[go: up one dir, main page]

 
 
Sign in to use this feature.

Years

Between: -

Subjects

remove_circle_outline
remove_circle_outline
remove_circle_outline
remove_circle_outline
remove_circle_outline
remove_circle_outline

Journals

Article Types

Countries / Regions

Search Results (202)

Search Parameters:
Keywords = TRPC1

Order results
Result details
Results per page
Select all
Export citation of selected articles as:
29 pages, 1624 KiB  
Review
Are Aminoglycoside Antibiotics TRPing Your Metabolic Switches?
by Alfredo Franco-Obregón and Yee Kit Tai
Cells 2024, 13(15), 1273; https://doi.org/10.3390/cells13151273 - 29 Jul 2024
Viewed by 1475
Abstract
Transient receptor potential (TRP) channels are broadly implicated in the developmental programs of most tissues. Amongst these tissues, skeletal muscle and adipose are noteworthy for being essential in establishing systemic metabolic balance. TRP channels respond to environmental stimuli by supplying intracellular calcium that [...] Read more.
Transient receptor potential (TRP) channels are broadly implicated in the developmental programs of most tissues. Amongst these tissues, skeletal muscle and adipose are noteworthy for being essential in establishing systemic metabolic balance. TRP channels respond to environmental stimuli by supplying intracellular calcium that instigates enzymatic cascades of developmental consequence and often impinge on mitochondrial function and biogenesis. Critically, aminoglycoside antibiotics (AGAs) have been shown to block the capacity of TRP channels to conduct calcium entry into the cell in response to a wide range of developmental stimuli of a biophysical nature, including mechanical, electromagnetic, thermal, and chemical. Paradoxically, in vitro paradigms commonly used to understand organismal muscle and adipose development may have been led astray by the conventional use of streptomycin, an AGA, to help prevent bacterial contamination. Accordingly, streptomycin has been shown to disrupt both in vitro and in vivo myogenesis, as well as the phenotypic switch of white adipose into beige thermogenic status. In vivo, streptomycin has been shown to disrupt TRP-mediated calcium-dependent exercise adaptations of importance to systemic metabolism. Alternatively, streptomycin has also been used to curb detrimental levels of calcium leakage into dystrophic skeletal muscle through aberrantly gated TRPC1 channels that have been shown to be involved in the etiology of X-linked muscular dystrophies. TRP channels susceptible to AGA antagonism are critically involved in modulating the development of muscle and adipose tissues that, if administered to behaving animals, may translate to systemwide metabolic disruption. Regenerative medicine and clinical communities need to be made aware of this caveat of AGA usage and seek viable alternatives, to prevent contamination or infection in in vitro and in vivo paradigms, respectively. Full article
Show Figures

Graphical abstract

Graphical abstract
Full article ">Figure 1
<p>Schematic of how the aminoglycoside antibiotics (AGAs) interrupt muscle–adipose crosstalk by inhibiting Ca<sup>2+</sup> entry via transient receptor potential (TRP) cation channel classes, causing systemic metabolic disruption. Semicircular arrows indicate the direction of muscle-adipose paracrine crosstalk ranging between anti-inflammatory (green; left) to pro-inflammatory (red; right) character.</p>
Full article ">Figure 2
<p>Streptomycin present during exposure suffices to preclude developmental responses to biophysical stimuli. Streptomycin (100 μg/mL) added to the bathing media 15 min before, but not 15 min after, exposure to light (optical), magnetic fields, or concurrent optical and magnetic field stimulation (COMS), inhibited the typical stimulation of myoblast proliferation. (<b>A</b>) Experimental timeline depicting when streptomycin was administered in the “During” (top/blue) or “After” (below/red) scenarios with reference to electromagnetic exposure (green shaded region). (<b>B</b>) Murine myoblast proliferation as aforementioned, with streptomycin added to the bathing media (100 μg/mL) either 15 min before (blue-shaded region) or after (red-shaded region) exposure to the different exposure paradigms. The shaded areas indicate either the absence of streptomycin (gray) or its application before and during (blue) or after exposure (red) to the indicated conditions. (<b>C</b>) Tabulated fold changes of live cell number for the different exposure conditions relative to sham (no biophysical exposure, no streptomycin). In summary, a criterion for streptomycin antagonism of electromagnetic regenerative responses is that it be present at the time of exposure. AGA: aminoglycoside antibiotic. Data were analysed using one-way ANOVA followed by multiple comparison tests. Significance levels are indicated as follows: ** <span class="html-italic">p</span> &lt; 0.01, and **** <span class="html-italic">p</span> &lt; 0.0001. The error bars represent the standard error of the mean. Adapted from Iversen et al. 2024 [<a href="#B43-cells-13-01273" class="html-bibr">43</a>].</p>
Full article ">Figure 3
<p>TRPC1-dependent transcriptional and enzymatic pathways co-activated by gravity-mediated (mechanical loading) or electromagnetic stimuli impinging on muscle. The orange circles represent Ca<sup>2+</sup> and the red dots represent ROS. TRPC1 activation enhances mitochondrial biogenesis and myogenesis towards the oxidative (slow) phenotype by contributing to the mitohormetic adaptations associated with resistance to oxidative stress. The TRPC1-mediated activation of these same pathways should forestall the development of sarcopenia and be reversed by aminoglycoside antibiotics, such as streptomycin. Abbreviations: ROS: reactive oxygen species; NFAT: Nuclear Factor of Activated T-Cells; PGC-1α: Peroxisome Proliferator-Activated Receptor Gamma Coactivator 1-alpha; Nrf2: Nuclear Factor Erythroid 2-Related Factor 2; MyoD: Myoblast Determination Protein 1. For further details, kindly refer to references [<a href="#B34-cells-13-01273" class="html-bibr">34</a>,<a href="#B96-cells-13-01273" class="html-bibr">96</a>].</p>
Full article ">
18 pages, 6676 KiB  
Article
Selective Assembly of TRPC Channels in the Rat Retina during Photoreceptor Degeneration
by Elena Caminos, Susana López-López and Juan R. Martinez-Galan
Int. J. Mol. Sci. 2024, 25(13), 7251; https://doi.org/10.3390/ijms25137251 - 30 Jun 2024
Viewed by 870
Abstract
Transient receptor potential canonical (TRPC) channels are calcium channels with diverse expression profiles and physiological implications in the retina. Neurons and glial cells of rat retinas with photoreceptor degeneration caused by retinitis pigmentosa (RP) exhibit basal calcium levels that are above those detected [...] Read more.
Transient receptor potential canonical (TRPC) channels are calcium channels with diverse expression profiles and physiological implications in the retina. Neurons and glial cells of rat retinas with photoreceptor degeneration caused by retinitis pigmentosa (RP) exhibit basal calcium levels that are above those detected in healthy retinas. Inner retinal cells are the last to degenerate and are responsible for maintaining the activity of the visual cortex, even after complete loss of photoreceptors. We considered the possibility that TRPC1 and TRPC5 channels might be associated with both the high calcium levels and the delay in inner retinal degeneration. TRPC1 is known to mediate protective effects in neurodegenerative processes while TRPC5 promotes cell death. In order to comprehend the implications of these channels in RP, the co-localization and subsequent physical interaction between TRPC1 and TRPC5 in healthy retina (Sprague-Dawley rats) and degenerating (P23H-1, a model of RP) retina were detected by immunofluorescence and proximity ligation assays. There was an overlapping signal in the innermost retina of all animals where TRPC1 and TRPC5 physically interacted. This interaction increased significantly as photoreceptor loss progressed. Both channels function as TRPC1/5 heteromers in the healthy and damaged retina, with a marked function of TRPC1 in response to retinal degenerative mechanisms. Furthermore, our findings support that TRPC5 channels also function in partnership with STIM1 in Müller and retinal ganglion cells. These results suggest that an increase in TRPC1/5 heteromers may contribute to the slowing of the degeneration of the inner retina during the outer retinal degeneration. Full article
(This article belongs to the Special Issue TRP Channels in Physiology and Pathophysiology 2.0)
Show Figures

Graphical abstract

Graphical abstract
Full article ">Figure 1
<p>This is a figure. Schemes follow the same formatting. Confocal images showing TRPC1 and TRPC5 immunoreactivities in SD and P23H rat retinas. (<b>A</b>–<b>E</b>) Retinal sections with TRPC1 immunoreactivity (magenta) prominently distributed throughout all retinas in all animals. (<b>F</b>–<b>J</b>) Confocal images showing TRPC5 immunoreactivity (green) distributed in the inner retina of all animals. The boxes in the second column contain images of whole retinas taken at the level of the ganglion cell layer. (<b>K</b>–<b>O</b>) Merged images showing TRPC1/TRPC5 co-localization (arrows) distributed throughout the inner retina. (<b>E</b>,<b>J</b>,<b>O</b>) Magnification of the GCL showing a detail of TRPC1 and TRPC5 overlapping in the P23H retina at P30. Nuclei are stained with DAPI (blue). GCL, ganglion cell layer; INL, inner nuclear layer; IPL, inner plexiform layer; NFL, nerve fiber layer. Scale bars: 20 μm.</p>
Full article ">Figure 2
<p>Representative images from retinal sections demonstrating the TRPC1–TRPC5 interaction (red dots) in the inner rat retina by proximity ligation assay. Nuclei are stained with DAPI (blue). The TRPC1–TRPC5 PLA signal was detected on the inner retinal layers of healthy retinas (<b>A</b>,<b>B</b>) and retinas with retinitis pigmentosa (<b>C</b>,<b>D</b>). (<b>E</b>) Representative image from control reaction where a cross-reaction between secondary antibodies did not show red PLA signals. (<b>F</b>) Quantification of the PLA signal, where the Y-axis is the number of red spots per 210 μm of retina. Statistically significant differences with <span class="html-italic">p</span> &lt; 0.01 (**) and <span class="html-italic">p</span> &lt; 0.001 (***), and no statistically significant differences (ns) according to Dunn’s test. GCL, ganglion cell layer; INL, inner nuclear layer; IPL, inner plexiform layer. Scale bar: 20 μm.</p>
Full article ">Figure 3
<p>Relative TRPC5 protein levels in the control rat retinas (SD) and the retinas with photoreceptor degeneration (P23H-1) at different ages. GAPDH is the internal standard. (<b>A</b>) Western Blot showing TRPC5 in rat retinas. (<b>B</b>) Boxplot chart showing a comparison of the relative TRPC5 expression levels between the healthy retinas and P23H rat retinas. No statistically significant differences were observed between the groups, as determined by Kruskal–Wallis’s test and the subsequent Dunn’s test. Values are expressed as the ratio of the optimal density on the TRPC5 bands according to that on the GAPDH bands (TRPC5/GAPDH). Data were obtained from six independent determinations.</p>
Full article ">Figure 4
<p>Identification of astrocytes and Müller cells immunoreactive for TRPC5 and STIM1 in SD and P23H rat retinas. (<b>A</b>–<b>D’</b>) Retinal sections immunostained for STIM1 (green) and GFAP (red). (<b>E</b>–<b>H’</b>) Retinal sections immunostained for TRPC5 (green) and GFAP (red). Overlap is observed in astrocytes and Müller endfeet and vertical processes (arrowheads). (<b>I</b>–<b>K</b>) Wholemounted retinas from aged P23H rats showing co-localization of GFAP (red) with TRPC1 (<b>I</b>), TRPC5 (<b>J</b>), and STIM1 (<b>K</b>). Images were taken at the level of the GCL and NFL, where overlap is observed in Müller cells and astrocytes (yellow). Retinal ganglion cells are immunoreactive for TRPC1, TRPC5, and STIM1 (arrows). Nuclei are labeled with DAPI (blue). IPL, inner plexiform layer. Scale bar: 20 μm.</p>
Full article ">Figure 5
<p>Confocal images showing STIM1 and TRPC1 immunolocalization in retinal sections of the inner retina from SD (<b>A</b>) and P23H (<b>B</b>) rat retinas and in a whole retina from a healthy rat (<b>C</b>). Exceptional co-localization of STIM1/TRPC1 in retinal ganglion cells (arrows and magnification in boxes in (<b>C</b>), asterisk) and STIM1 immunoreactivity around RGCs (arrowheads). Nuclei are stained with DAPI (blue). GCL, ganglion cell layer; IPL, inner plexiform layer; NFL, nerve fiber layer. Scale bar: 20 μm.</p>
Full article ">Figure 6
<p>Confocal images showing TRPC5 and STIM1 immunolocalization in retinal sections of the inner retina from SD (<b>A</b>) and P23H (<b>B</b>) rat retinas and in a whole retina from a healthy rat (<b>C</b>). STIM1 (green) and TRPC5 (red) and the co-localization of STIM1/TRPC5 (yellow) are shown in the merged image. Extensive co-localization is observed in retinal ganglion cells (arrows) and glial cells (arrowheads). Nuclei are stained with DAPI (blue). GCL, ganglion cell layer; IPL, inner plexiform layer; NFL, nerve fiber layer. Scale bar: 20 μm.</p>
Full article ">
13 pages, 3201 KiB  
Article
Synergistic Cellular Responses Conferred by Concurrent Optical and Magnetic Stimulation Are Attenuated by Simultaneous Exposure to Streptomycin: An Antibiotic Dilemma
by Jan Nikolas Iversen, Jürg Fröhlich, Yee Kit Tai and Alfredo Franco-Obregón
Bioengineering 2024, 11(7), 637; https://doi.org/10.3390/bioengineering11070637 - 21 Jun 2024
Cited by 2 | Viewed by 1417
Abstract
Concurrent optical and magnetic stimulation (COMS) combines extremely low-frequency electromagnetic and light exposure for enhanced wound healing. We investigated the potential mechanistic synergism between the magnetic and light components of COMS by comparing their individual and combined cellular responses. Lone magnetic field exposure [...] Read more.
Concurrent optical and magnetic stimulation (COMS) combines extremely low-frequency electromagnetic and light exposure for enhanced wound healing. We investigated the potential mechanistic synergism between the magnetic and light components of COMS by comparing their individual and combined cellular responses. Lone magnetic field exposure produced greater enhancements in cell proliferation than light alone, yet the combined effects of magnetic fields and light were supra-additive of the individual responses. Reactive oxygen species were incrementally reduced by exposure to light, magnetics fields, and their combination, wherein statistical significance was only achieved by the combined COMS modality. By contrast, ATP production was most greatly enhanced by magnetic exposure in combination with light, indicating that mitochondrial respiratory efficiency was improved by the combination of magnetic fields plus light. Protein expression pertaining to cell proliferation was preferentially enhanced by the COMS modality, as were the protein levels of the TRPC1 cation channel that had been previously implicated as part of a calcium–mitochondrial signaling axis invoked by electromagnetic exposure and necessary for proliferation. These results indicate that light facilitates functional synergism with magnetic fields that ultimately impinge on mitochondria-dependent developmental responses. Aminoglycoside antibiotics (AGAs) have been previously shown to inhibit TRPC1-mediated magnetotransduction, whereas their influence over photomodulation has not been explored. Streptomycin applied during exposure to light, magnetic fields, or COMS reduced their respective proliferation enhancements, whereas streptomycin added after the exposure did not. Magnetic field exposure and the COMS modality were capable of partially overcoming the antagonism of proliferation produced by streptomycin treatment, whereas light alone was not. The antagonism of photon-electromagnetic effects by streptomycin implicates TRPC1-mediated calcium entry in both magnetotransduction and photomodulation. Avoiding the prophylactic use of AGAs during COMS therapy will be crucial for maintaining clinical efficacy and is a common concern in most other electromagnetic regenerative paradigms. Full article
Show Figures

Graphical abstract

Graphical abstract
Full article ">Figure 1
<p><b>Incremental myoblast proliferation induction upon exposure to light, magnetic fields, or their combination in the absence of streptomycin.</b> (<b>A</b>) Live cell count of murine C2C12 myoblasts in response to sham (black), light (red), magnetic fields (blue), or combined COMS (hatched blue/red) exposure. (<b>B</b>) Table of fold changes in live cell count relative to the sham condition. The gray shaded area represents the absence of streptomycin at all times during exposure to the indicated conditions. Statistical analyses were performed minimally in three independent biological replicates. Data were analyzed using one-way ANOVA followed by multiple comparison tests. Significance levels are indicated as follows: * <span class="html-italic">p</span> &lt; 0.05, ** <span class="html-italic">p</span> &lt; 0.01, *** <span class="html-italic">p</span> &lt; 0.001, and **** <span class="html-italic">p</span> &lt; 0.0001. Error bars represent the standard error of the mean (SEM).</p>
Full article ">Figure 2
<p><b>The timing of streptomycin administration produced differential effects over myogenic proliferation induction in response to light and magnetic field combinations.</b> (<b>A</b>) Schematic illustration of when streptomycin (100 µg/mL) was added to the bathing media of cells in the “Strep During” (blue shaded) and “Strep After” (red shaded) paradigms prior to cell enumeration at 24 h after the indicated interventions. (<b>B</b>) Live cell count of mouse murine myoblasts in response to the different stimuli as aforementioned with streptomycin antibiotic supplementation (100 µg/mL) added 15 min before (left side, blue shaded box) or after (right side, red shaded box) exposure to the intervention. (<b>C</b>) Table showing the fold change of live cell count over sham condition (without streptomycin administration). The shaded areas indicate the absence of streptomycin (gray) or its application before and during (blue) or after exposure (red) to the indicated conditions. All data collected were from cells of the same plating. Statistical analyses were performed minimally in three independent biological replicates. Data were analyzed using one-way ANOVA followed by multiple comparison tests. Significance levels are indicated as follows: ** <span class="html-italic">p</span> &lt; 0.01, and **** <span class="html-italic">p</span> &lt; 0.0001. The error bars represent the standard error of the mean (SEM).</p>
Full article ">Figure 3
<p><b>Myogenic proliferation associated protein expression in response to exposure to light, magnetic fields, and their combination.</b> (<b>A</b>) Protein expression of (<b>i</b>) cyclin D1, (<b>ii</b>) TRPC1, (<b>iii</b>) phosphorylated ERK, (<b>iv</b>) and cyclin B1 in response to the indicated exposure intervention (n = 3) in the absence of streptomycin. (<b>B</b>) Protein expressions of (<b>i</b>) cyclin D1, (<b>ii</b>) TRPC1, (<b>iii</b>) phosphorylated ERK, (<b>iv</b>) and cyclin B1 either in the presence of streptomycin (100 µg/mL) during (blue shaded box) or after (red shaded box) exposure as indicated (n = 3). The shaded areas indicate the absence of streptomycin (gray) or its application before and during (blue) or after exposure (red) to the indicated conditions. All data shown in panel (<b>B</b>) were collected from cells of the same plating and represent independent cell samples as those used in panel (<b>A</b>). Statistical analyses were performed minimally in three independent biological replicates. Data were analyzed using one-way ANOVA followed by multiple comparison tests. Significance levels are indicated as follows: * <span class="html-italic">p</span> &lt; 0.05 and ** <span class="html-italic">p</span> &lt; 0.01. The error bars represent the standard error of the mean (SEM).</p>
Full article ">Figure 4
<p><b>Dichotomous changes in ROS and ATP production following COMS exposure</b>. (<b>A</b>) Bar chart showing the immediate ROS response of cells (fluorescent intensity averaged during initial 3 min of reading) for the indicated interventions (n = 4). (<b>B</b>) ROS level of cells expressed as fluorescent intensity fold change at 17 min (t<sub>17</sub>) over time 0, (t<sub>0</sub>), (n = 5). (<b>C</b>) The bar chart shows the ATP levels of cells (expressed as fold change over Sham) at t<sub>17</sub> (n = 5, with six technical replicates each). In all cases, cells were exposed for 5 min to the indicated exposure modality at the start of device activation before measurements were commenced. The presented data were generated in the absence of streptomycin. Data were analyzed using one-way ANOVA followed by multiple comparison tests with * <span class="html-italic">p</span> &lt; 0.05 and *** <span class="html-italic">p</span> &lt; 0.001. The error bars represent the standard error of the mean (SEM).</p>
Full article ">
10 pages, 882 KiB  
Article
Genetic Foundation of Male Spur Length and Its Correlation with Female Egg Production in Chickens
by Anqi Chen, Xiaoyu Zhao, Xiurong Zhao, Gang Wang, Xinye Zhang, Xufang Ren, Yalan Zhang, Xue Cheng, Xiaofan Yu, Huie Wang, Menghan Guo, Xiaoyu Jiang, Xiaohan Mei, Guozhen Wei, Xue Wang, Runshen Jiang, Xing Guo, Zhonghua Ning and Lujiang Qu
Animals 2024, 14(12), 1780; https://doi.org/10.3390/ani14121780 - 13 Jun 2024
Viewed by 701
Abstract
Spurs, which mainly appear in roosters, are protrusions near the tarsometatarsus on both sides of the calves of chickens, and are connected to the tarsometatarsus by a bony core. As a male-biased morphological characteristic, the diameter and length of spurs vary significantly between [...] Read more.
Spurs, which mainly appear in roosters, are protrusions near the tarsometatarsus on both sides of the calves of chickens, and are connected to the tarsometatarsus by a bony core. As a male-biased morphological characteristic, the diameter and length of spurs vary significantly between different individuals, mainly related to genetics and age. As a specific behavior of hens, egg-laying also varies greatly between individuals in terms of traits such as age at first egg (AFE), egg weight (EW), and so on. At present, there are few studies on chicken spurs. In this study, we investigated the inheritance pattern of the spur trait in roosters with different phenotypes and the correlations between spur length, body weight at 18 weeks of age (BW18), shank length at 18 weeks of age (SL18), and the egg-laying trait in hens (both hens and roosters were from the same population and were grouped according to their family). These traits related to egg production included AFE, body weight at first egg (BWA), and first egg weight (FEW). We estimated genetic parameters based on pedigree and phenotype data, and used variance analysis to calculate broad-sense heritability for correcting the parameter estimation results. The results showed that the heritability of male left and right spurs ranged from 0.6 to 0.7. There were significant positive correlations between left and right spur length, BW18, SL18, and BWA, as well as between left and right spur length and AFE. We selected 35 males with the longest spurs and 35 males with the shortest spurs in the population, and pooled them into two sets to obtain the pooled genome sequencing data. After genome-wide association and genome divergency analysis by FST, allele frequency differences (AFDs), and XPEHH methods, we identified 7 overlapping genes (CENPE, FAT1, FAM149A, MANBA, NFKB1, SORBS2, UBE2D3) and 14 peak genes (SAMD12, TSPAN5, ENSGALG00000050071, ENSGALG00000053133, ENSGALG00000050348, CNTN5, TRPC6, ENSGALG00000047655,TMSB4X, LIX1, CKB, NEBL, PRTFDC1, MLLT10) related to left and right spur length through genome-wide selection signature analysis and a genome-wide association approach. Our results identified candidate genes associated with chicken spurs, which helps to understand the genetic mechanism of this trait and carry out subsequent research around it. Full article
(This article belongs to the Section Animal Genetics and Genomics)
Show Figures

Figure 1

Figure 1
<p>The location of the spur and method to measure spur length.</p>
Full article ">Figure 2
<p>Manhattan plot and Q–Q plot for spur length by Pool-GWAS (<b>A</b>), FST (<b>B</b>), AFD (<b>C</b>), and XPEHH (<b>D</b>). Each point in the graph corresponds to the SNPs or regions in the dataset. The line in (<b>A</b>) represents the FDR value less than 1 × 10<sup>−16</sup> (<span class="html-italic">p</span> value &lt; 5.42 × 10<sup>−20</sup>), the lines in (<b>B</b>–<b>D</b>) represent the FST value of 0.098, the AFD of 0.346, and the XPEHH value of 1.94. The vertical axis (<span class="html-italic">y</span>-axis) of the Manhattan plot represents −log10 observed <span class="html-italic">p</span>-values of SNPs and the FST, AFD, and XPEHH values of SNPs compared between two groups (long spur, short spur). The horizontal axes (<span class="html-italic">x</span>-axes) all represent the position of these SNPs on the chromosome.</p>
Full article ">
18 pages, 3974 KiB  
Review
The Two Levels of Podocyte Dysfunctions Induced by Apolipoprotein L1 Risk Variants
by Etienne Pays
Kidney Dial. 2024, 4(2), 126-143; https://doi.org/10.3390/kidneydial4020010 - 7 Jun 2024
Viewed by 1130
Abstract
Apolipoprotein L1 (APOL1) nephropathy results from several podocyte dysfunctions involving morphological and motility changes, mitochondrial perturbations, inflammatory stress, and alterations in cation channel activity. I propose that this phenotype results from increased hydrophobicity of the APOL1 risk variants, which induces two distinct types [...] Read more.
Apolipoprotein L1 (APOL1) nephropathy results from several podocyte dysfunctions involving morphological and motility changes, mitochondrial perturbations, inflammatory stress, and alterations in cation channel activity. I propose that this phenotype results from increased hydrophobicity of the APOL1 risk variants, which induces two distinct types of podocyte dysfunctions. On one hand, increased hydrophobic interactions with APOL3 cause intracellular variant isoforms to impair both APOL3 control of Golgi PI(4)P kinase-B (PI4KB) activity and APOL3 control of mitochondrial membrane fusion, triggering actomyosin reorganisation together with mitophagy and apoptosis inhibition (hit 1). On the other hand, increased hydrophobic interactions with the podocyte plasma membrane may cause the extracellular variant isoforms to activate toxic Ca2+ influx and K+ efflux by the TRPC6 and BK channels, respectively (hit 2), presumably due to APOL1-mediated cholesterol clustering in microdomains. I propose that hit 2 depends on low HDL-C/high extracellular APOL1 ratio, such as occurs in cell culture in vitro, or during type I-interferon (IFN-I)-mediated inflammation. Full article
Show Figures

Figure 1

Figure 1
<p>Folding and interactions of APOL1 and APOL3. Models of <span class="html-italic">cis</span>- and <span class="html-italic">trans</span>-interactions of APOL1, APOL1 C-terminal variants and APOL3 (HC = hydrophobic cluster; Hel = helix; LZ = leucine zipper; MAD = membrane-addressing domain; TM = TM span (<span class="html-fig-inline" id="kidneydial-04-00010-i001"><img alt="Kidneydial 04 00010 i001" src="/kidneydial/kidneydial-04-00010/article_deploy/html/images/kidneydial-04-00010-i001.png"/></span>: in APOL1, TM insertion strictly requires acidic conditions; <span class="html-fig-inline" id="kidneydial-04-00010-i002"><img alt="Kidneydial 04 00010 i002" src="/kidneydial/kidneydial-04-00010/article_deploy/html/images/kidneydial-04-00010-i002.png"/></span>: C-terminal G1 and G2 mutations; <span class="html-fig-inline" id="kidneydial-04-00010-i003"><img alt="Kidneydial 04 00010 i003" src="/kidneydial/kidneydial-04-00010/article_deploy/html/images/kidneydial-04-00010-i003.png"/></span>: under calcium-free conditions). For the sake of clarity, each HC-LZ tandem is presented as a linear structure, but these tandems, as well as MAD, are probably folded as double-stranded hairpins. The functions of proteins interacting with APOLs are defined in the text.</p>
Full article ">Figure 2
<p>Hypothetical model of the APOL3 and APOL1 activities [<a href="#B7-kidneydial-04-00010" class="html-bibr">7</a>,<a href="#B17-kidneydial-04-00010" class="html-bibr">17</a>]. At the <span class="html-italic">trans</span>-Golgi, APOL3 is associated with PI4KB and PI4KB controllers ARF1, NCS1 and CALN1 to regulate membrane fission for secretion. APOL3 also weakly interacts with APOL1, which is associated with the NM2A myosin and mitophagy receptor PHB2. IFN-I signalling affects APOL3 interaction with PI4KB following the binding of activated ARF1 (ARF1*) to PI4KB. This allows PI4KB delocalisation in Golgi-derived ATG9A vesicles. Similarly, APOL1 C-terminal variants can induce APOL3 dissociation from PI4KB, partially mimicking APOL3 KO. PI4KB-carrying ATG9A vesicles are targeted to MERCS, and this may involve association of APOL1 with both NM2A and PHB2. The PI4KB-ARF1 complex of ATG9A vesicles triggers mitochondrial fission and mitophagy, whereas APOL3 promotes mitophagosome fusion with endosomes. In either APOL3-KO cells or APOL1 variant-expressing cells, mitochondrial membrane fission and fusion are reduced, due to PI4KB and APOL3 inactivation, respectively [<a href="#B17-kidneydial-04-00010" class="html-bibr">17</a>] (<span class="html-fig-inline" id="kidneydial-04-00010-i002"><img alt="Kidneydial 04 00010 i002" src="/kidneydial/kidneydial-04-00010/article_deploy/html/images/kidneydial-04-00010-i002.png"/></span><span class="html-fig-inline" id="kidneydial-04-00010-i001"><img alt="Kidneydial 04 00010 i001" src="/kidneydial/kidneydial-04-00010/article_deploy/html/images/kidneydial-04-00010-i001.png"/></span>: see <a href="#kidneydial-04-00010-f001" class="html-fig">Figure 1</a>).</p>
Full article ">Figure 3
<p>The APOL1 and APOL3 killing activities. In trypanosomes, APOL1 can insert into endosomal membranes owing to the acidic conditions, and APOL1-induced trypanosome lysis results from apoptotic-like megapore activity in the mitochondrial membrane [<a href="#B8-kidneydial-04-00010" class="html-bibr">8</a>]. In podocytes, APOL1 and APOL3 are crucially involved in inflammation-induced apoptosis, through induction of APOL3 trafficking to the mitochondrial membrane and APOL3-induced megapore activity, respectively [<a href="#B7-kidneydial-04-00010" class="html-bibr">7</a>]. For both APOL1 and APOL3, alternative hypotheses can account for death induction: either APOL oligomerisation in megapores or BH3-mediated APOL interaction with BCL2-like proteins (in green), which would trigger activation of pro-apoptotic BCL2 pores for the release of <span class="html-italic">T. brucei</span> endonuclease G (TbEndoG) or cytochrome C, in trypanosomes and podocytes, respectively. Whereas no evidence supports the eventual APOL oligomerisation in megapores (question marks), the interaction between murine mAPOL7 and anti-apoptotic BCL-XL in dendritic cells [<a href="#B6-kidneydial-04-00010" class="html-bibr">6</a>] suggests that APOLs could inhibit anti-apoptotic activity (double arrows: induction of apoptosis through inhibition of anti-apoptotic activity). Accordingly, trypanosomatids contain transmembrane BCL-XL-interacting proteins exhibiting a BH3-like motif (termed TbBCL2 here) [<a href="#B33-kidneydial-04-00010" class="html-bibr">33</a>].</p>
Full article ">Figure 4
<p>The role of cholesterol in APOL1 trafficking and toxic activity. APOL1 is associated with cholesterol in plasma HDL-C carrier particles or with the <span class="html-italic">Trypanosoma brucei</span> cholesterol carrier TbKIFC1 kinesin [<a href="#B69-kidneydial-04-00010" class="html-bibr">69</a>]. I propose a model of podocyte cation channel activation resulting from cholesterol clustering in surface microdomains by HDL-unbound APOL1. Both Ca<sup>2+</sup> TRPC6 channels and Ca<sup>2+</sup>-dependent big K<sup>+</sup> conductance BK channels are activated by cholesterol-dependent microdomain formation involving podocin and nephrin.</p>
Full article ">Figure 5
<p>The APOL1 membrane-addressing domain (MAD). The MAD domain promotes APOL1 membrane association [<a href="#B20-kidneydial-04-00010" class="html-bibr">20</a>], probably through the binding of positively charged residues (blue) to anionic phospholipids (CL = cardiolipin, PIP = phosphoinositides). This domain is also present in APOL3. The sequence highlighted in red is a putative APOL1-specific cholesterol-binding motif [CRAC, for cholesterol recognition amino acid consensus: (L/V)-X1–5-(Y)-X1–5-(K/R)] [<a href="#B38-kidneydial-04-00010" class="html-bibr">38</a>]. In APOL3, the CRAC motif is disrupted by the residue highlighted in yellow. Conversion of N264 (pink, encircled) to K is strongly correlated with reduced G2 toxicity [<a href="#B94-kidneydial-04-00010" class="html-bibr">94</a>,<a href="#B95-kidneydial-04-00010" class="html-bibr">95</a>]. In silico modelling (<a href="https://zhanggroup.org/I-TASSER/" target="_blank">https://zhanggroup.org/I-TASSER/</a>, accessed on 13 March 2024) predicts a hairpin-like double helix structure, but interactions with membrane lipids in the outer plasma membrane layer could result in deep MAD insertion as shown in the model at the bottom, with N264 facing the CRAC. Thus, the N264K mutation may impair cholesterol recognition, linking APOL1-induced cytotoxicity to cholesterol binding.</p>
Full article ">Figure 6
<p>The HC2 cholesterol-binding site. The APOL1 and APOL3 HC2 regions share putative phospholipid-binding sequences (blue boxes), but only APOL1 contains a putative CRAC motif (red box), which is the second of the two APOL1 CRAC motifs (highlighted in yellow in APOL3 sequence: the CRAC-disrupting residue). Encircled in violet, V353 is the C-terminal amino acid of APOL1Δ, which lacks the ability to induce cytotoxicity at the surface of podocytes [<a href="#B74-kidneydial-04-00010" class="html-bibr">74</a>]. The APOL1 HC2 CRAC is expected to be only accessible in C-terminal APOL1 variants, owing to disruption of the interaction between the two HC-LZ tandems (<b>right panel</b>).</p>
Full article ">
11 pages, 3078 KiB  
Article
Investigating Contributions of Canonical Transient Receptor Potential Channel 3 to Hippocampal Hyperexcitability and Seizure-Induced Neuronal Cell Death
by Kevin D. Phelan, U Thaung Shwe, Hong Wu and Fang Zheng
Int. J. Mol. Sci. 2024, 25(11), 6260; https://doi.org/10.3390/ijms25116260 - 6 Jun 2024
Viewed by 572
Abstract
Canonical transient receptor potential channel 3 (TRPC3) is the most abundant TRPC channel in the brain and is highly expressed in all subfields of the hippocampus. Previous studies have suggested that TRPC3 channels may be involved in the hyperexcitability of hippocampal pyramidal neurons [...] Read more.
Canonical transient receptor potential channel 3 (TRPC3) is the most abundant TRPC channel in the brain and is highly expressed in all subfields of the hippocampus. Previous studies have suggested that TRPC3 channels may be involved in the hyperexcitability of hippocampal pyramidal neurons and seizures. Genetic ablation of TRPC3 channel expression reduced the intensity of pilocarpine-induced status epilepticus (SE). However, the underlying cellular mechanisms remain unexplored and the contribution of TRPC3 channels to SE-induced neurodegeneration is not determined. In this study, we investigated the contribution of TRPC3 channels to the electrophysiological properties of hippocampal pyramidal neurons and hippocampal synaptic plasticity, and the contribution of TRPC3 channels to seizure-induced neuronal cell death. We found that genetic ablation of TRPC3 expression did not alter basic electrophysiological properties of hippocampal pyramidal neurons and had a complex impact on epileptiform bursting in CA3. However, TRPC3 channels contribute significantly to long-term potentiation in CA1 and SE-induced neurodegeneration. Our results provided further support for therapeutic potential of TRPC3 inhibitors and raised new questions that need to be answered by future studies. Full article
Show Figures

Figure 1

Figure 1
<p>Comparison of resting membrane potential (<b>A</b>), cell input resistance (<b>B</b>), and firing threshold (<b>C</b>) of CA1 pyramidal neurons in WT (<span class="html-italic">n</span> = 13) and TRPC3KO mice (<span class="html-italic">n</span> = 10). Note that there was no significant difference between WT and TRPC3KO mice (unpaired <span class="html-italic">t</span>-test).</p>
Full article ">Figure 2
<p>Epileptiform burst firing induced by mGluR agonist in CA1 pyramidal neurons is normal in TRPC3KO mice. (<b>A</b>) Representative current-clamp recordings showing epileptiform burst firing induced by 30 µM 1S,3R-ACPD in CA1 pyramidal neurons in adult WT and TRPC3KO mice. (<b>B</b>) The amplitude of the plateau underlying the burst is comparable in WT and TRPC3KO mice. Amplitudes were measured for three randomly selected bursts in each neuron and then averaged. Pooled data (mean ± SEM) was plotted (<span class="html-italic">n</span> = 5, 4 for WT and TRPC3KO mice). (<b>C</b>) The duration of each burst was quantified by the number of action potentials within each burst and three random bursts from each CA1 pyramidal neuron were analyzed to obtain the average number of spikes per burst. Pooled data (mean ± SEM) were plotted (<span class="html-italic">n</span> = 5, 4 for WT and TRPC3KO mice).</p>
Full article ">Figure 3
<p>Comparison of resting membrane potential (<b>A</b>), cell input resistance (<b>B</b>), and firing threshold (<b>C</b>) of CA3 pyramidal neurons in WT (<span class="html-italic">n</span> = 9) and TRPC3KO mice (<span class="html-italic">n</span> = 8). Note that there was no significant difference between WT and TRPC3KO mice (unpaired <span class="html-italic">t</span>-test).</p>
Full article ">Figure 4
<p>Spontaneous epileptiform burst firing induced by bicuculine in CA3 pyramidal neurons. (<b>A</b>) Representative traces of spontaneous epileptiform bursts occurred after 30 min bath application of Bicuculline in 2 WT and 2 TRPC3KO mice. (<b>B</b>) Comparison of spontaneous burst frequency after washout of bicuculine in WT (<span class="html-italic">n</span> = 4) and TRPC3KO mice (<span class="html-italic">n</span> = 4) (**: <span class="html-italic">p</span> &lt; 0.01, unpaired <span class="html-italic">t</span>-test).</p>
Full article ">Figure 5
<p>Comparison of evoked epileptiform burst discharges in WT and TRPC3KO mice. (<b>A</b>) Representative traces showing evoked epileptiform bursts by mossy fiber (MF) stimulation in CA3 pyramidal neurons after bath application of bicuculline for 30 min. (<b>B</b>–<b>D</b>) Quantitative analysis of evoked epileptiform burst firing by MF stimulations in CA3 pyramidal neurons after bath application of bicuculline for 30 min (<span class="html-italic">n</span> = 5, 4 for WT and TRPC3KO). There was no statistically significant difference between wildtype and TRPC3KO mice (unpaired <span class="html-italic">t</span>-test).</p>
Full article ">Figure 6
<p>Normal paired-pulse facilitation at Schaffer collateral-CA1 synapses in TRPC3KO. (<b>A</b>) Representative traces of paired-pulse facilitation (PPF) of Schaffer collateral field EPSP in WT and TRPC3KO mice. A pair of electric stimuli with increasing intervals (40, 80, 120, 160, 200, 240, 280 and 320 ms) was delivered at 10 s intervals and the resulting pair of field EPSPs was recorded. (<b>B</b>) The averaged PPF ratios (the peak of the second EPSP over the peak of the first EPSP in each pair) and standard errors were plotted (<span class="html-italic">n</span> = 10, 6 for WT and TRPC3KO). Note that the peak of PPF occurs around a 40 ms interval and the subsequent decays at greater intervals. There was no statistically significant difference between wildtype and TRPC3KO mice (Two-way ANOVA).</p>
Full article ">Figure 7
<p>Reduced high-frequency stimuli-induced long-term potentiation at Schaffer collateral-CA1 synapses in TRPC3KO mice. (<b>A</b>) Representative traces of Schaffer collateral field EPSP recorded before and 30 min after high-frequency stimuli (HFS; 100 Hz, 1 s; repeated three times at 20-s intervals) in WT and TRPC3KO mice. Traces shown are the average of 12 consecutive recordings collected at 0.2 Hz. (<b>B</b>) Field EPSP slopes for each minute were determined by averaging 12 consecutive field EPSP recordings in each mouse, and the normalized means and standard errors were plotted (<span class="html-italic">p</span> &lt; 0.01 for genotype effects, Two-way ANOVA; <span class="html-italic">n</span> = 14, 7 for WT and TRPC3KO mice). (<b>C</b>) Average field EPSP slope 30 min after 100 Hz HFS in WT (<span class="html-italic">n</span> = 14) and TRPC3KO mice (<span class="html-italic">n</span> = 7). Note the significantly reduced LTP in TRPC3KO mice (**: <span class="html-italic">p</span> &lt; 0.01, unpaired <span class="html-italic">t</span>-test).</p>
Full article ">Figure 8
<p>Comparison of SE-induced neuronal cell death in WT and TRPC3KO mice. (<b>A</b>,<b>B</b>) Representative images of FJC staining (<b>a</b>–<b>c</b>) and Nissl staining (<b>d</b>–<b>f</b>) of coronal hippocampal sections from WT (<b>A</b>) and TRPC3KO mice (<b>B</b>) (two-day survival; WT: 175 mg/kg pilocarpine, i.p.; TRPC3KO: 222 mg/kg pilocarpine, i.p.). CA1 (<b>b</b>,<b>e</b>) and CA3 (<b>c</b>,<b>f</b>) regions were shown at higher magnification. Scale bar: 0.2 mm for (<b>a</b>,<b>d</b>), 0.1 mm for (<b>b</b>,<b>c</b>,<b>e</b>,<b>f</b>). (<b>C</b>) Comparison of neuronal survival in the hippocampal subfields in WT (<span class="html-italic">n</span> = 6) and TRPC3KO mice (<span class="html-italic">n</span> = 9) using stereology. Note there are increases in neuronal survival in both CA1 and CA3 areas in TRPC3KO mice, but these increases are not statistically significant (<span class="html-italic">p</span> &gt; 0.05, unpaired <span class="html-italic">t</span>-test). The increase in neuronal survival in the hilar region in TRPC3KO mice is statistically significant (*: <span class="html-italic">p</span> &lt; 0.05, unpaired <span class="html-italic">t</span>-test).</p>
Full article ">
15 pages, 2153 KiB  
Article
CYP24A1 and TRPC3 Gene Expression in Kidneys and Their Involvement in Calcium and Phosphate Metabolism in Laying Hens
by Letícia Alves Salmória, Adriana Mércia Guaratini Ibelli, Fernando Castro Tavernari, Jane Oliveira Peixoto, Marcos Antônio Zanella Morés, Débora Ester Petry Marcelino, Karine Daenquele Silva Pinto, Arlei Coldebella, Diego Surek, Vicky Lilge Kawski and Mônica Corrêa Ledur
Animals 2024, 14(10), 1407; https://doi.org/10.3390/ani14101407 - 8 May 2024
Viewed by 774
Abstract
Ca and P homeostasis across the egg-laying cycle is a complex process involving absorption in the small intestine, reabsorption/excretion in the kidneys, and eggshell gland secretion. Diets with inadequate calcium and phosphorus can interfere with their absorption and digestibility, resulting in eggshell quality [...] Read more.
Ca and P homeostasis across the egg-laying cycle is a complex process involving absorption in the small intestine, reabsorption/excretion in the kidneys, and eggshell gland secretion. Diets with inadequate calcium and phosphorus can interfere with their absorption and digestibility, resulting in eggshell quality losses and reduced productive life, affecting egg production and welfare. A better understanding of gene expression profiles in the kidneys of laying hens during the late egg-laying period could clarify the renal role in mineral metabolism at this late stage. Therefore, the performance, egg quality and bone integrity-related traits, and expression profiles of kidney candidate genes were evaluated in 73-week-old laying hens receiving different Ca and P ratios in their diet: a high Ca/P ratio (HR, 22.43), a low ratio (LR, 6.71), and a medium ratio (MR, 11.43). The laying hens receiving the HR diet had improved egg production and eggshell quality traits compared to the other two groups. Humerus length was shorter in the HR than in the other groups. The CYP24A1 and TRPC3 genes were differentially expressed (p.adj ≤ 0.05) among the groups. Therefore, their expression profiles could be involved in calcium and phosphate transcellular transport in 73-week-old laying hens as a way to keep mineral absorption at adequate levels. Full article
(This article belongs to the Special Issue Genetic Analysis of Important Traits in Poultry)
Show Figures

Figure 1

Figure 1
<p>Kidney histopathology stained with hematoxylin and eosin (HE). (<b>A</b>,<b>B</b>) Normal histology, with no histopathological changes in samples from the high Ca/P ratio diet group. (<b>C</b>,<b>D</b>) Samples from the medium Ca/P ratio group with mild lymphocytic interstitial infiltration (1), collecting tubule lumen (2), and lymphocytic infiltrate in the collecting tubule mucosa (3). (<b>E</b>,<b>F</b>) Samples from the low Ca/P ratio group with lymphocytic infiltration in the collecting tubule mucosa (4), collecting tubule lumen (5), collecting tubule epithelial cells (6), and mild interstitial lymphocytic infiltration in the renal cortex (7).</p>
Full article ">Figure 2
<p>Relative expression of candidate genes performed in kidney of 73-week-old laying hens fed diets with different Ca/P ratios. HR: high Ca/P ratio (22.43 Ca/P), LR: low Ca/P ratio (6.71 Ca/P ratio), MR: medium Ca/P ratio (11.43 Ca/P ratio). * <span class="html-italic">p.adj</span> ≤ 0.05. The Figure was prepared using the ggplot2 package from R.</p>
Full article ">
14 pages, 3938 KiB  
Article
The Rise in Tubular pH during Hypercalciuria Exacerbates Calcium Stone Formation
by Farai C. Gombedza, Samuel Shin, Jaclyn Sadiua, George B. Stackhouse and Bidhan C. Bandyopadhyay
Int. J. Mol. Sci. 2024, 25(9), 4787; https://doi.org/10.3390/ijms25094787 - 27 Apr 2024
Viewed by 911
Abstract
In calcium nephrolithiasis (CaNL), most calcium kidney stones are identified as calcium oxalate (CaOx) with variable amounts of calcium phosphate (CaP), where CaP is found as the core component. The nucleation of CaP could be the first step of CaP+CaOx (mixed) stone formation. [...] Read more.
In calcium nephrolithiasis (CaNL), most calcium kidney stones are identified as calcium oxalate (CaOx) with variable amounts of calcium phosphate (CaP), where CaP is found as the core component. The nucleation of CaP could be the first step of CaP+CaOx (mixed) stone formation. High urinary supersaturation of CaP due to hypercalciuria and an elevated urine pH have been described as the two main factors in the nucleation of CaP crystals. Our previous in vivo findings (in mice) show that transient receptor potential canonical type 3 (TRPC3)-mediated Ca2+ entry triggers a transepithelial Ca2+ flux to regulate proximal tubular (PT) luminal [Ca2+], and TRPC3-knockout (KO; -/-) mice exhibited moderate hypercalciuria and microcrystal formation at the loop of Henle (LOH). Therefore, we utilized TRPC3 KO mice and exposed them to both hypercalciuric [2% calcium gluconate (CaG) treatment] and alkalineuric conditions [0.08% acetazolamide (ACZ) treatment] to generate a CaNL phenotype. Our results revealed a significant CaP and mixed crystal formation in those treated KO mice (KOT) compared to their WT counterparts (WTT). Importantly, prolonged exposure to CaG and ACZ resulted in a further increase in crystal size for both treated groups (WTT and KOT), but the KOT mice crystal sizes were markedly larger. Moreover, kidney tissue sections of the KOT mice displayed a greater CaP and mixed microcrystal formation than the kidney sections of the WTT group, specifically in the outer and inner medullary and calyceal region; thus, a higher degree of calcifications and mixed calcium lithiasis in the kidneys of the KOT group was displayed. In our effort to find the Ca2+ signaling pathophysiology of PT cells, we found that PT cells from both treated groups (WTT and KOT) elicited a larger Ca2+ entry compared to the WT counterparts because of significant inhibition by the store-operated Ca2+ entry (SOCE) inhibitor, Pyr6. In the presence of both SOCE (Pyr6) and ROCE (receptor-operated Ca2+ entry) inhibitors (Pyr10), Ca2+ entry by WTT cells was moderately inhibited, suggesting that the Ca2+ and pH levels exerted sensitivity changes in response to ROCE and SOCE. An assessment of the gene expression profiles in the PT cells of WTT and KOT mice revealed a safeguarding effect of TRPC3 against detrimental processes (calcification, fibrosis, inflammation, and apoptosis) in the presence of higher pH and hypercalciuric conditions in mice. Together, these findings show that compromise in both the ROCE and SOCE mechanisms in the absence of TRPC3 under hypercalciuric plus higher tubular pH conditions results in higher CaP and mixed crystal formation and that TRPC3 is protective against those adverse effects. Full article
(This article belongs to the Special Issue Calcium Homeostasis of Cells in Health and Disease 2.0)
Show Figures

Figure 1

Figure 1
<p>CaP and mixed crystal growth analysis in CaG+ACZ-treated WT (WTT) and TRPC3 KO (KOT) mice urine. Alizarin Red (AR); (<b>A</b>) pH 4.3 and (<b>B</b>) 6.8 staining were performed in WTT and KOT mice urine to detect CaP or CaOx crystals, respectively. (<b>C</b>) Time-lapsed measurements of (<b>D</b>) day-to-day WTT and KOT mice urine collection are shown in the depicted images and graph. All graphs in mean + SEM. Urine crystals were collected and measured from <span class="html-italic">n = 8</span> mice. *, <span class="html-italic">p</span> &lt; 0.05; **, <span class="html-italic">p</span> &lt; 0.01.</p>
Full article ">Figure 2
<p>CaG+ACZ-treated WT (WTT) and TRPC3 KO (KOT) kidney section staining to determine fibrosis and calcification. WTT and KOT mice kidney sections were stained with (<b>A</b>) Alizarin Red (AR) pH 4.3, (<b>B</b>) AR pH 6.8, (<b>C</b>) Von Kossa, or (<b>D</b>) Massan’s Trichrome staining. Stained kidney sections were retrieved from <span class="html-italic">n</span> = 8 mice. Bar diagrams in mean + SEM. *, <span class="html-italic">p</span> &lt; 0.05.</p>
Full article ">Figure 3
<p>CaG+ACZ-treated TRPC3 KO mice PT cells (KOT) depicted greater ROCE and SOCE compared to their WT counterparts (WTT). Fura-2 ratiometric (340/380 nm) traces of (<b>A</b>–<b>C</b>) WTT and (<b>D</b>–<b>F</b>) KOT mice PT cells in a Ca<sup>2+</sup>-free cell bath were obtained. During the experiment, neomycin was implemented and the Ca<sup>2+</sup> was adjusted to 2 mM. (<b>B</b>,<b>E</b>) Ca<sup>2+</sup> release and (<b>C</b>,<b>F</b>) Ca<sup>2+</sup> entry components are depicted in bar diagrams as mean + SEM. PT cells were extracted from <span class="html-italic">n</span> = 8 mice. *, <span class="html-italic">p</span> &lt; 0.05; **, <span class="html-italic">p</span> &lt; 0.01.</p>
Full article ">Figure 4
<p>Gene expression profile of CaG + ACZ-treated WT and TRPC3 KO PT cells. Densitometric analysis of (<b>A</b>) Calcification genes (OPN, BMP2, BMP6, RUNX2, VCAM), (<b>B</b>) inflammation (IL-6, IL-1b, NFκβ, NLRP3, MCP1), (<b>C</b>) apoptosis (BAX1, BCL2), and (<b>D</b>) fibrosis (a-SMa, FN-1) genes were performed in CaG+ACZ-treated WT and TRPC3 KO mice PT cells. Gene expression was conducted in extracted PT cells from <span class="html-italic">n</span> = 4 mice. Bar diagrams are depicted in mean + SEM. *, <span class="html-italic">p</span> &lt; 0.05; **, <span class="html-italic">p</span> &lt; 0.01. GAPDH was utilized as an internal control gene.</p>
Full article ">Figure 5
<p>Gene expression profile of CaG + ACZ-treated WT (WTT) and TRPC3 KO (KOT) urine cell debris. (<b>A</b>) Gene expression was measured from cellular debris in mice urine from <span class="html-italic">n = 4</span> mice Megalin and CD13. (<b>B</b>) Densitometric analysis of Megalin and CD13 is depicted in bar diagrams in mean + SEM. **, <span class="html-italic">p</span> &lt; 0.01. GAPDH was utilized as an internal control gene.</p>
Full article ">
12 pages, 1560 KiB  
Article
TRPC3 Is Downregulated in Primary Hyperparathyroidism
by Emilie Kirstein, Dirk Schaudien, Mathias Wagner, Coline M. Diebolt, Alessandro Bozzato, Thomas Tschernig and Colya N. Englisch
Int. J. Mol. Sci. 2024, 25(8), 4392; https://doi.org/10.3390/ijms25084392 - 16 Apr 2024
Cited by 2 | Viewed by 757
Abstract
Transient receptor potential canonical sub-family channel 3 (TRPC3) is considered to play a critical role in calcium homeostasis. However, there are no established findings in this respect with regard to TRPC6. Although the parathyroid gland is a crucial organ in calcium household regulation, [...] Read more.
Transient receptor potential canonical sub-family channel 3 (TRPC3) is considered to play a critical role in calcium homeostasis. However, there are no established findings in this respect with regard to TRPC6. Although the parathyroid gland is a crucial organ in calcium household regulation, little is known about the protein distribution of TRPC channels—especially TRPC3 and TRPC6—in this organ. Our aim was therefore to investigate the protein expression profile of TRPC3 and TRPC6 in healthy and diseased human parathyroid glands. Surgery samples from patients with healthy parathyroid glands and from patients suffering from primary hyperparathyroidism (pHPT) were investigated by immunohistochemistry using knockout-validated antibodies against TRPC3 and TRPC6. A software-based analysis similar to an H-score was performed. For the first time, to our knowledge, TRPC3 and TRPC6 protein expression is described here in the parathyroid glands. It is found in both chief and oxyphilic cells. Furthermore, the TRPC3 staining score in diseased tissue (pHPT) was statistically significantly lower than that in healthy tissue. In conclusion, TRPC3 and TRPC6 proteins are expressed in the human parathyroid gland. Furthermore, there is strong evidence indicating that TRPC3 plays a role in pHPT and subsequently in parathyroid hormone secretion regulation. These findings ultimately require further research in order to not only confirm our results but also to further investigate the relevance of these channels and, in particular, that of TRPC3 in the aforementioned physiological functions and pathophysiological conditions. Full article
(This article belongs to the Section Molecular Endocrinology and Metabolism)
Show Figures

Figure 1

Figure 1
<p>Immunohistochemical staining with an anti-TRPC3 antibody of the human parathyroid gland with and without image analysis. The upper two panels (<b>A</b>,<b>B</b>) illustrate microphotographs from healthy human parathyroid tissue. The lower two panels (<b>C</b>,<b>D</b>) present microphotographs from human parathyroid tissue from a patient with primary hyperparathyroidism (pHPT). Panels (<b>A</b>,<b>C</b>) are without image analysis and panels (<b>B</b>,<b>D</b>) are with image analysis. Panels (<b>B</b>,<b>D</b>) show semiquantitative diaminobenzidine tetrahydrochloride (DAB) color scoring. The color red represents a DAB score of 3, orange a DAB score of 2, and yellow a DAB score of 1. The color green represents subtracted connective and fatty tissues. The color blue shows the subtracted hematoxylin area. Both were attributed as DAB-negative. Panels (<b>A</b>–<b>D</b>) have 40× software magnification and a scale bar of 50 µm.</p>
Full article ">Figure 2
<p>Hematoxylin and eosin staining of human parathyroid tissue. The left column (<b>A</b>,<b>C</b>,<b>E</b>) displays microphotographs from healthy parathyroid tissue and the right column (<b>B</b>,<b>D</b>,<b>F</b>) shows tissue from patients with primary hyperparathyroidism (pHPT). Panel (<b>A</b>) features an overview of healthy parathyroid tissue with a pool of chief cells on the right side and a pool of oxyphilic cells on the left side of the demarcation (scale bar: 100 µm). Panels (<b>C</b>,<b>E</b>) display respectively chief (black arrow) and oxyphilic (white arrow) cells at higher magnification (scale bar: 50 µm). Panel (<b>B</b>) displays an overview of human parathyroid tissue from a patient with pHPT (scale bar: 100 µm). Panels (<b>D</b>,<b>F</b>) display respectively chief (black arrow) and oxyphilic (white arrow) cells at higher magnification (scale bar: 50 µm).</p>
Full article ">Figure 3
<p>Immunohistochemical staining with an anti-TRPC3 antibody of the human parathyroid gland. The left column (<b>A</b>,<b>C</b>,<b>E</b>) illustrates microphotographs from healthy human parathyroid tissue. The right column (<b>B</b>,<b>D</b>,<b>F</b>) presents microphotographs from human parathyroid tissue from patients with primary hyperparathyroidism (pHPT). Panel (<b>A</b>) displays an overview of healthy parathyroid tissue with a pool of chief cells on the right side and a pool of oxyphilic cells on the left side of the demarcation (scale bar: 100 µm). Panel (<b>C</b>) displays both chief (black arrow) and oxyphilic (white arrow) cells at higher magnification (scale bar: 50 µm). Panel (<b>B</b>) displays an overview of human parathyroid tissue from a patient with pHPT (scale bar: 100 µm). Panel (<b>D</b>) displays chief cells (black arrow) at higher magnification (scale bar: 50 µm). Panel (<b>E</b>) displays an overview of the negative control staining of healthy parathyroid tissue. Chief (black arrow), oxyphilic (white arrow), and fat cells (asterisk) are displayed (scale bar: 100 µm). Panel (<b>F</b>) displays an overview of the negative control staining of parathyroid tissue from a patient with pHPT (scale bar: 100 µm).</p>
Full article ">Figure 4
<p>Immunohistochemical staining with an anti-TRPC6 antibody of the human parathyroid gland. The left column (<b>A</b>,<b>C</b>,<b>E</b>) illustrates microphotographs from healthy human parathyroid tissue. The right column (<b>B</b>,<b>D</b>,<b>F</b>) presents microphotographs from human parathyroid tissue from patients with primary hyperparathyroidism (pHPT). Panel (<b>A</b>) displays an overview of healthy parathyroid tissue with a pool of chief cells on the right upper side and a pool of oxyphilic cells on the left lower side of the demarcation (scale bar: 100 µm). Panel (<b>C</b>) displays both chief (black arrow) and oxyphilic (white arrow) cells at higher magnification (scale bar: 50 µm). Panel (<b>B</b>) displays an overview of human parathyroid tissue from a patient with pHPT (scale bar: 100 µm). Panel (<b>D</b>) displays chief cells (black arrow) at higher magnification (scale bar: 50 µm). Panel (<b>E</b>) displays an overview of negative control staining of healthy parathyroid tissue (scale bar: 100 µm). Panel (<b>F</b>) displays an overview of negative control staining of parathyroid tissue from a patient with pHPT (scale bar: 100 µm).</p>
Full article ">Figure 5
<p>Immunohistochemical staining with an anti-CaSR antibody of the human parathyroid gland. The left column (<b>A</b>,<b>C</b>,<b>E</b>) illustrates microphotographs from healthy human parathyroid tissue. The right column (<b>B</b>,<b>D</b>,<b>F</b>) presents microphotographs from human parathyroid tissue from patients with primary hyperparathyroidism (pHPT). Panel (<b>A</b>) displays an overview of healthy parathyroid tissue with a pool of chief cells on the right side and a pool of oxyphilic cells on the left side of the demarcation (scale bar: 100 µm). Panel (<b>C</b>) displays both chief (black arrow) and oxyphilic (white arrow) cells at higher magnification (scale bar: 50 µm). Panel (<b>B</b>) displays an overview of human parathyroid tissue from a patient with pHPT (scale bar: 100 µm). Panel (<b>D</b>) displays chief cells (black arrow) at higher magnification (scale bar: 50 µm). Panel (<b>E</b>) displays an overview of the negative control staining of healthy parathyroid tissue. Chief (black arrow) and oxyphilic (white arrow) cells are displayed (scale bar: 100 µm). Panel (<b>F</b>) displays an overview of the negative control staining of parathyroid tissue from a patient with pHPT (scale bar: 100 µm).</p>
Full article ">
17 pages, 4125 KiB  
Article
Transient Receptor Potential Canonical 5 (TRPC5): Regulation of Heart Rate and Protection against Pathological Cardiac Hypertrophy
by Pratish Thakore, James E. Clark, Aisah A. Aubdool, Dibesh Thapa, Anna Starr, Paul A. Fraser, Keith Farrell-Dillon, Elizabeth S. Fernandes, Ian McFadzean and Susan D. Brain
Biomolecules 2024, 14(4), 442; https://doi.org/10.3390/biom14040442 - 4 Apr 2024
Viewed by 1374
Abstract
TRPC5 is a non-selective cation channel that is expressed in cardiomyocytes, but there is a lack of knowledge of its (patho)physiological role in vivo. Here, we examine the role of TRPC5 on cardiac function under basal conditions and during cardiac hypertrophy. Cardiovascular parameters [...] Read more.
TRPC5 is a non-selective cation channel that is expressed in cardiomyocytes, but there is a lack of knowledge of its (patho)physiological role in vivo. Here, we examine the role of TRPC5 on cardiac function under basal conditions and during cardiac hypertrophy. Cardiovascular parameters were assessed in wild-type (WT) and global TRPC5 knockout (KO) mice. Despite no difference in blood pressure or activity, heart rate was significantly reduced in TRPC5 KO mice. Echocardiography imaging revealed an increase in stroke volume, but cardiac contractility was unaffected. The reduced heart rate persisted in isolated TRPC5 KO hearts, suggesting changes in basal cardiac pacing. Heart rate was further investigated by evaluating the reflex change following drug-induced pressure changes. The reflex bradycardic response following phenylephrine was greater in TRPC5 KO mice but the tachycardic response to SNP was unchanged, indicating an enhancement in the parasympathetic control of the heart rate. Moreover, the reduction in heart rate to carbachol was greater in isolated TRPC5 KO hearts. To evaluate the role of TRPC5 in cardiac pathology, mice were subjected to abdominal aortic banding (AAB). An exaggerated cardiac hypertrophy response to AAB was observed in TRPC5 KO mice, with an increased expression of hypertrophy markers, fibrosis, reactive oxygen species, and angiogenesis. This study provides novel evidence for a direct effect of TRPC5 on cardiac function. We propose that (1) TRPC5 is required for maintaining heart rate by regulating basal cardiac pacing and in response to pressure lowering, and (2) TRPC5 protects against pathological cardiac hypertrophy. Full article
(This article belongs to the Special Issue TRP Channels in Cardiovascular and Inflammatory Disease)
Show Figures

Figure 1

Figure 1
<p>TRPC5 KO mice display a reduced heart rate phenotype. (<b>A</b>) Mean arterial pressure (MAP), (<b>B</b>) heart rate, and (<b>C</b>) activity recordings were obtained from radio-telemetered mice over a 2 light/3 dark period, and shaded regions depict dark phase recordings and bar charts illustrate the average for the light and dark periods (<span class="html-italic">n</span> = 19–20). (<b>D</b>) Stroke volume, (<b>E</b>) cardiac output, (<b>F</b>) ejection fraction, and (<b>G</b>) fractional shortening were assessed via 2-dimensional echocardiography (<span class="html-italic">n</span> = 8–10). Time-domain heart rate variability analysis was performed on the blood pressure pulse waveform and (<b>H</b>) pulse interval, (<b>I</b>) standard deviation (SD) of the pulse interval, (<b>J</b>) square root of the mean of squared successive differences between adjacent intervals (rMSSD), and the percentage of consecutive interval differences (pNN) exceeding (<b>K</b>) 10 msec, (<b>L</b>) 20 msec, and (<b>M</b>) 30 msec were determined and separated into light and dark phases (<span class="html-italic">n</span> = 7). Data shown as mean ± s.e.m. * <span class="html-italic">p</span> &lt; 0.05 as determined by two-ANOVA followed by Bonferroni post-hoc correction or unpaired Student <span class="html-italic">t</span>-test.</p>
Full article ">Figure 2
<p>Enhanced cardiac parasympathetic response in TRPC5 KO mice. Changes in mean arterial pressure (MAP) and heart rate were evaluated following phenylephrine i.v. (<span class="html-italic">n</span> = 6–8). Black shows responses from TRPC5WT mice and red from TRPC5KO mice. Representative traces depicting changes in mean arterial pressure (MAP) and heart rate following (<b>A</b>) 10 µg/kg and (<b>B</b>) 50 µg/kg phenylephrine. Peak changes in (<b>C</b>) MAP and (<b>D</b>) heart rate were assessed, and (<b>E</b>) ratio of heart rate over MAP was used a measure of overall baroreflex sensitivity. (<b>F</b>) Linear regression for the bradycardic responses following phenylephrine administration. (<b>G</b>) Baseline heart rate of isolated hearts on a Langendorff perfusion setup (<span class="html-italic">n</span> = 11). Peak change in heart rate in isolated hearts following perfusion with (<b>H</b>) isoprenaline (<span class="html-italic">n</span> = 5) and (<b>I</b>) carbachol (<span class="html-italic">n</span> = 6). Data shown as mean ± s.e.m. * <span class="html-italic">p</span> &lt; 0.05 as determined by two-ANOVA followed by Bonferroni post-hoc correction.</p>
Full article ">Figure 3
<p>Exacerbated pathological cardiac hypertrophy phenotype in TRPC5 KO mice subjected to pressure overload. WT and TRPC5 KO mice were subjected to either 4 weeks (<span class="html-italic">n</span> = 6–9) or 10 weeks (<span class="html-italic">n</span> = 10–12) of abdominal aortic banding (AAB) or sham. (<b>A</b>) Heart weight and (<b>B</b>) left ventricle (LV) mass were normalised to tibia length (TL). (<b>C</b>) Lung wet/dry ratio was used to determine presence of pulmonary oedema. LV posterior wall (LVPW) dimensions during (<b>D</b>) diastole and (<b>E</b>) systole were assessed via 2-dimensional ultrasound echocardiography. Representative images of left ventricular cardiomyocytes stained with the plasma membrane-binding wheat germ agglutinin (WGA) in transverse cardiac sections from mice subjected to (<b>F</b>) 4 weeks and (<b>G</b>) 10 weeks of pressure overload. Scale bar represents 50 µm (<b>H</b>) Cross-sectional surface area (CSA) was calculated for whole cardiomyocytes using ImageJ software (version 1.48). (<b>I</b>) Ejection fraction and (<b>J</b>) fractional shortening were also determined by echocardiography. Data shown as mean ± s.e.m. * <span class="html-italic">p</span> &lt; 0.05 as determined by two-ANOVA followed by Bonferroni post-hoc correction.</p>
Full article ">Figure 4
<p>Increased expression of hypertrophy and ROS markers in left ventricles of TRPC5 KO mice following pressure overload. (<b>A</b>) <span class="html-italic">Trpc5</span>, (<b>B</b>) <span class="html-italic">Nppa</span> (atrial natriuretic peptide), (<b>C</b>) <span class="html-italic">Mhy7</span> (β-myosin heavy chain), (<b>D</b>) <span class="html-italic">Acta1</span> (skeletal α-actin), and (<b>H</b>) <span class="html-italic">Tgfb1</span> (transforming growth factor-β1) gene expression were assessed via RT-qPCR in cardiac left ventricles from WT and TRPC5 KO mice subjected to 4 weeks (<span class="html-italic">n</span> = 5–7) and 10 weeks (<span class="html-italic">n</span> = 10–12) of abdominal aortic banding (AAB) or sham. Data were normalised to the reference genes <span class="html-italic">B2m</span> (β-2-microglobulin) and <span class="html-italic">Hprt</span> (hypoxanthine guanine phosphoribosyl transferase) and expressed as copies/µL. Representative immunoblots and densitometry analysis of (<b>E</b>) gp91(phox), (<b>F</b>) nitrotyrosine, and (<b>G</b>) heme oxygenase 1 (HO-1) protein expression from cardiac left ventricles. Data shown as mean ± s.e.m. * <span class="html-italic">p</span> &lt; 0.05 as determined by two-ANOVA followed by Bonferroni post-hoc correction. (<b>E</b>–<b>G</b>) Original western blots can be found at <a href="#app1-biomolecules-14-00442" class="html-app">Figures S4–S6</a>.</p>
Full article ">Figure 5
<p>Increased fibrotic remodelling of cardiac left ventricles of TRPC5 mice subjected to pressure overload. (<b>A</b>) <span class="html-italic">Col1a1</span> (collagen type I α1) and (<b>B</b>) <span class="html-italic">Col1a2</span> (collagen type I α2) gene expression were assessed via RT-qPCR in cardiac left ventricles from WT and TRPC5 KO mice subjected to 4 weeks (<span class="html-italic">n</span> = 5–7) and 10 weeks (<span class="html-italic">n</span> = 10–12) of abdominal aortic banding (AAB) or sham. Data were normalised to the reference genes <span class="html-italic">B2m</span> (β-2-microglobulin) and <span class="html-italic">Hprt</span> (hypoxanthine guanine phosphoribosyl transferase) and expressed as copies/µL. Representative images of (<b>C</b>,<b>D</b>) interstitial and (<b>E</b>,<b>F</b>) perivascular fibrosis of (<b>C</b>,<b>E</b>) for 4-week (<span class="html-italic">n</span> = 5–7) and (<b>D</b>,<b>F</b>) 10-week (<span class="html-italic">n</span> = 10–12) transverse cardiac left ventricle sections stained with Picro-Sirius Red, viewed under phase light (left) and circular polarised light (right); bar represents (<b>C</b>,<b>D</b>) 100 µm and (<b>E</b>,<b>F</b>) 50 µm. (<b>G</b>) Quantification of interstitial fibrosis was determined as percentage per field of view. (<b>H</b>) Cross-sectional surface area (CSA) of perivascular fibrosis was calculated and expressed as a ratio of total vessel cross-sectional surface area. All image analysis was performed using ImageJ software (version 1.48). Data shown as mean ± s.e.m. * <span class="html-italic">p</span> &lt; 0.05 as determined by two-ANOVA followed by Bonferroni post-hoc correction.</p>
Full article ">Figure 6
<p>Increased capillary density of cardiac left ventricles of TRPC5 mice subjected to pressure overload. Representative images of capillary density in transverse cardiac left ventricle sections from WT and TRPC5 KO mice subjected to (<b>A</b>) 4 weeks (<span class="html-italic">n</span> = 5–7) and (<b>B</b>) 10 weeks (<span class="html-italic">n</span> = 10–12) of abdominal aortic banding (AAB) or sham. Endothelial cells were stained with Isolectin B<sub>4</sub> (IB<sub>4</sub>), and cardiomyocyte plasma membranes with wheat germ agglutinin (WGA). Colour image overlays were created in Adobe Photoshop CC software; IB<sub>4</sub> green and WGA red with overlay in yellow; bar represents 50 µm. (<b>C</b>) Quantification was determined as total number of capillaries as a ratio to total number of cardiomyocytes per field of view using ImageJ software (version 1.48). Data shown as mean ± s.e.m. * <span class="html-italic">p</span> &lt; 0.05 as determined by two-ANOVA followed by Bonferroni post-hoc correction.</p>
Full article ">
14 pages, 2859 KiB  
Article
The Xanthine Derivative KMUP-1 Inhibits Hypoxia-Induced TRPC1 Expression and Store-Operated Ca2+ Entry in Pulmonary Arterial Smooth Muscle Cells
by Zen-Kong Dai, Yi-Chen Chen, Su-Ling Hsieh, Jwu-Lai Yeh, Jong-Hau Hsu and Bin-Nan Wu
Pharmaceuticals 2024, 17(4), 440; https://doi.org/10.3390/ph17040440 - 29 Mar 2024
Viewed by 1034
Abstract
Exposure to hypoxia results in the development of pulmonary arterial hypertension (PAH). An increase in the intracellular Ca2+ concentration ([Ca2+]i) in pulmonary artery smooth muscle cells (PASMCs) is a major trigger for pulmonary vasoconstriction and proliferation. This study [...] Read more.
Exposure to hypoxia results in the development of pulmonary arterial hypertension (PAH). An increase in the intracellular Ca2+ concentration ([Ca2+]i) in pulmonary artery smooth muscle cells (PASMCs) is a major trigger for pulmonary vasoconstriction and proliferation. This study investigated the mechanism by which KMUP-1, a xanthine derivative with phosphodiesterase inhibitory activity, inhibits hypoxia-induced canonical transient receptor potential channel 1 (TRPC1) protein overexpression and regulates [Ca2+]i through store-operated calcium channels (SOCs). Ex vivo PASMCs were cultured from Sprague-Dawley rats in a modular incubator chamber under 1% O2/5% CO2 for 24 h to elucidate TRPC1 overexpression and observe the Ca2+ release and entry. KMUP-1 (1 μM) inhibited hypoxia-induced TRPC family protein encoded for SOC overexpression, particularly TRPC1. KMUP-1 inhibition of TRPC1 protein was restored by the protein kinase G (PKG) inhibitor KT5823 (1 μM) and the protein kinase A (PKA) inhibitor KT5720 (1 μM). KMUP-1 attenuated protein kinase C (PKC) activator phorbol 12-myristate 13-acetate (PMA, 1 μM)-upregulated TRPC1. We suggest that the effects of KMUP-1 on TRPC1 might involve activating the cyclic guanosine monophosphate (cGMP)/PKG and cyclic adenosine monophosphate (cAMP)/PKA pathways and inhibiting the PKC pathway. We also used Fura 2-acetoxymethyl ester (Fura 2-AM, 5 μM) to measure the stored calcium release from the sarcoplasmic reticulum (SR) and calcium entry through SOCs in hypoxic PASMCs under treatment with thapsigargin (1 μM) and nifedipine (5 μM). In hypoxic conditions, store-operated calcium entry (SOCE) activity was enhanced in PASMCs, and KMUP-1 diminished this activity. In conclusion, KMUP-1 inhibited the expression of TRPC1 protein and the activity of SOC-mediated Ca2+ entry upon SR Ca2+ depletion in hypoxic PASMCs. Full article
(This article belongs to the Section Pharmacology)
Show Figures

Figure 1

Figure 1
<p>Effects of KMUP-1 in normoxic and hypoxic pulmonary arterial smooth muscle cells (PASMCs). (<b>A</b>) Structure of KMUP-1. (<b>B</b>) TRPC1 protein showed no significant differences in KMUP-1 (1, 10, 100 μM)-treated PASMCs under normoxic conditions. (<b>C</b>) Various concentrations of KMUP-1 (1, 10, 100 μM) inhibited hypoxia-induced TRPC1 protein overexpression. The quantitation of this protein is shown in the lower panel. Results were presented as the mean ± SE, n = 7. <sup>#</sup> <span class="html-italic">p</span> &lt; 0.05 compared with control (normoxia), r = 0.66; ** <span class="html-italic">p</span> &lt; 0.01, *** <span class="html-italic">p</span> &lt;0.001 compared with hypoxia, r = 0.81. C: control represents normoxia. r: correlation coefficient.</p>
Full article ">Figure 2
<p>KMUP-1 inhibited hypoxia-enhanced TRPC1 expression via the cGMP/PKG pathway. PASMCs pretreated with KMUP-1 (1 μM), KT5823 (1 μM), 8-Br-cGMP (100 μM), KT5823+KMUP-1, and 8-Br-cGMP+KMUP-1 under hypoxic states. The quantitation of these proteins is shown in the lower panel. Results are presented as the mean ± SE, n = 6. <sup>#</sup> <span class="html-italic">p</span> &lt; 0.05 compared with normoxia, r = 0.47; ** <span class="html-italic">p</span> &lt; 0.01, *** <span class="html-italic">p</span> &lt; 0.001 compared with hypoxia, r = 0.72; <sup>†</sup> <span class="html-italic">p</span> &lt; 0.05 compared with KMUP-1 group, r = 0.56. 8-Br-cGMP: membrane-permeable analog of cGMP. r: correlation coefficient.</p>
Full article ">Figure 3
<p>KMUP-1 inhibited hypoxia-enhanced TRPC1 expression via the cAMP/PKA pathway. PASMCs pretreated with KMUP-1 (1 μM), KT5720 (1 μM), 8-Br-cAMP (100 μM), KT5720+KMUP-1, and 8-Br-cAMP+KMUP-1 under hypoxic states. The quantitation of these proteins is shown in the lower panel. Results are presented as the mean ± SE, n = 6. <sup>#</sup> <span class="html-italic">p</span> &lt; 0.05 compared with normoxia, r = 0.59; ** <span class="html-italic">p</span> &lt; 0.01, *** <span class="html-italic">p</span> &lt; 0.001 compared with hypoxia, r = 0.68; <sup>†</sup> <span class="html-italic">p</span> &lt; 0.05 compared with KMUP-1 group, r = 0.61. 8-Br-cAMP: membrane-permeable analog of cAMP. r: correlation coefficient.</p>
Full article ">Figure 4
<p>KMUP-1 inhibited hypoxia-enhanced TRPC1 expression via PKC pathway. PASMCs pretreated with KMUP-1 (1 μM), PMA (1 μM), chelerythrine (1 μM), PMA+KMUP-1, and chelerythrine+KMUP-1 under hypoxic states. The quantitation of these proteins is shown in the lower panel. Results are presented as the mean ± SE, n = 6. <sup>#</sup> <span class="html-italic">p</span> &lt; 0.05 compared with normoxia, r = 0.48; ** <span class="html-italic">p</span> &lt; 0.01, *** <span class="html-italic">p</span> &lt; 0.001 compared with hypoxia, r = 0.70; <sup>†</sup> <span class="html-italic">p</span> &lt; 0.05 compared with the PMA group, r = 0.52. PMA: phorbol 12-myristate 13-acetate. r: correlation coefficient.</p>
Full article ">Figure 5
<p>Effects of exposure to hypoxia on capacitative calcium entry in PASMCs. (<b>A</b>) Change in [Ca<sup>2+</sup>]<sub>i</sub> in pulmonary arterial smooth muscle cells (PASMCs) from control or hypoxia and subjected to the restoration of extracellular Ca<sup>2+</sup> following store depletion with TG (1 μM). (<b>B</b>) Bar graph illustrates mean ± SE change in [Ca<sup>2+</sup>]<sub>i</sub> (∆[Ca<sup>2+</sup>]<sub>i</sub>) in response to TG and Ca<sup>2+</sup> restoration. All experiments were performed in the presence of nifedipine, n = 12 for control and n = 13 for hypoxia. *** <span class="html-italic">p</span> &lt; 0.001 compared with control (normoxia). TG: thapsigargin.</p>
Full article ">Figure 6
<p>Effects of KMUP-1 on capacitative calcium entry in PASMCs. (<b>A</b>) Change in [Ca<sup>2+</sup>]i in pulmonary arterial smooth muscle cells from hypoxic or pretreated KMUP-1 groups and subjected to the restoration of extracellular Ca<sup>2+</sup> following store depletion with TG (1 μM). (<b>B</b>) Bar graph illustrates mean ± SEM change in [Ca<sup>2+</sup>]<sub>i</sub> (∆[Ca<sup>2+</sup>]<sub>i</sub>) in response to TG and Ca<sup>2+</sup> restoration. All experiments were performed in the presence of nifedipine, n = 7–13 of independent experiments. * <span class="html-italic">p</span> &lt; 0.05, ** <span class="html-italic">p</span> &lt; 0.01, *** <span class="html-italic">p</span> &lt; 0.001 compared with hypoxic group. TG: thapsigargin.</p>
Full article ">Figure 7
<p>Effects of KT5823, KT5720, and PMA on KMUP-1-inhibited capacitative calcium entry in hypoxic PASMCs. (<b>A</b>): Change in [Ca<sup>2+</sup>]<sub>i</sub> in pulmonary arterial smooth muscle cells (PASMCs) from pretreated KMUP-1 (1 μM), KT5823 (1 μM) with KMUP-1, KT5720 (1 μM) with KMUP-1, PMA (1 μM) with KMUP-1 and then subjected to the restoration of extracellular Ca<sup>2+</sup> following store depletion with TG (1 μM). (<b>B</b>): Bar graph illustrates mean ± SE change in [Ca<sup>2+</sup>]<sub>i</sub> (∆[Ca<sup>2+</sup>]<sub>i</sub>) in response to TG and Ca<sup>2+</sup> restoration. All experiments were performed in the presence of nifedipine, n = 6 of independent experiments. <sup>#</sup> <span class="html-italic">p</span> &lt; 0.05 compared with hypoxic group; * <span class="html-italic">p</span> &lt; 0.05 compared with KMUP-1 group in hypoxic PASMCs. PMA: phorbol 12-myristate 13-acetate; TG: thapsigargin.</p>
Full article ">Figure 8
<p>Diagram summarizing the actions of KMUP-1 on the store-operated calcium channel (SOCs) in hypoxic PASMCs. Data suggest that KMUP-1 inhibits TRPC1 encodes for SOCs, activates the PKA/PKG pathway, and inhibits the PKC pathway. SOCs: store-operated calcium channels; SR: sarcoplasmic reticulum; SERCA: sarco/endoplasmic reticulum Ca<sup>2+</sup>-ATPase; TG: thapsigargin; PMA: phorbol 12-myristate 13-acetate; 8-Br-cAMP: membrane-permeable analog of cAMP; 8-Br-cGMP: membrane-permeable analog of cGMP; VOCC: voltage-operated calcium channels.</p>
Full article ">
15 pages, 2249 KiB  
Article
Pharmacological Activation of TRPC6 Channel Prevents Colitis Progression
by Kazuhiro Nishiyama, Yuri Kato, Akiyuki Nishimura, Xinya Mi, Ryu Nagata, Yasuo Mori, Yasu-Taka Azuma and Motohiro Nishida
Int. J. Mol. Sci. 2024, 25(4), 2401; https://doi.org/10.3390/ijms25042401 - 18 Feb 2024
Cited by 1 | Viewed by 1315
Abstract
We recently reported that transient receptor potential canonical (TRPC) 6 channel activity contributes to intracellular Zn2+ homeostasis in the heart. Zn2+ has also been implicated in the regulation of intestinal redox and microbial homeostasis. This study aims to investigate the role [...] Read more.
We recently reported that transient receptor potential canonical (TRPC) 6 channel activity contributes to intracellular Zn2+ homeostasis in the heart. Zn2+ has also been implicated in the regulation of intestinal redox and microbial homeostasis. This study aims to investigate the role of TRPC6-mediated Zn2+ influx in the stress resistance of the intestine. The expression profile of TRPC1-C7 mRNAs in the actively inflamed mucosa from inflammatory bowel disease (IBD) patients was analyzed using the GEO database. Systemic TRPC3 knockout (KO) and TRPC6 KO mice were treated with dextran sulfate sodium (DSS) to induce colitis. The Zn2+ concentration and the mRNA expression levels of oxidative/inflammatory markers in colon tissues were quantitatively analyzed, and gut microbiota profiles were compared. TRPC6 mRNA expression level was increased in IBD patients and DSS-treated mouse colon tissues. DSS-treated TRPC6 KO mice, but not TRPC3 KO mice, showed severe weight loss and increased disease activity index compared with DSS-treated WT mice. The mRNA abundances of antioxidant proteins were basically increased in the TRPC6 KO colon, with changes in gut microbiota profiles. Treatment with TRPC6 activator prevented the DSS-induced colitis progression accompanied by increasing Zn2+ concentration. We suggest that TRPC6-mediated Zn2+ influx activity plays a key role in stress resistance against IBD, providing a new strategy for treating colitis. Full article
(This article belongs to the Special Issue TRP Channels in Physiology and Pathophysiology 2.0)
Show Figures

Figure 1

Figure 1
<p>Increase in TRPC6 mRNA expression level in inflamed mucosa of IBD patients and colon tissue of DSS-treated mice: (<b>A</b>) The expression levels of TRPC1-7 genes were analyzed using a Gene Expression Omnibus (GEO) dataset (GSE83687) containing the expression profile of actively inflamed mucosa from colitis patients. Data are shown as the mean ± SEM (control; n = 60, CD; n = 42, UC; n = 32). <span class="html-italic">p</span> &lt; 0.05, one-way ANOVA followed Tukey’s comparison test. (<b>B</b>) Quantification of TRPC1-7 mRNAs in colon tissue from each group of mice as measured by quantitative PCR and normalized against 18s rRNA (n = 5 mice in each group). Data are shown as the mean ± SEM; <span class="html-italic">p</span> &lt; 0.05, significantly different as indicated; unpaired <span class="html-italic">t</span>-test.</p>
Full article ">Figure 2
<p>TRPC6 deficiency aggravates DSS-induced colitis progression: (<b>A</b>) Body weight changes and (<b>B</b>) disease activity index (DAI) in WT, TRPC3 KO, and TRPC6 KO mice treated with 5% DSS (n = 5 mice in each group). (<b>C</b>) Quantification of IL-6, TNFα, and IL-1β mRNAs in mouse colon tissues normalized by 18s rRNA (n = 5 mice in each group). (<b>D</b>) Zn<sup>2+</sup> concentrations of colon tissues (n = 5 mice in each group). (<b>E</b>,<b>F</b>) Immunohistochemical staining of TRPC6 in the colon of mice. (<b>E</b>) Wide field of view. Scale bars; 100 μm. (<b>F</b>) Enlarged view. Scale bars; 50 μm. Data are shown as the mean ± SEM; <span class="html-italic">p</span> &lt; 0.05, significance was determined using two-way ANOVA followed by Tukey’s comparison test (<b>A</b>,<b>B</b>) and unpaired <span class="html-italic">t</span>-test (<b>C</b>,<b>D</b>).</p>
Full article ">Figure 3
<p>TRPC6 deficiency increased antioxidant proteins. Quantification of mRNA expression levels of antioxidants ((<b>A</b>) Nrf2; (<b>B</b>) SOD1; (<b>C</b>) Prdx3; (<b>D</b>) SRXN1) in mouse colon tissues (control WT: n = 5; control TRPC6 KO: n = 5; DSS WT: n = 5; DSS TRPC6 KO: n = 5). Data are shown as the mean ± SEM; <span class="html-italic">p</span> &lt; 0.05, significance was determined using two-way ANOVA followed by Sidak’s comparison test.</p>
Full article ">Figure 4
<p>TRPC6 deficiency changes in the gut microbiota profile: (<b>A</b>) Taxonomic distribution at the phylum level, showing individual samples. Taxonomic changes in the intestinal microbiota. (<b>B</b>) Relative abundance of S24-7 is presented at the family level. (<b>C</b>) The relative abundance of Anaerotruncus is presented at the genus level. (<b>D</b>) The relative abundance of Parabacteroides distasonis is presented at the species level (n = 5 mice in each group). Taxa with LDA scores &gt; 2.0 and <span class="html-italic">p</span> &lt; 0.05, determined using the Wilcoxon signed-rank test, are shown.</p>
Full article ">Figure 5
<p>Treatment of mice with PPZ2 attenuates DSS-induced colitis. C57BL/6J mice were administered with DSS. Osmotic pumps including PPZ2 (2.5 mg/kg/day), or vehicle were implanted intraperitoneally 5 days before DSS administration: (<b>A</b>) Body weight changes, (<b>B</b>) DAI, and (<b>C</b>) colon length in mice treated with 3% DSS (n = 5 mice in each group). (<b>D</b>) Quantification of IL-6 mRNA in colon tissues (n = 5). (<b>E</b>) Zn<sup>2+</sup> concentrations of colon tissues (n = 5 mice in each group). Data are shown as the mean ± SEM; <span class="html-italic">p</span> &lt;0.05, significance was imparted using two-way ANOVA followed by Tukey’s comparison test (<b>A</b>,<b>B</b>) and unpaired <span class="html-italic">t</span>-test (<b>C</b>–<b>E</b>).</p>
Full article ">Figure 6
<p>Schema of the role of TRPC6-mediated Zn<sup>2+</sup> influx in the intestine. TRPC6-mediated Zn<sup>2+</sup> influx plays a key role in stress resistance through the maintenance of redox homeostasis and gut microbiota.</p>
Full article ">
20 pages, 3850 KiB  
Article
TRPV4 Channels Promote Pathological, but Not Physiological, Cardiac Remodeling through the Activation of Calcineurin/NFAT and TRPC6
by Laia Yáñez-Bisbe, Mar Moya, Antonio Rodríguez-Sinovas, Marisol Ruiz-Meana, Javier Inserte, Marta Tajes, Montserrat Batlle, Eduard Guasch, Aleksandra Mas-Stachurska, Elisabet Miró, Nuria Rivas, Ignacio Ferreira González, Anna Garcia-Elias and Begoña Benito
Int. J. Mol. Sci. 2024, 25(3), 1541; https://doi.org/10.3390/ijms25031541 - 26 Jan 2024
Viewed by 1425
Abstract
TRPV4 channels, which respond to mechanical activation by permeating Ca2+ into the cell, may play a pivotal role in cardiac remodeling during cardiac overload. Our study aimed to investigate TRPV4 involvement in pathological and physiological remodeling through Ca2+-dependent signaling. TRPV4 [...] Read more.
TRPV4 channels, which respond to mechanical activation by permeating Ca2+ into the cell, may play a pivotal role in cardiac remodeling during cardiac overload. Our study aimed to investigate TRPV4 involvement in pathological and physiological remodeling through Ca2+-dependent signaling. TRPV4 expression was assessed in heart failure (HF) models, induced by isoproterenol infusion or transverse aortic constriction, and in exercise-induced adaptive remodeling models. The impact of genetic TRPV4 inhibition on HF was studied by echocardiography, histology, gene and protein analysis, arrhythmia inducibility, Ca2+ dynamics, calcineurin (CN) activity, and NFAT nuclear translocation. TRPV4 expression exclusively increased in HF models, strongly correlating with fibrosis. Isoproterenol-administered transgenic TRPV4−/− mice did not exhibit HF features. Cardiac fibroblasts (CFb) from TRPV4+/+ animals, compared to TRPV4−/−, displayed significant TRPV4 overexpression, elevated Ca2+ influx, and enhanced CN/NFATc3 pathway activation. TRPC6 expression paralleled that of TRPV4 in all models, with no increase in TRPV4−/− mice. In cultured CFb, the activation of TRPV4 by GSK1016790A increased TRPC6 expression, which led to enhanced CN/NFATc3 activation through synergistic action of both channels. In conclusion, TRPV4 channels contribute to pathological remodeling by promoting fibrosis and inducing TRPC6 upregulation through the activation of Ca2+-dependent CN/NFATc3 signaling. These results pose TRPV4 as a primary mediator of the pathological response. Full article
(This article belongs to the Section Molecular Biology)
Show Figures

Figure 1

Figure 1
<p>Differential features between physiological (green) and pathological (red) remodeling. From (<b>A</b>–<b>G</b>), results of the four experimental groups generated in mice (N = 6–12/group). (<b>A</b>) Echocardiographic parameters of interventricular septum thickness in diastole (IVSd), left ventricular diameter in diastole (LVDd), and ejection fraction (EF). (<b>B</b>) Heart weight-to-tibial length (HW/TL) ratio. On the right, representative images visually depict the size of the heart. (<b>C</b>) Overall quantification of cardiomyocyte cross-sectional area (CSA) with representative photomicrographs of the four study groups stained with hematoxylin and eosin (H&amp;E). Scale bar represents 50 µm. (<b>D</b>) Percentage of fibrosis measured by collagen deposition with representative images of the four study groups stained with picrosirius red. Scale bar corresponds to 200 µm. (<b>E</b>) Gene expression of fibrotic markers in the four study groups. (<b>F</b>) Gene and protein expression of TRPV4 in the four study groups. (<b>G</b>) Gene expression of TRPV4 in the four experimental groups studied in rats (see the Methods section, N = 12/group). Ex: exercise; Sed: sedentary; HF(iso): HF induced by isoproterenol infusion; HF(TAC): HF induced by transverse aortic constriction. * <span class="html-italic">p</span> &lt; 0.05; ** <span class="html-italic">p</span> &lt; 0.01; *** <span class="html-italic">p</span> &lt; 0.001; and **** <span class="html-italic">p</span> &lt; 0.0001.</p>
Full article ">Figure 2
<p>Time-course changes during the development of pathological remodeling induced by isoproterenol infusion (N = 5–18/group). (<b>A</b>) Echocardiographic parameters of interventricular septum thickness in diastole (IVSd), left ventricular diameter in diastole (LVDd), and ejection fraction (EF). (<b>B</b>) Heart weight-to-tibial length ratio (HW/TL, left) and cardiomyocyte cross-sectional area (CSA, right). (<b>C</b>) Fibrosis quantification by percentage of collagen deposition. (<b>D</b>) Representative microphotographs stained with picrosirius red at each timepoint. Scale bar corresponds to 200 µm. (<b>E</b>) Protein expression of TRPV4 channels over time. (<b>F</b>) Correlation of TRPV4 expression with cardiomyocyte CSA (left) and collagen deposition (right) at all timepoints. Iso 3d, 7d, 14d, and 28d represent mice subjected to isoproterenol infusion for 3, 7, 14 and 28 days, respectively. * <span class="html-italic">p</span> &lt; 0.05; ** <span class="html-italic">p</span> &lt; 0.01; *** <span class="html-italic">p</span> &lt; 0.001; and **** <span class="html-italic">p</span> &lt; 0.0001.</p>
Full article ">Figure 3
<p>Induction of pathological remodeling in TRPV4+/+ and TRPV4−/− mice (N = 6–12/group). (<b>A</b>) Gene and protein expression of TRPV4 channels in the four study groups. (<b>B</b>) Echocardiographic parameters of interventricular septum thickness in diastole (IVSd), left ventricular diameter in diastole (LVDd), and ejection fraction (EF). (<b>C</b>) Heart weight-to-tibial length (HW/TL) ratio. (<b>D</b>) Overall quantification of cardiomyocyte cross-sectional area (CSA), with representative hematoxylin and eosin (H&amp;E) stained images of all experimental groups. Scale bar represents 50 µm. (<b>E</b>) Fibrosis quantification by percentage of collagen deposition with representative microphotographs stained with picrosirius red. Scale bar corresponds to 200 µm. (<b>F</b>) Arrhythmia inducibility under normoxia and ischemia. The number of total ventricular tachyarrhythmias (VTA) are shown for each condition. HF(iso): HF induced by isoproterenol infusion; nd: not detected; * <span class="html-italic">p</span> &lt; 0.05; ** <span class="html-italic">p</span> &lt; 0.01; and **** <span class="html-italic">p</span> &lt; 0.0001.</p>
Full article ">Figure 4
<p>Functional dynamics of Ca<sup>2+</sup> in response to specific TRPV4 activation and/or inhibition (N = 4–8 replicates per group from 3–4 independent experiments). (<b>A</b>) Relative expression of TRPV4 in cardiomyocytes (CM) and fibroblasts (FB). (<b>B</b>) Calcium influx recorded in FB from TRPV4+/+ and TRPV4−/− mice (groups HF(iso) and sham) in response to the selective TRPV4 activator GSK1016790A (GSK10, 100 nM) in the absence (left) or the presence (right) of the TRPV4 inhibitor HC067047 (HC, 10 µM). (<b>C</b>) Calcium influx in response to a hypotonic solution (140 mOsm) in the absence (left) or the presence (right) of the TRPV4 inhibitor HC067047 (HC, 10 µM). F/F0 = ratio of fluorescence intensity relative to time 0; HF(iso): HF induced by isoproterenol infusion. In (<b>B</b>,<b>C</b>), all Ca<sup>2+</sup> measures are expressed as the change with respect to baseline values, which have been normalized to 1. * <span class="html-italic">p</span> &lt; 0.05.</p>
Full article ">Figure 5
<p>(<b>A</b>) Calcineurin (CN) activity and (<b>B</b>) protein expression in isolated FB from TRPV4+/+ and TRPV4−/− mice of the HF(iso) and sham groups (N = 4–5/group). (<b>C</b>) NFAT cytoplasmatic and nuclear fluorescence quantification expressed as the MGV N/C, the ratio of mean fluorescence intensity (mean gray value) between the nucleus (N) and the cytoplasm (C), n = 202–244. (<b>D</b>) Representative confocal microscopy images of all experimental groups. From top to bottom, each row represents NFAT staining (red), nuclear staining (blue), and merging. The negative control was not incubated with the primary antibody. The positive control was obtained by activation of CN following incubation with a high-calcium medium (4 mM). Scale bar corresponds to 150 µm. (<b>E</b>) Gene expression of col1a1 and acta2 in the four study groups. (<b>F</b>) NFAT nuclear translocation in isolated FB from TRPV4+/+ mice (WT), expressed as the MGV N/C, after specific TRPV4 stimulation with GSK101679A (GSK10 (100 nM), purple), and GSK + pre-incubation with the TRPV4 inhibitor HC067047 (HC (10 µM), gray) in isolated FB. * <span class="html-italic">p</span> &lt; 0.05; ** <span class="html-italic">p</span> &lt; 0.01; and **** <span class="html-italic">p</span> &lt; 0.0001.</p>
Full article ">Figure 6
<p>(<b>A</b>) TRPC6 protein expression in the HF and Ex models induced in mice (n = 5/group). (<b>B</b>) Gene expression of TRPC6 in the four experimental groups studied in rats (N = 12/group). (<b>C</b>) Time-course protein expression of TRPC6 channels during the development of pathological remodeling (N = 5–11/group). (<b>D</b>) TRPC6 protein expression in TRPV4+/+ and TRPV4−/− mice receiving isoproterenol (HF(iso)) or saline (sham) (N = 6–8/group). (<b>E</b>) Effects of Ca<sup>2+</sup> overload and calcineurin inhibition with cyclosporine A (CsA) on NFATc3 nuclear translocation (left) and TRPC6 gene expression (right) in FB from TRPV4+/+ and TRPV4−/− mice from the HF(iso) and sham groups. (<b>F</b>) Proximity ligation assay (PLA) for heteromeric channel formation of TRPV4-TRPC1 and TRPV4-TRPC6. PLA was performed with a combination of anti-TRPV4, anti-TRPC1, and anti-TRPC6 antibodies conjugated to PLA PLUS or MINUS probes. The negative control was only incubated with PLA PLUS and MINUS probes. (<b>G</b>) In TRPV4+/+ mice (WT), TRPC6 (right) and TRPV4 (left) expression following exposure to the specific TRPV4 activator GSK10 (100 nM) without and with pre-incubation with CsA (1 µM), a CN inhibitor (N = 5–6/group). (<b>H</b>) NFAT nuclear translocation as a surrogate of activation of the CN/NFAT pathway in cardiac FB from TRPV4+/+ (WT) animals; experimental conditions were as follows: TRPV4 activation by GSK1016790A (GSK10, 100 nM), TRPC6 activation by GSK1702934A (GSK17, 1 µM), simultaneous activation of TRPV4 and TRPC6 (GSK10, 100 nM + GSK17, 1 µM), TRPV4 activation with previous inhibition of TRPC6 (GSK10, 100 nM + BI-749327, 1 µM), TRPC6 activation with previous inhibition of TRPV4 (GSK17, 1 µM + HC067047, 10 µM), and control. All conditions were significantly different (<span class="html-italic">p</span> &lt; 0.0001) compared to controls. Ex: exercise; Sed: sedentary; HF(iso): HF induced by isoproterenol infusion; HF(TAC): HF induced by transaortic constriction; * <span class="html-italic">p</span> &lt; 0.05; ** <span class="html-italic">p</span> &lt; 0.01; *** <span class="html-italic">p</span> &lt; 0.001; and **** <span class="html-italic">p</span> &lt; 0.0001.</p>
Full article ">
25 pages, 1472 KiB  
Review
Cytoskeleton Rearrangement in Podocytopathies: An Update
by Sijia Ma, Yang Qiu and Chun Zhang
Int. J. Mol. Sci. 2024, 25(1), 647; https://doi.org/10.3390/ijms25010647 - 4 Jan 2024
Cited by 3 | Viewed by 2333
Abstract
Podocyte injury can disrupt the glomerular filtration barrier (GFB), leading to podocytopathies that emphasize podocytes as the glomerulus’s key organizer. The coordinated cytoskeleton is essential for supporting the elegant structure and complete functions of podocytes. Therefore, cytoskeleton rearrangement is closely related to the [...] Read more.
Podocyte injury can disrupt the glomerular filtration barrier (GFB), leading to podocytopathies that emphasize podocytes as the glomerulus’s key organizer. The coordinated cytoskeleton is essential for supporting the elegant structure and complete functions of podocytes. Therefore, cytoskeleton rearrangement is closely related to the pathogenesis of podocytopathies. In podocytopathies, the rearrangement of the cytoskeleton refers to significant alterations in a string of slit diaphragm (SD) and focal adhesion proteins such as the signaling node nephrin, calcium influx via transient receptor potential channel 6 (TRPC6), and regulation of the Rho family, eventually leading to the disorganization of the original cytoskeletal architecture. Thus, it is imperative to focus on these proteins and signaling pathways to probe the cytoskeleton rearrangement in podocytopathies. In this review, we describe podocytopathies and the podocyte cytoskeleton, then discuss the molecular mechanisms involved in cytoskeleton rearrangement in podocytopathies and summarize the effects of currently existing drugs on regulating the podocyte cytoskeleton. Full article
(This article belongs to the Special Issue Molecular Advances in Glomerular Diseases)
Show Figures

Figure 1

Figure 1
<p>The structure of podocytes and the network of podocyte cytoskeletons. Podocytes consist of cell bodies, primary processes, and secondary foot processes; adjacent foot processes form the slit diaphragm. The cell bodies and primary processes are based on microtubules and intermediate filaments, while the foot processes mainly depend on the actin filaments. Abbreviations: FPs: foot processes; SD: slit diaphragm; GBM: glomerular basement membrane; MTs: microtubules; IFs: intermediate filaments; AFs: actin filaments.</p>
Full article ">Figure 2
<p>Modulation of the actin cytoskeleton within the foot processes. At the SD, nephrin acts as a signaling hub to interact with other proteins. Nephrin interacts with CD2AP with a subunit of PI3K and then activates the Akt kinase pathway, leading to actin cytoskeleton rearrangement. The adaptor protein Nck binds to phosphorylated nephrin and N-WASP, stimulating the Arp2/3 complex, which triggers actin polymerization. The scaffolding protein IQGAP1 interacts with nephrin and regulates actin dynamin. In DKD, insulin modulates actin cytoskeleton rearrangement by activating TRPC6, affecting AMPKα and PKGIα, and then regulating RhoA and Rac1. These interactions can regulate the PAK/cofilin-dependent signaling pathway. Dynamin functions in endocytosis and regulates the scission and release of vesicles. Rabs are localized in close proximity to vesicles within FPs and serve as predominant mediators of vesicle trafficking while closely interacting with cytoskeleton proteins. Abbreviations: FPs: foot processes; SD: slit diaphragm; GBM: glomerular basement membrane; AFs: actin filaments; TRPC6: Transient receptor potential channel 6; CD2AP: CD2-associated protein; CAPZ: F-actin capping protein; PAK: p21-activated kinases; Nck: Non-catalytic region of tyrosine kinase; N-WASP: neuronal Wiskott–Aldrich syndrome protein; Arp2/3: actin-related proteins 2/3; ACTN4: α-actinin-4; NEPH1: nephrin-like protein 1; NHEFR 2: Na<sup>+</sup>/H<sup>+</sup> exchange regulatory cofactor 2; IR: insulin receptor; AMPKα: AMP-activated protein kinase α; PKGIα: protein kinase G type Iα; ROCK: Rho-associated protein kinase; IQGAP1: IQ domain GTPase-activating protein 1; PI3K: phosphoinositide 3-kinase; PKB: protein kinase B; Ca<sup>2+</sup>: calcium; F-actin: filamentous actin.</p>
Full article ">
18 pages, 3158 KiB  
Article
Disruption of Atrial Rhythmicity by the Air Pollutant 1,2-Naphthoquinone: Role of Beta-Adrenergic and Sensory Receptors
by Antonio G. Soares, Simone A. Teixeira, Pratish Thakore, Larissa G. Santos, Walter dos R. P. Filho, Vagner R. Antunes, Marcelo N. Muscará, Susan D. Brain and Soraia K. P. Costa
Biomolecules 2024, 14(1), 57; https://doi.org/10.3390/biom14010057 - 31 Dec 2023
Viewed by 1647
Abstract
The combustion of fossil fuels contributes to air pollution (AP), which was linked to about 8.79 million global deaths in 2018, mainly due to respiratory and cardiovascular-related effects. Among these, particulate air pollution (PM2.5) stands out as a major risk factor for heart [...] Read more.
The combustion of fossil fuels contributes to air pollution (AP), which was linked to about 8.79 million global deaths in 2018, mainly due to respiratory and cardiovascular-related effects. Among these, particulate air pollution (PM2.5) stands out as a major risk factor for heart health, especially during vulnerable phases. Our prior study showed that premature exposure to 1,2-naphthoquinone (1,2-NQ), a chemical found in diesel exhaust particles (DEP), exacerbated asthma in adulthood. Moreover, increased concentration of 1,2-NQ contributed to airway inflammation triggered by PM2.5, employing neurogenic pathways related to the up-regulation of transient receptor potential vanilloid 1 (TRPV1). However, the potential impact of early-life exposure to 1,2-naphthoquinone (1,2-NQ) on atrial fibrillation (AF) has not yet been investigated. This study aims to investigate how inhaling 1,2-NQ in early life affects the autonomic adrenergic system and the role played by TRPV1 in these heart disturbances. C57Bl/6 neonate male mice were exposed to 1,2-NQ (100 nM) or its vehicle at 6, 8, and 10 days of life. Early exposure to 1,2-NQ impairs adrenergic responses in the right atria without markedly affecting cholinergic responses. ECG analysis revealed altered rhythmicity in young mice, suggesting increased sympathetic nervous system activity. Furthermore, 1,2-NQ affected β1-adrenergic receptor agonist-mediated positive chronotropism, which was prevented by metoprolol, a β1 receptor blocker. Capsazepine, a TRPV1 blocker but not a TRPC5 blocker, reversed 1,2-NQ-induced cardiac changes. In conclusion, neonate mice exposure to AP 1,2-NQ results in an elevated risk of developing cardiac adrenergic dysfunction, potentially leading to atrial arrhythmia at a young age. Full article
(This article belongs to the Section Cellular Biochemistry)
Show Figures

Figure 1

Figure 1
<p>Schematic overview of experimental design for premature exposure to pollutant 1,2-NQ and cardiovascular assessments.</p>
Full article ">Figure 2
<p>Impact of 1,2–NQ on isolated murine right atria. (<b>A</b>) The concentration–response curve of 1,2–NQ induces positive chronotropism in isolated murine right atria. (<b>B</b>) The atrial force contraction triggered by ouabain (OUA, white column) demonstrates a significant increase in the presence of 1,2–NQ (100 nM; black column). (<b>C</b>) 1,2–NQ does not alter ouabain-induced atrial frequency (BPM). ** <span class="html-italic">p</span> &lt; 0.05 vs. ouabain. BPM (beats per minute), 1,2–NQ - 1,2–naphthoquinone.</p>
Full article ">Figure 3
<p>Impact of 1,2-NQ on mean arterial blood pressure. Exposure to 1,2-NQ during the neonatal stage (6, 8, and 10th of life) does not affect mean arterial blood pressure compared to the vehicle group (white column). Data are expressed as mean ± s.e.m. for <span class="html-italic">n</span> = 5 in each group.</p>
Full article ">Figure 4
<p>Impaired atrial rhythmicity in young mice previously exposed to 1,2–NQ as neonates. In panel (<b>A</b>), neonate exposure to 1,2–NQ (red squares) results in impaired atrial rhythmicity evoked by noradrenaline compared to the control vehicle group (black circles). In panel (<b>B</b>), the negative chronotropic response of carbachol remains unaffected by premature exposure of 1,2–NQ (red squares) compared to the control vehicle (black circles). Data expressed as mean ± s.e.m. for <span class="html-italic">n</span> = 5 in each group. Heart rate changes (ΔBPM) are presented as beats per minute evoked by noradrenaline, while %BPM represents beats per minute evoked by carbachol. 1,2–NQ - 1,2–naphthoquinone.</p>
Full article ">Figure 5
<p>Impaired atrial rhythmicity with 1,2–NQ incubation in isolated right atria. In comparison to the vehicle (black circles), incubation with 1,2–NQ (100 nM; red squares) adversely affects atrial rhythmicity in response to noradrenaline (<b>A</b>), isoproterenol (<b>B</b>), and salbutamol (<b>C</b>) concentration–response curves. The data are expressed as mean ± s.e.m. for <span class="html-italic">n</span> = 5 in each group. BPM (beats per minute) is expressed as the rate evoked by the agonist. 1,2–NQ — 1,2–naphthoquinone.</p>
Full article ">Figure 6
<p>Impact of 1,2–NQ on forskolin-induced positive chronotropism and effects of metoprolol on 1,2-NQ-induced increased BPM in isolated right atria. Panels (<b>A</b>) illustrate that the positive chronotropic effect induced by forskolin remains unaffected by 1,2–NQ (100 nM; red squares) compared to the vehicle (black circles). Conversely, panel (<b>B</b>) shows that metoprolol incubation (400 nM; red squares) completely inhibits the positive chronotropic response evoked by 1,2-NQ when compared to vehicle control (black circles). Data are expressed as mean ± s.e.m. for <span class="html-italic">n</span> = 5 in each group. BPM (beats per minute). 1,2–NQ—1,2–naphthoquinone.</p>
Full article ">Figure 7
<p>Reversal of atrial rhythmicity impairment in young mice previously exposed to 1,2–NQ by capsazepine, a TRPV1 antagonist. Concentration–response curves for noradrenaline (<b>A</b>) and carbachol (<b>B</b>) were conducted on isolated right atria from the following groups: Vehicle (black circles), 1,2–NQ (red squares), capsazepine (CZP, green diamonds), and 1,2–NQ + CZP (purple squares). The data is expressed as mean ± s.e.m. Heart rate (BPM) is presented as beats per minute evoked by noradrenaline, and %BPM is presented as beats per minute evoked by carbachol. 1,2–NQ – 1,2–naphthoquinone.</p>
Full article ">Figure 8
<p>Blockade of TRPC5 on 1,2–NQ-induced atrial rhythmicity impairment. Concentration–response curves to noradrenaline (<b>A</b>) and carbachol (<b>B</b>) were performed on isolated right atria from the following groups treated as a neonate with vehicle (black circles), 1,2–NQ (red squares), TRPC5 antagonist (ML204, brown triangles), and 1,2–NQ plus ML204 (blue triangles). Data are expressed as mean ± s.e.m. for <span class="html-italic">n</span> = 5 in each group. Beats per minute (BPM) are presented as responses to noradrenaline, while %BPM represents responses to carbachol. 1,2–NQ—1,2–naphthoquinone.</p>
Full article ">Figure 9
<p>Early-life exposure to 1,2-naphthoquinone (1,2-NQ) significantly heightens the likelihood of cardiac adrenergic dysfunction, manifesting as atrial fibrillation, electrocardiogram irregularities (ECG), and diminished heart rate. This heightened risk potentially predisposes individuals to atrial arrhythmia during their youth. The underlying mechanisms hinge on the potential disruption of β-adrenoceptors and the impairment of atrial maturation or proper heart development (desensitization or genotoxicity). This process is partially mediated by TRPV1 activation but not by TRPC5.</p>
Full article ">
Back to TopTop