[go: up one dir, main page]
More Web Proxy on the site http://driver.im/
You seem to have javascript disabled. Please note that many of the page functionalities won't work as expected without javascript enabled.
 
 

The Role of Ion Channels and Transporters in Human Health and Diseases

A special issue of Biomedicines (ISSN 2227-9059). This special issue belongs to the section "Molecular and Translational Medicine".

Deadline for manuscript submissions: 31 December 2024 | Viewed by 11558

Special Issue Editor


E-Mail Website
Guest Editor
Department of Health Science, Laboratory of Physiology and Neuropharmacology, University Magna Græcia of Catanzaro, 88100 Catanzaro, Italy
Interests: synaptic plasticity; NMDAreceptor; autism; endothelial cell; neurovascular coupling; Ca2+ signaling; nitric oxide

Special Issue Information

Dear Colleagues,

The human body relies on a delicate balance of various elements, including the precise movement of ions, electrically charged atoms, or molecules, across cell membranes. This critical task requires two crucial players: ion channels and transporters. An intricate network of membrane proteins orchestrates the flow of ions, dictating their entry and exit from cells and their release or sequestration by endogenous organelles. This meticulously controlled movement underpins numerous physiological processes, forming the very foundation of human health. Ion channels act as selective pores, opening and closing in response to various stimuli like voltage changes or the presence of specific molecules. They allow specific ions, like sodium, potassium, chloride, and calcium, to pass through the membrane, generating electrical signals and enabling communication within and between cells. In addition, water channels, known as aquaporin, are pivotal to fine-tune cellular volume during osmotic challenges. On the other hand, transporters actively move ions and solutes according to or against their concentration gradient, ensuring their proper distribution and controlling their concentration within cells. This precise control is vital for maintaining homeostasis, the stable internal environment required for cellular function. Transporters also play a key role in nutrient absorption, facilitating the uptake of essential molecules from the gut into the bloodstream for distribution throughout the body. The importance of ion channels and transporters cannot be overstated. Their proper function is vital for countless physiological processes, including synaptic transmission and plasticity, muscle contraction, fluid balance, secretion, and absorption. When these proteins malfunction, it can lead to a cascade of consequences, disrupting the delicate cellular equilibrium and causing various illnesses. Therefore, understanding the intricacies of ion channels and transporters is critical for comprehending human health and disease. The following sections will explore the specific roles these proteins play in various diseases and the potential for therapeutic interventions based on this knowledge.

I am pleased to invite you to participate in this Special Issue. Experimental papers and up-to-date review articles are all welcome.

Dr. Teresa Soda
Guest Editor

Manuscript Submission Information

Manuscripts should be submitted online at www.mdpi.com by registering and logging in to this website. Once you are registered, click here to go to the submission form. Manuscripts can be submitted until the deadline. All submissions that pass pre-check are peer-reviewed. Accepted papers will be published continuously in the journal (as soon as accepted) and will be listed together on the special issue website. Research articles, review articles as well as short communications are invited. For planned papers, a title and short abstract (about 100 words) can be sent to the Editorial Office for announcement on this website.

Submitted manuscripts should not have been published previously, nor be under consideration for publication elsewhere (except conference proceedings papers). All manuscripts are thoroughly refereed through a single-blind peer-review process. A guide for authors and other relevant information for submission of manuscripts is available on the Instructions for Authors page. Biomedicines is an international peer-reviewed open access monthly journal published by MDPI.

Please visit the Instructions for Authors page before submitting a manuscript. The Article Processing Charge (APC) for publication in this open access journal is 2600 CHF (Swiss Francs). Submitted papers should be well formatted and use good English. Authors may use MDPI's English editing service prior to publication or during author revisions.

Keywords

  • ion channels
  • transporters
  • ion pumps
  • voltage-gated channels
  • ionotropic receptors
  • TRP channels
  • store-operated channels
  • Piezo channels
  • NALCN channels
  • inositol-1,4,5-receptors
  • ryanodine receptors
  • two-pore channels
  • synaptic transmission
  • channelopathies
  • neurodegenerative disorders
  • cardiovascular disorders

Benefits of Publishing in a Special Issue

  • Ease of navigation: Grouping papers by topic helps scholars navigate broad scope journals more efficiently.
  • Greater discoverability: Special Issues support the reach and impact of scientific research. Articles in Special Issues are more discoverable and cited more frequently.
  • Expansion of research network: Special Issues facilitate connections among authors, fostering scientific collaborations.
  • External promotion: Articles in Special Issues are often promoted through the journal's social media, increasing their visibility.
  • e-Book format: Special Issues with more than 10 articles can be published as dedicated e-books, ensuring wide and rapid dissemination.

Further information on MDPI's Special Issue polices can be found here.

Published Papers (6 papers)

Order results
Result details
Select all
Export citation of selected articles as:

Research

Jump to: Review

16 pages, 1838 KiB  
Article
SGLT-2 Inhibitors’ and GLP-1 Receptor Agonists’ Influence on Neuronal and Glial Damage in Experimental Stroke
by Anna Murasheva, Oksana Fuks, Natalya Timkina, Arina Mikhailova, Timur Vlasov, Konstantin Samochernykh and Tatiana Karonova
Biomedicines 2024, 12(12), 2797; https://doi.org/10.3390/biomedicines12122797 - 10 Dec 2024
Viewed by 370
Abstract
Background: SGLT-2 inhibitors (SGLT-2i) and GLP-1 receptor agonists (GLP-1RA) have demonstrated nephro- and cardioprotective effects, but their neuroprotective properties, especially concerning stroke severity, and mechanisms are not unambiguous. We aimed to study the influence of SGLT-2i with different selectivity and GLP-1RA on brain [...] Read more.
Background: SGLT-2 inhibitors (SGLT-2i) and GLP-1 receptor agonists (GLP-1RA) have demonstrated nephro- and cardioprotective effects, but their neuroprotective properties, especially concerning stroke severity, and mechanisms are not unambiguous. We aimed to study the influence of SGLT-2i with different selectivity and GLP-1RA on brain damage volume and neurological status in non-diabetic and diabetic rats and to investigate the underlying mechanisms. Methods: Non-diabetic Wistar rats were divided into five groups (n = 10 each) and received empagliflozin, canagliflozin, or dulaglutide as study drugs and metformin as comparison drug. Control animals were administered 0.9% NaCl for 7 days before stroke. At 48 h after stroke, we assessed neurological deficit, neuronal and astroglial damage markers, and brain damage volume. We also modeled type 2 DM in Wistar rats using the high-fat diet+nicotinamide/streptozotocin method and established similar treatment groups. After 8 weeks, rats were subjected to stroke with further neurological deficit, neuroglial damage markers, and brain necrosis volume measurement. Results: In non-diabetic rats, all the drugs showed an infarct-limiting effect; SGLT-2i and dulaglutide were more effective than metformin. DULA improved neurological status compared with MET and SGLT-2i treatment. All the drugs decreased neurofilament light chains (NLCs) level and neuronal damage markers, but none of them decreased the glial damage marker S100BB. In DM, similarly, all the drugs had infarct-limiting effects. Neurological deficit was most pronounced in the untreated diabetic rats and was reduced by all study drugs. All the drugs reduced NLC level; dulaglutide and empagliflozin, but not canagliflozin, also decreased S100BB. None of the drugs affected neuron-specific enolase. Conclusions: SGLT-2i and GLP-1RA are neuroprotective in experimental stroke. GLP-1RA might be more effective than SGLT-2i as in non-diabetic conditions it influences both brain damage volume and neurological status. All study drugs decrease neuronal damage, while GLP-1RA and highly selective SGLT-2i EMPA, but not low-selective CANA, also have an impact on neuroglia in diabetic conditions. Full article
Show Figures

Figure 1

Figure 1
<p>Neurological deficit 48 h after MCAO in non-diabetic rats. # <span class="html-italic">p</span> &lt; 0.05 in comparison with the “MET” group, ¶ <span class="html-italic">p</span> &lt; 0.05 in comparison with the “DULA” group. MCAO—middle cerebral artery occlusion. Metformin, empagliflozin, and canagliflozin administration for 7 days prior to MCAO did not decrease neurological deficit. Neurological deficit was significantly smaller in the “DULA” group in comparison with all the other treatment groups.</p>
Full article ">Figure 2
<p>Ischemia–reperfusion injury-induced brain damage in non-diabetic rats. (<b>A</b>) Brain damage volume measurement results presented as dot plots with median values. (<b>B</b>) Representative images of brain slices stained with triphenyltetrazolium chloride. * <span class="html-italic">p</span> &lt; 0.05 in comparison with the “Control” group, # <span class="html-italic">p</span> &lt; 0.05 in comparison with the “MET” group. Metformin, empagliflozin, canagliflozin, and dulaglutide reduce brain damage volume in comparison with the control group without treatment 48 h after MCAO in non-diabetic rats. The infarct-limiting effect of empagliflozin and dulaglutide is more prominent than that of metformin.</p>
Full article ">Figure 3
<p>Concentration of neuronal and glial damage markers after MCAO in non-diabetic rats. (<b>A</b>) Neurofilament light chains level 48 h after stroke. (<b>B</b>) Neuron-specific enolase level 48 h after stroke. (<b>C</b>) S100BB level 48 h after stroke. * <span class="html-italic">p</span> &lt; 0.05 in comparison with the “Control” group, ¶ <span class="html-italic">p</span> &lt; 0.05 in comparison with the “DULA” group. Dashed line—normal value. Ischemic stroke is characterized by NLC, NSE, and S100BB elevation. Metformin, empagliflozin, canagliflozin, and dulaglutide decreased NLC compared with the “Control” group. None of the drugs significantly influenced NSE or S100BB level.</p>
Full article ">Figure 4
<p>Glycemia dynamics in diabetic rats receiving variable glucose-lowering drugs. Strept.—nicotinamide/streptozotocin administration. * <span class="html-italic">p</span> &lt; 0.05 in comparison with the “Control” group, § <span class="html-italic">p</span> &lt; 0.05 in comparison with the “DM” group. Eight-week metformin, empagliflozin, canagliflozin, and dulaglutide treatment in diabetic rats caused similar glycemic profile improvement in comparison with untreated diabetes mellitus.</p>
Full article ">Figure 5
<p>Neurological deficit 48 h after MCAO in diabetic rats. * <span class="html-italic">p</span> &lt; 0.05 in comparison with the “Control” group, § <span class="html-italic">p</span> &lt; 0.05 in comparison with the “DM” group, # <span class="html-italic">p</span> &lt; 0.05 in comparison with the “DM+MET” group, ¶ <span class="html-italic">p</span> &lt; 0.05 in comparison with the “DM+DULA” group. MCAO—middle cerebral artery occlusion. Neurological deficit in diabetic rats without treatment was more serious than in the “Control” group. Metformin, empagliflozin, canagliflozin, and dulaglutide improved neurological status. There was no significant difference in empagliflozin and canagliflozin effectiveness, whereas the positive effect of canagliflozin was more prominent than that of metformin and dulaglutide.</p>
Full article ">Figure 6
<p>Ischemia–reperfusion injury-induced brain damage in diabetic rats. (<b>A</b>) Brain damage volume measurement results, presented as dot plots with median values. (<b>B</b>) Representative images of brain slices stained with triphenyltetrazolium chloride. *: <span class="html-italic">p</span> &lt; 0.05 in comparison with the “Control” group, §: <span class="html-italic">p</span> &lt; 0.05 in comparison with the “DM” group, #: <span class="html-italic">p</span> &lt; 0.05 in comparison with the “DM+MET” group. Brain damage volume in the diabetic rats without treatment was as large as that in the “Control” group. All study drugs diminished necrosis volume in comparison with the “DM” group. There was no difference in the infarct-limiting effects of empagliflozin, canagliflozin, and dulaglutide.</p>
Full article ">Figure 7
<p>Concentration of neuronal and glial damage markers after MCAO in diabetic rats. (<b>A</b>) Neurofilament light chains level 48 h after stroke. (<b>B</b>) Neuron-specific enolase level 48 h after stroke. (<b>C</b>) S100BB level 48 h after stroke. *: <span class="html-italic">p</span> &lt; 0.05 in comparison with the “Control” group, §: <span class="html-italic">p</span> &lt; 0.05 in comparison with the “DM” group, #: <span class="html-italic">p</span> &lt; 0.05 in comparison with the “DM+MET” group. Dashed line—normal value. NLC concentration was elevated in both the “DM” and the “Control” groups, but metformin, empagliflozin, canagliflozin, and dulaglutide caused similar decreases in it. NSE levels were similar in all the study groups. S100BB was similarly elevated in the “DM” and “Control” groups. Empagliflozin and dulaglutide caused its significant decrease, while metformin and canagliflozin did not influence it.</p>
Full article ">
19 pages, 4438 KiB  
Article
The Functional Interaction of KATP and BK Channels with Aquaporin-4 in the U87 Glioblastoma Cell
by Fatima Maqoud, Laura Simone, Domenico Tricarico, Giulia Maria Camerino, Marina Antonacci and Grazia Paola Nicchia
Biomedicines 2024, 12(8), 1891; https://doi.org/10.3390/biomedicines12081891 - 19 Aug 2024
Viewed by 3759
Abstract
K+ channels do play a role in cell shape changes observed during cell proliferation and apoptosis. Research suggested that the dynamics of the aggregation of Aquaporin-4 (AQP4) into AQP4-OAP isoforms can trigger cell shape changes in malignant glioma cells. Here, we investigated [...] Read more.
K+ channels do play a role in cell shape changes observed during cell proliferation and apoptosis. Research suggested that the dynamics of the aggregation of Aquaporin-4 (AQP4) into AQP4-OAP isoforms can trigger cell shape changes in malignant glioma cells. Here, we investigated the relationship between AQP4 and some K+ channels in the malignant glioma U87 line. The U87 cells transfected with the human M1-AQP4 and M23-AQP4 isoforms were investigated for morphology, the gene expression of KCNJ8, KCNJ11, ABCC8, ABCC9, KCNMA1, and Cyclin genes by RT-PCR, recording the whole-cell K+ ion currents by patch-clamp experiments. AQP4 aggregation into OAPs increases the plasma membrane functional expression of the Kir6.2 and SUR2 subunits of the KATP channels and of the KCNMA1 of the BK channels in U87 cells leading to a large increase in inward and outward K+ ion currents. These changes were associated with changes in morphology, with a decrease in cell volume in the U87 cells and an increase in the ER density. These U87 cells accumulate in the mitotic and G2 cell cycle. The KATP channel blocker zoledronic acid reduced cell proliferation in both M23 AQP4-OAP and M1 AQP4-tetramer-transfected cells, leading to early and late apoptosis, respectively. The BK channel sustains the efflux of K+ ions associated with the M23 AQP4-OAP expression in the U87 cells, but it is downregulated in the M1 AQP4-tetramer cells. The KATP channels are effective in the M1 AQP4-tetramer and M23 AQP4-OAP cells. Zoledronic acid can be effective in targeting pathogenic M1 AQP4-tetramer cell phenotypes inhibiting KATP channels and inducing early apoptosis. Full article
Show Figures

Figure 1

Figure 1
<p>Expression of AQP4 in U87 cells. (<b>A</b>) Epifluorescence images of U87wt and U87 expressing M1 AQP4-tetramers or M23 AQP4-OAPs. AQP4 staining is shown in red and DAPI in blue. Scale bar 20 μm. (<b>B</b>) Immunoblot detection of AQP4 expression levels in U87 cells transfected with M1-AQP4 (AQP4-tetramers) and M23-AQP4 (AQP4-OAPs). GFAP and GAPDH were used to normalize for equal loading.</p>
Full article ">Figure 2
<p>Characterization of inward and outward macroscopic K<sup>+</sup> ion currents recorded in U87wt cells. The currents were recorded using a whole cell configuration under physiological concentration of K<sup>+</sup> ions in the bath and pipette and were obtained in response to voltage pulses from −120 to +120 mV in increments of 20 mV, starting at HP = −60 mV (Vm) and intracellular-free Ca<sup>2+</sup> ions (1.6 × 10<sup>−6</sup> M). (<b>A</b>) TEA (5 × 10<sup>−3</sup> M) suppressed the outward K<sup>+</sup> ion currents, and TEA/Ba<sup>2+</sup> (5 × 10<sup>−3</sup> M) fully suppressed also the inward K<sup>+</sup> ion currents. (<b>B</b>) The selective BK channel blocker IbTX (4 × 10<sup>−7</sup> M) reduced the outward K<sup>+</sup> ion currents that were fully reduced by TEA (5 × 10<sup>−3</sup> M). (<b>C</b>) KATP currents recorded in U87wt cells in low intracellular ATP 1 × 10<sup>−3</sup> M. Glibenclamide (5 × 10<sup>−8</sup> M) suppressed the inward currents, suggesting the presence of KATP channels in the cells. Cells of the same size were selected for patch-clamp experiments. Each point represented the mean ± SEM (N patches = 10<sup>–15</sup>). (<b>D</b>) Percentage of blocking of K<sup>+</sup> ion currents in the presence of the antagonists with respect to the control condition at 60 mV and −60 mV (Vm).</p>
Full article ">Figure 3
<p>Effects of antagonists and agonists of the K<sup>+</sup> channels on the proliferation of U87wt cells. * <span class="html-italic">p</span> &lt; 0.05 and ** <span class="html-italic">p</span> &lt; 0.01 when compared to the experimental group not treated after 48 h of incubation. Each point represented the mean ± SEM.</p>
Full article ">Figure 4
<p>Characterization of macroscopic inward and outward K<sup>+</sup> ion currents recorded in U87wt cells and after transfection with M1 AQP4-tetramer or M23 AQP4-OAP. Whole cell currents were recorded under physiological concentration of K<sup>+</sup> ions in the bath and pipette and were obtained in response to voltage pulses from −120 to +120 mV in 20 mV increments, starting at HP = −60 mV (Vm). (<b>A</b>) Macroscopic K<sup>+</sup> ion currents recorded in AQP4-OAP-transfected U87 cells. The presence of AQP4-OAPs caused a significant increase in the inward and outward currents that were reduced by TEA and BaCl<sub>2</sub>. (<b>B</b>) Macroscopic K<sup>+</sup> ion currents recorded in AQP4-tetramer-transfected U87 cells. The presence of M1 AQP4-tetramers caused a significant increase in the inward currents; however, the outward K<sup>+</sup> ion currents were reduced in the amplitude. (<b>C</b>) The presence of M23 AQP4-OAPs caused a large increase in the currents at negative and positive membrane potentials vs. not transfected cell. Instead, the presence of AQP4-tetramers led to an increase in the currents at negative membrane potentials; conversely, the outward K<sup>+</sup> ion currents decreased at positive membrane potentials. Data were pooled from N patches = 10–12. (<b>D</b>) Percentage of reduction in the K<sup>+</sup> ion currents in the presence of the antagonist TEA with respect to the control condition at 60 mV and −60 mV (Vm). Each point represents the mean ± SEM.</p>
Full article ">Figure 5
<p>Expression profile of <span class="html-italic">KCNMA1</span>, <span class="html-italic">KCNJ11</span>, <span class="html-italic">KCNJ8</span>, <span class="html-italic">ABCC8</span>, <span class="html-italic">ABCC9</span>, and <span class="html-italic">TRPV1</span> genes in U87 glioma cells in the presence of the malignant M1 AQP4-tetramer and M23 AQP4-OAP aggregation vs. wt condition. * <span class="html-italic">p</span> &lt; 0.05 when compared to the experimental group not treated (WT) after 48 h of incubation. Each point represented the mean ± SEM.</p>
Full article ">Figure 6
<p>Effect of pharmacological inhibition of the Kir6.2-SUR2 channel activity with zoledronic acid on AQP4s expressing U87 cells. (<b>A</b>) Epifluorescence images of U87 cells expressing AQP4-tetramers or AQP4-OAPs treated with diazoxide or zoledronic acid. AQP4 staining is shown in red, and DAPI for nuclear staining is in blue. Phalloidin (in green) was used to visualize F-actin. The white arrowheads indicate the round-shaped cells, the arrows indicate irregular-shaped cells, and the blue arrowheads indicate the apoptotic beads. Scale bar 100 μm. ((<b>B</b>) top) Drawing/diagram showing the morphological change of U87 cells expressing AQP4-OAPs after pharmacological treatment with zoledronic acid according Coffin hypothesis and relative epifluorescence images. ((<b>B</b>) bottom) Epifluorescence images of U87 expressing AQP4-OAPs treated with zoledronic, showing the irregular-shaped cell, the round-shaped cell, and the apoptosis. AQP4 staining is shown in red. Phalloidin (in green) was used to visualize F-actin. Scale bar 10 µm. A histogram was created to display the percentage of round-shaped cells per field in U87 cells and those expressing AQP4-OAPs after treatment under different conditions. The conditions include treatment with zoledronic acid (an inhibitor), control conditions (vehicle), and diazoxide (an agonist). The histogram compares these percentages across the different treatment groups, highlighting the effects of the inhibitor and agonist on cell morphology. Values are expressed as mean ± SEM of percentage of cells with altered cell morphology out of the total number of transfected cells per field. ** <span class="html-italic">p</span> &lt; 0.005; ((<b>C</b>) top) Representative image of expressing AQP4-OAPs in control condition or after treatment with zoledronic or diazoxide as indicated and stained with DAPI to visualize nuclei. Scale bar 50 µm. ((<b>C</b>) bottom) Dot plot showing the analysis of % of condensed nuclei for field and the nuclear area of images in A. Values are expressed in µm<sup>2</sup> and represent mean ± SEM. * <span class="html-italic">p</span> &lt; 0.05, *** <span class="html-italic">p</span> &lt; 0.001, <span class="html-italic">n</span> = 3, two-way ANOVA/Tukey’s tests.</p>
Full article ">Figure 7
<p>RT PCR expression of the cyclin genes (E, A, and B1) in the U87 cells following M23 AQP4-OAP transfection. * <span class="html-italic">p</span> &lt; 0.05 when compared to the experimental group not treated (WT) after 48 h of incubation. • represent the average of each single experiment. Each point represented the mean of at least three experiments± SEM.</p>
Full article ">
13 pages, 3173 KiB  
Article
Chronic Mexiletine Administration Increases Sodium Current in Non-Diseased Human Induced Pluripotent Stem Cell-Derived Cardiomyocytes
by Giovanna Nasilli, Arie O. Verkerk, Molly O’Reilly, Loukia Yiangou, Richard P. Davis, Simona Casini and Carol Ann Remme
Biomedicines 2024, 12(6), 1212; https://doi.org/10.3390/biomedicines12061212 - 29 May 2024
Cited by 1 | Viewed by 1499
Abstract
A sodium current (INa) reduction occurs in the setting of many acquired and inherited conditions and is associated with cardiac conduction slowing and increased arrhythmia risks. The sodium channel blocker mexiletine has been shown to restore the trafficking of mutant sodium [...] Read more.
A sodium current (INa) reduction occurs in the setting of many acquired and inherited conditions and is associated with cardiac conduction slowing and increased arrhythmia risks. The sodium channel blocker mexiletine has been shown to restore the trafficking of mutant sodium channels to the membrane. However, these studies were mostly performed in heterologous expression systems using high mexiletine concentrations. Moreover, the chronic effects on INa in a non-diseased cardiomyocyte environment remain unknown. In this paper, we investigated the chronic and acute effects of a therapeutic dose of mexiletine on INa and the action potential (AP) characteristics in human induced pluripotent stem cell-derived cardiomyocytes (hiPSC-CMs) of a healthy individual. Control hiPSC-CMs were incubated for 48 h with 10 µM mexiletine or vehicle. Following the wash-out of mexiletine, patch clamp analysis and immunocytochemistry experiments were performed. The incubation of hiPSC-CMs for 48 h with mexiletine (followed by wash-out) induced a significant increase in peak INa of ~75%, without any significant change in the voltage dependence of (in)activation. This was accompanied by a significant increase in AP upstroke velocity, without changes in other AP parameters. The immunocytochemistry experiments showed a significant increase in membrane Nav1.5 fluorescence following a 48 h incubation with mexiletine. The acute re-exposure of hiPSC-CMs to 10 µM mexiletine resulted in a small but significant increase in AP duration, without changes in AP upstroke velocity, peak INa density, or the INa voltage dependence of (in)activation. Importantly, the increase in the peak INa density and resulting AP upstroke velocity induced by chronic mexiletine incubation was not counteracted by the acute re-administration of the drug. In conclusion, the chronic administration of a clinically relevant concentration of mexiletine increases INa density in non-diseased hiPSC-CMs, likely by enhancing the membrane trafficking of sodium channels. Our findings identify mexiletine as a potential therapeutic strategy to enhance and/or restore INa and cardiac conduction. Full article
Show Figures

Figure 1

Figure 1
<p>Incubation of hiPSC-CMs for 48 h with 10 µM mexiletine increases I<sub>Na</sub> density. (<b>A</b>) Schematic representation of the experimental approach. (<b>B</b>) Representative I<sub>Na</sub> traces for hiPSC-CMs recorded after a 48 h incubation with the vehicle or mexiletine (10 µM). Average current-voltage relationships in hiPSC-CMs (<b>C</b>) and voltage dependence of activation (<b>D</b>) and inactivation (<b>E</b>). Insets: voltage clamp protocols; <span class="html-italic">n</span>, number of cells; * <span class="html-italic">p</span> &lt; 0.05, two-way RM ANOVA followed by a Holm-Sidak test for post hoc analysis.</p>
Full article ">Figure 2
<p>Immunostaining of hiPSC-CMs treated with vehicle or 10 µM mexiletine for 48 h. (<b>A</b>) Example images of hiPSC-CMs stained for DAPI (blue), α-actinin (red), and Na<sub>V</sub>1.5 (green). (<b>B</b>) Quantification of Na<sub>V</sub>1.5 membrane fluorescence intensity. Scale bar is 25 µm; <span class="html-italic">n</span> represents the number of cells analyzed. * <span class="html-italic">p</span> &lt; 0.05, unpaired Student’s <span class="html-italic">t</span>-test.</p>
Full article ">Figure 3
<p>Acute effect of mexiletine on I<sub>Na</sub> density and I<sub>Na</sub> voltage dependence of (in)activation in hiPSC-CMs. (<b>A</b>) Schematic representation of the experimental approach. (<b>B</b>,<b>C</b>) Average sodium current-voltage relationships of hiPSC-CMs (previously incubated for 48 h with the vehicle (<b>B</b>), or mexiletine (<b>C</b>)) at baseline and after the acute (re-)administration (5 min) of 10 µM mexiletine. (<b>D</b>–<b>G</b>) Voltage dependence of activation (<b>D</b>,<b>E</b>) and inactivation (<b>F</b>,<b>G</b>) measured in hiPSC-CMs (previously incubated for 48 h with the vehicle (<b>D</b>,<b>F</b>), or mexiletine (<b>E</b>,<b>G</b>)) at baseline and after the acute (re-)administration (5 min) of 10 µM mexiletine. Insets: voltage clamp protocols; <span class="html-italic">n</span>, number of cells. V<sub>1/2</sub> and slope factor <span class="html-italic">k</span> values are listed in <a href="#biomedicines-12-01212-t002" class="html-table">Table 2</a>.</p>
Full article ">Figure 4
<p>Incubation of hiPSC-CMs for 48 h with 10 µM mexiletine increases action potential upstroke velocity without affecting repolarization. (<b>A</b>) Schematic representation of the experimental approach. (<b>B</b>) Representative action potential (AP) traces of hiPSC-CMs incubated for 48 h with the vehicle or mexiletine (10 µM). Insets: first derivative of the AP upstroke. (<b>C</b>) Average values for the AP upstroke velocity (V<sub>max</sub>), AP amplitude (APA), resting membrane potential (RMP), and AP duration at 20%, 50% and 90% repolarization (APD<sub>20</sub>, APD<sub>50</sub>, and APD<sub>90</sub>, respectively), measured at a pacing frequency of 1 Hz. <span class="html-italic">n</span>, number of cells. * <span class="html-italic">p</span> &lt; 0.05, unpaired Student’s <span class="html-italic">t</span>-test.</p>
Full article ">Figure 5
<p>Acute effect of mexiletine on action potential properties in hiPSC-CMs. (<b>A</b>) Schematic representation of the experimental approach. (<b>B</b>) Average data for the maximal upstroke velocity (V<sub>max</sub>), AP amplitude (APA), resting membrane potential (RMP), AP duration at 20%, 50%, and 90% repolarization (APD<sub>20</sub>, APD<sub>50</sub>, and APD<sub>90</sub>, respectively) in hiPSC-CMs after a 48 h incubation with the vehicle (baseline), after 5 min of the acute administration of 10 µM mexiletine (wash-in), and after the wash-out of the drug. (<b>C</b>) Average data for V<sub>max</sub>, APA, RMP, APD<sub>20</sub>, APD<sub>50</sub>, and APD<sub>90</sub> in hiPSC-CMs after a 48 h incubation with 10 µM mexiletine (baseline), after 5 min of the acute re-administration of 10 µM mexiletine (wash-in), and after wash-out. <span class="html-italic">n</span>, number of cells; * <span class="html-italic">p</span> &lt; 0.05, one-way RM ANOVA followed by a Holm-Sidak test for post hoc analysis or a one-way RM ANOVA on Ranks (Friedman test) followed by Tukey’s test for post hoc analysis when the data were not normally distributed.</p>
Full article ">

Review

Jump to: Research

29 pages, 3011 KiB  
Review
Sodium Chloride Cotransporter in Hypertension
by Annalisa Castagna, Gabriele Mango, Nicola Martinelli, Luigi Marzano, Sara Moruzzi, Simonetta Friso and Francesca Pizzolo
Biomedicines 2024, 12(11), 2580; https://doi.org/10.3390/biomedicines12112580 - 11 Nov 2024
Viewed by 1444
Abstract
The sodium chloride cotransporter (NCC) is essential for electrolyte balance, blood pressure regulation, and pathophysiology of hypertension as it mediates the reabsorption of ultrafiltered sodium in the renal distal convoluted tubule. Given its pivotal role in the maintenance of extracellular fluid volume, the [...] Read more.
The sodium chloride cotransporter (NCC) is essential for electrolyte balance, blood pressure regulation, and pathophysiology of hypertension as it mediates the reabsorption of ultrafiltered sodium in the renal distal convoluted tubule. Given its pivotal role in the maintenance of extracellular fluid volume, the NCC is regulated by a complex network of cellular pathways, which eventually results in either its phosphorylation, enhancing sodium and chloride ion absorption from urines, or dephosphorylation and ubiquitination, which conversely decrease NCC activity. Several factors could influence NCC function, including genetic alterations, hormonal stimuli, and pharmacological treatments. The NCC’s central role is also highlighted by several abnormalities resulting from genetic mutations in its gene and consequently in its structure, leading to dysregulation of blood pressure control. In the last decade, among other improvements, the acquisition of knowledge on the NCC and other renal ion channels has been favored by studies on extracellular vesicles (EVs). Dietary sodium and potassium intake are also implicated in the tuning of NCC activity. In this narrative review, we present the main cornerstones and recent evidence related to NCC control, focusing on the context of blood pressure pathophysiology, and promising new therapeutical approaches. Full article
Show Figures

Graphical abstract

Graphical abstract
Full article ">Figure 1
<p>Sodium reabsorption along the nephron. Most of the sodium reabsorption occurs in the proximal tubule (pink) via NHE3, which mediates about 70% of the Na<sup>+</sup> transport, followed by NKCC2 in the thick ascending limb (green), which is responsible for 25%. About 5% of Na<sup>+</sup> reabsorption occurs in the distal tubule, which includes the distal convoluted tubule (purple) and the collecting duct (magenta). The NCC is located only in the proximal part of the distal convoluted tubule, while ENaC is present in both distal convoluted tubule and collecting duct. A representation of the NCC dimeric structure is also illustrated in the blow-up. Figure created in <a href="http://BioRender.com" target="_blank">BioRender.com</a> [<a href="#B6-biomedicines-12-02580" class="html-bibr">6</a>].</p>
Full article ">Figure 2
<p>Three-dimensional structure of the NCC dimer embedded in the apical cell membrane. This schematic representation highlights the functional domains of the transporter: (1) the extracellular domain (green); (2) the transmembrane domain (red); and (3) the cytosolic domain (blue). A close-up of one ion binding pocket, located within the transmembrane domain, shows three binding sites: one for Na<sup>+</sup> (magenta); one for Cl<sup>−</sup> (cyan); and one for polythiazide, a thiazide diuretic (yellow). The representation is based on the structural prediction provided by the Orientations of Proteins in Membranes (OPM) database [<a href="#B22-biomedicines-12-02580" class="html-bibr">22</a>] of the NCC inward-facing conformation obtained by Fan et al. [<a href="#B20-biomedicines-12-02580" class="html-bibr">20</a>] (PDB ID: 8FHO). PyMOL (Schrödinger, LLC) was used to visualize the structure and generate the image. Figure created in <a href="http://BioRender.com" target="_blank">BioRender.com</a> [<a href="#B6-biomedicines-12-02580" class="html-bibr">6</a>].</p>
Full article ">Figure 3
<p>Post-translational modifications regulating NCC activity in the DCT. NCC phosphorylation (yellow) is primarily mediated by the WNK-SPAK/OSR1 kinase cascade. WNK1 and WNK4 (with-no-lysine kinases) activate SPAK (STE20/SPS1-related proline/alanine-rich kinase) and OSR1 (oxidative stress-response protein 1), which, in turn, phosphorylate the NCC, increasing its transport activity in the apical membrane. WNK autophosphorylation is stimulated by low Cl<sup>−</sup> intracellular levels, which result from a decrease in K<sup>+</sup> intracellular concentration. These changes are, respectively, mediated by basolateral chlorine (light blue) and potassium channels (pink). Cab39 (calcium binding protein 39) is another regulator of NCC phosphorylation since it is required for SPAK activation. NCC dephosphorylation (red) leads to a reduction in its activity and is carried out by protein phosphatases PP1, PP3, and PP4. PP1 is regulated by protein phosphatase 1 inhibitor-1 (I-1) through the cAMP/PKA pathway. A rise in cAMP levels activates PKA (protein kinase A), which phosphorylates I-1, preventing PP1 from dephosphorylating the NCC. Ubiquitination (purple) is another key regulatory mechanism that affects NCC function. WNK abundance is controlled by an E3 ubiquitin ligase complex composed of KLHL3 (Kelch-like protein 3) and CUL3 (Cullin 3). This complex tags the NCC with ubiquitin (Ub), marking it for proteasomal degradation. KLHL3 acts as adapter for WNK, and its phosphorylation prevents this binding. PP3 is responsible for KLHL3 activation through dephosphorylation. Additionally, NEDD4-2, another ubiquitin ligase, directly mediates NCC degradation through ubiquitination. Figure created in <a href="http://BioRender.com" target="_blank">BioRender.com</a> [<a href="#B6-biomedicines-12-02580" class="html-bibr">6</a>].</p>
Full article ">Figure 4
<p>Physiological stimuli that modulate NCC activity. The key signaling pathways involved in NCC regulation by various hormones are illustrated. Arginine vasopressin (AVP, blue) binds to the V<sub>2</sub> receptor (V<sub>2</sub>R) on the basolateral membrane, activating adenylate cyclase 6 (AC6), which increases cyclic AMP (cAMP) levels. Elevated cAMP stimulates protein kinase A (PKA), leading to: (a) stimulation of WNK-SPAK/OSR1-NCC cascade through WNK kinase phosphorylation; (b) inhibition of protein phosphatase 1 (PP1) by phosphorylating its inhibitor, I-1, maintaining the phosphorylated state of the NCC; (c) phosphorylation and inactivation of Kelch-like protein 3 (KLHL3), diminishing WNK ubiquitination and consequent degradation. Aldosterone (ALDO, orange) enhances the NCC in both a mineralocorticoid receptor (MR)-dependent and MR-independent manner. Binding to MR, it could stimulate serum- and glucocorticoid-regulated kinase 1 (SGK1), which activates WNKs and simultaneously inhibits NEDD4-2-mediated ubiquitination. On the other hand, aldosterone stimulates ENaC-mediated potassium secretion, eventually inducing hypokalemia. Low extracellular K<sup>+</sup> levels are sensed by DCT cells through Kir4.1/Kir5.1 K<sup>+</sup> channels, which, once active, lead to a drop in K<sup>+</sup> intracellular concentration, stimulating basolateral Cl<sup>−</sup> transport through ClC-Kb. A decrease in intracellular Cl<sup>−</sup> levels releases the inhibition of WNK kinases, enabling their autophosphorylation and thus stimulating WNK-SPAK/OSR1-NCC pathway. Cortisol (CORT, red) may exert the same effect as aldosterone, likely through the interaction with the glucocorticoid receptor (GR). Angiotensin II (ANGII, purple) binds to the angiotensin II type 1 receptor (AT1R), activating protein kinase C (PKC), which catalyzes both WNK and KLHL3 phosphorylation. Angiotensin II activates the NCC via AngII receptor type 1 (AT1R), stimulating the WNK4-SPAK kinase cascade and the phosphorylation of KLHL. Angiotensin II can also induce aldosterone and ENaC-mediated hypokalemia, thereby triggering potassium sensing through Kir4.1/Kir5.1 K<sup>+</sup> channels. Insulin (INS, green) activates the phosphoinositide 3-kinase (PI3K) pathway, leading to the activation of protein kinase B (AKT). AKT stimulates NCC activity through the phosphorylation of both WNK kinases and KLHL3. Figure created in <a href="http://BioRender.com" target="_blank">BioRender.com</a> [<a href="#B6-biomedicines-12-02580" class="html-bibr">6</a>].</p>
Full article ">Figure 5
<p>Pharmacological modulation of NCC function in the DCT. Inhibitors are shown in red, while enhancers are shown in light blue. Thiazide diuretics directly inhibit NCC blocking ion translocation, while mineralocorticoid receptor antagonists (MRA) act indirectly, blocking the MR-mediated effect of aldosterone. Loop diuretics (LD) activate the NCC, increasing the amount of sodium delivered to the DCT and the secretion of aldosterone, which increases NCC activity in both an MR-dependent and an MR-independent manner. Calcineurin inhibitors (CNI) decrease PP3-mediated dephosphorylation of NCC, basolateral K<sup>+</sup> channel, and KLHL3, thus stimulating NCC activity at various levels. Salbutamol (SALB) could activate basolateral K<sup>+</sup> channel through a signaling pathway likely mediated by β<sub>2</sub>-adrenergic receptor. A drop in K<sup>+</sup> intracellular concentration eventually triggers WNK-SPAK/OSR1-NCC cascade. SGLT2 inhibitors (SGLT2i, shown in purple) may influence NCC activity, increasing the delivery of glucose and fructose to the DCT and possibly stimulating a signaling pathway involving the calcium-sensing receptor (CaSR) and protein kinase C (PKC). Figure created in <a href="http://BioRender.com" target="_blank">BioRender.com</a> [<a href="#B6-biomedicines-12-02580" class="html-bibr">6</a>].</p>
Full article ">
34 pages, 3310 KiB  
Review
Two Signaling Modes Are Better than One: Flux-Independent Signaling by Ionotropic Glutamate Receptors Is Coming of Age
by Valentina Brunetti, Teresa Soda, Roberto Berra-Romani, Giovambattista De Sarro, Germano Guerra, Giorgia Scarpellino and Francesco Moccia
Biomedicines 2024, 12(4), 880; https://doi.org/10.3390/biomedicines12040880 - 16 Apr 2024
Cited by 3 | Viewed by 1620
Abstract
Glutamate is the major excitatory neurotransmitter in the central nervous system. Glutamatergic transmission can be mediated by ionotropic glutamate receptors (iGluRs), which mediate rapid synaptic depolarization that can be associated with Ca2+ entry and activity-dependent change in the strength of synaptic transmission, [...] Read more.
Glutamate is the major excitatory neurotransmitter in the central nervous system. Glutamatergic transmission can be mediated by ionotropic glutamate receptors (iGluRs), which mediate rapid synaptic depolarization that can be associated with Ca2+ entry and activity-dependent change in the strength of synaptic transmission, as well as by metabotropic glutamate receptors (mGluRs), which mediate slower postsynaptic responses through the recruitment of second messenger systems. A wealth of evidence reported over the last three decades has shown that this dogmatic subdivision between iGluRs and mGluRs may not reflect the actual physiological signaling mode of the iGluRs, i.e., α-amino-3-hydroxy-5-methyl-4-isoxasolepropionic acid (AMPA) receptors (AMPAR), kainate receptors (KARs), and N-methyl-D-aspartate (NMDA) receptors (NMDARs). Herein, we review the evidence available supporting the notion that the canonical iGluRs can recruit flux-independent signaling pathways not only in neurons, but also in brain astrocytes and cerebrovascular endothelial cells. Understanding the signaling versatility of iGluRs can exert a profound impact on our understanding of glutamatergic synapses. Furthermore, it may shed light on novel neuroprotective strategies against brain disorders. Full article
Show Figures

Graphical abstract

Graphical abstract
Full article ">Figure 1
<p>Flux-independent signaling pathways activated by AMPARs in neurons. AMPARs can signal in a flux-independent manner to recruit several signaling pathways. AMPARs can interact with G<sub>i/o</sub> proteins to either inhibit voltage-gated Ca<sub>V</sub>2.1 channels or activate voltage-gated Na<sup>+</sup> channels (Na<sub>V</sub>). Extracellular Na<sup>+</sup> entry through Na<sub>V</sub> channels can promote mitochondrial Ca<sup>2+</sup> release through the mitochondrial Na<sup>+</sup>/Ca<sup>2+</sup> exchanger (mNCX). In addition, AMPARs can signal through G<sub>i/o</sub> proteins to accelerate vesicle recycling or induce <span class="html-italic">BDNF</span> gene expression through ERK activation, which is mediated by the interaction between AMPARs and the tyrosine kinase, Lyn. Alternately, AMPARs can activate the ERK phosphorylation cascade to promote cell survival via the G<sub>i/o</sub> protein-dependent recruitment of the PI3K/Akt pathway. Finally, AMPARs can also signal through G<sub>i/o</sub> proteins to inhibit the expression of the <span class="html-italic">Arc</span> gene.</p>
Full article ">Figure 2
<p>Flux-independent signaling by KARs modulates the intracellular Cl<sup>−</sup> concentration ([Cl<sup>−</sup>]<sub>i</sub>) and the reversal potential for GABA (E<sub>GABA</sub>). (<b>A</b>) Pharmacological stimulation of postsynaptic KARs with KA (green circle) reduces the [Cl<sup>−</sup>]<sub>i</sub> in hippocampal CA3 pyramidal neurons. Upon permissive cPKC-dependent phosphorylation, KARs promote the recycling of the K<sup>+</sup>–Cl<sup>−</sup> cotransporter 2 (KCC2) from Rab11-containing vesicles to the PM; the tight interaction between KAR and KCC2 stimulates K<sup>+</sup> and Cl<sup>−</sup> efflux into the extracellular milieu, thereby reducing the [Cl<sup>−</sup>]<sub>i</sub> and increasing extracellular Cl<sup>−</sup> influx through GABA<sub>A</sub>Rs. (<b>B</b>) Synaptically released glutamate (red circle) stimulates flux-independent signaling by KARs, which engage cPKC to phosphorylate extrasynaptic GABA<sub>A</sub>Rs and thereby increase extracellular Cl<sup>−</sup> influx in hippocampal CA1 pyramidal neurons.</p>
Full article ">Figure 3
<p>Flux-independent signaling by KARs modulates neuronal excitability and synaptic plasticity. Flux-independent signaling by KAR can increase neuronal excitability by inhibiting I<sub>sAHP</sub> and I<sub>mAHP</sub> via G<sub>i/o</sub> proteins and cPKC. Additionally, KARs can stimulate the exocytosis of AMPARs from Rab1-containing vesicles on the PM of dendritic spines. KARs trigger a signaling cascade involving G<sub>i/o</sub> proteins and phospholipase C (PLC). PLC, in turn, synthesizes the two intracellular second messengers: DAG, which activates cPKC, and InsP<sub>3</sub>, which induces ER Ca<sup>2+</sup> release through InsP<sub>3</sub> receptors (InsP<sub>3</sub>Rs). The combined effect of cPKC and Ca<sup>2+</sup> release results in the recruitment of Rab11 endosomes to the PM, thereby increasing the surface expression of AMPARs.</p>
Full article ">Figure 4
<p>Flux-independent signaling by NMDARs in neurons. (<b>A</b>) Agonist and co-agonist binding leads to the dephosphorylation of GluN1 Y837 and GluN2A Y842, resulting in clathrin-dependent internalization via the adaptor protein AP-2. High concentrations (10 µM) of glycine alone increase the interaction with AP-2, thereby priming NMDARs for dynamin-dependent endocytosis upon agonist binding. (<b>B</b>) Agonist and co-agonist binding to GluN1-containing NMDARs can stimulate the p38 MAPK in a flux-independent manner to promote AMPAR endocytosis, resulting in Ca<sup>2+</sup>-independent LTD induction. (<b>C</b>) Flux-independent signaling by NMDARs regulates spine morphology. In the presence of weak Ca<sup>2+</sup> entry, NMDARs promote dendritic spine shrinkage by triggering a signaling pathway that requires interaction between nNOS and NOS1P, involving p38 MAPK, MK2, and cofilin, which promotes actin depolymerization. This signaling pathway is supported by mTORC1, which is likely to drive new protein synthesis. In the presence of strong Ca<sup>2+</sup> influx (highlighted in red), the Ca<sup>2+</sup>-dependent recruitment of CaMKII leads to dendritic spine growth via inducing actin polymerization. (<b>D</b>) Flux-independent signaling by postsynaptic NMDARs regulates glutamate release by stimulating Src kinase to activate pannexin 1 (PANX1) channels, which clear synaptic anandamide (AEA) and prevent AEA-induced activation of presynaptic TRPV1 channels. This causes a reduction in [Ca<sup>2+</sup>]<sub>i</sub> at the presynaptic terminal and therefore decreases Ca<sup>2+</sup>-dependent glutamate release. By contrast, upon agonist and co-agonist binding, presynaptic NMDARs promote glutamate release via the JNK2-dependent signaling pathway. In all the panels, the red circle indicates the agonist, whereas the blue circle indicates the co-agonist. ↑: increase.</p>
Full article ">Figure 5
<p>Flux-independent signaling by NMDARs in brain disorders. (<b>A</b>) Flux-independent signaling by NMDARs leads to neuronal excitotoxicity upon massive glutamatergic stimulation. NMDARs trigger InsP<sub>3</sub>-induced ER Ca<sup>2+</sup> release, which inhibits the EF-2 protein and interferes with protein synthesis. Furthermore, NMDARs can stimulate Ca<sup>2+</sup> entry through PANX1 channels via Src-dependent phosphorylation of PANX1. Excessive Ca<sup>2+</sup> entry leads to mitochondrial Ca<sup>2+</sup> overload, mPTP opening, and apoptosis. Finally, agonist binding to GluN2B-containing NMDARs can remove the p85 regulatory subunit from the catalytic domain of PI3K, thereby inducing PI3K-dependent NADPH oxidase-2 (NOX2) activation. NOX2 can also be activated via Ca<sup>2+</sup> entry through ionotropic NMDARs and lead to cytotoxic superoxide production. The red circle indicates the agonist, whereas the blue circle indicates the co-agonist. (<b>B</b>) The Aβ protein binds to GluN2B-containing NMDARs to induce a flux-independent signaling pathway that leads to AMPAR removal from the PM and dendritic spine shrinkage via the p38 MAPK signaling pathway. ↓: decrease; ↑: increase.</p>
Full article ">
16 pages, 2385 KiB  
Review
Transporters, Ion Channels, and Junctional Proteins in Choroid Plexus Epithelial Cells
by Masaki Ueno, Yoichi Chiba, Ryuta Murakami, Yumi Miyai, Koichi Matsumoto, Keiji Wakamatsu, Toshitaka Nakagawa, Genta Takebayashi, Naoya Uemura, Ken Yanase and Yuichi Ogino
Biomedicines 2024, 12(4), 708; https://doi.org/10.3390/biomedicines12040708 - 22 Mar 2024
Cited by 3 | Viewed by 1807
Abstract
The choroid plexus (CP) plays significant roles in secreting cerebrospinal fluid (CSF) and forming circadian rhythms. A monolayer of epithelial cells with tight and adherens junctions of CP forms the blood–CSF barrier to control the movement of substances between the blood and ventricles, [...] Read more.
The choroid plexus (CP) plays significant roles in secreting cerebrospinal fluid (CSF) and forming circadian rhythms. A monolayer of epithelial cells with tight and adherens junctions of CP forms the blood–CSF barrier to control the movement of substances between the blood and ventricles, as microvessels in the stroma of CP have fenestrations in endothelial cells. CP epithelial cells are equipped with several kinds of transporters and ion channels to transport nutrient substances and secrete CSF. In addition, junctional components also contribute to CSF production as well as blood–CSF barrier formation. However, it remains unclear how junctional components as well as transporters and ion channels contribute to the pathogenesis of neurodegenerative disorders. In this manuscript, recent findings regarding the distribution and significance of transporters, ion channels, and junctional proteins in CP epithelial cells are introduced, and how changes in expression of their epithelial proteins contribute to the pathophysiology of brain disorders are reviewed. Full article
Show Figures

Graphical abstract

Graphical abstract
Full article ">Figure 1
<p>Representative electron microscopic images of epithelial cells from mouse CP tissues fixed with a mixed solution composed of paraformaldehyde and glutaraldehyde, followed by osmium tetroxide solution, and embedded in epoxy resin. (<b>A</b>) CPEs with junctions on the lateral side and microvilli on the apical side are seen. CPEs facing the ventricle are bound by tight and adherens junctions (indicated by arrows), whereas the lateral intercellular space (LIS) and basal labyrinth (BL) are present in junctional clefts. (<b>B</b>,<b>C</b>) Enlarged images show fenestrations (<b>B</b>: dotted arrow) in fenestrated endothelial cells and tight and adherens junctions (<b>C</b>: arrows). Scale bars indicate 1 μm. AJ: adherens junction; BL: basal labyrinth; FC: fenestrated capillar; FN: fenestration; LIS: lateral intercellular space; Mt: mitochondria; MV: microvilli; TJ: tight junction.</p>
Full article ">Figure 2
<p>Representative immunoreactivity for CD34, ABCG2, and GLUT1 in microvessels of the brain parenchyma with BBB (<b>A</b>–<b>C</b>) and those in the CP stroma (<b>D</b>–<b>F</b>) of human brains. Arrows indicate immunoreactivity for CD34 (<b>A</b>,<b>D</b>), ABCG2 (<b>B</b>,<b>E</b>), and GLUT1 (<b>C</b>), whereas dotted arrows indicate no immunoreactivity for GLUT1 in the CP stroma (<b>F</b>). Arrowheads in (<b>F</b>) indicate positive immunostaining for GLUT1 on the basolateral surface of CPEs. These are compatible with findings reported in previous papers [<a href="#B24-biomedicines-12-00708" class="html-bibr">24</a>,<a href="#B26-biomedicines-12-00708" class="html-bibr">26</a>,<a href="#B33-biomedicines-12-00708" class="html-bibr">33</a>,<a href="#B34-biomedicines-12-00708" class="html-bibr">34</a>]. Scale bars indicate 10 μm.</p>
Full article ">Figure 3
<p>Polarized distribution of representative ion transporters/channels and proteins implicated in CSF secretion in the cytoplasmic membrane of CPEs. AQP1, a water channel, is expressed in large quantities in the luminal membrane, whereas it is also present less frequently in the basolateral membrane. On the apical side of the cytoplasmic membrane of CPEs, sodium potassium ATPase Na<sup>+</sup>,K<sup>+</sup>-ATPase, the sodium hydrogen exchanger 1 NHE1, sodium bicarbonate cotransporter e2 NBCe2, sodium potassium chloride cotransporter 1 NKCC1, potassium chloride cotransporter 4 KCC4, inward rectifying potassium channel Kir7.1, and transient receptor potential vanilloid TRPV4 (a non-selective cation channel) are expressed. On the basal side of the cytoplasmic membrane of CPEs, anion exchange protein AE2 (a chloride bicarbonate exchanger), potassium chloride cotransporter KCC1, Na<sup>+</sup> dependent HCO<sub>3</sub><sup>−</sup> transporter NBCn1, and Na<sup>+</sup>-dependent Cl<sup>−</sup>/HCO<sub>3</sub><sup>−</sup> exchanger <b>Ncbe</b> are expressed. In the junctional space, claudin-2, a component of the tight junction, is involved in the permeation of water as well as monovalent cations, whereas cadherins, components of the adherens junction, have extracellular calcium ion-binding domains. Among MCTs, proton-coupled lactate transporters, MCT1 and MCT2 are immunohistochemically expressed on the apical side of CPEs, whereas MCT3 and MCT4 are considered to be on their basal side. (#1): Directions of Na<sup>+</sup>, K<sup>+</sup>, and Cl<sup>−</sup> through NKCC1 were unidirectional (outward flow) according to some studies [<a href="#B7-biomedicines-12-00708" class="html-bibr">7</a>,<a href="#B62-biomedicines-12-00708" class="html-bibr">62</a>], whereas they were bidirectional according to others [<a href="#B17-biomedicines-12-00708" class="html-bibr">17</a>,<a href="#B63-biomedicines-12-00708" class="html-bibr">63</a>]. Steffensen et al. [<a href="#B62-biomedicines-12-00708" class="html-bibr">62</a>] reported that mouse CP has the ability to transport water against an osmotic gradient in a K<sup>+</sup>-induced, NKCC1-mediated manner [<a href="#B62-biomedicines-12-00708" class="html-bibr">62</a>]. (#2): TRPV4 may have a significant role in controlling ion and water flux. (#3): Cadherins have extracellular calcium ion-binding domains and depend on calcium ions to function. (#4): Claudin−2 likely contributes to the transport of water as well as monovalent cations in TJs. (#5, #6): Transmembranous directions of lactate and proton have not been determined at MCT1 and MCT2 on the apical side of CPEs or at MCT3 and MCT4 on the basal side of CPEs [<a href="#B25-biomedicines-12-00708" class="html-bibr">25</a>,<a href="#B41-biomedicines-12-00708" class="html-bibr">41</a>,<a href="#B49-biomedicines-12-00708" class="html-bibr">49</a>,<a href="#B50-biomedicines-12-00708" class="html-bibr">50</a>]. AJ/TJ: adherens and tight junctions; BL: basal labyrinth; LIS: lateral intercellular space.</p>
Full article ">Figure 4
<p>A schematic illustration of CP and ependymal cells in the normal (left side) and aged (right side) human brain (<b>A</b>), and images of CP stained with hematoxylin and eosin in human brains (<b>B</b>–<b>E</b>). (<b>A</b>) CPEs are bound by TJ/AJ, which are composed of BCSFB. Ependymal cells, mainly bound by gap junctions, are located between the ventricle and brain parenchyma. ISF can pass through the ependymal cell layer into the ventricle and is mixed with CSF. In aged human brains, psammoma bodies are frequently present in the stroma of CPEs, whereas Biondi ring tangles can be occasionally observed in the CP cytoplasm. Fenestrated capillaries without BBB in the CP stroma show CD34 (+), ABCG2 (+), and GLUT1 (−), whereas non-fenestrated capillaries with BBB in the BBB area show CD34 (+), ABCG2 (+), and GLUT1 (+). (<b>B</b>–<b>E</b>) Images stained with hematoxylin and eosin in CPEs of human brains of male patients in their 40 s (<b>B</b>), 80 s (<b>C</b>), and 70 s (<b>D</b>,<b>E</b>) are shown. Thick epithelial cells exhibit a normal-looking morphology, and the stroma is filled with capillaries (<b>B</b>). Epithelial cells (indicated by a long arrow) covering the fibrous stroma are very thin or have disappeared (<b>C</b>). Epithelial cells (indicated by a single arrow) covering the psammoma body (indicated by double arrows) are very thin or have disappeared (<b>D</b>). Biondi ring tangles (indicated by a short arrow) are seen in the cytoplasm of CPEs (<b>E</b>). Scale bars indicate 10 μm.</p>
Full article ">
Back to TopTop