiB  
Review
Angiotensin II as a Vasopressor for Perioperative Hypotension in Solid Organ Transplant
by Scott T. Benken, Riya Thomas, Dustin R. Fraidenburg and Jamie J. Benken
Biomedicines 2024, 12(8), 1817; https://doi.org/10.3390/biomedicines12081817 - 9 Aug 2024
Viewed by 292
Abstract
During the perioperative period of transplantation, patients experience hypotension secondary to the side effects of anesthesia, surgical stress, inflammatory triggering, and intraoperative fluid shifts, among others causes. Vasopressor support, in this context, must reverse systemic hypotension, but ideally, the agents used should benefit [...] Read more.
During the perioperative period of transplantation, patients experience hypotension secondary to the side effects of anesthesia, surgical stress, inflammatory triggering, and intraoperative fluid shifts, among others causes. Vasopressor support, in this context, must reverse systemic hypotension, but ideally, the agents used should benefit allograft function and avoid the adverse events commonly seen after transplantation. Traditional therapies to reverse hypotension include catecholamine vasopressors (norepinephrine, epinephrine, dopamine, and phenylephrine), but their utility is limited when considering allograft complications and adverse events such as arrhythmias with agents with beta-adrenergic properties. Synthetic angiotensin II (AT2S–[Giapreza]) is a novel vasopressor indicated for distributive shock with a unique mechanism of action as an angiotensin receptor agonist restoring balance to an often-disrupted renin angiotensin aldosterone system. Additionally, AT2S provides a balanced afferent and efferent arteriole vasoconstriction at the level of the kidney and could avoid the arrhythmic complications of a beta-adrenergic agonist. While the data, to date, are limited, AT2S has demonstrated safety in case reports, pilot studies, and small series in the kidney, liver, heart, and lung transplant populations. There are physiologic and hemodynamic reasons why AT2S could be a more utilized agent in these populations, but further investigation is warranted. Full article
Show Figures

Figure 1

Figure 1
<p>Depiction of the renin–angiotensin–aldosterone System (RAAS) and classic metabolic pathway. Low systemic blood pressure (BP) will stimulate kidney renin release to promote angiotensin I (AngI) cleavage from liver-expressed angiotensinogen. This is followed by converting AngI to angiotensin II (AngII) via angiotensin-converting enzyme (ACE). AngII then acts to increase BP via the angiotensin II type 1 receptor through direct vasoconstrictive actions and secondary signaling increasing vasopressin and aldosterone release. The production of AngII then inhibits further renin release through negative biofeedback (shown as a dotted line).</p>
Full article ">Figure 2
<p>Schematic of the mechanism of action of different vasopressors during hypotension surrounding kidney transplant (KT). Panel (<b>A</b>) demonstrates the primary effect of the different vasopressors on the microcirculation of the kidney. Subpanel (i.) shows a primary afferent arteriole vasoconstriction with alpha-1 adrenergic agents. Subpanel (ii.) shows a primary efferent arteriole vasoconstriction with angiotensin II (AT2S). The difference is vital during kidney transplant as the increased endothelin, reduced nitric oxide production, and increased endothelial reactivity to vasoactive substances lead to afferent arteriole vasoconstriction, which could be worsened if using alpha-1 adrenergic vasopressors. Panel (<b>B</b>), subpanel (iii.) shows that while catecholamine vasopressors work through the alpha-1 adrenergic receptor to cause arteriole vasoconstriction, they do not restore imbalances to the renin aldosterone angiotensin system, which could lead to increased amounts of vasodilating byproduct of angiotensin I (AngI) metabolism, angiotensin-1-7 (Ang1-7), leading to ineffective action as a vasopressor. Subpanel (iv.) demonstrates both the vasoconstrictor properties of AT2S via the angiotensin II type I receptor and the restoration of AngI to Ang II balance, leading to a decrease in vasodilatory byproduct of Ang I metabolism, Ang1-7. Figure created with BioRender.com.</p>
Full article ">Figure 3
<p>Alternative angiotensin metabolic pathway. Renin cleaves angiotensinogen to angiotensin I (AngI). This is then converted to angiotensin II (AngII) by angiotensin-converting enzyme (ACE). Angiotensin-(1-7) is created from AngI through the endopeptidases neprilysin (NEP) and thimet oligopeptidase (TOP) and AngII conversion via ACE2. Creating angiotensin-(1-7) leads to vasodilation, a decrease in fibrosis, inflammation, and increased nitric oxide levels. Angiotensin II can be created through tissue chymase via an intermediate metabolite angiotensin-(1-12). The dotted lines represent alternative metabolic pathways.</p>
Full article ">
12 pages, 1428 KiB  
Communication
β-Adrenoceptor Agonists Attenuate Thrombin-Induced Impairment of Human Lung Endothelial Cell Barrier Function and Protect the Lung Vascular Barrier during Resuscitation from Hemorrhagic Shock
by Michelle Y. McGee, Ololade Ogunsina, Sadia N. Boshra, Xianlong Gao and Matthias Majetschak
Biomedicines 2024, 12(8), 1813; https://doi.org/10.3390/biomedicines12081813 - 9 Aug 2024
Viewed by 350
Abstract
β-adrenoceptor (β-AR) agonists are known to antagonize thrombin-induced impairment (TII) of bovine and ovine lung endothelial barrier function. The effects of adrenoceptor agonists and other vasoactive agents on human lung microvascular endothelial cell (HULEC-5a) barrier function upon thrombin exposure have not been studied. [...] Read more.
β-adrenoceptor (β-AR) agonists are known to antagonize thrombin-induced impairment (TII) of bovine and ovine lung endothelial barrier function. The effects of adrenoceptor agonists and other vasoactive agents on human lung microvascular endothelial cell (HULEC-5a) barrier function upon thrombin exposure have not been studied. Furthermore, it is unknown whether the in vitro effects of adrenoceptor agonists translate to lung protective effects in vivo. We observed that epinephrine, norepinephrine, and phenylephrine enhanced normal and prevented TII of HULEC-5a barrier function. Arginine vasopressin and angiotensin II were ineffective. α1B-, α2A/B-, and β1/2-ARs were detectable in HULEC-5a by RT-PCR. Propranolol but not doxazosin blocked the effects of all adrenoceptor agonists. Phenylephrine stimulated β2-AR-mediated Gαs activation with 13-fold lower potency than epinephrine. The EC50 to inhibit TII of HULEC-5a barrier function was 1.8 ± 1.9 nM for epinephrine and >100 nM for phenylephrine. After hemorrhagic shock and fluid resuscitation in rats, Evans blue extravasation into the lung increased threefold (p < 0.01 vs. sham). Single low-dose (1.8 μg/kg) epinephrine administration at the beginning of resuscitation had no effects on blood pressure and reduced Evans blue extravasation by 60% (p < 0.05 vs. vehicle). Our findings confirm the effects of β-adrenoceptor agonists in HULEC-5a and suggest that low-dose β-adrenoceptor agonist treatment protects lung vascular barrier function after traumatic hemorrhagic shock. Full article
(This article belongs to the Special Issue Molecular Mechanisms and Therapeutics in Hemorrhagic Shock)
Show Figures

Figure 1

Figure 1
<p>Effects of adrenoceptor agonists, arginine vasopressin (AVP), and angiotensin II (ATII) on normal (<b>A</b>/<b>B</b>) and thrombin-stimulated (<b>C</b>/<b>D</b>) HULEC-5a barrier function. RFU (%): Relative fluorescence units (RFU) in % of the RFU measured in untreated cell monolayers after 1 h. Bars and error bars indicate mean ± SE. Open circles show the RFU (%) from duplicate measurements from each experiment. (<b>A</b>) HULEC-5a cell monolayers were treated with vehicle (ctrl.) or 5 μM of epinephrine (EPI), norepinephrine (NE), or phenylephrine (PE), and FITC-dextran permeability measured after 1 h and 4 h. <span class="html-italic">n</span> = 3 independent experiments. *: <span class="html-italic">p</span> &lt; 0.05 vs. vehicle-treated cells after 4 h. (<b>B</b>) HULEC-5a cell monolayers were treated with vehicle (ctrl.) or 5 μM of AVP or ATII and FITC-dextran permeability was measured after 1 h and 4 h. <span class="html-italic">n</span> = 3 independent experiments. (<b>C</b>) HULEC-5a cell monolayers were treated with 25 nM thrombin (+) or vehicle (−) plus vehicle drug (−) or 5 μM of EPI, NE or PE, and FITC-dextran permeability measured after 1 h and 4 h. <span class="html-italic">n</span> = 3 independent experiments. *: <span class="html-italic">p</span> &lt; 0.05 vs. cells treated with thrombin plus vehicle at the corresponding time point. #: <span class="html-italic">p</span> &lt; 0.05 vs. EPI and NE at the corresponding time point. (<b>D</b>) HULEC-5a cell monolayers were treated with 25 nM thrombin (+) or vehicle (−) plus vehicle drug (−) or 5 μM AVP or ATII, and FITC-dextran permeability was measured after 1 h and 4 h. <span class="html-italic">n</span> = 3 independent experiments.</p>
Full article ">Figure 2
<p>mRNA expression of adrenergic receptors in HULEC-5a cells detected by reverse transcription (RT)-PCR. Images show the agarose gel electrophoresis of cDNAs amplified by PCR and represent <span class="html-italic">n</span> = 3 independent experiments. M: molecular marker; RT: reverse transcription.</p>
Full article ">Figure 3
<p>Propranolol but not doxazosin antagonizes the effects of EPI and PE on thrombin-induced impairment of HULEC-5a barrier function. RFU (%): Relative fluorescence units (RFU) in % of the RFU measured in untreated cell monolayers after 1 h. Bars and error bars indicate mean ± SE. Open circles show the RFU (%) from duplicate measurements from each experiment. (<b>A</b>/<b>D</b>) HULEC-5a cell monolayers were treated with 25 nM thrombin (+) or vehicle (−) plus vehicle drug (−) or 5 μM of doxazosin (Dox, <b>A</b>) or propranolol (Prop, <b>D</b>) and FITC-dextran permeability measured after 1 h and 4 h. <span class="html-italic">n</span> = 3–4 independent experiments in duplicate. (<b>B</b>) HULEC-5a cell monolayers were treated with 25 nM thrombin (+) or vehicle (−) plus vehicle drug (−) or 5 μM of doxazosin and 5 μM of EPI, and FITC-dextran permeability measured after 1 h and 4 h. <span class="html-italic">n</span> = 3 independent experiments in duplicate. *: <span class="html-italic">p</span> &lt; 0.05 vs. cells treated with thrombin plus vehicle. (<b>C</b>) HULEC-5a cell monolayers were treated with 25 nM thrombin (+) or vehicle (−) plus vehicle drug (−) or 5 μM of doxazosin and 5 μM of PE, and FITC-dextran permeability measured after 1 h and 4 h. <span class="html-italic">n</span> = 3 independent experiments in duplicate. *: <span class="html-italic">p</span> &lt; 0.05 vs. cells treated with thrombin plus vehicle. (<b>E</b>) HULEC-5a cell monolayers were treated with 25 nM thrombin (+) or vehicle (−) plus vehicle drug (−) or 5 μM of propranolol and 5 μM of EPI, and FITC-dextran permeability measured after 1 h and 4 h. <span class="html-italic">n</span> = 3 independent experiments in duplicate. (<b>F</b>) HULEC-5a cell monolayers were treated with 25 nM thrombin (+) or vehicle (−) plus vehicle drug (−) or 5 μM of propranolol and 5 μM of PE, and FITC-dextran permeability measured after 1 h and 4 h. <span class="html-italic">n</span> = 3 independent experiments in duplicate.</p>
Full article ">Figure 4
<p>(<b>A</b>) β<sub>2</sub>-AR mediated Gαs activation monitored by BRET. HEK293T cells were transfected to express β<sub>2</sub>-AR together with GαsS-Rluc8, Gβ3, and Gγ9-GFP2. Activation of β<sub>2</sub>-AR by epinephrine (EPI) or phenylephrine (PE) leads to dissociation of GαsS from Gγ9 and thereby to the reduction of BRET. Data are the mean ± SE from <span class="html-italic">n</span> = 3 independent experiments in duplicate. (<b>B</b>/<b>C</b>) RFU (%): Relative fluorescence units (RFU) in % of the RFU measured in untreated cell monolayers after 1 h. Bars and error bars indicate mean ± SE. Open circles show the RFU (%) from duplicate measurements from each experiment. (<b>B</b>) HULEC-5a cell monolayers were treated with 25 nM thrombin (+) or vehicle (−) in the absence or presence of various concentrations of PE and FITC-dextran permeability measured after 1 h and 4 h. <span class="html-italic">n</span> = 3 independent experiments in duplicate. (<b>C</b>) HULEC-5a cell monolayers were treated with 25 nM thrombin (+) or vehicle (−) in the absence or presence of various concentrations of EPI and FITC-dextran permeability measured after 1 h and 4 h. <span class="html-italic">n</span> = 3 independent experiments in duplicate. *: <span class="html-italic">p</span> &lt; 0.05 vs. cells incubated with thrombin plus vehicle.</p>
Full article ">Figure 5
<p>A single low-dose EPI treatment protects the lung vascular barrier during resuscitation from hemorrhagic shock. Animals were hemorrhaged to a MAP of 30 mmHg for 60 min, followed by crystalloid fluid resuscitation with 1.5 times the shed blood volume. At t = 60 min, animals were treated with EPI (grey symbols, <span class="html-italic">n</span> = 4) or vehicle (open symbols, <span class="html-italic">n</span> = 4). Evans blue was injected at t = 90 min. Animals were euthanized at t = 120 min, and lungs were harvested. Data are the mean ± SE. (<b>A</b>) MAP, mmHg. (<b>B</b>) Hemorrhage volumes (mL/kg) and fluid resuscitation volumes (mL/kg). (<b>C</b>) Typical appearance of lungs from sham-treated animals (left) and animals after hemorrhage and fluid resuscitation with vehicle (hem vehicle, center) or EPI (hem EPI, right) treatment. (<b>D</b>/<b>E</b>) Quantification of Evans blue extravasation into the right (<b>D</b>) and left (<b>E</b>) lungs (μg/mg). Sham, <span class="html-italic">n</span> = 3. Hem vehicle, <span class="html-italic">n</span> = 4. Hem EPI, <span class="html-italic">n</span> = 4. Bars and error bars indicate mean ± SE. Open circles show the individual measurements. The levels of statistical significance are shown in the graphs.</p>
Full article ">
16 pages, 5630 KiB  
Article
Angiotensin II Alters Mitochondrial Membrane Potential and Lipid Metabolism in Rat Colonic Epithelial Cells
by Darby D. Toth, Christopher L. Souder, Sarah Patuel, Cole D. English, Isaac Konig, Emma Ivantsova, Wendi Malphurs, Jacqueline Watkins, Kaylie Anne Costa, John A. Bowden, Jasenka Zubcevic and Christopher J. Martyniuk
Biomolecules 2024, 14(8), 974; https://doi.org/10.3390/biom14080974 - 9 Aug 2024
Viewed by 336
Abstract
An over-active renin-angiotensin system (RAS) is characterized by elevated angiotensin II (Ang II). While Ang II can promote metabolic and mitochondrial dysfunction in tissues, little is known about its role in the gastrointestinal system (GI). Here, we treated rat primary colonic epithelial cells [...] Read more.
An over-active renin-angiotensin system (RAS) is characterized by elevated angiotensin II (Ang II). While Ang II can promote metabolic and mitochondrial dysfunction in tissues, little is known about its role in the gastrointestinal system (GI). Here, we treated rat primary colonic epithelial cells with Ang II (1–5000 nM) to better define their role in the GI. We hypothesized that Ang II would negatively affect mitochondrial bioenergetics as these organelles express Ang II receptors. Ang II increased cellular ATP production but reduced the mitochondrial membrane potential (MMP) of colonocytes. However, cells maintained mitochondrial oxidative phosphorylation and glycolysis with treatment, reflecting metabolic compensation with impaired MMP. To determine whether lipid dysregulation was evident, untargeted lipidomics were conducted. A total of 1949 lipids were detected in colonocytes spanning 55 distinct (sub)classes. Ang II (1 nM) altered the abundance of some sphingosines [So(d16:1)], ceramides [Cer-AP(t18:0/24:0)], and phosphatidylcholines [OxPC(16:0_20:5(2O)], while 100 nM Ang II altered some triglycerides and phosphatidylserines [PS(19:0_22:1). Ang II did not alter the relative expression of several enzymes in lipid metabolism; however, the expression of pyruvate dehydrogenase kinase 2 (PDK2) was increased, and PDK2 can be protective against dyslipidemia. This study is the first to investigate the role of Ang II in colonic epithelial cell metabolism. Full article
Show Figures

Figure 1

Figure 1
<p>Cytotoxicity of Ang II to colonocytes at 72 h. (<b>A</b>) Cytotoxicity, (<b>B</b>) Cell viability. The lysis control was used as a positive control for the assay (induces cell death of colonocytes). The columns represent the mean relative fluorescence ± standard deviation. Different letters denote significant differences from the media-only control (One-way ANOVA, Dunnett multiple comparison test, n = 4/experiment, significance determined at <span class="html-italic">p</span> &lt; 0.05).</p>
Full article ">Figure 2
<p>ATP levels after exposure to Ang II at 72 h. Carbonyl cyanide-4-phenylhydrazone (FCCP) was used as a positive control. Relative luminescence is graphed for each experimental group (horizontal bar represents mean relative luminescence ± standard deviation). Asterisks (****) denotes significant differences from the media-only control (One-way ANOVA followed by a Dunnett multiple comparison test, n = 4/experiment, significance determined at <span class="html-italic">p</span> &lt; 0.001).</p>
Full article ">Figure 3
<p>Mitochondrial membrane potential (MMP) after exposure to Ang II at 72 h. Carbonyl cyanide-4-phenylhydrazone (FCCP) was used as a positive control as it acts as an uncoupling agent for mitochondrial membranes. Relative fluorescence is based on the red/green signal intensity, and all data are normalized to the media-only control (mean relative fluorescence ± standard deviation). Asterisk denotes significant differences compared to the media-only control (one-way ANOVA followed by a Dunnett multiple comparison test, n = 4/experiment, significance determined at * <span class="html-italic">p</span> &lt; 0.05, ** <span class="html-italic">p</span> &lt; 0.01, *** <span class="html-italic">p</span> &lt; 0.001, **** <span class="html-italic">p</span> &lt; 0.0001).</p>
Full article ">Figure 4
<p>Normalized oxygen consumption rate and extracellular acidification rate for rat epithelial colonocytes after a 24 h exposure to Ang II. (<b>A</b>) Oxygen consumption rates over time (<b>B</b>) Acidification rates over time. Data are represented as mean ± standard deviation (one-way ANOVA followed by a Dunnett multiple comparison test, n = 4 replicates/groups, significance determined at <span class="html-italic">p</span> &lt; 0.05).</p>
Full article ">Figure 5
<p>Lipid abundance and categorical classification of lipids in rat epithelial colonocytes (all lipids detected in all treatments). Abbreviations: triglycerides (TG), plasmanyl-TG (plasmanyl-triglycerides), phosphatidylcholine (PC), phosphatidylethanolamines (PE), ceramide (Cer), diacylglycerol (DG), plasmanyl-PC (plasmanyl-phosphatidylcholine), plasmenyl-PE (plasmenyl- phosphatidylethanolamines), phosphatidylserines (PS), plasmenyl-PS (plasmenyl-phosphatidylethanolamines), oxidized phosphatidylcholines (OxPC), phosphatidylglycerol (PG), phosphoinositide (PI), oxidized phosphatidylethanolamines (PE), dimethyl-phosphatidylethanolamine (DMPE), hemibismonoacylglycerophosphate (HBMP), plasmenyl-PC (plasmenyl-phosphatidylcholine), polyethylene glycol (PEG), oxidized lysophosphatidylcholines (OxLPC), lysophosphatidylcholines (LPC), oxidized triglycerides (OxTG), cardiolipins (CL), monomethyl-phosphatidylethanolamine (MMPE), lysophosphatidylethanolamine (LPE), and glucosylceramide non-hydroxyfatty acid-sphingosine (HexCer-NS).</p>
Full article ">Figure 6
<p>(<b>A</b>-Top graph) Principal component analysis scores plot for rat colonocyte lipids with each point representing the lipids in a single sample, the ellipses representing the 95% confidence interval, and the colored groups representing the three different treatments (blue = control, red = low Ang II, and green = high Ang II). (<b>B</b>-bottom graph) Heatmap showing significant changes in the levels of lipids following exposure to Ang II. Data were subjected to ANOVA followed by Fisher’s least significant difference method (Fisher’s LSD), and significant changes were set at <span class="html-italic">p</span> &lt; 0.05.</p>
Full article ">Figure 7
<p>Volcano plots of (<b>A</b>) 1 nM of Ang II and (<b>B</b>) 100 nM of Ang II and the differentially abundant lipids (<span class="html-italic">p</span> &lt; 0.05) outlined in red and blue.</p>
Full article ">Figure 8
<p>Relative concentrations of lipid abundance by sample weight in rat epithelial colonocytes exposed to 1 nM of Ang II (<b>left panel</b>). The most abundant lipids measured include So(d16:1) and Cer-AP(t18:0/24:0). Abbreviations: sphingosine (So), alpha-hydroxy-fatty acid phytosphingosine ceramide (Cer-AP), oxidized lysophosphatidylcholines (OxLPC), and phosphatidylethanolamines (PE). Relative concentrations of lipid abundance by sample weight in rat epithelial colonocytes exposed to 100 nM of Ang II (<b>right panel</b>). The most abundant lipid measured was So(d16:1). Abbreviations: sphingosine (So), oxidized phosphatidylcholines (OxPC), phosphatidylethanolamines (PE), and phosphatidylserines (PS). The black dots represent the metabolite levels in all samples, and the yellow diamond represents the average value.</p>
Full article ">Figure 9
<p>Relative gene expression for rat colonocytes after exposure to Ang II. (<b>a</b>) <span class="html-italic">PDK1,</span> (<b>b</b>) <span class="html-italic">PDK2</span>, (<b>c</b>) <span class="html-italic">PDK4</span>. Data are represented as mean ± standard deviation. Asterisks (**) denote significant differences from the media-only control (data were evaluated using a Mann–Whitney U test, n = 4/experiment, significance determined at <span class="html-italic">p</span> &lt; 0.01).</p>
Full article ">
18 pages, 1606 KiB  
Review
mTOR Dysregulation, Insulin Resistance, and Hypertension
by Silviu Marcel Stanciu, Mariana Jinga, Daniela Miricescu, Constantin Stefani, Remus Iulian Nica, Iulia-Ioana Stanescu-Spinu, Ileana Adela Vacaroiu, Maria Greabu and Silvia Nica
Biomedicines 2024, 12(8), 1802; https://doi.org/10.3390/biomedicines12081802 - 8 Aug 2024
Viewed by 314
Abstract
Worldwide, diabetes mellitus (DM) and cardiovascular diseases (CVDs) represent serious health problems associated with unhealthy diet and sedentarism. Metabolic syndrome (MetS) is characterized by obesity, dyslipidemia, hyperglycemia, insulin resistance (IR) and hypertension. The mammalian target of rapamycin (mTOR) is a serine/threonine kinase with [...] Read more.
Worldwide, diabetes mellitus (DM) and cardiovascular diseases (CVDs) represent serious health problems associated with unhealthy diet and sedentarism. Metabolic syndrome (MetS) is characterized by obesity, dyslipidemia, hyperglycemia, insulin resistance (IR) and hypertension. The mammalian target of rapamycin (mTOR) is a serine/threonine kinase with key roles in glucose and lipid metabolism, cell growth, survival and proliferation. mTOR hyperactivation disturbs glucose metabolism, leading to hyperglycemia and further to IR, with a higher incidence in the Western population. Metformin is one of the most used hypoglycemic drugs, with anti-inflammatory, antioxidant and antitumoral properties, having also the capacity to inhibit mTOR. mTOR inhibitors such as rapamycin and its analogs everolimus and temsirolimus block mTOR activity, decrease the levels of glucose and triglycerides, and reduce body weight. The link between mTOR dysregulation, IR, hypertension and mTOR inhibitors has not been fully described. Therefore, the main aim of this narrative review is to present the mechanism by which nutrients, proinflammatory cytokines, increased salt intake and renin–angiotensin–aldosterone system (RAAS) dysregulation induce mTOR overactivation, associated further with IR and hypertension development, and also mTOR inhibitors with higher potential to block the activity of this protein kinase. Full article
Show Figures

Figure 1

Figure 1
<p>Phosphatidylinositol 3-kinase (PI3K) protein kinases B (AKT)/mammalian target of rapamycin (mTOR) pathway in healthy conditions: Nutrients, growth factors, cytokines and insulin bind to tyrosine kinases receptors (RTKs), leading to insulin receptors substrate 1 or 2 (IRS1/IRS2) activation and further AKT activation by phosphorylation. Once activated, AKT will phosphorylate other protein kinases such as mTOR, composed of the two complexes mTOR complex 1 (mTORC1) and mTOR complex 2 (mTORC2). mTORC1 activates sterol response element binding protein (SREBP) and eukaryotic translation initiator factor 4E binding protein (4EBP) and 70Ka ribosomal protein S56 kinase 1 (p70S6K1), leading to lipid and protein synthesis, respectively. mTORC1 inhibits the activity of unc-51-like kinase 1 (ULK1) and autophagy-related gene 13 (ATG13) blocking autophagy. Inactivation of AKT substrate 160 (AS160) and glycogen synthase 3 (GSK3) induces plasma membrane GLUT translocation. mTORC2 activates other protein kinases such as A, G and C, which positively regulate cellular metabolism. Activation of PI3K/AKT/mTOR will be correlated with cell growth, survival and proliferation. “+” activation.</p>
Full article ">Figure 2
<p>Phosphatidylinositol 3-kinase (PI3K) protein kinase B (AKT)/mammalian target of rapamycin (mTOR) pathway, insulin resistance and hypertension. Hypercaloric diet, branched-chain amino acids (BCAAs), proinflammatory cytokines, free fatty acids (FFAs), increased salt intake, the renin–angiotensin–aldosterone system (RAAS) and salt-inducible kinase (SIK) induce mTOR hyperactivation via RTKs or insulin receptor substrates (IRS1/2). IRS phosphorylation produced by angiotensin II (Ang II) and aldosterone decreases nitric oxide (NO) synthesis. Activation of 70Ka ribosomal protein S6 kinase 1 (p70S6K1) and glycogen synthase 3 (GSK3) inhibits IRS conducing to an increased blood glucose level because GLUT will be blocked inside the cell. mTOR complex 2 (mTORC2) activates serum/glucocorticoid-regulated kinase 1 (SGK1) stimulating Na transport. mTOR over-activation is associated with synthesis of advanced end products (AGEs), reactive oxygen species (ROS), reactive nitrogen species (RNS) and lipids. Metformin has the capacity to inhibit mTOR, while rapamycin, everolimus, temsirolimus and sodium-glucose transporter protein 2 (SGLT2) block mTORC1. All these events will lead to insulin resistance (IR) and further to hypertension. “+” activation; “↓” decrease; “↑” increase.</p>
Full article ">Figure 3
<p>Hyperglycemia is correlated with the activation of the polyol pathway, where glucose is reduced to sorbitol by aldolase reductase (AR) and NADPH. Further, sorbitol will be oxidized to fructose by the enzyme sorbitol dehydrogenase (SDH), and NADH is generated. Fructose will be metabolized into ketone bodies, triose phosphate or carbonylic compounds such as glyoxal, methylglyoxal and 3-deoxyglucose. The last three compounds will contribute to irreversible advanced glycation end product (AGE) formation. NADPH and NADH represent sources for reactive species generation. In hyperglycemic conditions, the hexosamine biosynthesis pathway (HBP) is also activated, leading to the formation of uridine diphosphate N-acetylglucosamine (UDP-GlcNAc), which can induce protein damage. Protein kinase C (PKC) is activated by hyperglycemia, which has the capacity to activate polyol and HBP pathways. PKC activation is associated with decreased levels of nitric oxide (NO) biosynthesis and elevated levels of vascular endothelial growth factor (VEGF). All these molecular events will induce, in the end, mTOR dysregulation. “+” activation; “↑” increase; “↓” decrease.</p>
Full article ">
14 pages, 3416 KiB  
Article
Lactoferrin Supplementation during Pregnancy and Lactation Protects Adult Male Rat Offspring from Hypertension Induced by Maternal Adenine Diet
by You-Lin Tain, Chih-Yao Hou, Wei-Ling Chen, Wei-Ting Liao and Chien-Ning Hsu
Nutrients 2024, 16(16), 2607; https://doi.org/10.3390/nu16162607 - 8 Aug 2024
Viewed by 396
Abstract
Lactoferrin, a glycoprotein derived from breastmilk, is recognized for its health benefits in infants and children; however, its protective effects when administered during gestation and lactation against offspring hypertension remain unclear. This study aimed to investigate whether maternal lactoferrin supplementation could prevent hypertension [...] Read more.
Lactoferrin, a glycoprotein derived from breastmilk, is recognized for its health benefits in infants and children; however, its protective effects when administered during gestation and lactation against offspring hypertension remain unclear. This study aimed to investigate whether maternal lactoferrin supplementation could prevent hypertension in offspring born to mothers with chronic kidney disease (CKD), with a focus on nitric oxide (NO), renin–angiotensin system (RAS) regulation, and alterations in gut microbiota and short-chain fatty acids (SCFAs). Prior to pregnancy, female rats were subjected to a 0.5% adenine diet for 3 weeks to induce CKD. During pregnancy and lactation, pregnant rats received one of four diets: normal chow, 0.5% adenine diet, 10% lactoferrin diet, or adenine diet supplemented with lactoferrin. Male offspring were euthanized at 12 weeks of age (n = 8 per group). Supplementation with lactoferrin during gestation and lactation prevented hypertension in adult offspring induced by a maternal adenine diet. The maternal adenine diet caused a decrease in the index of NO availability, which was restored by 67% with maternal LF supplementation. Additionally, LF was related to the regulation of the RAS, as evidenced by a reduced renal expression of renin and the angiotensin II type 1 receptor. Combined maternal adenine and LF diets altered beta diversity, shifted the offspring’s gut microbiota, decreased propionate levels, and reduced the renal expression of SCFA receptors. The beneficial effects of lactoferrin are likely mediated through enhanced NO availability, rebalancing the RAS, and alterations in gut microbiota composition and SCFAs. Our findings suggest that maternal lactoferrin supplementation improves hypertension in offspring in a model of adenine-induced CKD, bringing us closer to potentially translating lactoferrin supplementation clinically for children born to mothers with CKD. Full article
(This article belongs to the Special Issue Breastmilk for Healthy Development)
Show Figures

Figure 1

Figure 1
<p>Effects of maternal adenine diet (CKD) and lactoferrin (LF) on systolic blood pressure in offspring from Week 3 to 12. N = 8/group. * <span class="html-italic">p</span> &lt; 0.05 vs. CN; # <span class="html-italic">p</span> &lt; 0.05 vs. CKD.</p>
Full article ">Figure 2
<p>Effects of maternal adenine diet (CKD) and lactoferrin (LF) on the renin–angiotensin system at Week 12. N = 8/group. * <span class="html-italic">p</span> &lt; 0.05 vs. CN.</p>
Full article ">Figure 3
<p>The evaluation of gut microbial biodiversity in offspring born to dams fed an adenine (CKD) or lactoferrin (LF) diet. (<b>A</b>) Faith’s phylogenic diversity (pd), (<b>B</b>) Shannon index, and (<b>C</b>) principal coordinate analysis (PCoA). Outliers are denoted by dots. Each color corresponds to a different group, with each data point representing one sample.</p>
Full article ">Figure 4
<p>Linear discriminant analysis effect size (LEfSe) with an LDA score &gt; 4 identified significantly differential taxa between groups. The respective group is denoted by the color of the horizontal bar.</p>
Full article ">Figure 5
<p>Bar plots showing the genus-level discrimination between the CKD and CKDLF groups.</p>
Full article ">Figure 6
<p>The comparison of relative abundance of genus <span class="html-italic">Robinsoniella</span> among the four groups. Outliers are denoted by dots. ** <span class="html-italic">p</span> &lt; 0.01. **** <span class="html-italic">p</span> &lt; 0.001.</p>
Full article ">Figure 7
<p>Plasma concentrations of (<b>A</b>) acetate, (<b>B</b>) propionate, and (<b>C</b>) butyrate, and (<b>D</b>) renal mRNA expression of SCFA receptors at Week 12. N = 8/group. * <span class="html-italic">p</span> &lt; 0.05 vs. CN.</p>
Full article ">
19 pages, 1547 KiB  
Review
The Significance of Endothelial Dysfunction in Long COVID-19 for the Possible Future Pandemic of Chronic Kidney Disease and Cardiovascular Disease
by Hidekatsu Yanai, Hiroki Adachi, Mariko Hakoshima, Hisayuki Katsuyama and Akahito Sako
Biomolecules 2024, 14(8), 965; https://doi.org/10.3390/biom14080965 - 8 Aug 2024
Viewed by 437
Abstract
Various symptoms have been reported to persist beyond the acute phase of severe acute respiratory syndrome coronavirus 2 (SARS-CoV-2) infection, which is referred to as long coronavirus disease 19 (long COVID-19). Over 65 million individuals suffer from long COVID-19. However, the causes of long [...] Read more.
Various symptoms have been reported to persist beyond the acute phase of severe acute respiratory syndrome coronavirus 2 (SARS-CoV-2) infection, which is referred to as long coronavirus disease 19 (long COVID-19). Over 65 million individuals suffer from long COVID-19. However, the causes of long COVID-19 are largely unknown. Since long COVID-19 symptoms are observed throughout the body, vascular endothelial dysfunction is a strong candidate explaining the induction of long COVID-19. The angiotensin-converting enzyme 2 (ACE2), the entry receptor for SARS-CoV-2, is ubiquitously expressed in endothelial cells. We previously found that the risk factors for atherosclerotic cardiovascular disease (ASCVD) and a history of ASCVD raise the risk of severe COVID-19, suggesting a contribution of pre-existing endothelial dysfunction to severe COVID-19. Here, we show a significant association of endothelial dysfunction with the development of long COVID-19 and show that biomarkers for endothelial dysfunction in patients with long COVID-19 are also crucial players in the development of ASCVD. We consider the influence of long COVID-19 on the development of chronic kidney disease (CKD) and ASCVD. Future assessments of the outcomes of long COVID-19 in patients resulting from therapeutic interventions that improve endothelial function may imply the significance of endothelial dysfunction in the development of long COVID-19. Full article
(This article belongs to the Special Issue New Insights into Cardiometabolic Diseases)
Show Figures

Figure 1

Figure 1
<p>The mechanisms underlying the development of severe COVID-19 in patients with metabolic syndrome. Up arrows indicate increases in the expression of molecules, such as ACE2, angiotensin-converting enzyme 2; IL-6, interleukin-6; PAI-1, plasminogen activator inhibitor-1; TNF-α, tumor necrosis factor-alpha; VWF, von Willebrand factor.</p>
Full article ">Figure 2
<p>Various atherogenic factors induced by vascular endothelial dysfunction caused by infection with SARS-CoV-2. Up arrows indicate an increase in degree of atherogenic and thrombogenic factors. ACE2, angiotensin-converting enzyme 2; IL-6, interleukin-6; LDL, low-density lipoprotein; MCP-1, monocyte chemoattractant protein-1; NETs, neutrophil extracellular traps; NOX2, NADPH oxidase 2; PAI-1, plasminogen activator inhibitor-1; ROS, reactive oxygen species; SR, scavenger receptor; TF, tissue factor; tPA, tissue plasminogen activator; VWF, von Willebrand factor.</p>
Full article ">Figure 3
<p>Possible induction of widespread CKD, cardiovascular disease, and ischemic stroke due to endothelial dysfunction in long COVID-19, and the prevention of such problems using therapeutic interventions for long COVID-19 considering endothelial dysfunction as the therapeutic target. ACE2, angiotensin-converting enzyme 2; CAD, coronary artery disease; CKD, chronic kidney disease; ESRD, end-stage renal disease; HF, heart failure.</p>
Full article ">
17 pages, 1223 KiB  
Review
Sex-Specific Differences in Kidney Function and Blood Pressure Regulation
by Eleni Stamellou, Viktor Sterzer, Jessica Alam, Stefanos Roumeliotis, Vassilios Liakopoulos and Evangelia Dounousi
Int. J. Mol. Sci. 2024, 25(16), 8637; https://doi.org/10.3390/ijms25168637 - 8 Aug 2024
Viewed by 300
Abstract
Premenopausal women generally exhibit lower blood pressure and a lower prevalence of hypertension than men of the same age, but these differences reverse postmenopause due to estrogen withdrawal. Sexual dimorphism has been described in different components of kidney physiology and pathophysiology, including the [...] Read more.
Premenopausal women generally exhibit lower blood pressure and a lower prevalence of hypertension than men of the same age, but these differences reverse postmenopause due to estrogen withdrawal. Sexual dimorphism has been described in different components of kidney physiology and pathophysiology, including the renin–angiotensin–aldosterone system, endothelin system, and tubular transporters. This review explores the sex-specific differences in kidney function and blood pressure regulation. Understanding these differences provides insights into potential therapeutic targets for managing hypertension and kidney diseases, considering the patient’s sex and hormonal status. Full article
(This article belongs to the Special Issue Molecular Pathology, Diagnostics, and Therapeutics of Kidney Disease)
Show Figures

Figure 1

Figure 1
<p>Sex-specific pathophysiological differences in kidney function and blood pressure regulation.</p>
Full article ">Figure 2
<p>Overview of suggested sex differences in the renin–angiotensin–aldosterone system. Males typically have a more active ACE/Angiotensin II/AT1R pathway, increasing the risk of hypertension, while females tend to have a more active AT2R/MasR pathway, offering protective effects against hypertension.</p>
Full article ">
11 pages, 5872 KiB  
Communication
Interaction of Receptor-Binding Domain of the SARS-CoV-2 Omicron Variant with hACE2 and Actin
by Ai Fujimoto, Haruki Kawai, Rintaro Kawamura and Akira Kitamura
Cells 2024, 13(16), 1318; https://doi.org/10.3390/cells13161318 - 7 Aug 2024
Viewed by 498
Abstract
The omicron variant of severe acute respiratory syndrome coronavirus 2 (SARS-CoV-2) was identified in 2021 as a variant with heavy amino acid mutations in the spike protein, which is targeted by most vaccines, compared to previous variants. Amino acid substitutions in the spike [...] Read more.
The omicron variant of severe acute respiratory syndrome coronavirus 2 (SARS-CoV-2) was identified in 2021 as a variant with heavy amino acid mutations in the spike protein, which is targeted by most vaccines, compared to previous variants. Amino acid substitutions in the spike proteins may alter their affinity for host viral receptors and the host interactome. Here, we found that the receptor-binding domain (RBD) of the omicron variant of SARS-CoV-2 exhibited an increased affinity for human angiotensin-converting enzyme 2, a viral cell receptor, compared to the prototype RBD. Moreover, we identified β- and γ-actin as omicron-specific binding partners of RBD. Protein complex predictions revealed that many omicron-specific amino acid substitutions affected the affinity between RBD of the omicron variant and actin. Our findings indicate that proteins localized to different cellular compartments exhibit strong binding to the omicron RBD. Full article
Show Figures

Figure 1

Figure 1
<p>Interaction analysis between angiotensin-converting enzyme 2 (ACE2) and receptor-binding domain (RBD) of omicron/prototype using fluorescence cross-correlation spectroscopy (FCCS). (<b>a</b>) Western blotting of the purified recombinant eGFP monomer, hACE2-eGFP, mCherry monomer, ER-mCherry-RBD<sup>Wh1</sup>, and ER-mCherry-RBD<sup>Omic</sup> using anti-GFP and anti-mCherry antibodies (<span class="html-italic">left</span> and <span class="html-italic">right</span>, respectively). Uncropped and unedited blots were provided in <a href="#app1-cells-13-01318" class="html-app">Figures S1 and S2</a>. (<b>b</b>) Sodium dodecyl sulfate-polyacrylamide gel electrophoresis (SDS-PAGE) gel followed by silver staining of all purified samples shown in (<b>a</b>). Numbers on the left side of the gel image indicate the positions of the molecular weight markers. (<b>c</b>) Typical normalized cross-correlation functions ([<span class="html-italic">G</span><sub>c</sub>(τ)−1/<span class="html-italic">G</span><sub>c</sub>(0)−1]) of the mixtures of purified mCherry monomers (Ctrl; gray), ER-mCherry-RBD<sup>Wh1</sup> (Wh1; green), and ER-mCherry-RBD<sup>Omic</sup> (Omic; magenta) with hACE2-eGFP. <span class="html-italic">X</span>-axis shows the lag time (τ). (<b>d</b>) Relative cross-correlation amplitude (RCA) of the indicated two fluorescent color mixtures. (<b>e</b>) Counts per molecule (CPM) of eGFP-tagged proteins via FCCS. (<b>f</b>) CPM of mCherry-tagged proteins via FCCS. (<b>d</b>–<b>f</b>) Mo indicates the GFP or mCherry monomer. Bars indicate the mean ± standard error (SE). Dots indicate the independent values. *** <span class="html-italic">p</span> &lt; 0.001; NS, not significant (<span class="html-italic">p</span> ≥ 0.05). The source data for the graphs (<b>c</b>–<b>f</b>) are provided in <a href="#app1-cells-13-01318" class="html-app">Supplementary Data</a>.</p>
Full article ">Figure 1 Cont.
<p>Interaction analysis between angiotensin-converting enzyme 2 (ACE2) and receptor-binding domain (RBD) of omicron/prototype using fluorescence cross-correlation spectroscopy (FCCS). (<b>a</b>) Western blotting of the purified recombinant eGFP monomer, hACE2-eGFP, mCherry monomer, ER-mCherry-RBD<sup>Wh1</sup>, and ER-mCherry-RBD<sup>Omic</sup> using anti-GFP and anti-mCherry antibodies (<span class="html-italic">left</span> and <span class="html-italic">right</span>, respectively). Uncropped and unedited blots were provided in <a href="#app1-cells-13-01318" class="html-app">Figures S1 and S2</a>. (<b>b</b>) Sodium dodecyl sulfate-polyacrylamide gel electrophoresis (SDS-PAGE) gel followed by silver staining of all purified samples shown in (<b>a</b>). Numbers on the left side of the gel image indicate the positions of the molecular weight markers. (<b>c</b>) Typical normalized cross-correlation functions ([<span class="html-italic">G</span><sub>c</sub>(τ)−1/<span class="html-italic">G</span><sub>c</sub>(0)−1]) of the mixtures of purified mCherry monomers (Ctrl; gray), ER-mCherry-RBD<sup>Wh1</sup> (Wh1; green), and ER-mCherry-RBD<sup>Omic</sup> (Omic; magenta) with hACE2-eGFP. <span class="html-italic">X</span>-axis shows the lag time (τ). (<b>d</b>) Relative cross-correlation amplitude (RCA) of the indicated two fluorescent color mixtures. (<b>e</b>) Counts per molecule (CPM) of eGFP-tagged proteins via FCCS. (<b>f</b>) CPM of mCherry-tagged proteins via FCCS. (<b>d</b>–<b>f</b>) Mo indicates the GFP or mCherry monomer. Bars indicate the mean ± standard error (SE). Dots indicate the independent values. *** <span class="html-italic">p</span> &lt; 0.001; NS, not significant (<span class="html-italic">p</span> ≥ 0.05). The source data for the graphs (<b>c</b>–<b>f</b>) are provided in <a href="#app1-cells-13-01318" class="html-app">Supplementary Data</a>.</p>
Full article ">Figure 2
<p>Co-precipitation of beta- and gamma-actin with omicron RBD. (<b>a</b>) SDS-PAGE gel followed by silver staining of purified ER-mCherry-RBD<sup>Wh1</sup> and ER-mCherry-RBD<sup>Omic</sup>. Asterisk indicates the band specifically co-precipitated with the omicron variant. (<b>b</b>,<b>c</b>) Western blotting of the recombinant mCherry monomer, ER-mCherry-RBD<sup>Wh1</sup>, and ER-mCherry-RBD<sup>Omic</sup> using anti-β-actin and anti-γ-actin antibodies (<b>b</b>,<b>c</b>, respectively). Uncropped and unedited blots were provided in <a href="#app1-cells-13-01318" class="html-app">Figures S3 and S4</a>. Images on the right show the Coomassie brilliant blue (CBB)-stained membranes after antibody detection. Arrowheads indicate the position of β- and γ-actin.</p>
Full article ">Figure 2 Cont.
<p>Co-precipitation of beta- and gamma-actin with omicron RBD. (<b>a</b>) SDS-PAGE gel followed by silver staining of purified ER-mCherry-RBD<sup>Wh1</sup> and ER-mCherry-RBD<sup>Omic</sup>. Asterisk indicates the band specifically co-precipitated with the omicron variant. (<b>b</b>,<b>c</b>) Western blotting of the recombinant mCherry monomer, ER-mCherry-RBD<sup>Wh1</sup>, and ER-mCherry-RBD<sup>Omic</sup> using anti-β-actin and anti-γ-actin antibodies (<b>b</b>,<b>c</b>, respectively). Uncropped and unedited blots were provided in <a href="#app1-cells-13-01318" class="html-app">Figures S3 and S4</a>. Images on the right show the Coomassie brilliant blue (CBB)-stained membranes after antibody detection. Arrowheads indicate the position of β- and γ-actin.</p>
Full article ">Figure 3
<p>Confocal fluorescence images of actin in HeLa cells expressing mCherry-tagged RBD. (<b>a</b>) Fluorescence images of HeLa cells expressing ER-mCherry-RBD<sup>Wh1</sup> or ER-mCherry-RBD<sup>Omic</sup> (mCherry-RBD; magenta) with eGFP-β-actin (eGFP-actin; green). Bar = 10 μm. eGFP and mCherry monomers-expressing HeLa cells were used as a control (<span class="html-italic">upper left</span>). mCherry monomers- and eGFP-actin-expressing cells were also a control (<span class="html-italic">upper right</span>). (<b>b</b>) Fluorescence images of HeLa cells expressing mCherry monomers, ER-mCherry-RBD<sup>Wh1</sup>, or ER-mCherry-RBD<sup>Omic</sup> (mCherry-RBD; magenta) stained with phalloidin-iFluor 488 (Phalloidin; green) and Hoechst 33342 for the nucleus (Hoechst; cyan). Bar = 10 μm.</p>
Full article ">Figure 3 Cont.
<p>Confocal fluorescence images of actin in HeLa cells expressing mCherry-tagged RBD. (<b>a</b>) Fluorescence images of HeLa cells expressing ER-mCherry-RBD<sup>Wh1</sup> or ER-mCherry-RBD<sup>Omic</sup> (mCherry-RBD; magenta) with eGFP-β-actin (eGFP-actin; green). Bar = 10 μm. eGFP and mCherry monomers-expressing HeLa cells were used as a control (<span class="html-italic">upper left</span>). mCherry monomers- and eGFP-actin-expressing cells were also a control (<span class="html-italic">upper right</span>). (<b>b</b>) Fluorescence images of HeLa cells expressing mCherry monomers, ER-mCherry-RBD<sup>Wh1</sup>, or ER-mCherry-RBD<sup>Omic</sup> (mCherry-RBD; magenta) stained with phalloidin-iFluor 488 (Phalloidin; green) and Hoechst 33342 for the nucleus (Hoechst; cyan). Bar = 10 μm.</p>
Full article ">Figure 4
<p>Key amino acids in the predicted complex of omicron RBD and β-actin. (<b>a</b>) Predicted protein complex of the omicron RBD (RBD<sup>Omic</sup>; light magenta) and β-actin (light green). Yellow spheres indicate the missense mutations in the omicron RBD. (<b>b</b>) Enlarged view of the complex of omicron RBD and actin with higher transparency. Black letters indicate the clustered missense amino acids in the RBD within 6 Å from the actin surface.</p>
Full article ">
25 pages, 29683 KiB  
Article
Identification and Validation of the miR/RAS/RUNX2 Autophagy Regulatory Network in AngII-Induced Hypertensive Nephropathy in MPC5 Cells Treated with Hydrogen Sulfide Donors
by Qing Ye, Mi Ren, Di Fan, Yicheng Mao and Yi-Zhun Zhu
Antioxidants 2024, 13(8), 958; https://doi.org/10.3390/antiox13080958 - 7 Aug 2024
Viewed by 356
Abstract
The balanced crosstalk between miRNAs and autophagy is essential in hypertensive nephropathy. Hydrogen sulfide donors have been reported to attenuate renal injury, but the mechanism is unclear. We aimed to identify and verify the miRNAs and autophagy regulatory networks in hypertensive nephropathy treated [...] Read more.
The balanced crosstalk between miRNAs and autophagy is essential in hypertensive nephropathy. Hydrogen sulfide donors have been reported to attenuate renal injury, but the mechanism is unclear. We aimed to identify and verify the miRNAs and autophagy regulatory networks in hypertensive nephropathy treated with hydrogen sulfide donors through bioinformatics analysis and experimental verification. From the miRNA dataset, autophagy was considerably enriched in mice kidney after angiotensin II (AngII) and combined hydrogen sulfide treatment (H2S_AngII), among which there were 109 differentially expressed miRNAs (DEMs) and 21 hub ADEGs (autophagy-related differentially expressed genes) in the AngII group and 70 DEMs and 13 ADEGs in the H2S_AngII group. A miRNA–mRNA–transcription factors (TFs) autophagy regulatory network was then constructed and verified in human hypertensive nephropathy samples and podocyte models. In the network, two DEMs (miR-98-5p, miR-669b-5p), some hub ADEGs (KRAS, NRAS), and one TF (RUNX2) were altered, accompanied by a reduction in autophagy flux. However, significant recovery occurred after treatment with endogenous or exogenous H2S donors, as well as an overexpression of miR-98-5p and miR-669b-5p. The miR/RAS/RUNX2 autophagy network driven by H2S donors was related to hypertensive nephropathy. H2S donors or miRNAs increased autophagic flux and reduced renal cell injury, which could be a potentially effective medical therapy. Full article
Show Figures

Figure 1

Figure 1
<p>The flowchart of screening strategy. Ctrl, control kidney at 28 days. AngII, mice treated with AngII via mini osmotic pump for 28 days. AngII_H<sub>2</sub>S, mice treated with AngII via mini osmotic pump and hydrogen sulfide donor via i.p. injection for 28 days. DEMs, differentially expressed miRNAs. ADEGs, potential autophagy differentially expressed genes. GO, Gene Ontology. KEGG, Kyoto Encyclopedia of Genes and Genomes. PPI, protein–protein interaction. TF, transcription factor.</p>
Full article ">Figure 2
<p>Screening of DEMs in AngII- or H<sub>2</sub>S-treated mouse kidney. Clustered heatmap of DEMs between AngII and Ctrl groups (<b>A</b>) as well as comparison of AngII and H<sub>2</sub>S donor treatments (AngII_H<sub>2</sub>S) and AngII alone group (<b>B</b>). Green label represents Ctrl group, fuchsia represents AngII group, and khaki represents AngII_H<sub>2</sub>S group. The miRNA expression values were processed by log base 2 and are represented from blue to red in ascending order. (<b>C</b>) The volcano plot of DEMs between AngII and Ctrl groups. (<b>D</b>) The volcano plot of DEMs between AngII_H<sub>2</sub>S and AngII groups. Each of the top five miRNAs with the largest difference in forward and reverse changes are marked. Down means downregulated miRNAs, and these are shown as blue dots. Up, presented as red dots, means upregulated miRNAs. The gray dots show the not significantly changed miRNAs. The criteria were absolute FC &gt; 1.5 and <span class="html-italic">p</span> value &lt; 0.05.</p>
Full article ">Figure 3
<p>Hub target genes and ADEGs were selected for DEMs associated with AngII alone or AngII plus H<sub>2</sub>S donor treatment. Venn diagrams showing predicted target genes (red) and validated target genes (sky blue) corresponding to DEMs in comparison between AngII vs. Ctrl (<b>A</b>) and AngII_H<sub>2</sub>S vs. AngII (<b>B</b>). PPI network displaying the interaction of hub target genes from comparison between AngII vs. Ctrl (<b>C</b>), and AngII_H<sub>2</sub>S vs. AngII (<b>D</b>). The interacting genes between hub target genes and autophagy-related gene set in AngII vs. Ctrl (<b>E</b>) and AngII_H<sub>2</sub>S vs. AngII (<b>F</b>) comparison.</p>
Full article ">Figure 4
<p>Enrichment plots from GO and KEGG functional analysis of targeted ADEGs. Barplots of the enrichment analysis for each of the top 10 biological processes (green), cellular component (orange-red), and molecular function (blue) from AngII vs. Ctrl (<b>A</b>) and AngII_H<sub>2</sub>S vs. AngII (<b>B</b>). In addition, the barplots are arranged from smallest to largest according to the negative logarithm of the adjusted <span class="html-italic">p</span> value. The KEGG pathway enrichment analysis results of AngII vs. Ctrl (<b>C</b>) and AngII_H<sub>2</sub>S vs. AngII (<b>D</b>) are shown as dotplots. Among them, the size of the dots represents the number of ADEGs hitting the corresponding pathway, and the adjusted <span class="html-italic">p</span> value of each pathway is displayed from red to blue in ascending order. CellP., cellular processes; EnvIP., environmental information processing; HumaD., human diseases; OrgaS., organismal systems.</p>
Full article ">Figure 5
<p>Protein–protein interaction networks and core modules of ADEGs constructed by STRING and Cytoscape. PPI analysis of targeted ADEGs from AngII vs. Ctrl (<b>A</b>) and AngII_H<sub>2</sub>S vs. AngII (<b>B</b>). The network illustration of top 10 hub ADEGs explored by CytoHubba in comparison between AngII vs. Ctrl (<b>C</b>) and AngII_H<sub>2</sub>S vs. AngII (<b>D</b>). Hub ADEGs are ranked by Degree value from dark red to orange red. Core modules analyzed by molecular complex detection network clustering analysis in AngII vs. Ctrl (<b>E</b>) and AngII_H<sub>2</sub>S vs. AngII (<b>F</b>).</p>
Full article ">Figure 6
<p>Regulatory networks of the DEMs, target hub ADEGs, and transcription factors. (<b>A</b>) The miRNA-mRNA autophagy regulatory network in AngII-induced hypertensive nephropathy and hydrogen sulfide donor drug therapy. Octagons, rectangles, and ovals represent diseases/drugs, DEMs, and target hub ADEGs, respectively. The red fonts represent miRNAs and mRNAs co-regulated by AngII and H<sub>2</sub>S donors. The miRNA-mRNA-TF autophagy regulatory network of miR-98-5p (<b>B</b>) and miR-669b-5p (<b>C</b>). Green octagons represent transcription factors, blue ovals indicate target genes regulated by miRNAs, and purple ovals show target genes regulated by both miRNAs and transcription factors.</p>
Full article ">Figure 7
<p>H<sub>2</sub>S pathway in AngII-treated MPC5 cells. (<b>A</b>) Western blot analysis of nephrin, podocin, and GAPDH in MPC5 cells under different stimuli. <span class="html-italic">n</span> = 3. (<b>B</b>) Normalized mRNA expression of Tnf-α, il-1β, and il-6. <span class="html-italic">n</span> = 3. (<b>C</b>) H<sub>2</sub>S content using spectrophotometer; <span class="html-italic">n</span> = 4. (<b>D</b>) H<sub>2</sub>S levels using fluorescent probe WSP-1. Bar = 50 μm; <span class="html-italic">n</span> = 5. (<b>E</b>) Western blot analysis of CBS, CSE, MST, and GAPDH in MPC5 cells under indicated stimuli; <span class="html-italic">n</span> = 3. (<b>F</b>) Normalized mRNA expression of <span class="html-italic">Cbs</span>, <span class="html-italic">Cse</span>, and <span class="html-italic">Mst</span>; <span class="html-italic">n</span> = 4. # <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 vs. AngII group, * <span class="html-italic">p</span> &lt; 0.05, ** <span class="html-italic">p</span> &lt; 0.01, *** <span class="html-italic">p</span> &lt; 0.001 vs. Ctrl group.</p>
Full article ">Figure 8
<p>Validation of co-regulated DEMs, ADEGs, and TFs in hypertensive nephropathy clinical samples and AngII-treated MPC5 cells. (<b>A</b>) Normalized mRNA expression of <span class="html-italic">KRAS</span>, <span class="html-italic">MAPK1</span>, <span class="html-italic">NRAS</span>, and <span class="html-italic">PTEN</span> in the tubulointerstitial region (left) and glomeruli (right) of human renal biopsies. Data were extracted from GSE37455 and GSE37460. # <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 vs. healthy living donor group using two-sided unpaired <span class="html-italic">t</span>-test; <span class="html-italic">n</span> = 21 and 20 (<b>left</b>) or 27 and 15 (<b>right</b>) in the healthy and hypertensive group. (<b>B</b>,<b>C</b>) RNA expression of miR-98-5p and miR-669b-5p were measured in MPC5 cells by qPCR. (<b>D</b>) The mRNA levels of <span class="html-italic">Kras</span>, <span class="html-italic">Mapk1</span>, <span class="html-italic">Nras</span>, <span class="html-italic">Pten</span>, <span class="html-italic">Pik3ca</span>, <span class="html-italic">Pik3r3</span>, <span class="html-italic">Mapk9</span>, and <span class="html-italic">Runx2</span> in MPC5 cells were calculated using qPCR. For (<b>B</b>,<b>D</b>), the MPC5 cells were pre-incubated with or without 2mM PAG for 30 min and then pre-treated with NaHS (50 μM) or SPRC (50 μM) for 1 h before exposure to AngII (1 μM) treatment. NaHS, SPRC, and SPRC + PAG groups were all treated with AngII. # <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 vs. Ctrl group, * <span class="html-italic">p</span> &lt; 0.05, ** <span class="html-italic">p</span> &lt; 0.01, *** <span class="html-italic">p</span> &lt; 0.001 vs. AngII group, and @@ <span class="html-italic">p</span> &lt; 0.01, @@@ <span class="html-italic">p</span> &lt; 0.001 vs. SPRC group. For (<b>C</b>), MPC5 cells were pre-stimulated with or without losartan (300 μM) for 12 h and then co-incubated with or without AngII (1 μM). ### <span class="html-italic">p</span> &lt; 0.001 vs. Ctrl group, ** <span class="html-italic">p</span> &lt; 0.01 vs. AngII group. One-way ANOVA coupled with Tukey’s multiple comparison post hoc test was used. Data are shown as mean ± SD, <span class="html-italic">n</span> = 3. ns, not significant.</p>
Full article ">Figure 9
<p>Overexpression of miR-98 and miR-669b. (<b>A</b>) RNA expression of miR-98-5p and miR-669b-5p were measured in MPC5 cells by qPCR. (<b>B</b>) The mRNA levels of <span class="html-italic">Kras</span>, <span class="html-italic">Mapk1</span>, <span class="html-italic">Nras</span>, <span class="html-italic">Pten</span>, <span class="html-italic">Pik3ca</span>, <span class="html-italic">Pik3r3</span>, <span class="html-italic">Mapk9</span>, and <span class="html-italic">Runx2</span> in MPC5 were calculated using qPCR. MPC5 cells were transfected with miR-NC, miR-98-5p mimic, and miR-669b-5p mimic. Forty-eight hours after infection, cells were incubated with or without AngII (1 μM) to detect RNA expression. Both miR-98-5p and miR-669b-5p mimic groups were incubated with AngII. # <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 vs. miR-NC group, * <span class="html-italic">p</span> &lt; 0.05, ** <span class="html-italic">p</span> &lt; 0.01, *** <span class="html-italic">p</span> &lt; 0.001 vs. AngII group. One-way ANOVA coupled with Tukey’s multiple-comparison post hoc test was used. Data are shown as mean ± SD, <span class="html-italic">n</span> = 3.</p>
Full article ">Figure 10
<p>Validation of co-regulated DEMs, ADEGs, and TFs in protein levels. (<b>A</b>) Western blot analysis of PTEN, KRAS, NRAS, MAPK1, RUNX2, P62, and GAPDH protein expression levels in MPC5 cells under different stimuli. Cells were pre-incubated with or without PAG and then pre-treated with NaHS or SPRC for 1 h before exposure to AngII treatment. NaHS, SPRC, and SPRC + PAG groups were all treated with AngII. Expression levels were calculated by densitometric analysis and normalized to GAPDH. Protein quantification is shown in bar charts. # <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 vs. Ctrl group, * <span class="html-italic">p</span> &lt; 0.05, ** <span class="html-italic">p</span> &lt; 0.01, *** <span class="html-italic">p</span> &lt; 0.001 vs. AngII group, and @@ <span class="html-italic">p</span> &lt; 0.01, @@@ <span class="html-italic">p</span> &lt; 0.001 vs. SPRC group. For (<b>B</b>–<b>D</b>), MPC5 cells were transfected with miR-NC, miR-98-5p mimic, and miR-669b-5p mimic and then incubated with AngII to detect the expression of autophagy markers P62 and LC3B (<b>B</b>), renal injury markers nephrin and podocin (<b>C</b>), and hub targets such as KRAS, NRAS, and MAPK1 as well as the TF RUNX2 (<b>D</b>) by Western blot analysis. Both miR-98-5p and miR-669b-5p mimic groups were incubated with AngII. With GAPDH as a normalized reference, protein quantification is shown in bar charts. Dark green, red, blue, and orange represent miR-NC, AngII, miR-98-5p, and miR-669b-5p mimic groups, respectively. # <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 vs. miR-NC group, * <span class="html-italic">p</span> &lt; 0.05, ** <span class="html-italic">p</span> &lt; 0.01, *** <span class="html-italic">p</span> &lt; 0.001 vs. AngII group. One-way ANOVA coupled with Tukey’s multiple comparison post hoc test was used. Data are shown as mean ± SD, <span class="html-italic">n</span> = 3.</p>
Full article ">
17 pages, 1278 KiB  
Article
Evaluation of In Vitro Antihypertensive and Anti-Inflammatory Properties of Dairy By-Products
by Eleni Dalaka, Georgios C. Stefos, Ioannis Politis and Georgios Theodorou
Appl. Sci. 2024, 14(16), 6885; https://doi.org/10.3390/app14166885 - 6 Aug 2024
Viewed by 449
Abstract
Sweet whey (SW) and yogurt acid whey (YAW) are dairy by-products of the cheese-making process and Greek-style yogurt production, respectively. Both of them are considered pollutants with huge volumes of SW and YAW produced due to the growing demand for dairy products worldwide. [...] Read more.
Sweet whey (SW) and yogurt acid whey (YAW) are dairy by-products of the cheese-making process and Greek-style yogurt production, respectively. Both of them are considered pollutants with huge volumes of SW and YAW produced due to the growing demand for dairy products worldwide. Moreover, whey-derived peptides, resulting from fermentation as well as from further hydrolysis during digestion, have been associated with various biological activities. In the present study, the angiotensin-converting enzyme (ACE)-inhibitory activity of 48 SW samples and 33 YAW samples from bovine, ovine, caprine, and ovine/caprine milk obtained were evaluated. Additionally, the SW and YAW digestates and two of their fractions (smaller than 10 kDa, SW-D-P10 and YAW-D-P10, and smaller than 3 kDa, SW-D-P3 and YAW-D-P3), which were obtained after in vitro digestion and subsequent ultrafiltration, were also subjected to evaluation. Our data indicated that the D-P10 and D-P3 fractions exhibited higher ACE-inhibitory activity compared to the corresponding values before digestion. The ACE-inhibitory capacity after in vitro digestion was higher for the ovine SW samples compared to their bovine and caprine counterparts. The effect of the D-P3 fraction on the inhibition of nitric oxide (NO) production and the expression of a selected panel of immune-response-related genes in LPS-stimulated RAW 264.7 macrophages was also evaluated. Fractions from both dairy by-products inhibited NO production in LPS-stimulated RAW 264.7 cells. Especially, ovine SW-D-P3 showed a strong NO inhibitory activity and suppressed inducible nitric oxide synthase (Nos2) mRNA levels. However, YAW-D-P3 could not trigger neither the gene expression of inflammatory macrophage mediators Nos2 and cyclooxygenase-2 (Ptgs2) nor tumor necrosis factor-α (Tnf) and interleukin 6 (Il6) in LPS-stimulated murine macrophages regardless of animal origin. These findings suggest that in vitro digestion could enhance the production of ACE-inhibitory peptides in both dairy by-products, while SW from ovine origin displays higher potential as an anti-inflammatory agent, effectively preventing excessive NO production. Full article
(This article belongs to the Special Issue Innovation in Dairy Products)
Show Figures

Figure 1

Figure 1
<p>Effect of animal origin of sweet whey (SW) on angiotensin-converting enzyme (ACE)-inhibitory activity (%) before and after in vitro digestion. The vertical dashed lines separate the three different groups as follows: SW: sweet whey before digestion; SW-D-P10: digested fraction below 10 kDa after digestion; SW-D-P3: digested fraction below 3 kDa. Data are presented as means ± SEM of three independent experiments. Columns with different letters significantly differ within the same group (<span class="html-italic">p</span> &lt; 0.05). * indicates a significant difference between two groups (<span class="html-italic">p</span> &lt; 0.05); ns indicates no significant difference between two groups (<span class="html-italic">p</span> &gt; 0.05).</p>
Full article ">Figure 2
<p>Effect of animal origin of yogurt acid whey (YAW) on angiotensin-converting enzyme (ACE)-inhibitory activity (%) before and after in vitro digestion. The vertical dashed lines separate the three different groups as follows: YAW: yogurt acid whey before digestion; YAW-D-P10: digested fraction below 10 kDa after digestion; YAW-D-P3: digested fraction below 3 kDa. Data are presented as means ± SEM of three independent experiments. Columns with different letters significantly differ within the same group (<span class="html-italic">p</span> &lt; 0.05). * indicates a significant difference between two groups (<span class="html-italic">p</span> &lt; 0.05); ns indicates no significant difference between two groups (<span class="html-italic">p</span> &gt; 0.05).</p>
Full article ">Figure 3
<p>Effect of SW-D-P3 on LPS-induced NO production in RAW 264.7 cells. RAW 264.7 cells were treated as follows: LPS, cells stimulated only with 1 μg/mL LPS; D-P3, cells stimulated with LPS and subsequently incubated in the presence of SW-D-P3 (caprine, ovine, bovine, and mix; 0.038% <span class="html-italic">w</span>/<span class="html-italic">v</span> refers to starting SW-D protein) or BL-D-P3 (blank) for 24 h. Data are presented as means ± SEM of four independent experiments. Columns with different letters are significantly different (<span class="html-italic">p</span> &lt; 0.05).</p>
Full article ">Figure 4
<p>Effect of YAW-D-P3 on LPS-induced NO production in RAW 264.7 cells. RAW 264.7 cells were treated as follows: LPS, cells stimulated only with 1 μg/mL LPS; D-P3, cells stimulated with LPS and subsequently incubated in the presence of YAW-D-P3 (caprine, ovine, and bovine; 0.011% <span class="html-italic">w</span>/<span class="html-italic">v</span> refers to starting YAW-D protein) or BL-D-P3 (blank) for 24 h. Data are presented as means ± SEM of four independent experiments. Columns with different letters are significantly different (<span class="html-italic">p</span> &lt; 0.05).</p>
Full article ">Figure 5
<p>Effect of SW-D-P3 on LPS-induced mRNA expression in RAW 264.7 cells. RAW 264.7 cells were treated as follows: LPS, cells stimulated only with 1 μg/mL LPS; D-P3, cells stimulated with LPS and subsequently incubated in the presence of SW-D-P3 (caprine, ovine, bovine, and mix; 0.038% <span class="html-italic">w</span>/<span class="html-italic">v</span> refers to starting SW-D protein) or BL-D-P3 (blank) for 24 h. The expression levels of (<b>a</b>) <span class="html-italic">Nos2</span>, (<b>b</b>) <span class="html-italic">Ptgs2</span>, (<b>c</b>) <span class="html-italic">Tnf</span>, and (<b>d</b>) <span class="html-italic">Il6</span> were measured using qPCR and were normalized to housekeeping genes (Actb and Cyc1). Data are presented as mean ± SEM of two independent experiments. Columns with different letters within the same panel are significantly different (<span class="html-italic">p</span> &lt; 0.05); ns = not significant (<span class="html-italic">p</span> &gt; 0.05).</p>
Full article ">Figure 6
<p>Effect of YAW-D-P3 on LPS-induced mRNA expression in RAW 264.7 cells. RAW 264.7 cells were treated as follows: LPS, cells stimulated only with 1 μg/mL LPS; D-P3, cells stimulated with LPS and subsequently incubated in the presence of YAW-D-P3 (caprine, ovine, and bovine; 0.011% <span class="html-italic">w</span>/<span class="html-italic">v</span> refers to starting YAW-D protein) or BL-D-P3 (blank) for 24 h. The expression levels of (<b>a</b>) <span class="html-italic">Nos2</span>, (<b>b</b>) <span class="html-italic">Ptgs2</span>, (<b>c</b>) <span class="html-italic">Tnf</span>, and (<b>d</b>) <span class="html-italic">Il6</span> were measured using qPCR and were normalized to housekeeping genes (Actb and Cyc1). Data are presented as mean ± SEM of two independent experiments. ns = not significant (<span class="html-italic">p</span> &gt; 0.05).</p>
Full article ">
12 pages, 1229 KiB  
Article
Effects of Angiotensin-Converting Enzyme Inhibition on the Recurrence and Internal Structure of Chronic Subdural Hematomas
by Michael Veldeman, Hani Ridwan, Mohamed Alzaiyani, Rastislav Pjontek, Benedikt Kremer, Anke Hoellig, Hans Clusmann and Hussam Hamou
J. Clin. Med. 2024, 13(16), 4591; https://doi.org/10.3390/jcm13164591 - 6 Aug 2024
Viewed by 442
Abstract
Background/Objectives: Chronic subdural hematoma (cSDH) is a common disease of growing significance due to the increasing use of antithrombotic drugs and population aging. There exists conflicting observational evidence that previous treatment with angiotensin-converting enzyme (ACE) inhibitors reduces the rate of cSDH recurrence. This [...] Read more.
Background/Objectives: Chronic subdural hematoma (cSDH) is a common disease of growing significance due to the increasing use of antithrombotic drugs and population aging. There exists conflicting observational evidence that previous treatment with angiotensin-converting enzyme (ACE) inhibitors reduces the rate of cSDH recurrence. This study assesses the hypothesis that ACE inhibitors may affect recurrence rates by altering hematoma membrane formation. Methods: All patients with chronic subdural hematoma who were operated upon in a single university hospital between 2015 and 2020 were considered for inclusion. Hematomas were classified according to their structural appearance in computed tomography (CT) imaging into one of eight subtypes. Patients’ own medication, prior to hospitalization for cSDH treatment, was noted, and the use of ACI-inhibitors was identified. Results: Of the included 398 patients, 142 (35.9%) were treated with ACE inhibitors before admission for cSDH treatment. Of these, 115 patients (81.0%) received ramipril, 13 received patients lisinopril (11.3%), and 11 patients (9.6%) received enalapril. Reflecting cardiovascular comorbidity, patients on ACE inhibitors were more often simultaneously treated with antithrombotics (63.4% vs. 42.6%; p ≤ 0.001). Hematomas with homogenous hypodense (OR 11.739, 95%CI 2.570 to 53.612; p = 0.001), homogenous isodense (OR 12.204, 95%CI 2.669 to 55.798; p < 0.001), and homogenous hyperdense (OR 9.472, 95%CI 1.718 to 52.217; p < 0.001) architectures, as well as the prior use of ACE inhibitors (OR 2.026, 95%CI 1.214 to 3.384; p = 0.007), were independently associated with cSDH recurrence. Conclusions: Once corrected for hematoma architecture, type of surgery, and use of antithrombotic medication, preoperative use of ACE inhibitors was associated with a twofold increase in the likelihood of hematoma recurrence. Full article
(This article belongs to the Section Brain Injury)
Show Figures

Figure 1

Figure 1
<p>Eight chronic subdural hematoma types based on the appearance of internal architecture in computed tomography imaging (adapted from Hamou et al. [<a href="#B11-jcm-13-04591" class="html-bibr">11</a>]).</p>
Full article ">Figure 2
<p>Flowchart of patients’ inclusion. ACI, angiotensin-converting enzyme; cSDH, chronic subdural hematoma.</p>
Full article ">Figure 3
<p>The literature overview of factors associated with the recurrence of chronic subdural hematoma in patients with or without prior treatment with angiotensin-converting enzyme inhibitors. <span class="html-italic">p</span>-values below the preset α-level of 0.05 are marked in bold [<a href="#B8-jcm-13-04591" class="html-bibr">8</a>,<a href="#B9-jcm-13-04591" class="html-bibr">9</a>,<a href="#B10-jcm-13-04591" class="html-bibr">10</a>,<a href="#B14-jcm-13-04591" class="html-bibr">14</a>].</p>
Full article ">
21 pages, 1442 KiB  
Review
Receptors Involved in COVID-19-Related Anosmia: An Update on the Pathophysiology and the Mechanistic Aspects
by Noor N. Al-Saigh, Amani A. Harb and Shtaywy Abdalla
Int. J. Mol. Sci. 2024, 25(15), 8527; https://doi.org/10.3390/ijms25158527 - 5 Aug 2024
Viewed by 542
Abstract
Olfactory perception is an important physiological function for human well-being and health. Loss of olfaction, or anosmia, caused by viral infections such as severe acute respiratory syndrome coronavirus 2 (SARS-CoV-2), has received considerable attention, especially in persistent cases that take a long time [...] Read more.
Olfactory perception is an important physiological function for human well-being and health. Loss of olfaction, or anosmia, caused by viral infections such as severe acute respiratory syndrome coronavirus 2 (SARS-CoV-2), has received considerable attention, especially in persistent cases that take a long time to recover. This review discusses the integration of different components of the olfactory epithelium to serve as a structural and functional unit and explores how they are affected during viral infections, leading to the development of olfactory dysfunction. The review mainly focused on the role of receptors mediating the disruption of olfactory signal transduction pathways such as angiotensin converting enzyme 2 (ACE2), transmembrane protease serine type 2 (TMPRSS2), neuropilin 1 (NRP1), basigin (CD147), olfactory, transient receptor potential vanilloid 1 (TRPV1), purinergic, and interferon gamma receptors. Furthermore, the compromised function of the epithelial sodium channel (ENaC) induced by SARS-CoV-2 infection and its contribution to olfactory dysfunction are also discussed. Collectively, this review provides fundamental information about the many types of receptors that may modulate olfaction and participate in olfactory dysfunction. It will help to understand the underlying pathophysiology of virus-induced anosmia, which may help in finding and designing effective therapies targeting molecules involved in viral invasion and olfaction. To the best of our knowledge, this is the only review that covered all the receptors potentially involved in, or mediating, the disruption of olfactory signal transduction pathways during COVID-19 infection. This wide and complex spectrum of receptors that mediates the pathophysiology of olfactory dysfunction reflects the many ways in which anosmia can be therapeutically managed. Full article
(This article belongs to the Section Molecular Pathology, Diagnostics, and Therapeutics)
Show Figures

Figure 1

Figure 1
<p>Organization of the olfactory epithelium and its communication with the olfactory bulb (prepared with <a href="http://Biorender.com" target="_blank">http://Biorender.com</a>).</p>
Full article ">Figure 2
<p>(<b>A</b>) Entry of SARS-CoV-2 into the sustentacular cell and the molecular interaction of the virus spike protein and the host cellular receptor ACE2, which depends on spike protein activation by TMPRSS2 and/or furin [<a href="#B46-ijms-25-08527" class="html-bibr">46</a>,<a href="#B49-ijms-25-08527" class="html-bibr">49</a>,<a href="#B53-ijms-25-08527" class="html-bibr">53</a>,<a href="#B59-ijms-25-08527" class="html-bibr">59</a>]. (<b>B</b>) Viral invasion activates Panx1, promoting the release of ATP from the cell [<a href="#B60-ijms-25-08527" class="html-bibr">60</a>]. (<b>C</b>) Extracellular ATP binds to P2 receptors, causing an increase in intracellular Ca<sup>2+</sup> and ROS, which in turn results in DNA damage and cell death. The elevated intracellular Ca<sup>2+</sup> and DNA damage collectively stimulate inflammatory responses, including the synthesis of cytokines like TNF-α and IL-1ß [<a href="#B61-ijms-25-08527" class="html-bibr">61</a>]. (<b>D</b>) Intercellular communication of SC with OSN through gap junctions, creating Ca<sup>2+</sup> waves in the OSN. (<b>E</b>) Increasing intracellular Ca<sup>2+</sup> in OSN activates Ca<sup>2+</sup>-dependent K<sup>+</sup> channels that may mediate a hyperpolarization and subsequently inhibit the odor response or activate Cl<sup>−</sup> channels, reducing the negativity in the cell and causing a depolarization. ACE2, angiotensin-converting enzyme 2 receptor; AC, adenylyl cyclase; ATP, adenosine triphosphate; cAMP, cyclic adenosine monophosphate; CD147, basigin; CNG, cyclic nucleotide-gated channel; IL-1ß, interleukin 1 ß; NRP-1, neuropilin-1; OR, olfactory receptor; OSN, olfactory sensory neuron; PANX1, pannexin 1; P2X4, P2X7, purinergic receptors; ROS, reactive oxygen species; SC, sustentacular cell; TMPRSS 2, transmembrane protease serine type 2; TNF-α, tumor necrosis factor-alpha. This figure was created with <a href="http://Biorender.com" target="_blank">http://Biorender.com</a>.</p>
Full article ">Figure 3
<p>TRPV1 participates in anosmia associated with SARS-CoV-2 infection. (1) SARS-CoV-2 infection activates the immune response; (2) inflammatory cells produce proinflammatory cytokines, such as TNF-α, which binds to its receptor (TNF-αR) expressed on the trigeminal ganglion neuron in the OE; (3) TNF-α activates intracellular cascades, including MAPK, p38, and other translational factors, to increase CGRP synthesis and increases the expression and trafficking of TRPV1 to the plasma membrane, thus triggering CGRP release. (4) CGRP binds to its receptor (CGRPR) expressed on OSN. (5) CGRPR activates adenylyl cyclase, which in turn converts ATP to cAMP; (6) cAMP activates a CNG channel, allowing Ca<sup>2+</sup> influx; and (7) Ca<sup>2+</sup> activates a Ca<sup>2+</sup>-dependent K<sup>+</sup> channel (K<sub>Ca</sub>), leading to (8) hyperpolarization and inhibition of odor detection. TGN: trigeminal neuron. AC: adenylyl cyclase; CGRP: calcitonin-related gene peptide; CNG: cyclic nucleotide gated channel; TRPV: transient receptor potential vanilloid; TNFR: TNF-α receptor. TGN: trigeminal nerve ending; OSN: olfactory sensory neuron.</p>
Full article ">
Back to TopTop