Nothing Special   »   [go: up one dir, main page]

You seem to have javascript disabled. Please note that many of the page functionalities won't work as expected without javascript enabled.
 
 
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

Article Types

Countries / Regions

remove_circle_outline
remove_circle_outline
remove_circle_outline
remove_circle_outline
remove_circle_outline
remove_circle_outline

Search Results (1,607)

Search Parameters:
Keywords = calcium channels

Order results
Result details
Results per page
Select all
Export citation of selected articles as:
10 pages, 1914 KiB  
Article
Nitric Oxide and Small and Intermediate Calcium-Activated Potassium Channels Mediate the Vasodilation Induced by Apigenin in the Resistance Vessels of Hypertensive Rats
by Lislaine Maria Klider, Maria Luiza Fidelis da Silva, Gustavo Ratti da Silva, João Ricardo Cray da Costa, Marcia Alessandra Arantes Marques, Emerson Luiz Botelho Lourenço, Francislaine Aparecida dos Reis Lívero, Jane Manfron and Arquimedes Gasparotto Junior
Molecules 2024, 29(22), 5425; https://doi.org/10.3390/molecules29225425 (registering DOI) - 18 Nov 2024
Viewed by 227
Abstract
Background: Apigenin (4′,5,7-trihydroxyflavone), a flavonoid with potential cardiovascular benefits, has unclear mechanisms of action. This study investigates its effects on vascular function in Spontaneously Hypertensive Rats (SHRs). Methods: Mesenteric vascular beds (MVBs) were isolated from SHRs and perfused with increasing doses of apigenin [...] Read more.
Background: Apigenin (4′,5,7-trihydroxyflavone), a flavonoid with potential cardiovascular benefits, has unclear mechanisms of action. This study investigates its effects on vascular function in Spontaneously Hypertensive Rats (SHRs). Methods: Mesenteric vascular beds (MVBs) were isolated from SHRs and perfused with increasing doses of apigenin after pre-contraction with phenylephrine. To explore the mechanisms, different MVBs were pre-perfused with antagonists and inhibitors, including indomethacin, L-NAME, and potassium channel blockers (tetraethylammonium, a non-specific potassium channel blocker; glibenclamide, an ATP-sensitive potassium channel blocker; 4-aminopyridine, a voltage-gated potassium channel blocker; charybdotoxin a selective intermediate-conductance calcium-activated potassium channel blocker; and apamin, a selective small-conductance calcium-activated potassium channel blocker). Results: Apigenin induced a dose-dependent reduction in perfusion pressure in MVBs with intact endothelium, an effect abolished by endothelium removal. L-NAME reduced apigenin-induced vasodilation by approximately 40%. The vasodilatory effect was blocked by potassium chloride and tetraethylammonium. The inhibition of small and intermediate calcium-activated potassium channels with charybdotoxin and apamin reduced apigenin-induced vasodilation by 50%, and a combination of these blockers with L-NAME completely inhibited the effect. Conclusions: Apigenin promotes vasodilation in resistance arteries through endothelial nitric oxide and calcium-activated potassium channels. These findings suggest that apigenin could have therapeutic potential in cardiovascular disease, warranting further clinical research. Full article
(This article belongs to the Special Issue Analyses and Applications of Phenolic Compounds in Food—2nd Edition)
Show Figures

Figure 1

Figure 1
<p>Apigenin induces vasodilation in resistance arteries of rats. Molecular structure of apigenin (<b>A</b>). Effects of apigenin on the mesenteric vascular beds of WKY and SHRs (<b>B</b>,<b>C</b>). Typical representative record of the administration of acetylcholine (1 nmol) and apigenin (0.1, 0.3, and 1 µmol) in preparations of mesenteric vascular bed of SHRs (<b>D</b>). Values represent the mean ± standard error of the mean (<span class="html-italic">n</span> = 6 preparations). + <span class="html-italic">p</span> &lt; 0.05 compared with the previously administered dose. All experiments were performed in endothelium-intact preparations.</p>
Full article ">Figure 2
<p>The vasodilatory effects of apigenin are dependent on vascular endothelium and nitric oxide release. Vasodilatory action of apigenin in the presence (End+) and absence of endothelium (End-) are presented (<b>A</b>). Apigenin’s vasodilatory action was investigated in the presence of the nitric oxide synthase inhibitor (L-NAME) (<b>B</b>), regarding the formation of cGMP (<b>C</b>), or during the inhibition of the enzyme cyclooxygenase (indomethacin) (<b>D</b>). Values represent the mean ± standard error of the mean (<span class="html-italic">n</span> = 6 preparations). * <span class="html-italic">p</span> &lt; 0.05 compared with preparations in the presence of endothelium (<b>A</b>) or after treatment only with vehicle (<b>B</b>,<b>C</b>). + <span class="html-italic">p</span> &lt; 0.05 compared with the respective previous dose.</p>
Full article ">Figure 3
<p>The vasodilatory effects of apigenin depend on potassium channels. Apigenin’s vasodilatory action was investigated in the presence of 40 mM KCl (<b>A</b>), tetraethylammonium (TEA) (<b>B</b>), glibenclamide (<b>C</b>), 4-aminopyridine (4-AP) (<b>D</b>), charybdotoxin (ChTx) (<b>E</b>), or apamin (Apm) (<b>F</b>). The results show the mean ± S.E.M. of six preparations per group. * Indicate <span class="html-italic">p</span> &lt; 0.05 compared with the effects of apigenin on the vehicle group. + indicates <span class="html-italic">p</span> &lt; 0.05 compared with the respective previous dose. All experiments were performed in endothelium-intact preparations.</p>
Full article ">Figure 4
<p>The vasodilator effects of apigenin are dependent on the release of nitric oxide and the activation of potassium channels. Apigenin’s vasodilatory action was investigated in the presence of charybdotoxin (ChTx) plus apamin (Apm) (<b>A</b>), and charybdotoxin (ChTx) plus apamin (Apm) plus L-NAME (<b>B</b>). The results show the mean ± S.E.M. of six preparations per group. * Indicate <span class="html-italic">p</span> &lt; 0.05 compared with the effects of apigenin on the vehicle group. + indicates <span class="html-italic">p</span> &lt; 0.05 compared with the respective previous dose. All experiments were performed in endothelium-intact preparations.</p>
Full article ">
20 pages, 2243 KiB  
Article
New Pharmacological Insight into Etanercept and Pregabalin in Allodynia and Nociception: Behavioral Studies in a Murine Neuropathic Pain Model
by Loulwah Alothman, Emad Alhadlaq, Asma Alhussain, Alwaleed Alabdulkarim, Youssef Sari and Shakir D. AlSharari
Brain Sci. 2024, 14(11), 1145; https://doi.org/10.3390/brainsci14111145 - 15 Nov 2024
Viewed by 301
Abstract
Background/Objectives: Neuropathic pain is challenging to treat, often resistant to current therapies, and associated with significant side effects. Pregabalin, an anticonvulsant that modulates calcium channels, is effective but can impair mental and motor functions, especially in older patients. To improve patient outcomes, reducing [...] Read more.
Background/Objectives: Neuropathic pain is challenging to treat, often resistant to current therapies, and associated with significant side effects. Pregabalin, an anticonvulsant that modulates calcium channels, is effective but can impair mental and motor functions, especially in older patients. To improve patient outcomes, reducing the doses of pregabalin and combining it with other drugs targeting different neuropathic pain mechanisms may be beneficial. TNF-α blockers such as etanercept have shown potential in addressing neuropathic pain by affecting sodium channels, synaptic transmission, and neuroinflammation. This study evaluates the efficacy and safety of combining low doses of etanercept and pregabalin in allodynia and nociceptive tests. Materials and Methods: Male C57/BL6 mice underwent chronic constriction injury (CCI) of the sciatic nerve to induce neuropathic pain. They were divided into seven groups: sham control, CCI control, low and high doses of pregabalin, low and high doses of etanercept, and a combination of low doses of both drugs. Behavioral tests, including von Frey, hot-plate, and rotarod tests, were used to assess pain responses and motor activity. Results: The results indicated that a high dose of pregabalin significantly reduced mechanical allodynia and thermal hyperalgesia but impaired motor function. Conversely, low doses of etanercept alone had no significant effect. However, the combination of low doses of etanercept (20 mg/kg) and pregabalin (5 mg/kg) effectively alleviated pain without compromising locomotor activity. Conclusions: These results suggest a novel therapeutic strategy for neuropathic pain, enhancing analgesic efficacy while minimizing adverse effects. Full article
(This article belongs to the Section Neuroscience of Pain)
Show Figures

Figure 1

Figure 1
<p>The flowchart of the study design. Baseline pain-related behaviors and motor co-ordination tests were performed on each study group. Neuropathic pain was induced using the chronic constriction of the sciatic nerve (CCI) technique. On day 3 post-surgery, pain-related behavior and motor co-ordination tests were conducted. On day 4 post-surgery, tactile allodynia was assessed using the von Frey test at 60, 120, and 180 min following the intraperitoneal (i.p.) administration of low and high doses of pregabalin, low and high doses of etanercept, a combination of low doses of pregabalin and etanercept, and saline (control). On day 5 post-surgery, the plantar thermal stimulation test and rotarod test were performed 60 min after drug and saline administration. On day 6 post-surgery, both the hot-plate test and locomotor test were conducted 60 min after drug and saline administration. Figure created in BioRender. Alothman, L. (2024) <a href="https://BioRender.com/n71a150" target="_blank">https://BioRender.com/n71a150</a> (accessed on 31 October 2024).</p>
Full article ">Figure 2
<p>Mean withdrawal thresholds in the von Frey test for postoperative day 4, at different time points (60, 120, and 180 min after i.p. drug administration) among seven groups: sham control, CCI control, pregabalin at 5 mg/kg, pregabalin at 30 mg/kg, etanercept at 20 mg/kg, etanercept at 40 mg/kg, and combination treatment of low doses of pregabalin and etanercept. The data are expressed as the mean ± SD (<span class="html-italic">n</span> = 8 per group). Statistical significance was assessed using two-way ANOVA followed by Tukey’s post hoc test. * Statistically significant (<span class="html-italic">p</span> &lt; 0.05) for groups “combination treatment” and “pregabalin at 30 mg/kg” (filled legends) compared to other groups at time 120 min. # Statistically significant (<span class="html-italic">p</span> &lt; 0.05) for the groups of combination treatment, pregabalin at 30 mg/kg (filled legends), and the sham control compared to other groups at time 180 min.</p>
Full article ">Figure 3
<p>The comparison of mean values of paw withdrawal latency (in seconds) at three time points (baseline, day 0, and day 5) among seven groups: sham control, CCI control, pregabalin at 5 mg/kg, pregabalin at 30 mg/kg, etanercept at 20 mg/kg, etanercept at 40 mg/kg, and combination treatment. All drug treatments were administered intraperitoneally (i.p.). The data are expressed as the mean ± SD of eight mice, and the two-way ANOVA measurement followed by Tukey’s post hoc test was used to determine statistical significance. * Statistically significant (<span class="html-italic">p</span> &lt; 0.05) for the sham control group compared to all other groups at day 0. # Statistically significant (<span class="html-italic">p</span> &lt; 0.05) for combination treatment and pregabalin at 30 mg/kg compared to CCI control, pregabalin at 5 mg/kg, etanercept at 20 mg/kg, and etanercept at 40 mg/kg at day 5. @ Statistically significant (<span class="html-italic">p</span> &lt; 0.05) for pregabalin at 5 mg/kg and etanercept at 40 mg/kg compared to CCI control and etanercept at 20 mg/kg at day 5.</p>
Full article ">Figure 4
<p>The comparison of mean paw withdrawal latency (in seconds) at three time points (baseline, day 0, and day 6) among seven groups: sham control, CCI control, pregabalin at 5 mg/kg, pregabalin at 30 mg/kg, etanercept at 20 mg/kg, etanercept at 40 mg/kg, and combination treatment of low doses of each drug. All drug treatments were administered intraperitoneally (i.p.). The data are expressed as the mean ± SD of eight mice, and the two-way ANOVA measurement followed by Tukey’s post hoc test was used to determine statistical significance. * Statistically significant (<span class="html-italic">p</span> &lt; 0.05) for the sham control group compared to all other groups at day 0. @ Statistically significant (<span class="html-italic">p</span> &lt; 0.05) for combination treatment, pregabalin at 5 mg/kg, etanercept at 40 mg/kg, and sham control compared to CCI control and etanercept at 20 mg/kg at day 6. # Statistically significant (<span class="html-italic">p</span> &lt; 0.05) for pregabalin at 30 mg/kg compared to sham control, CCI control, etanercept at 20 mg/kg, and etanercept at 40 mg/kg at day 6.</p>
Full article ">Figure 5
<p>The comparison of mean locomotor counts at three time points (baseline, day 0, and day 6) among seven groups: sham control, CCI control, pregabalin at 5 mg/kg, pregabalin at 30 mg/kg, etanercept at 20 mg/kg, etanercept at 40 mg/kg, and combination treatment of pregabalin and etanercept. All drug treatments were administered intraperitoneally (i.p.). The data are expressed as the mean ± SD of eight mice, and the two-way ANOVA measurement followed by Tukey’s post hoc test was used to determine statistical significance. * Statistically significant (<span class="html-italic">p</span> &lt; 0.05) for the sham control group compared to all other groups at day 0. @ Statistically significant (<span class="html-italic">p</span> &lt; 0.05) for combination treatment and sham control compared to CCI control and pregabalin at 30 mg/kg. # Statistically significant (<span class="html-italic">p</span> &lt; 0.05) for pregabalin at 5 mg/kg, etanercept at 20 mg/kg, and etanercept at 40 mg/kg compared to pregabalin at 30 mg/kg at day 6.</p>
Full article ">Figure 6
<p>The comparison of mean latency to fall (in seconds) at three time points (baseline, day 0, and day 4) among seven groups: sham control, CCI control, pregabalin at 5 mg/kg, pregabalin at 30 mg/kg, etanercept at 20 mg/kg, etanercept at 40 mg/kg, and combination treatment. All drug treatments were administered intraperitoneally (i.p.). The data are expressed as the mean ± SD of eight mice, and the two-way ANOVA measurement followed by Tukey’s post hoc test was used to determine statistical significance. * Statistically significant (<span class="html-italic">p</span> &lt; 0.05) for the sham control group compared to all other groups at day 0 and day 5. @ Statistically significant (<span class="html-italic">p</span> &lt; 0.05) for combination treatment and pregabalin at 5 mg/kg compared to CCI control, etanercept at 20 mg/kg, etanercept at 40 mg/kg, and pregabalin at 30 mg/kg at day 5. # Statistically significant (<span class="html-italic">p</span> &lt; 0.05) for pregabalin at 30 mg/kg and etanercept at 40 mg/kg compared to CCI control and etanercept at 20 mg/kg at day 5.</p>
Full article ">
14 pages, 1015 KiB  
Review
Connexins and Aging-Associated Respiratory Disorders: The Role in Intercellular Communications
by Tatiana Zubareva, Ekaterina Mironova, Anna Panfilova, Yulia Krylova, Gianluigi Mazzoccoli, Maria Greta Pia Marasco, Igor Kvetnoy and Peter Yablonsky
Biomedicines 2024, 12(11), 2599; https://doi.org/10.3390/biomedicines12112599 - 13 Nov 2024
Viewed by 347
Abstract
This article reviews the contemporary understanding of the functional role of connexins in intercellular communications, their involvement in maintaining cellular and tissue homeostasis, and in aging-associated respiratory disease pathogenesis. Connexins are discussed as potential therapeutic targets. The review particularly focuses on the involvement [...] Read more.
This article reviews the contemporary understanding of the functional role of connexins in intercellular communications, their involvement in maintaining cellular and tissue homeostasis, and in aging-associated respiratory disease pathogenesis. Connexins are discussed as potential therapeutic targets. The review particularly focuses on the involvement of gap junction connexins and hemichannels in the transfer of calcium ions, metabolite molecules, ATP, and mitochondria through the cell membrane. Various disorders in the regulation of intercellular communication can heavily contribute to the pathogenesis of multiple diseases, including respiratory system diseases. A deeper understanding of molecular mechanisms underlying the activities of various connexins in gap junction channels will enable the prospective development of therapeutic approaches by either inhibiting or stimulating the activities of a certain connexin, while considering its critical functions in intercellular communications on the whole. Full article
Show Figures

Figure 1

Figure 1
<p>The intracellular stages of gap junction formation. The connexins are synthesized as monomers in the endoplasmic reticulum, then transported to the Goldgi apparatus, where they assemble into hexameric structures, connexons, which are then exported to the cell membrane, forming a hemichannel. The connexons of the neighboring cells form an intercellular channel of gap junctions by connecting with each other.</p>
Full article ">Figure 2
<p>The regulation of gap junction operation. The opening and closing of gap junctions are regulated by extracellular and intracellular factors such as pH, calcium ion concentration, and connexin phosphorylation. Gap junctions serve to move ions and small molecules up to 1.2 kDa between adjacent cells. The cells can exchange molecules through gap junctions such as carbohydrates, nucleotides, second messengers (cAMP or cGMP), small peptides, and RNA.</p>
Full article ">Figure 3
<p>Intracellular localization and intercellular functioning of connexins. Abbreviation: AG—Apparatus of Goldgi, N—Nucleus, ER—Endoplasmic Reticulum, GJ—Gap Junction Channel, PM—Plasma Membrane.</p>
Full article ">
22 pages, 4662 KiB  
Article
An Immune-Independent Mode of Action of Tacrolimus in Promoting Human Extravillous Trophoblast Migration Involves Intracellular Calcium Release and F-Actin Cytoskeletal Reorganization
by Ahmad J. H. Albaghdadi, Wei Xu and Frederick W. K. Kan
Int. J. Mol. Sci. 2024, 25(22), 12090; https://doi.org/10.3390/ijms252212090 - 11 Nov 2024
Viewed by 397
Abstract
We have previously reported that the calcineurin inhibitor macrolide immunosuppressant Tacrolimus (TAC, FK506) can promote the migration and invasion of the human-derived extravillous trophoblast cells conducive to preventing implantation failure in immune-complicated gestations manifesting recurrent implantation failure. Although the exact mode of action [...] Read more.
We have previously reported that the calcineurin inhibitor macrolide immunosuppressant Tacrolimus (TAC, FK506) can promote the migration and invasion of the human-derived extravillous trophoblast cells conducive to preventing implantation failure in immune-complicated gestations manifesting recurrent implantation failure. Although the exact mode of action of TAC in promoting implantation has yet to be elucidated, the integral association of its binding protein FKBP12 with the inositol triphosphate receptor (IP3R) regulated intracellular calcium [Ca2+]i channels in the endoplasmic reticulum (ER), suggesting that TAC can mediate its action through ER release of [Ca2+]i. Using the immortalized human-derived first-trimester extravillous trophoblast cells HTR8/SVneo, our data indicated that TAC can increase [Ca2+]I, as measured by fluorescent live-cell imaging using Fluo-4. Concomitantly, the treatment of HTR8/SVneo with TAC resulted in a major dynamic reorganization in the actin cytoskeleton, favoring a predominant distribution of cortical F-actin networks in these trophoblasts. Notably, the findings that TAC was unable to recover [Ca2+]i in the presence of the IP3R inhibitor 2-APB indicate that this receptor may play a crucial role in the mechanism of action of TAC. Taken together, our results suggest that TAC has the potential to influence trophoblast migration through downstream [Ca2+]i-mediated intracellular events and mechanisms involved in trophoblast migration, such as F-actin redistribution. Further research into the mono-therapeutic use of TAC in promoting trophoblast growth and differentiation in clinical settings of assisted reproduction is warranted. Full article
(This article belongs to the Special Issue Physiology and Pathophysiology of Placenta 2.0)
Show Figures

Figure 1

Figure 1
<p>Tacrolimus spiked [Ca<sup>2+</sup>]i release in live HTR8/Svneo cells. (<b>A</b>–<b>H</b>): Representative photomicrographs of individual live HTR8/SVneo cells visualized by intravital confocal microscopy depicting the intracellular contents of Ca<sup>2+</sup> in DMSO-treated control ((<b>A</b>): green), ionomycin ((<b>B</b>): blue), TAC-treated ((<b>C</b>): red), 2-APB + TAC ((<b>D</b>): yellow), U73122 + TAC ((<b>E</b>): purple), Wortmannin + TAC ((<b>F</b>): orange), BAPTA + TAC ((<b>G</b>): turquoise) and EGTA + TAC ((<b>H</b>): green), respectively. The timeframe for Ca<sup>2+</sup> imaging was for a total of 540 s. [Ca<sup>2+</sup>]i release was spiked at the end of minute 3 of live-image tracing (depicted by the blue arrow) by the addition of ionomycin (<b>B</b>), or TAC alone (<b>C</b>) or in the presence of other inhibitors of [Ca<sup>2+</sup>]i release and Ca<sup>2+</sup> chelators (<b>D</b>–<b>G</b>). Compared to the DMSO-treated (<b>A</b>) and Ionomycin-treated controls (<b>B</b>), the single use of TAC (10 ng/mL) resulted in a significant increase in [Ca<sup>2+</sup>]i in the HTR8/SVneo cells (compare the intensity of red color in (<b>C</b>) before and after the addition of TAC). The inability of TAC to spike [Ca<sup>2+</sup>]i-release in the presence of the IP3R antagonist 2-APB (<b>D</b>) suggests a crucial role for IP3R in mediating the [Ca<sup>2+</sup>]i modulatory actions of TAC. Unexpectedly, the TAC-induced [Ca<sup>2+</sup>]i-release in HTR8/SVneo cells was unaffected by inhibitory actions of the PLC inhibitor U73122 (<b>E</b>) as well as the potent and specific phosphatidylinositol 3-kinase (PI3-K) inhibitor Wortmannin (<b>F</b>). The intracellular source of the TAC-induced [Ca<sup>2+</sup>]i-release is confirmed by the use of intracellular [Ca<sup>2+</sup>]i chelator BPATA (<b>G</b>) and EGTA (<b>H</b>). Scale bars: (<b>A</b>–<b>H</b>) 30 µm; [Ca<sup>2+</sup>]i is depicted in pseudocolors in (<b>A</b>–<b>H</b>).</p>
Full article ">Figure 2
<p>Real-time tracing of [Ca<sup>2+</sup>]i-release in live HTR8/SVneo cells. (<b>A</b>–<b>F</b>): Fold change in Mean Fluorescence Intensity (MFI) of Fluo-4 as a measurement of [Ca<sup>2+</sup>]i-release in HTR8/SVneo cells receiving TAC (<b>A</b>) in the presence and absence of the [Ca<sup>2+</sup>]i-release inhibitors (2-APB: (<b>B</b>), U73122: (<b>C</b>), Wortmannin: (<b>D</b>)), as well as the Ca<sup>2+</sup>chelators BAPTA (<b>E</b>), and EGTA (<b>F</b>). (<b>A</b>) depicts the influence of TAC administration on spiking [Ca<sup>2+</sup>]i-release (black arrow) in a comparable intensity to that of the calcium ionophore Ionomycin. (<b>B</b>) depicts pre-incubation with the IP3R inhibitor 2-APB significantly suppressed TAC-induced [Ca<sup>2+</sup>]i-release in HTR8/SVneo cells (<span class="html-italic">p</span> &lt; 0.05). (<b>C</b>,<b>D</b>): Unlike the suppressive effects of the 2-APB on TAC-induced [Ca<sup>2+</sup>]i-release, the inhibition of the phospholipase C (PLC) or the phosphatidylinositol 3-kinase (PI3K) by use of the compound U73122 (<b>C</b>) and Wortmannin (<b>D</b>) did not restrict this cellular action of TAC. In (<b>E</b>,<b>F</b>), the spiked TAC-induced [Ca<sup>2+</sup>]i-release in HTR8/SVneo cells concomitantly treated with the intracellular and extracellular calcium chelators BAPTA (<b>E</b>) and EGTA (<b>F</b>) suggests the release of [Ca<sup>2+</sup>]i is from the intracellular stores, namely the endoplasmic reticulum. In (<b>A</b>–<b>F</b>), HTR8/SVneo cells were pre-incubated for 10 min with the [Ca<sup>2+</sup>]i-release inhibitors prior to the administration of TAC. Real-time tracing of [Ca<sup>2+</sup>]i-release in Ionomycin-treated HTR8/SVneo cells was included in (<b>A</b>) for a comparison. Black arrows in (<b>A</b>–<b>F</b>) depict spiked [Ca<sup>2+</sup>]i-release. The time recording was 10 min.</p>
Full article ">Figure 3
<p>Tacrolimus influences F-actin cytoskeletal re-arrangement in the human-derived first-trimester extravillous trophoblast cells. (<b>A</b>–<b>C</b>): Single cell confocal images of control (<b>A</b>) and TAC-treated HTR8/SVneo cells (<b>B</b>,<b>C</b>) labeled with the CellMask Green<sup>TM</sup> Actin tracking stain. F-actin is mostly distributed in the form of stress fibers running across the cell body of untreated cells (white arrows in (<b>A1</b>,<b>A2</b>)). 10 min pre-incubation with TAC resulted in a global reorganization of the F-actin filaments manifested in the formation of cortical fibers (white arrows in (<b>B1</b>,<b>B2</b>)). Notably, filopodia-like structures (white arrows in (<b>C1</b>,<b>C2</b>)) were observed among TAC-treated HTR8/SVneo cells evidently demonstrating a tangible outcome of the influence of TAC on F-actin cytoskeletal reorganization suggestive of cell motility. Green: CellMask Green<sup>TM</sup>-labled F-actin, Blue: DAPI-stained nuclei. Scale bars: (<b>A</b>–<b>C</b>) 10 µm. Nuclei were counterstained with DAPI in (<b>A</b>–<b>C</b>).</p>
Full article ">Figure 4
<p>The influence of [Ca<sup>2+</sup>]i-release inhibitors and chelators on the structural distribution of F-actin in TAC-treated HTR8/SVneo cells. (<b>A</b>–<b>F</b>): Time-dependent cytoskeletal reorganization of F-actin in the HTR8/SVneo cells in response to Ionomycin (<b>A1</b>,<b>A2</b>), TAC (<b>B1</b>,<b>B2</b>), 2-APB (<b>C1</b>,<b>C2</b>), U73122 (<b>D1</b>,<b>D2</b>), Wortmannin (<b>E1</b>,<b>E2</b>) and PABTA (<b>F1</b>,<b>F2</b>), respectively. Note the characteristic distribution of the stress fibers throughout the cytoplasm (solid white arrows) versus the peripherally condensed cortical fiber (dashed white arrows). Failure of F-actin cellular reorganization in the 2-APB-inhibited cells ((<b>C1</b>) vs. (<b>C2</b>)) indicates the dependence of TAC actions on the functional IP3R-signaling pathway. Distinctly, unlike pre-incubation with U73122 (<b>D1</b>,<b>D2</b>) and Wortmannin (<b>E1</b>,<b>E2</b>), the presence of 2-APB (<b>C1</b>,<b>C2</b>) and PABTA (<b>F1</b>,<b>F2</b>) compromised the structural integrity and consequently the visualization of the F-actin cytoskeleton in HTR8/SVneo cells. The recording after the addition of TAC to the inhibitor pre-treated cells was 6 min. Green: CellMask Green<sup>TM</sup>-labled F-actin, Blue: DAPI-stained nuclei. Scale bars: (<b>A1</b>–<b>F2</b>) 45 µm.</p>
Full article ">Figure 5
<p>Tacrolimus negatively influenced the co-localization of IP3R and FKBP12 in HTR8/SVneo cells. (<b>A</b>–<b>E</b>): Representative confocal images for the immunofluorescent detection of IP3R-1 ((<b>A1</b>–<b>D1</b>): red *) and FKBP12 ((<b>A2</b>–<b>D2</b>): green *) and their co-localization ((<b>A3</b>–<b>D4</b>); firey orange, and co-localization analysis bar graphs in (<b>E</b>)) in PFA-fixed monolayers of control (DMSO-treated) and TAC-treated HTR8/SVneo cells depicting a significant reduction (<span class="html-italic">p</span> = 0.025) in the Pearson’s correlation coefficient of mean fluorescent intensities (MFI) of the co-localization of these two protein components of the ER [Ca<sup>2+</sup>]i-release channels after 1 h of exposure to TAC (<b>E</b>). Indeed, Pearson’s correlation quantification revealed that IP3R-I and FKBP12 are co-expressed in the same pixel in control cells more than in TAC-treated cells **. Note that the characteristic perinuclear distribution of these two proteins in HTR8/SVeno cells (white arrows in (<b>A4</b>–<b>D4</b>), respectively). (<b>F</b>): representative Western blot (Wb) detection of IP3R and FKBP12 in control (experimental repeats C1–C3 in lanes C1–C3) and TAC-treated (experimental repeats T1–T9 in lanes T1–T9) HTR8/SVneo cells, revealing a preservative effect of TAC in maintaining the levels of these two protein components of the ER [Ca<sup>2+</sup>]i-release channels in trophoblasts as measured by the lack of a significant fold change in their protein band intensities compared to untreated control cells (<span class="html-italic">p</span> &gt; 0.05). The IP3R-1 bar graphs in (<b>F</b>) exclusively depict the intensities of the Wb protein bands at the 210 kDa molecular weight. Images in (<b>A4</b>–<b>D4</b>) are higher magnifications of the yellow-boxed cellular areas in (<b>A3</b>–<b>D3</b>), demonstrating the peri-nuclear distribution and co-localization of IP3R-1 and FKBP12 in control and TAC-treated HTR8/SVneo cells, respectively. (<b>A</b>,<b>B</b>): Representative brightfield images of control and TAC-treated HTR8/SVEneo cells depicting their general morphology and their blue-colored DAPI-stained nuclei, respectively. Scale bars: (<b>A</b>–<b>D4</b>) 5 µm. Nu (<b>A4</b>–<b>D4</b>): Nuclei. TAC: Tacrolimus. ns in (<b>F</b>): Not statistically significant (<span class="html-italic">p</span> &gt; 0.05). *: Alexa-Fluor 790-conjugated anti-FKBP12 (anti-FKBP12-AF790; mouse anti-human) and Alexa-Fluor 680-conjugated IP3R-I (anti-IP3R-1-AF680; mouse anti-human) primary antibodies were used for the detection of their respective proteins labeled in the confocal images of HTR8/SVneo cells shown in (<b>A1</b>–<b>D4</b>). Due to current technical limitations with the wavelength detection of the 790 fluorochrome, anti-FKBP12-AF790-labeled monolayers of these cells were allowed a brief incubation with FITC-labeled goat-anti-mouse antibody suspension prior to re-incubation with the anti-IP3R-1-AF680 as described in the methods section. **: The quantification was performed by comparing all the individual frames (one cell per frame; four cells per experiment; a total of six experiments; therefore, 24 cells per treatment). Scatter blots in (<b>E</b>) represent the average rate of four cells imaged in randomly selected 5 high-power fields (HPFs) in an experiment (n = 6 plates (30 mm) per treatment).</p>
Full article ">Figure 6
<p>(<b>A</b>): Schematic depicting of the inositol triphosphate receptor (IP3R), which is an intracellular Ca<sup>2+</sup>-release channel located on the membrane of the endoplasmic reticulum (ER), and which belongs to the same family of the ryanodine receptors (RyRs). The conserved and widely abundant immunophilin FKBP12, which is a primary receptor for the immunosuppressant actions of TAC (FK506), has been demonstrated to physiologically interact with the inositol 1,4,5-trisphosphate receptor (IP3R) via a leucyl-prolyl dipeptide epitope that structurally resembles TAC (FK506). Here, we are postulating that TAC binding to FKBP12, likely through its structural mimicry to dipeptide epitopes on the FKBP12, sequesters this immunophilin from the IP3R, thus structurally destabilizing the channel conducive to a spiked release of [Ca<sup>2+</sup>]i from ER stores (arrow). Abbreviations: TAC (FK506): tacrolimus; IP3R: inositol triphosphate receptor; ER: endoplasmic reticulum. (<b>B</b>): Schematic depicting of the inositol triphosphate receptor (IP3R) [Ca<sup>2+</sup>]i-release pathway in trophoblast cells. The illustration depicts a potential mechanism through which TAC may influence [Ca<sup>2+</sup>]i-release along the IP3R pathway and its putative intracellular signal transduction pathways involved in F-actin cytoskeletal reorganization in trophoblast cells. [Ca<sup>2+</sup>]i-release in trophoblasts is normally a function of the G-protein-coupled receptor (GPCR)-mediated activation of phospholipase C (PLC) and the membrane-bound PI3K (which produces inositol triphosphate (IP3)). IP3 is a ligand for the intracellular IP3R channel of the internal Ca<sup>2+</sup> stores of the endoplasmic reticulum (ER). It is postulated that TAC influences [Ca<sup>2+</sup>]i-release via its binding to the immunophilin FKBP12, plausibly resulting in the destabilization of the ER’s IP3R [Ca<sup>2+</sup>]i-release channels. The observation that TAC was unable to release [Ca<sup>2+</sup>]i in trophoblast cells in the presence of the IP3R inhibitor 2-APB suggests a major role for this RYR channel in the presently proposed mode of action of TAC. This notion is also supported by the ability of TAC to release [Ca<sup>2+</sup>]i in the presence of the PI3K inhibitor Wortmannin, and the PLC inhibitor U73122. Moreover, PLC activation can also lead to the production of diacylglycerol (DAG), which in turn activates protein kinase C (PKC), contributing to F-actin polymerization through the phosphorylation of a large library of intermediate targets of Ca<sup>2+</sup> binding proteins. Based on data obtained in the present study, it is presently unclear if TAC-induced [Ca<sup>2+</sup>]i-release can influence the activation of a multitude of Ca<sup>2+</sup>-binding proteins involved in the F-actin polymerization through the PKC signaling pathway. Abbreviations: TAC (FK506): tacrolimus; GPCR: G-coupled protein receptor; PLC: phospholipase C; IP3: inositol (1,4,5)3-phosphate; IP3R: inositol triphosphate receptor; PI3K: phosphatidylinositol 3-kinase; PKC: protein kinase C.</p>
Full article ">
16 pages, 10553 KiB  
Article
Evaluation of the Compatibility Between Formation and Injection Water in Ultra-Low Permeability Reservoirs
by Zhaobo Gong, Leilei Zhang, Tingting Zhang, Zhong Yan, Shuping Cong, Zhenyu Zhou and Debin Kong
Processes 2024, 12(11), 2475; https://doi.org/10.3390/pr12112475 - 7 Nov 2024
Viewed by 516
Abstract
This study focuses on the reservoir scaling and the under-injection issues of the water injection well during the water injection development of an ultra-low permeability reservoir in Xinjiang due to the complex composition of injected water. Microfluidic experiments were applied to visualize the [...] Read more.
This study focuses on the reservoir scaling and the under-injection issues of the water injection well during the water injection development of an ultra-low permeability reservoir in Xinjiang due to the complex composition of injected water. Microfluidic experiments were applied to visualize the flow channel changes during water flooding, indoor core flooding experiments were employed to analyze the permeability and ion concentration, and nuclear magnetic resonance (NMR) was used to evaluate the pore structure damage. Together, these experiments were used to clarify the scaling and precipitation characteristics as the injected water met the formation water in porous media and the effects on reservoir damage. The research results showed that the poor compatibility of the injected water with the formation water could easily produce calcium carbonate scaling. The scaling products exhibited a unique network structure of blocks and a radial distribution, mainly composed of calcium carbonate and aluminosilicate. The scaling in the porous media exhibited the characteristics of unstable crystal precipitation, migration, and repeated scaling following water mixing, while the scale crystal growth occurred in the pores and the throats. According to the scaling characteristics, the damage to the reservoir permeability by scaling can be divided into the induction, damage, and stabilization stages. The filling and clogging of the scale crystals enhanced the pore structure heterogeneity, with the median pore radius reduced by 21.61% and the permeability reduced by 50%. Full article
Show Figures

Figure 1

Figure 1
<p>Ultra-low permeability reservoir core casting sheet.</p>
Full article ">Figure 2
<p>Schematic diagram of micro-model pore network and pore distribution curve.</p>
Full article ">Figure 3
<p>Diagram of the microfluidic displacement setup.</p>
Full article ">Figure 4
<p>Diagram of the core displacement and NMR setup.</p>
Full article ">Figure 5
<p>Changes in (<b>a</b>) HCO<sub>3</sub><sup>−</sup>, (<b>b</b>) pH, and (<b>c</b>) calcium loss rate under different injected water (IW)–formation water (FW) mixing ratios.</p>
Full article ">Figure 6
<p>Changes in (<b>a</b>) saturation index, (<b>b</b>) stability index, and (<b>c</b>) Ca<sup>2+</sup> and SO<sub>4</sub><sup>2−</sup> molar concentrations under different injected water–formation water mixing ratios.</p>
Full article ">Figure 7
<p>Microscopic morphology of scaling products at different magnifications. (<b>a</b>) magnification 300×, (<b>b</b>) magnification 10,000×, and (<b>c</b>) magnification 30,000×.</p>
Full article ">Figure 8
<p>Energy spectra of scaling products at different locations.</p>
Full article ">Figure 8 Cont.
<p>Energy spectra of scaling products at different locations.</p>
Full article ">Figure 9
<p>Real-time microscopic views of the steady-state experiment (Scheme 1). (<b>a</b>) 4 h injection. (<b>b</b>) 8 h injection. (<b>c</b>) end of injection.</p>
Full article ">Figure 10
<p>Injection pressure curve of the steady-state experiment (Scheme 1).</p>
Full article ">Figure 11
<p>Real-time microscopic views of the steady-state experiment (Scheme 2). (<b>a</b>) 4 h injection. (<b>b</b>) 8 h injection. (<b>c</b>) end of injection.</p>
Full article ">Figure 12
<p>Injection pressure curve of the steady-state experiment (Scheme 2).</p>
Full article ">Figure 13
<p>Real-time microscopic views of the steady-state experiment (Scheme 3). (<b>a</b>) 4 h injection. (<b>b</b>) 8 h injection. (<b>c</b>) end of injection.</p>
Full article ">Figure 14
<p>Injection pressure curve of the steady-state experiment (Scheme 3).</p>
Full article ">Figure 15
<p>The local scaling process is at the main flow channel. (<b>a</b>) 1 h, (<b>b</b>) 2 h, (<b>c</b>) 3 h, (<b>d</b>) 4 h, (<b>e</b>) 5 h, (<b>f</b>) 6 h, (<b>g</b>) 7 h, and (<b>h</b>) 8 h.</p>
Full article ">Figure 16
<p>Real-time water injection pressure and permeability variation curves.</p>
Full article ">Figure 17
<p>Pore radius distribution curves before and after displacement.</p>
Full article ">
25 pages, 1200 KiB  
Review
Exploring the Landscape of Pre- and Post-Synaptic Pediatric Disorders with Epilepsy: A Narrative Review on Molecular Mechanisms Involved
by Giovanna Scorrano, Ludovica Di Francesco, Armando Di Ludovico, Francesco Chiarelli and Sara Matricardi
Int. J. Mol. Sci. 2024, 25(22), 11982; https://doi.org/10.3390/ijms252211982 - 7 Nov 2024
Viewed by 584
Abstract
Neurodevelopmental disorders (NDDs) are a group of conditions affecting brain development, with variable degrees of severity and heterogeneous clinical features. They include intellectual disability (ID), autism spectrum disorder (ASD), attention-deficit/hyperactivity disorder (ADHD), often coexisting with epilepsy, extra-neurological comorbidities, and multisystemic involvement. In recent [...] Read more.
Neurodevelopmental disorders (NDDs) are a group of conditions affecting brain development, with variable degrees of severity and heterogeneous clinical features. They include intellectual disability (ID), autism spectrum disorder (ASD), attention-deficit/hyperactivity disorder (ADHD), often coexisting with epilepsy, extra-neurological comorbidities, and multisystemic involvement. In recent years, next-generation sequencing (NGS) technologies allowed the identification of several gene pathogenic variants etiologically related to these disorders in a large cohort of affected children. These genes encode proteins involved in synaptic homeostasis, such as SNARE proteins, implicated in calcium-triggered pre-synaptic release of neurotransmitters, or channel subunit proteins, such as post-synaptic ionotropic glutamate receptors involved in the brain’s fast excitatory neurotransmission. In this narrative review, we dissected emerged molecular mechanisms related to NDDs and epilepsy due to defects in pre- and post-synaptic transmission. We focused on the most recently discovered SNAREopathies and AMPA-related synaptopathies. Full article
(This article belongs to the Special Issue Molecular Advances in Epilepsy and Seizures)
Show Figures

Figure 1

Figure 1
<p>Steps in synaptic vesicle fusion and critical proteins involved (realized with <a href="http://BioRender.com" target="_blank">BioRender.com</a>, accessed on 10 April 2024). The exocytosis of synaptic vesicles includes different stages, such as tethering, docking, priming, and fusion. In the docking step, SNARE proteins interact with each other via SNARE motifs. The matching progresses from the N-terminus of the SNARE proteins to the C-terminal transmembrane regions, where the membranes interact. The resulting trans-SNARE complex firmly pulls the two membranes together (that of the synaptic vesicle and that of the plasma membrane of the nervous cell), providing the energy to fuse the lipid bilayers.</p>
Full article ">Figure 2
<p>Organization of AMPA and NMDA receptors at the postsynaptic density (realized with <a href="http://BioRender.com" target="_blank">BioRender.com</a>, accessed on 10 April 2024).</p>
Full article ">
17 pages, 5134 KiB  
Article
Research on the Blocking Mechanism of Stagnant Water and the Prediction of Scaling Trend in Fractured Reservoirs in Keshen Gas Field
by Qi Mao, Licheng Lu, Yejing Gong, Libin Zhao, Zihao Yang, Hongzhi Song and Rui Han
Processes 2024, 12(11), 2427; https://doi.org/10.3390/pr12112427 - 4 Nov 2024
Viewed by 504
Abstract
In this paper, well Keshen 221 was taken as the research object. The stagnant water–rock static experiment showed that, after 8 weeks of the residual water–rock static reaction, the pore size of the inner profile of the rock slice increased from 5 μm [...] Read more.
In this paper, well Keshen 221 was taken as the research object. The stagnant water–rock static experiment showed that, after 8 weeks of the residual water–rock static reaction, the pore size of the inner profile of the rock slice increased from 5 μm to 90 μm, and calcium carbonate crystals were deposited in the hole. Combined with the microscopic visualization model, it is observed that the reservoir blockage mostly occurs at the pore throat diameter, and the small fracture (30 μm) is blocked first, then the large fracture (50 μm). So, it is inferred that the blockage of the reservoir flow channel is caused by the migration of the crystals precipitated by the interaction between the stagnant water and the reservoir rock. On this basis, the TOUGHREACT reservoir model was further constructed to simulate the scaling of the stagnant water in the reservoir matrix and used to compare the scaling of the fractures with 7% and 30% porosity and the retained water at 0.658 m and 768 m. The pre-results of reservoir scaling show that the scaling is more serious when the fractures occur in the far well zone than when the fractures occur in the well entry zone. At the same location, the deposition of large fractures is six times that of small fractures, and the scaling is more severe in large fractures. Full article
Show Figures

Figure 1

Figure 1
<p>Reservoir scaling law study device.</p>
Full article ">Figure 2
<p>Research device for scale migration law.</p>
Full article ">Figure 3
<p>Microscopic visualization model diagram.</p>
Full article ">Figure 4
<p>Schematic section of reservoir matrix–small fracture–large fracture model.</p>
Full article ">Figure 5
<p>FEI digital rock analysis map of Keshen 221 reservoir rock.</p>
Full article ">Figure 6
<p>Optical observation of reservoir rocks. (<b>a</b>) High-angle fracture profile of reservoir rock; (<b>b</b>) thin section diagram of high-angle seam casting; (<b>c</b>) XRD plot of pore filler (changed).</p>
Full article ">Figure 7
<p>Intra-rock profile before and after the reaction.</p>
Full article ">Figure 8
<p>Internal sections of rock sections with different reaction times.</p>
Full article ">Figure 9
<p>Experimental diagram of rock microscopic migration and blocking mechanism.</p>
Full article ">Figure 10
<p>Schematic diagram of rock crack blockage.</p>
Full article ">Figure 11
<p>Trend of reservoir temperature with distance at different mining times.</p>
Full article ">Figure 12
<p>Variation of reservoir matrix mineral formation with distance after one year of mining.</p>
Full article ">Figure 13
<p>Variation of reservoir matrix permeability and porosity with distance at different mining times (changed). (<b>a</b>) Variation of reservoir matrix permeability with distance at different mining times; (<b>b</b>) variation of reservoir matrix porosity with distance at different mining times.</p>
Full article ">Figure 14
<p>Variation of temperature with distances at 0.658 m and 786 m for small cracks at different mining times. (<b>a</b>) The temperature of small cracks at 0.658 m with different mining times as a function of distance; (<b>b</b>) the temperature of small cracks at 786 m with different mining times as a function of distance; (<b>c</b>) the temperature of the large crack at 0.658 m with different mining times as a function of distance; (<b>d</b>) the temperature of the large fracture at 768 m with different mining times as a function of distance.</p>
Full article ">Figure 15
<p>Variation of mineral formation with distance at 0.658 m and 768 m in small fractures after one year of mining (changed). (<b>a</b>) Variation of mineral formation at 0.658 m in small fractures with distance after one year of mining; (<b>b</b>) variation of mineral formation in small fractures at 768 m after one year of mining; (<b>c</b>) the variation of mineral formation at 0.658 m after one year of mining as a function of distance; (<b>d</b>) variation of mineral formation in the large fracture at 768 m after one year of mining.</p>
Full article ">
24 pages, 30802 KiB  
Article
Effect of Calcium on the Characteristics of Action Potential Under Different Electrical Stimuli
by Xuan Qiao and Wei Yao
AppliedMath 2024, 4(4), 1358-1381; https://doi.org/10.3390/appliedmath4040072 - 1 Nov 2024
Viewed by 501
Abstract
This study investigates the role of calcium ions in the release of action potentials by comparing two models based on the framework: the standard HH model and a HH + Ca model that incorporates calcium ion channels. Purkinje cells’ responses to four types [...] Read more.
This study investigates the role of calcium ions in the release of action potentials by comparing two models based on the framework: the standard HH model and a HH + Ca model that incorporates calcium ion channels. Purkinje cells’ responses to four types of electrical current stimuli—constant direct current, step current, square wave current, and sine current—were simulated to analyze the impact of calcium on action potential characteristics. The results indicate that, under the constant direct current stimulation, the action potential firing frequency of both models increased with the escalating current intensity, while the delay time of the first action potential decreased. However, when the current intensity exceeded a specific threshold, the peak amplitude of the action potential gradually diminished. The HH + Ca model exhibited a longer delay in the first action potential compared to the HH model but maintained an action potential release under stronger currents. In response to the step current, both models showed an increased action potential frequency with a higher current, but the HH + Ca model generated subthreshold oscillations under weak currents. With the square wave current, the action potential frequency increased, though the HH + Ca model experienced suppression under high-frequency weak currents. Under the sine current, the action potential frequency rose, with the HH + Ca model showing less depression near the sine peak due to calcium’s role in modulating membrane potential. These findings suggest that calcium ions contribute to a more stable action potential release under varying stimuli. Full article
Show Figures

Figure 1

Figure 1
<p>Schematic diagram of action potential.</p>
Full article ">Figure 2
<p>HH model equivalent circuit.</p>
Full article ">Figure 3
<p>HH + Ca model equivalent circuit.</p>
Full article ">Figure 4
<p>The release of action potentials in the HH and HH + Ca models under weak electrical stimulations.</p>
Full article ">Figure 5
<p>(<b>a</b>) The release of action potentials when the current intensity of the HH model increases from 0 pA to 2 pA. (<b>b</b>) The release of action potentials when the current intensity of the HH + Ca model increases from 6 pA to 8 pA.</p>
Full article ">Figure 6
<p>Peak variation. (<b>a</b>) The peak change in the action potential when the current intensity of the HH model increases from 0 pA to 10 pA. (<b>b</b>) The peak change in the action potential when the current intensity of the HH model increases from 0 pA to 20 pA. (<b>c</b>) The peak change in the action potential when the current intensity of the HH + Ca model increases from 0 pA to 10 pA. (<b>d</b>) The peak change in the action potential when the current intensity of the HH + Ca model increases from 0 pA to 20 pA.</p>
Full article ">Figure 7
<p>The release of action potentials in the HH and HH + Ca models under moderate to strong electrical stimulations.</p>
Full article ">Figure 8
<p>Comparison of the peak values of the first action potential.</p>
Full article ">Figure 9
<p>Delay times of the first action potential.</p>
Full article ">Figure 10
<p>Frequency of action potential release under different current intensities.</p>
Full article ">Figure 11
<p>Peak time interval of action potential under 20 pA electrical stimulation.</p>
Full article ">Figure 12
<p>Peak time interval of action potential under 40 pA electrical stimulation.</p>
Full article ">Figure 13
<p>Peak time interval of action potential under 60 pA electrical stimulation.</p>
Full article ">Figure 14
<p>Distribution of action potential under different current intensities within 20 ms.</p>
Full article ">Figure 15
<p>Phase plane trajectory under different current intensities within 20 ms.</p>
Full article ">Figure 16
<p>The release of action potentials in the HH and HH + Ca models under strong electrical stimulations.</p>
Full article ">Figure 17
<p>The release of action potentials in the HH model and HH + Ca model under step current stimulations. (<b>a</b>) Under the 0–0.5 pA step stimulation, the action potential firing characteristics of HH model and HH+Ca model. (<b>b</b>) Under the 0.5–3 pA step stimulation, the action potential firing characteristics of HH model and HH+Ca model. (<b>c</b>) Under the 4–9 pA step stimulation, the action potential firing characteristics of HH model and HH+Ca model. (<b>d</b>) Under the 10–60 pA step stimulation, the action potential firing characteristics of HH model and HH+Ca model.</p>
Full article ">Figure 18
<p>Comparison of the release frequency of action potentials under a direct current stimulation and step current stimulation. (<b>a</b>) Under the 10–40 pA DC stimulation, the frequency of the action potential release in the HH model. (<b>b</b>) Under the 10–40 pA step stimulation, the frequency of the action potential release in the HH model.</p>
Full article ">Figure 19
<p>The release of action potentials under different square wave current intensities and current frequencies. (<b>a</b>,<b>c</b>,<b>e</b>,<b>g</b>,<b>i</b>,<b>k</b>) The action potential release of the HH model under square wave current stimulations of different frequencies and current intensities. (<b>b</b>,<b>d</b>,<b>f</b>,<b>h</b>,<b>j</b>,<b>l</b>) The action potential release of the HH + Ca model under square wave current stimulations of different frequencies and current intensities.</p>
Full article ">Figure 19 Cont.
<p>The release of action potentials under different square wave current intensities and current frequencies. (<b>a</b>,<b>c</b>,<b>e</b>,<b>g</b>,<b>i</b>,<b>k</b>) The action potential release of the HH model under square wave current stimulations of different frequencies and current intensities. (<b>b</b>,<b>d</b>,<b>f</b>,<b>h</b>,<b>j</b>,<b>l</b>) The action potential release of the HH + Ca model under square wave current stimulations of different frequencies and current intensities.</p>
Full article ">Figure 20
<p>The release of action potentials under different sine current intensities and current frequencies. (<b>a</b>,<b>c</b>,<b>e</b>,<b>g</b>,<b>i</b>,<b>k</b>) The action potential release of the HH model under sine wave current stimulations of different frequencies and current intensities. (<b>b</b>,<b>d</b>,<b>f</b>,<b>h</b>,<b>j</b>,<b>l</b>) The action potential release of the HH + Ca model under sine wave current stimulations of different frequencies and current intensities.</p>
Full article ">Figure 20 Cont.
<p>The release of action potentials under different sine current intensities and current frequencies. (<b>a</b>,<b>c</b>,<b>e</b>,<b>g</b>,<b>i</b>,<b>k</b>) The action potential release of the HH model under sine wave current stimulations of different frequencies and current intensities. (<b>b</b>,<b>d</b>,<b>f</b>,<b>h</b>,<b>j</b>,<b>l</b>) The action potential release of the HH + Ca model under sine wave current stimulations of different frequencies and current intensities.</p>
Full article ">Figure 21
<p>The peak time interval of action potential under different frequency sine current stimulation of 20 pA. (<b>a</b>) Peak time interval of action potential at 10 Hz. (<b>b</b>) Peak time interval of action potential at 20 Hz. (<b>c</b>) Peak time interval of action potential at 30 Hz. (<b>d</b>) Peak time interval of action potential at 40 Hz. (<b>e</b>) Peak time interval of action potential at 50 Hz. (<b>f</b>) Peak time interval of action potential at 60 Hz.</p>
Full article ">
24 pages, 820 KiB  
Systematic Review
The Effect of Overweight/Obesity on Cutaneous Microvascular Reactivity as Measured by Laser-Doppler Fluxmetry: A Systematic Review
by Ally McIllhatton, Sean Lanting and Vivienne Chuter
Biomedicines 2024, 12(11), 2488; https://doi.org/10.3390/biomedicines12112488 - 30 Oct 2024
Viewed by 442
Abstract
Introduction: We sought to determine by systematic review the independent effect of overweight/obesity on cutaneous microvascular reactivity in adults as measured by laser-Doppler fluxmetry. Methods: CINAHL Complete, SPORTSDiscus, Embase, Medline, and Cochrane Library were searched until March 2024 to identify studies investigating cutaneous [...] Read more.
Introduction: We sought to determine by systematic review the independent effect of overweight/obesity on cutaneous microvascular reactivity in adults as measured by laser-Doppler fluxmetry. Methods: CINAHL Complete, SPORTSDiscus, Embase, Medline, and Cochrane Library were searched until March 2024 to identify studies investigating cutaneous microvascular reactivity in an overweight/obese but otherwise healthy group versus a lean/healthy weight. Reporting is consistent with the Preferred Reporting Items for Systematic Reviews and Meta-Analyses (PRISMA) statement. Quality appraisal of included studies was performed using the Strengthening the Reporting of Observational Studies in Epidemiology (STROBE) checklist. Results: Nineteen eligible articles reported on 1847 participants. Most articles reported impaired cutaneous microvascular reactivity in cohorts with overweight/obesity compared to cohorts with lean/healthy weight. Investigating reactivity via post-occlusive reactive hyperaemia (PORH) and iontophoresis of acetylcholine (ACh) has shown significance. No significant differences were reported between groups in response to local heating or to iontophoresis of methacholine or insulin, while findings of the effect of obesity on iontophoresis of sodium nitroprusside (SNP) were mixed. Conclusions: The pathophysiology of impaired cutaneous microvascular reactivity in overweight/obesity requires further investigation; however, impaired function of vasoactive substances, endothelial dysfunction, sensory nerves, and calcium-activated potassium channels may be implicated. Identifying these impaired microvascular responses should inform possible therapy targets in overweight and obesity.activated potassium channels may be implicated. Identifying these impaired microvascular responses should inform possible therapy targets in overweight and obesity. Full article
(This article belongs to the Special Issue Microcirculation in Health and Diseases)
Show Figures

Figure 1

Figure 1
<p>PRISMA flowchart.</p>
Full article ">
22 pages, 6439 KiB  
Article
Using a Failing Human Ventricular Cardiomyocyte Model to Re-Evaluate Ca2+ Cycling, Voltage Dependence, and Spark Characteristics
by Jerome Anthony E. Alvarez, Mohsin Saleet Jafri and Aman Ullah
Biomolecules 2024, 14(11), 1371; https://doi.org/10.3390/biom14111371 - 28 Oct 2024
Viewed by 528
Abstract
Previous studies have observed alterations in excitation–contraction (EC) coupling during end-stage heart failure that include action potential and calcium (Ca2+) transient prolongation and a reduction of the Ca2+ transient amplitude. Underlying these phenomena are the downregulation of potassium (K+ [...] Read more.
Previous studies have observed alterations in excitation–contraction (EC) coupling during end-stage heart failure that include action potential and calcium (Ca2+) transient prolongation and a reduction of the Ca2+ transient amplitude. Underlying these phenomena are the downregulation of potassium (K+) currents, downregulation of the sarcoplasmic reticulum Ca2+ ATPase (SERCA), increase Ca2+ sensitivity of the ryanodine receptor, and the upregulation of the sodium–calcium (Na=-Ca2+) exchanger. However, in human heart failure (HF), debate continues about the relative contributions of the changes in calcium handling vs. the changes in the membrane currents. To understand the consequences of the above changes, they are incorporated into a computational human ventricular myocyte HF model that can explore the contributions of the spontaneous Ca2+ release from the sarcoplasmic reticulum (SR). The reduction of transient outward K+ current (Ito) is the main membrane current contributor to the decrease in RyR2 open probability and L-type calcium channel (LCC) density which emphasizes its importance to phase 1 of the action potential (AP) shape and duration (APD). During current-clamp conditions, RyR2 hyperphosphorylation exhibits the least amount of Ca2+ release from the SR into the cytosol and SR Ca2+ fractional release during a dynamic slow–rapid–slow (0.5–2.5–0.5 Hz) pacing, but it displays the most abundant and more lasting Ca2+ sparks two-fold longer than a normal cell. On the other hand, under voltage-clamp conditions, HF by decreased SERCA and upregulated INCX show the least SR Ca2+ uptake and EC coupling gain, as compared to HF by hyperphosphorylated RyR2s. Overall, this study demonstrates that the (a) combined effect of SERCA and NCX, and the (b) RyR2 dysfunction, along with the downregulation of the cardiomyocyte’s potassium currents, could substantially contribute to Ca2+ mishandling at the spark level that leads to heart failure. Full article
Show Figures

Figure 1

Figure 1
<p>Comparisons of normal cardiac cell conditions vs. heart failure. To ensure steady state conditions, 20 s simulations at 1 Hz were performed. (<b>A</b>) APD prolongation with longer plateau phase due to influx of Ca<sup>2+</sup>. Early afterdepolarizations (EADs) were also observed through the oscillations of the membrane potential. (<b>B</b>) Ca<sup>2+</sup> concentration displays an incomplete extrusion from the myoplasm. (<b>C</b>) RyR2 open probability is decreased during heart failure, and small openings were observed during EAD occurrence. (<b>D</b>) SR Ca<sup>2+</sup> also displays incomplete recovery, mainly due to decreased SERCA activity and increased NCX expression.</p>
Full article ">Figure 2
<p>Prominent effects of I<sub>to</sub> among other K<sup>+</sup> currents. (<b>A</b>) APD prolongation from the downregulated repolarizing K<sup>+</sup> currents present in failing human cardiomyocytes with decreased I<sub>to</sub> (blue) showing a loss of the distinctive “notch” in the action potential shape. (<b>B</b>) Roughly two-fold increase in time-to-peak Ca<sup>2+</sup> transient from 0.02 to 0.04 s duration was observed. (<b>C</b>) I<sub>LCC</sub> peak density also decreased roughly 30–35%, respectively, by the effects of I<sub>to</sub> alone. (<b>D</b>) RyR open probability is reduced ~20–25% and (<b>E</b>) delayed kinetics of Ca<sup>2+</sup> release from the SR by I<sub>to</sub>.</p>
Full article ">Figure 3
<p>Dynamic slow–rapid–slow pacing (0.5–2.5–0.5 Hz) [Ca<sup>2+</sup>]<sub>myo</sub> concentrations between normal, HF with K<sup>+</sup> block (i.e., It<sub>o</sub>, I<sub>K1</sub>, I<sub>Kr</sub>, and I<sub>Ks</sub>), and HF without K<sup>+</sup> block. To ensure steady state conditions, 45 s simulations were performed. (<b>A</b>) Erratic Ca<sup>2+</sup> transient peaks were observed during 0.5 Hz pacing in failing hearts. HF with reduced K<sup>+</sup> currents (red) slightly elevated the [Ca<sup>2+</sup>]<sub>myo</sub> amplitude during high pacing (2.5 Hz) as compared to normal (black) and HF without K<sup>+</sup> reductions (yellow). (<b>B</b>) Predicted force using the same color schemes. HF with reduced K<sup>+</sup> currents (red) exhibit higher predicted force using peak systolic [Ca<sup>2+</sup>]<sub>myo</sub> concentrations in rapid pacing.</p>
Full article ">Figure 4
<p>Comparisons of normal cardiac cell conditions vs. heart failure. To ensure steady state conditions, 20 s simulations at 1 Hz were performed. (<b>A</b>) Peak RyR2 open probability and (<b>B</b>) L-type calcium channel is generally decreased with each HF-related change, i.e., (1) RyR hyperphosphorylation (blue), (2) downregulated SERCA and increased NCX (green), and (3) combination of both (red) mechanisms. (<b>C</b>) On average, RyR2 open probabilities and (<b>D</b>) L-type calcium channel densities are significantly different from normal conditions in HF-related changes (P<sub>ANOVA</sub> &lt; 0.05). (<b>E</b>) Reductions in SR Ca<sup>2+</sup> concentrations, both at systole and diastole, are observed in heart failure. In all cases except Normal, the K<sup>+</sup> changes reflect those seen in HF. The legend on the bottom shows the increase or decrease of the different components.</p>
Full article ">Figure 5
<p>SR Ca<sup>2+</sup> dynamic pacing (0.5–2.5–0.5 Hz) and fractional release. To ensure steady state conditions, 45 s simulations were performed. (<b>A</b>) Normal cardiomyocyte. (<b>B</b>) Due to impaired SERCA to re-sequester Ca<sup>2+</sup> back into the SR and accompanied with stronger extrusion of Ca<sup>2+</sup> by NCX, SR Ca<sup>2+</sup> concentrations stay at 800 µM during 2.5 Hz pacing. (<b>C</b>) Increased RyR sensitivity shows diminished amplitudes between systole and diastole, but still exhibits a staircase effect. (<b>D</b>) the combined effects of three HF-related Ca<sup>2+</sup> regulatory proteins display the absence of the Bowditch effect, which could be related to the deficient cardiac output seen in failing hearts. (<b>E</b>) SR Ca<sup>2+</sup> fractional release is generally increased in cases of HF, but hyperphosphorylation of RyR2 alone accounts for lower Ca<sup>2+</sup> release (blue). In all cases except Normal, the K<sup>+</sup> changes reflect those seen in HF. The legend on the bottom shows the increase or decrease of the different components. The computed fractional release from the diastolic and systolic [Ca<sup>2+</sup>]<sub>SR</sub> values from simulation results are in the <a href="#app1-biomolecules-14-01371" class="html-app">Supplemental Materials: HF Supplementary Materials.rar</a>.</p>
Full article ">Figure 6
<p>Excitation–contraction coupling gain. (<b>A</b>) The amount of SR Ca<sup>2+</sup> release (in µM)—hyperphosphorylated RyR2s (blue) generally increase release flux as compared with impaired SERCA and upregulated NCX (green). (<b>B</b>) Peak I<sub>CaL</sub> (in µA) observed—upregulation of I<sub>NCX</sub> decreases the amount of Ca<sup>2+</sup> current (green). However, as RyR2 is 50% more sensitive, more Ca<sup>2+</sup> is released from the SR (blue). (<b>C</b>) ECC gain is altered at negative (−40 mV to −25 mV) and at positive potentials (from 15 mV and above). (<b>D</b>) RyR open probabilities (P<sub>O,RyR</sub>) are statistically insignificant and are also observed in experimental findings by Jiang et al. in failing canine and human hearts [<a href="#B7-biomolecules-14-01371" class="html-bibr">7</a>]. (<b>E</b>) A 50% reduction in SERCA activity accounts for the least Ca<sup>2+</sup> sequestration into the SR. (<b>F</b>) Systolic SR Ca<sup>2+</sup> levels are decreased in HF due to the decreased SR Ca<sup>2+</sup> content and other relative HF conditions. The legend on the bottom shows the increase or decrease of the different components.</p>
Full article ">Figure 7
<p>Representative sample (0.1% of 20,000 CRUs) of detected calcium sparks in the local subspace from 20 CRUs. Superimposed action potential from pacing at 0.5 Hz depicting relationship between AP and Ca<sup>2+</sup> sparks (N<sub>sparks</sub>). (<b>A</b>) Ca<sup>2+</sup> spark behavior in a normally functioning human AP. Minimal Ca<sup>2+</sup> leak was observed in the late phase of the AP during extrusion of Ca<sup>2+</sup> ions in cell repolarization. (<b>B</b>) Ca<sup>2+</sup> spark behavior in heart failure with superimposed AP in the occurrence of early afterdepolarization (EAD). An increase in diastolic spark rate is evident in the late phase of the action potential, which causes APD prolongation and potential presence of EADs. The different colors in the plots indicate the Ca<sup>2+</sup> spark from different CRUs.</p>
Full article ">Figure 8
<p>Comparisons of spark characteristics between normal and heart failure conditions. (<b>A</b>) Systolic SR Ca<sup>2+</sup> levels are decreased in HF. (<b>B</b>) RyR hyperphosphorylation produces more sparks than normal and other HF conditions. (<b>C</b>) Similarly, increased RyR sensitivity prolongs spark duration. (<b>D</b>) Modest differences in average spark peaks in all stated underlying cardiac cell impairment.</p>
Full article ">Figure A1
<p>Early afterdepolarizations. Remodeled HF is paced at 0.5 Hz to induce EADs. Five EADs were observed throughout the 20 s simulated HF.</p>
Full article ">Figure A2
<p>Simulation results of various HF conditions. RyR hyperphosphorylation (blue) displays more distinctive characteristics, especially in the prolonged calcium transient.</p>
Full article ">Figure A3
<p>Comparisons of spark characteristics of normal versus each respective K<sup>+</sup> block. (<b>A</b>–<b>D</b>) The differences in spark characteristics are negligible in all cases of downregulated repolarizing K<sup>+</sup> currents.</p>
Full article ">
11 pages, 2367 KiB  
Article
High-Affinity Plasma Membrane Ca2+ Channel Cch1 Modulates Adaptation to Sodium Dodecyl Sulfate-Triggered Rise in Cytosolic Ca2+ Concentration in Ogataea parapolymorpha
by Maria Kulakova, Maria Pakhomova, Victoria Bidiuk, Alexey Ershov, Alexander Alexandrov and Michael Agaphonov
Int. J. Mol. Sci. 2024, 25(21), 11450; https://doi.org/10.3390/ijms252111450 - 25 Oct 2024
Viewed by 479
Abstract
The cytosolic calcium concentration ([Ca2+]cyt) in yeast cells is maintained at a low level via the action of different transporters sequestrating these cations in the vacuole. Among them, the vacuolar Ca2+ ATPase Pmc1 crucially contributes to this process. [...] Read more.
The cytosolic calcium concentration ([Ca2+]cyt) in yeast cells is maintained at a low level via the action of different transporters sequestrating these cations in the vacuole. Among them, the vacuolar Ca2+ ATPase Pmc1 crucially contributes to this process. Its inactivation in Ogataea yeasts was shown to cause sodium dodecyl sulfate (SDS) hypersensitivity that can be alleviated by the inactivation of the plasma membrane high-affinity Ca2+ channel Cch1. Here, we show that SDS at low concentrations induces a rapid influx of external Ca2+ into cells, while the plasma membrane remains impermeable for propidium iodide. The inactivation of Pmc1 disturbs efficient adaptation to this activity of SDS. The inactivation of Cch1 partially restores the ability of pmc1 mutant cells to cope with an increased [Ca2+]cyt that correlates with the suppression of SDS hypersensitivity. At the same time, Cch1 is unlikely to be directly involved in SDS-induced Ca2+ influx, since its inactivation does not decrease the amplitude of the rapid [Ca2+]cyt elevation in the pmc1-Δ mutant. The obtained data suggest that the effects of CCH1 inactivation on SDS sensitivity and coping with increased [Ca2+]cyt are related to an additional Cch1 function beyond its direct involvement in Ca2+ transport. Full article
(This article belongs to the Section Biochemistry)
Show Figures

Figure 1

Figure 1
<p>Change in CFU numbers (survival rate) after 10′ and 60′ incubation in presence of 0.01% SDS. Log., logarithmic cultures; Stat., stationary cultures; WT, DL5-LC strain; <span class="html-italic">cch1-Δ</span>, DL5-cch1 strain; <span class="html-italic">pmc1-Δ</span>, DL5-pmc1-LC strain; <span class="html-italic">pmc1-Δ cch1-Δ</span>, DL5-pmc1-cch1 strain. For details, see <a href="#sec4-ijms-25-11450" class="html-sec">Section 4</a>. *—<span class="html-italic">p</span>-value &lt; 0.005.</p>
Full article ">Figure 2
<p>Distribution of fluorescence (FL) of DL5-LC (WT) and DL5-pmc1-LC (<span class="html-italic">pmc1-Δ</span>) cells after PI staining. Red line “SDS”, cells after 1 h incubation with 0.01% SDS; blue line “C”, untreated control cells; black line “B”, cells inactivated by boiling. Comparison of three replicates is represented in <a href="#app1-ijms-25-11450" class="html-app">Supplementary Figure S1</a>.</p>
Full article ">Figure 3
<p>Dynamics of median FL<sub>450</sub>/FL<sub>525</sub> values of cells of DL5-LC (WT), DL5-cch1 (<span class="html-italic">cch1-Δ</span>), DL5-pmc1-LC (<span class="html-italic">pmc1-Δ</span>) and DL5-pmc1-cch1 strains expressing GEM-GECO. (<b>A</b>), response of DL5-LC cells in logarithmic culture to different SDS concentrations; (<b>B</b>), response of cells in logarithmic cultures to 100 mM CaCl<sub>2</sub>; (<b>C</b>), response of cells in logarithmic cultures to 0.01% SDS; (<b>D</b>), response of cells in stationary cultures to 0.01% SDS; (<b>E</b>), response of cells in logarithmic cultures to 0.004% SDS; (<b>F</b>), response of cells in stationary cultures to 0.004% SDS. Median FL<sub>450</sub>/FL<sub>525</sub> values of 10<sup>4</sup> cells obtained from 3 replicates were averaged, and standard deviations were calculated.</p>
Full article ">Figure 4
<p>Distributions of FL<sub>450</sub>/FL<sub>525</sub> in exponentially grown cultures of DL5-LC (WT) and DL5-pmc1-LC (<span class="html-italic">pmc1-Δ</span>) strains before (-) or after 10′ incubation with 0.01% SDS in absence (SDS) or presence of 10 mM EDTA (SDS + EDTA) or 20 mM EGTA (SDS + EGTA). Median values are indicated by bars.</p>
Full article ">Figure 5
<p>SDS-PAGE and immunoblotting of phosphorylated (p38PP) and total Hog1 (Hog1). Exponential cultures were supplemented with 0.004% SDS (4), 0.008% SDS (8) or 0.6 M NaCl and incubated for 3, 10 or 20 min. Untreated cells (-) were used as reference. <span class="html-italic">hog1-Δ</span> mutant was used as negative control. Panel “<span class="html-italic">PMC1</span>”—DL5-LC strain; panel “<span class="html-italic">pmc1-Δ</span>”—DL5-pmc1-LC strain; panel “<span class="html-italic">pmc1-Δ</span> vs. <span class="html-italic">PMC1</span>”—comparison of DL5-pmc1-LC and DL5-LC strains after 20 min incubation with SDS.</p>
Full article ">
15 pages, 2699 KiB  
Review
Voltage-Gated Ion Channel Compensatory Effect in DEE: Implications for Future Therapies
by Khadijeh Shabani, Johannes Krupp, Emilie Lemesre, Nicolas Lévy and Helene Tran
Cells 2024, 13(21), 1763; https://doi.org/10.3390/cells13211763 - 24 Oct 2024
Viewed by 495
Abstract
Developmental and Epileptic Encephalopathies (DEEs) represent a clinically and genetically heterogeneous group of rare and severe epilepsies. DEEs commonly begin early in infancy with frequent seizures of various types associated with intellectual disability and leading to a neurodevelopmental delay or regression. Disease-causing genomic [...] Read more.
Developmental and Epileptic Encephalopathies (DEEs) represent a clinically and genetically heterogeneous group of rare and severe epilepsies. DEEs commonly begin early in infancy with frequent seizures of various types associated with intellectual disability and leading to a neurodevelopmental delay or regression. Disease-causing genomic variants have been identified in numerous genes and are implicated in over 100 types of DEEs. In this context, genes encoding voltage-gated ion channels (VGCs) play a significant role, and part of the large phenotypic variability observed in DEE patients carrying VGC mutations could be explained by the presence of genetic modifier alleles that can compensate for these mutations. This review will focus on the current knowledge of the compensatory effect of DEE-associated voltage-gated ion channels and their therapeutic implications in DEE. We will enter into detailed considerations regarding the sodium channels SCN1A, SCN2A, and SCN8A; the potassium channels KCNA1, KCNQ2, and KCNT1; and the calcium channels CACNA1A and CACNA1G. Full article
(This article belongs to the Special Issue Nucleic Acid Therapeutics (NATs): Advances and Perspectives)
Show Figures

Figure 1

Figure 1
<p>Expression profile of the selected VGCs throughout the human lifespan. Among the sodium channels, the expression of (<b>A</b>) SCN1A is lowest prenatally, while the expression of (<b>B</b>) SCN8A and (<b>C</b>) SCN2A is slightly higher, followed by a rise in all three channels expression towards adulthood. Among potassium channels, (<b>D</b>) KCNA1 and (<b>E</b>) KCNT1 show a lower level of expression during early development and increase as development progresses. On the contrary, (<b>F</b>) KCNQ2 shows the highest expression prenatally, followed by a slight decrease postnatally. (<b>G</b>) CACNA1A has a higher postnatal expression, with the highest level in the cerebellar cortex. (<b>H</b>) CACNA1G shows fluctuation within the brain regions, with the highest expression in the cerebellar cortex. Nine windows (W) include fatal development (W1 to W4), birth (W5), infancy and childhood (W6 to W7), adolescence, and adulthood (W8 to W9) in different regions of the brain, including the neocortex (NCX), hippocampus (HIP), amygdala (AMY), striatum (STR), mediodorsal nucleus of the thalamus (MD), and cerebellar cortex (CBC). The graphs are extracted from the <span class="html-italic">PsychNode dataset</span>; Li et al., 2018 (<span class="html-italic">PsychEncode</span>) [<a href="#B61-cells-13-01763" class="html-bibr">61</a>].</p>
Full article ">Figure 2
<p>Differential expression profile of the selected voltage-gated ion channels in inhibitory (PV, SST, VIP) vs excitatory (EXC) neurons in adult mice cortex. (<b>A</b>) Scn1a, (<b>B</b>) Kcnt1, and (<b>C</b>) Cacna1g show enrichment in inhibitory neurons, whereas (<b>D</b>) Scn2a, (<b>E</b>) Scn8a, (<b>F</b>) Kcnq2, (<b>G</b>) Kcn1a, and (<b>H</b>) Cacna1a display higher enrichment in excitatory neurons (all the graphs extracted from <a href="http://research-pub.gene.com/NeuronSubtypeTranscriptomes" target="_blank">http://research-pub.gene.com/NeuronSubtypeTranscriptomes</a> (accessed on 8 September 2024); Huntley et al., 2020 [<a href="#B91-cells-13-01763" class="html-bibr">91</a>]).</p>
Full article ">Figure 3
<p>Model of the compensatory effect of certain VGCs in Dravet syndrome. (<b>A</b>) Enrichment pattern of selected voltage-gated ion channels in excitatory neurons (EN) and inhibitory neurons (IN) in healthy conditions. (<b>B</b>) SCN1A LOF in Dravet syndrome causes hypo-excitability in inhibitory neurons, leading to hyper-excitability of excitatory neurons. (<b>C</b>) Manipulating the excitation/inhibition (E/I) ratio by targeting different genes in Dravet syndrome. Inhibiting KCNT1 (Scenario 1), SCN8A (Scenario 2), and CACNA1G (Scenario 3) improves the seizure phenotype, probably by bringing the E/I ratio closer to its optimal level. In contrast, inducing SCN2A (Scenario 4), KCNQ2 (Scenario 5), and CACNA1A (Scenario 6) worsens the phenotype due to increased excitability in an already hyperexcitable network. Sodium, potassium, and calcium channels are shown in orange, purple, and green, respectively. PV, SST, and VIP show inhibitory neurons, and EN refers to excitatory neurons. The model is generated using BioRender (BioRender: Scientific Image and Illustration Software, <a href="http://www.biorender.com" target="_blank">www.biorender.com</a>).</p>
Full article ">
11 pages, 1507 KiB  
Article
Drugs That Induce Gingival Overgrowth Drive the Pro-Inflammatory Polarization of Macrophages In Vitro
by Annalisa Palmieri, Agnese Pellati, Dorina Lauritano, Alberta Lucchese, Francesco Carinci, Luca Scapoli and Marcella Martinelli
Int. J. Mol. Sci. 2024, 25(21), 11441; https://doi.org/10.3390/ijms252111441 - 24 Oct 2024
Viewed by 464
Abstract
Several attempts have been made to elucidate the pathogenesis of drug-induced gingival overgrowth (DIGO), which is triggered by the chronic use of certain drugs that fall into three main categories: anticonvulsants, immunosuppressants, and calcium channel blockers. Previous research suggests that cytokines and impaired [...] Read more.
Several attempts have been made to elucidate the pathogenesis of drug-induced gingival overgrowth (DIGO), which is triggered by the chronic use of certain drugs that fall into three main categories: anticonvulsants, immunosuppressants, and calcium channel blockers. Previous research suggests that cytokines and impaired cellular functions play a role in DIGO. Of particular interest are macrophages, immune cells that can switch between M1 (pro-inflammatory) and M2 (anti-inflammatory) phenotypes in response to exogenous signals and stimuli. An imbalance between M1 and M2 macrophage populations may underlie DIGO. M1 may contribute to the initial tissue damage in DIGO, while M2 may then attempt to repair the damage with anti-inflammatory mechanisms. To test the hypothesis that drugs associated with DIGO could influence macrophage polarization, human monocytes (precursors of macrophages) were induced to differentiate into M0-naïve macrophages and then exposed to drugs: diphenylhydantoin, gabapentin, mycophenolate, and amlodipine. Quantitative real-time PCR amplification was used to measure the expression of specific genes associated with macrophage polarization. All of the drugs tested induced M0 macrophages to overexpress genes typical of the M1 phenotype, such as CCL5, CXCL10, and IDO1. This investigation provides the first evidence of a link between drugs that cause DIGO and M1 pro-inflammatory macrophage polarization. The knowledge gained from this research could be valuable for future DIGO treatment strategies. Full article
(This article belongs to the Special Issue Oral Cancer and Disease in Humans and Animals)
Show Figures

Figure 1

Figure 1
<p>Immunofluorescence staining of the adhered macrophages after differentiation. Magnification 20X. (<b>a</b>,<b>b</b>) M0 macrophages negative for CD80 and CD163, respectively; (<b>c</b>) M1 macrophages positive for CD80 (red); (<b>d</b>) M1 macrophages negative for CD163; (<b>e</b>) M2 macrophages positive for CD163 (green); (<b>f</b>) M2 macrophages negative for CD80. Nuclei were stained with DAPI (blue).</p>
Full article ">Figure 2
<p>Gene expression profile in differentiated macrophage subtypes M1 and M2: (<b>a</b>) macrophages treated with LPS and IFNγ showed the overexpression of the M1 markers and downregulation of the M2 markers; (<b>b</b>) IL-4 treatment induces the overexpression of the M2 markers. The bars in the graph represent the fold change in gene expression on a logarithmic scale. The error bars indicate the standard deviation of fold changes calculated from two biological samples and three experimental replicates. The green line marks a fold change = 2; the red line marks a fold change = 0.5.</p>
Full article ">Figure 3
<p>The gene expression profile of macrophages treated with different drugs after 48 h. Panel (<b>a</b>) shows the results of gabapentin treatment, panel (<b>b</b>) of mycophenolate, panel (<b>c</b>) of amlodipine, and panel (<b>d</b>) of diphenylhydantoin. All the treatments induced M1 polarization, as evidenced by the significant upregulation of CCL5, CXCL10, and IDO1 genes. The bars in the graph represent the fold change in gene expression on a logarithmic scale. The error bars indicate the standard deviation of fold changes calculated from two biological samples and three experimental replications. The green line marks a fold change = 2; the red line marks a fold change = 0.5.</p>
Full article ">Figure 3 Cont.
<p>The gene expression profile of macrophages treated with different drugs after 48 h. Panel (<b>a</b>) shows the results of gabapentin treatment, panel (<b>b</b>) of mycophenolate, panel (<b>c</b>) of amlodipine, and panel (<b>d</b>) of diphenylhydantoin. All the treatments induced M1 polarization, as evidenced by the significant upregulation of CCL5, CXCL10, and IDO1 genes. The bars in the graph represent the fold change in gene expression on a logarithmic scale. The error bars indicate the standard deviation of fold changes calculated from two biological samples and three experimental replications. The green line marks a fold change = 2; the red line marks a fold change = 0.5.</p>
Full article ">Figure 4
<p>Cell viability of macrophages (M0) treated for 48 h with different concentrations of drugs, assessed by PrestoBlue™ reagent protocol. The viability of treated samples was normalized to untreated control; error bars represent standard errors calculated from three experimental replications.</p>
Full article ">
41 pages, 811 KiB  
Review
A Scoping Review of GLP-1 Receptor Agonists: Are They Associated with Increased Gastric Contents, Regurgitation, and Aspiration Events?
by Marvin G. Chang, Juan G. Ripoll, Ernesto Lopez, Kumar Krishnan and Edward A. Bittner
J. Clin. Med. 2024, 13(21), 6336; https://doi.org/10.3390/jcm13216336 - 23 Oct 2024
Viewed by 1004
Abstract
Background: The increased popularity and ubiquitous use of glucagon-like peptide-1 receptor agonists (GLP-1 RAs) for the treatment of diabetes, heart failure, and obesity has led to significant concern for increased risk for perioperative aspiration, given their effects on delayed gastric emptying. This concern [...] Read more.
Background: The increased popularity and ubiquitous use of glucagon-like peptide-1 receptor agonists (GLP-1 RAs) for the treatment of diabetes, heart failure, and obesity has led to significant concern for increased risk for perioperative aspiration, given their effects on delayed gastric emptying. This concern is highlighted by many major societies that have published varying guidance on the perioperative management of these medications, given limited data. We conducted a scoping review of the available literature regarding the aspiration risk and aspiration/regurgitant events related to GLP-1 RAs. Methods: A librarian-assisted search was performed using five electronic medical databases (PubMed, Embase, and Web of Science Platform Databases, including Web of Science Core Collection, KCI Korean Journal Database, MEDLINE, and Preprint Citation Index) from inception through March 2024 for articles that reported endoscopic, ultrasound, and nasogastric evaluation for increased residual gastric volume retained food contents, as well as incidences of regurgitation and aspiration events. Two reviewers independently screened titles, abstracts, and full text of articles to determine eligibility. Data extraction was performed using customized fields established a priori within a systematic review software system. Results: Of the 3712 citations identified, 24 studies met eligibility criteria. Studies included four prospective, six retrospective, five case series, and nine case reports. The GLP-1 RAs reported in the studies included semaglutide, liraglutide, lixisenatide, dulaglutide, tirzepatide, and exenatide. All studies, except one case report, reported patients with confounding factors for retained gastric contents and aspiration, such as a history of diabetes, cirrhosis, hypothyroidism, psychiatric disorders, gastric reflux, Barrett’s esophagus, Parkinson’s disease, dysphagia, obstructive sleep apnea, gastric polyps, prior abdominal surgeries, autoimmune diseases, pain, ASA physical status classification, procedural factors (i.e., thyroid surgery associated with risk for nausea, ketamine associated with nausea and secretions), and/or medications associated with delayed gastric emptying (opioids, anticholinergics, antidepressants, beta-blockers, calcium channel blockers, DPP-IV inhibitors, and antacids). Of the eight studies (three prospective and five retrospective) that evaluated residual contents in both GLP-1 users and non-users, seven studies (n = 7/8) reported a significant increase in residual gastric contents in GLP-1 users compared to non-users (19–56% vs. 5–20%). In the three retrospective studies that evaluated for aspiration events, there was no significant difference in aspiration events, with one study reporting aspiration rates of 4.8 cases per 10,000 in GLP-1 RA users compared to 4.6 cases per 10,000 in nonusers and the remaining two studies reporting one aspiration event in the GLP-1 RA user group and none in the non-user group. In one study that evaluated for regurgitation or reflux by esophageal manometry and pH, there was no significant difference in reflux episodes but a reduction in gastric acidity in the GLP-1 RA user group compared to the non-user group. Conclusions: There is significant variability in the findings reported in the studies, and most of these studies include confounding factors that may influence the association between GLP-1 RAs and an increased risk of aspiration and related events. While GLP-1 RAs do increase residual gastric contents in line with their mechanism of action, the currently available data do not suggest a significant increase in aspiration and regurgitation events associated with their use and the withholding of GLP-1 RAs to reduce aspiration and regurgitation events, as is currently recommended by many major societal guidelines. Large randomized controlled trials (RCTs) may be helpful in further elucidating the impact of GLP-1 RAs on perioperative aspiration risk. Full article
Show Figures

Figure 1

Figure 1
<p>Study extraction and inclusion diagram. EMBASE = Excerpta Medica dataBASE.</p>
Full article ">
12 pages, 216 KiB  
Systematic Review
Use of Beta-Blockers as a First-Line Treatment for Primary Hypertension
by Maryam Izadi, Shiva Shafabakhsh and Amir Mirnateghi
Hearts 2024, 5(4), 460-471; https://doi.org/10.3390/hearts5040033 - 22 Oct 2024
Viewed by 706
Abstract
Background: Even though beta-blockers had been used as a first-line therapy for hypertension, since the late 1960s, the Eighth Joint National Committee, JNC 8, decided to recommend them no longer. This decision was based on relatively weak evidence from previous studies, which [...] Read more.
Background: Even though beta-blockers had been used as a first-line therapy for hypertension, since the late 1960s, the Eighth Joint National Committee, JNC 8, decided to recommend them no longer. This decision was based on relatively weak evidence from previous studies, which found that first-line beta-blockers were less effective in reducing stroke and heart failure, the main outcomes of hypertension. Despite the general perception, the most common events caused by hypertension are death and MI, not stroke or heart failure. Therefore, this study aimed to clarify beta-blocker efficacy by incorporating the data from all relevant beta-blocker trials, using the composite outcome of major cardiovascular events. Method: A search was conducted on MEDLINE, PubMed, Embase, and the Cochrane Library, restricted to published, peer-reviewed, human, meta-analysis, and controlled clinical trials. The term words used were “beta-blockers or adrenergic beta antagonists”, “hypertension”, and “death or coronary heart disease or stroke or congestive heart failure or myocardial infarction”. For this research, we selected six randomized controlled trials, and three meta-analyses were also chosen. Results: The results showed that beta-blockers were as effective as other first-line medications in younger hypertensive patients. On the other hand, in the patients aged above 60, the results were mixed. Beta-blockers were more effective than diuretics, but inferior to angiotensin receptor blockers. Also, beta-blockers were as safe and effective as angiotensin-converting enzyme inhibitors in reducing coronary heart disease, myocardial infarction, heart failure, and sudden death. However, beta-blockers were inferior to calcium channel blockers in reducing strokes. Conclusions: Beta-blockers were found to be the most effective in many aspects except for strokes. Further studies are needed to assess beta-blockers’ effectiveness in treating primary hypertension. Full article
Back to TopTop