[go: up one dir, main page]

 
 
Sign in to use this feature.

Years

Between: -

Subjects

remove_circle_outline
remove_circle_outline
remove_circle_outline
remove_circle_outline
remove_circle_outline
remove_circle_outline
remove_circle_outline
remove_circle_outline
remove_circle_outline

Journals

Article Types

Countries / Regions

remove_circle_outline
remove_circle_outline
remove_circle_outline
remove_circle_outline

Search Results (284)

Search Parameters:
Keywords = TRPM2

Order results
Result details
Results per page
Select all
Export citation of selected articles as:
17 pages, 7749 KiB  
Article
Decoding Cold Therapy Mechanisms of Enhanced Bone Repair through Sensory Receptors and Molecular Pathways
by Matthew Zakaria, Justin Matta, Yazan Honjol, Drew Schupbach, Fackson Mwale, Edward Harvey and Geraldine Merle
Biomedicines 2024, 12(9), 2045; https://doi.org/10.3390/biomedicines12092045 - 9 Sep 2024
Viewed by 358
Abstract
Applying cold to a bone injury can aid healing, though its mechanisms are complex. This study investigates how cold therapy impacts bone repair to optimize healing. Cold was applied to a rodent bone model, with the physiological responses analyzed. Vasoconstriction was mediated by [...] Read more.
Applying cold to a bone injury can aid healing, though its mechanisms are complex. This study investigates how cold therapy impacts bone repair to optimize healing. Cold was applied to a rodent bone model, with the physiological responses analyzed. Vasoconstriction was mediated by an increase in the transient receptor protein channels (TRPs), transient receptor potential ankyrin 1 (TRPA1; p = 0.012), and transient receptor potential melastatin 8 (TRPM8; p < 0.001), within cortical defects, enhancing the sensory response and blood flow regulation. Cold exposure also elevated hypoxia (p < 0.01) and vascular endothelial growth factor expression (VEGF; p < 0.001), promoting angiogenesis, vital for bone regeneration. The increased expression of osteogenic proteins peroxisome proliferator-activated receptor gamma coactivator (PGC-1α; p = 0.039) and RNA-binding motif protein 3 (RBM3; p < 0.008) suggests that the reparative processes have been stimulated. Enhanced osteoblast differentiation and the presence of alkaline phosphatase (ALP) at day 5 (three-fold, p = 0.021) and 10 (two-fold, p < 0.001) were observed, along with increased osteocalcin (OCN) at day 10 (two-fold, p = 0.019), indicating the presence of mature osteoblasts capable of mineralization. These findings highlight cold therapy’s multifaceted effects on bone repair, offering insights for therapeutic strategies. Full article
(This article belongs to the Section Cell Biology and Pathology)
Show Figures

Figure 1

Figure 1
<p>Potential mechanisms activated in response to cold exposure.</p>
Full article ">Figure 2
<p>Identification of receptor proteins involved in vasoconstriction within a cortical defect following cold exposure. TRPM8 and TRPA1 detection (bright green) within the cortical defect region (4× magnification). Scale bar represents 500 µm. (<b>A</b>,<b>D</b>) are non-treated femurs serving as a baseline. (<b>B</b>,<b>E</b>) are cold-treated femurs. (<b>C</b>) TRPA1 staining expression analysis (<span class="html-italic">n</span> = 8) of positive staining for TRPA1 in the control group was 5.38 ± 2.32, while in the experimental group it was 9.03% ± 2.78 (* <span class="html-italic">p</span>-value = 0.012). (<b>F</b>) TRPM8 staining expression analysis (<span class="html-italic">n</span> = 8) of positive staining for TRPM8 in the control group was 3.86% ± 1.72, while in the experimental group it was 8.75% ± 1.68 (** <span class="html-italic">p</span>-value &lt; 0.001).</p>
Full article ">Figure 3
<p>Identification of angiogenic-related factors and endothelial cells within a cortical defect following cold exposure. VEGF and CD34 (bright green) detection within the cortical defect region (4× magnification). Scale bar represents 500 µm. (<b>A</b>,<b>D</b>,<b>G</b>) are the cold-treated femurs. (<b>B</b>,<b>E</b>,<b>H</b>) are the non-treated femurs serving as a baseline. (<b>C</b>) Hypoxia staining expression analysis (<span class="html-italic">n</span> = 10) of positive staining for hypoxia in the control group was 17.9% ± 3.8, while in the experimental group it was 23.5% ± 05.8 (* <span class="html-italic">p</span>-value &lt; 0.01). (<b>F</b>) VEGF staining expression analysis (<span class="html-italic">n</span> = 8) of positive staining for VEGF in the control group was 4.05% ± 1.83, while in the experimental group it was 9.69% ± 1.75 (* <span class="html-italic">p</span>-value &lt; 0.001). (<b>I</b>) CD34 staining expression analysis (<span class="html-italic">n</span> = 8) of positive staining for CD34 in the control group was 6.92% ± 2.77, while in the experimental group it was 8.00% ± 2.76 (<span class="html-italic">p</span>-value = 0.42).</p>
Full article ">Figure 4
<p>Identification of cold-shock proteins and hypoxia in regenerating bone following cold exposure. PGC-1a (brown), RBM3 (brown), and hypoxia (brown) detection within the cortical defect region (4x magnification). Scale bar represents 500 µm. (<b>A</b>,<b>D</b>) are cold-treated femurs. (<b>B</b>,<b>E</b>) are non-treated femurs serving as a baseline. (<b>C</b>) RBM3 staining expression analysis (<span class="html-italic">n</span> = 8) of positive staining for RBM3 in the control group was 4.47% ± 2.84, while in the experimental group it was 11.0% ± 4.26 (** <span class="html-italic">p</span>-value &lt; 0.008). (<b>F</b>) PGC-1a staining expression analysis (<span class="html-italic">n</span> = 8) of positive staining for PGC-1a in the control group was 2.57% ± 0.36, while in the experimental group it was 5.85% ± 1.01 (* <span class="html-italic">p</span>-value = 0.039).</p>
Full article ">Figure 5
<p>Prevalence of PGC-1a at the junction of old and new bone formation following cold therapy treatment. (<b>A</b>,<b>E</b>) at 2.5× magnification. Scale bar represents 500 µm. (<b>B</b>–<b>D</b>,<b>F</b>–<b>H</b>) at 20× magnification. Scale bar represents 50 µm. PGC-1a has been shown to be upregulated (<b>I</b>) at the junction of new and old bone. (<b>F</b>–<b>H</b>) Furthermore, PGC-1a expression within these junctions appears to be elevated following cold exposure (<b>B</b>–<b>D</b>) indicating PGC-1a upregulation. (<b>I</b>) Detected location of PGC-1α staining expression analysis (<span class="html-italic">n</span> = 8) of PGC-1α within cells at the junction of newly formed bone and old bone was 70.56% ± 7.08, while in non-junctional regions it was 29.44% ± 7.08 (* <span class="html-italic">p</span>-value &lt; 0.00001).</p>
Full article ">Figure 6
<p>Detection of ALP and OCN at day 5 and day 10 within isolated osteoblasts following cold exposure and mineralization within isolate osteoblasts following cold exposure for 14 days. (<b>A</b>) ALP activity after 5 days of cold exposure was 2.72 U/L (µmol/min/L) ± 0.85, while in the non-treated group it was 0.74 U/L ± 0.35 (* <span class="html-italic">p</span>-value = 0.021). (<b>B</b>) ALP activity after 10 days of cold exposure was 2.25 U/L ± 0.04, while in the non-treated group it was 1.0301 U/L ± 0.24 (** <span class="html-italic">p</span>-value &lt; 0.001). (<b>C</b>) OCN levels after 5 days of cold exposure was 0.34 mg/mL ± 0.10, while in the non-treated group it was 0.43 mg/mL ± 0.10 (<span class="html-italic">p</span>-value = 0.29). (<b>D</b>) OCN levels after 10 days of cold exposure was 0.84 mg/mL ± 0.17, while in the non-treated group it was 0.35 mg/mL ± 0.10 (*** <span class="html-italic">p</span>-value = 0.019). ARS (red) detection within isolated osteoblasts (10× magnification). Scale bar represents 250 µm. (<b>E</b>) Osteoblasts isolated from non-treated fractured femurs. (<b>F</b>) Osteoblasts isolated from daily cold-treated fractured femurs. (<b>G</b>) ARS staining analysis: Detection of ARS after 14 days of cold exposure was 1.04% ± 0.40, while in the non-treated group it was 0.27% ± 0.025 (*** <span class="html-italic">p</span>-value = 0.030).</p>
Full article ">Figure 7
<p>Cold-based activation of vasoconstrictive pathways and hypoxic impact. (<b>A</b>,<b>B</b>) TRPA1 and TRPM8 cold exposure mechanism leads to the activation of TRPM8 (8–28 °C) and TRPA1 (8–17 °C), initiating a contractive response leading to vasoconstriction. (<b>C</b>) Hypoxic impacts. A reduction in the influx of blood supply coincides with a decrease in oxygen availability, leading to the formation of an acute hypoxic microenvironment with respect to nearby cells. NA: norepinephrine.</p>
Full article ">Figure 8
<p>Hypoxic upregulation following vasoconstriction via local cold exposure and the impact on angiogenic pathways. Under normal oxygen conditions, HIF-1a is degraded via von Hippel–Lindau (VHL) disease after prolyl hydroxylation, but in hypoxia, HIF-1a forms a dimer with HIF-1B, nucleolocalizes with the aryl hydrocarbon receptor nuclear translocator (ARNT), leading to VEGF expression through the HRE pathway, which then activates VEGFR2 in endothelial cells, upregulating cellular proliferation and angiogenesis via the Ras/Raf signaling pathway.</p>
Full article ">Figure 9
<p>Cellular adaptive signaling pathways activated by cold exposure. RBM3 activation and downstream osteogenic potential following short-term cold exposure. (<b>A</b>) Cold exposure triggers a cellular cold-shock response, activating RBM3 via HSPs released by phosphorylated and nuclear-localized HSF1. (<b>B</b>) HSP70 activates TLR4, initiating the MyD88/NF-kB pathway, with phosphorylated NF-kB essential for RBM3 expression. (<b>C</b>) RBM3 activates the MAPK/ERK pathway, leading to Runx2 and osteocalcin (OCN) expression, crucial for osteoblast differentiation. PGC-1a activation and downstream osteogenic potential following short-term cold exposure. PGC-1α activation and osteogenic potential following short-term cold exposure. (<b>D</b>) Cold activation of adrenergic receptors and TRPs stimulates cAMP and calcium influx, leading to CREB phosphorylation via the PKA/Ca<sup>2+</sup>/CaMK pathway. (<b>E</b>) PGC-1α forms a complex with ERRα in osteoblastic cells, promoting differentiation via the Runx2/OCN pathway for mineralization. (<b>F</b>) PGC-1α and NRF-1/NRF-2 activate TFAM and NCMP, crucial for mitochondrial biogenesis and intracellular calcium storage.</p>
Full article ">Figure 10
<p>Overview of the mechanistic propensities of short-duration cold exposure and the potential impacts on a bone injury site. Blue: Observed impact of cold exposure on bone and vasculature morphology. Green: Observed impact of cold exposure on pathways involved in bone repair.</p>
Full article ">
27 pages, 13297 KiB  
Article
Complement Component C5a and Fungal Pathogen Induce Diverse Responses through Crosstalk between Transient Receptor Potential Channel (TRPs) Subtypes in Human Conjunctival Epithelial Cells
by Loreena Rech, Tina Dietrich-Ntoukas, Peter S. Reinach, Tobias Brockmann, Uwe Pleyer and Stefan Mergler
Cells 2024, 13(16), 1329; https://doi.org/10.3390/cells13161329 - 9 Aug 2024
Viewed by 625
Abstract
The conjunctiva has immune-responsive properties to protect the eye from infections. Its innate immune system reacts against external pathogens, such as fungi. The complement factor C5a is an important contributor to the initial immune response. It is known that activation of transient-receptor-potential-vanilloid 1 [...] Read more.
The conjunctiva has immune-responsive properties to protect the eye from infections. Its innate immune system reacts against external pathogens, such as fungi. The complement factor C5a is an important contributor to the initial immune response. It is known that activation of transient-receptor-potential-vanilloid 1 (TRPV1) and TRP-melastatin 8 (TRPM8) channels is involved in different immune reactions and inflammation in the human body. The aim of this study was to determine if C5a and mucor racemosus e voluminae cellulae (MR) modulate Ca2+-signaling through changes in TRPs activity in human conjunctival epithelial cells (HCjECs). Furthermore, crosstalk was examined between C5a and MR in mediating calcium regulation. Intracellular Ca2+-concentration ([Ca2+]i) was measured by fluorescence calcium imaging, and whole-cell currents were recorded using the planar-patch-clamp technique. MR was used as a purified extract. Application of C5a (0.05–50 ng/mL) increased both [Ca2+]i and whole-cell currents, which were suppressed by either the TRPV1-blocker AMG 9810 or the TRPM8-blocker AMTB (both 20 µM). The N-terminal peptide C5L2p (20–50 ng/mL) blocked rises in [Ca2+]i induced by C5a. Moreover, the MR-induced rise in Ca2+-influx was suppressed by AMG 9810 and AMTB, as well as 0.05 ng/mL C5a. In conclusion, crosstalk between C5a and MR controls human conjunctival cell function through modulating interactions between TRPV1 and TRPM8 channel activity. Full article
(This article belongs to the Section Cell Signaling)
Show Figures

Figure 1

Figure 1
<p>C5a increases the Ca<sup>2+</sup>-concentration in HCjEC at 0.05, 5, and 50 ng/mL. Data are means ± SEM. The reagents are added at the time points indicated by arrows. A control baseline measurement shows invariant Ca<sup>2+</sup> levels (<span class="html-italic">n</span> = 123; open circles). (<b>a</b>) 0.05 ng/mL C5a induces an irreversible increase in intracellular Ca<sup>2+</sup>-concentration (<span class="html-italic">n</span> = 313, red-filled circles). (<b>b</b>) Same experiment as shown in (<b>a</b>), but with 5 ng/mL C5a (<span class="html-italic">n</span> = 56, red-filled circles). (<b>c</b>) In the presence of 50 ng/mL C5a, Ca<sup>2+</sup> similarly increases (<span class="html-italic">n</span> = 357, red-filled circles). (<b>d</b>) Summary of the experiments with C5a in HCjEC. The dashed line is the reference line at 0.1. The asterisks (***) designate significant increases in [Ca<sup>2+</sup>]<sub>i</sub> with C5a (<span class="html-italic">t</span> = 400 s, 600 s; <span class="html-italic">n</span> = 56–357; <span class="html-italic">p</span> &lt; 0.001; paired tested) (gray bars) compared to controls (white bars) (<span class="html-italic">t</span> = 100 s). The hashtags (###) indicate statistically significant differences between C5a at different concentrations (0.05 ng/mL, 5 ng/mL, 50 ng/mL) (<span class="html-italic">t</span> = 400 s, 600 s; <span class="html-italic">n</span> = 56–357; ## <span class="html-italic">p</span> &lt; 0.01; ### <span class="html-italic">p</span> &lt; 0.001; unpaired tested).</p>
Full article ">Figure 2
<p>TRP channel blockers block the C5a-induced Ca<sup>2+</sup> increase. Data are means ± SEM. The reagents are added at the time points indicated by arrows. A control baseline is measured, showing an invariant Ca<sup>2+</sup> baseline (4 min) in the presence of the TRP channel blockers (<span class="html-italic">n</span> = 216–252). (<b>a</b>) Same experiment as shown in <a href="#cells-13-01329-f001" class="html-fig">Figure 1</a>a, but in the presence of AMG 9810 (20 µM) (<span class="html-italic">n</span> = 252). (<b>b</b>) AMTB (20 µM) had a similar inhibitory effect (<span class="html-italic">n</span> = 216). (<b>c</b>) Summary of the experiments with C5a and the TRP channel blockers in HCjEC. The dashed line is the reference line at 0.1. The asterisks (***) designate significant increases in [Ca<sup>2+</sup>]<sub>i</sub> without blocker (<span class="html-italic">t</span> = 400 s, 600 s; <span class="html-italic">n</span> = 313; <span class="html-italic">p</span> &lt; 0.001; paired tested) (gray bars) compared to controls (white bars) (<span class="html-italic">t</span> = 100 s). The hashtags (###) indicate statistically significant differences between C5a with and without the aforementioned TRP channel blockers (<span class="html-italic">t</span> = 400 s, 600 s; <span class="html-italic">n</span> = 216–313; <span class="html-italic">p</span> &lt; 0.001; unpaired tested).</p>
Full article ">Figure 3
<p>The N-terminal peptide fragment of C5L2 from human (h) of murine (m) origin (h/mC5L2p) suppresses the C5a-induced Ca<sup>2+</sup> increase in a dose-dependent manner. Data are means ± SEM. The reagents are added at the time points indicated by arrows. A control baseline was measured showing Ca<sup>2+</sup> invariance (4 min) in the presence of mC5L2p or hC5L2p (<span class="html-italic">n</span> = 96–313). (<b>a</b>) Same experiment as shown in <a href="#cells-13-01329-f001" class="html-fig">Figure 1</a>a, but in the presence of 20 ng/mL mouse C5L2p (<span class="html-italic">n</span> = 248). (<b>b</b>) 50 ng/mL mC5L2p abolishes the C5a-induced Ca<sup>2+</sup>-increase (<span class="html-italic">n</span> = 96). (<b>c</b>) 20 ng/mL hC5L2p (20 ng/mL) partially suppresses C5a-induced Ca<sup>2+</sup>-increase (<span class="html-italic">n</span> = 151). (<b>d</b>) 50 ng/mL human hC5L2p (50 ng/mL) abolishes the C5a-induced Ca<sup>2+</sup>-increase (<span class="html-italic">n</span> = 162). (<b>e</b>) Summary of the experiments with C5a and m/hC5L2p in HCjEC. The dashed line is the reference line at 0.1. The asterisks (***) designate significant change in [Ca<sup>2+</sup>]<sub>i</sub> with C5a (<span class="html-italic">t</span> = 600 s; <span class="html-italic">n</span> = 96–313; <span class="html-italic">p</span> &lt; 0.001; paired tested) compared to the control (<span class="html-italic">t</span> = 100 s). The hashtags (###) indicate statistically significant differences between C5a with and without m/hC5L2p at concentrations of 20 ng/mL and 50 ng/mL (<span class="html-italic">t</span> = 600 s; <span class="html-italic">n</span> = 96–313; <span class="html-italic">p</span> &lt; 0.001; unpaired tested).</p>
Full article ">Figure 4
<p>MR increases intracellular Ca<sup>2+</sup>. Data are means ± SEM. The reagents were added at the time points indicated by arrows. A control baseline measurement is invariant (<span class="html-italic">n</span> = 123; open circles). (<b>a</b>) 1 mg/mL MR induces an increase in intracellular Ca<sup>2+</sup>-concentration (<span class="html-italic">n</span> = 145, red filled circles) in HCjEC. (<b>b</b>) Same experiment as shown in (<b>a</b>), but with 1.5 ng/mL C5a (<span class="html-italic">n</span> = 92, red filled circles). (<b>c</b>) Summary of the experiments with MR in HCjEC. The dashed line is the reference line at 0.1. The asterisks (***) designate significant increases in [Ca<sup>2+</sup>]<sub>i</sub> with MR (<span class="html-italic">t</span> = 400 s, 600 s; <span class="html-italic">n</span> = 92, 145; <span class="html-italic">p</span> &lt; 0.001; paired tested) compared to control (<span class="html-italic">t</span> = 100 s). The hashtag (##) indicates statistically significant differences between MR at different concentrations of 1 mg/mL and 1.5 mg/mL); <span class="html-italic">t</span> = 400 s, 600 s; <span class="html-italic">n</span> = 92–145; <span class="html-italic">p</span> &lt; 0.01; unpaired tested).</p>
Full article ">Figure 5
<p>TRP channel blockers suppress the MR-induced rise in Ca<sup>2+</sup>. Data are means ± SEM. The reagents are added at the time points indicated by arrows. An invariant control baseline is evident (4 min) in the presence of the TRP channel blockers (<span class="html-italic">n</span> = 50–245). (<b>a</b>) Same experiment as shown in <a href="#cells-13-01329-f004" class="html-fig">Figure 4</a>a, but in the presence of La<sup>3+</sup> (1 mM), which completely blocks the MR-induced Ca<sup>2+</sup>-increase (<span class="html-italic">n</span> = 50). (<b>b</b>) AMG 9810 (20 µM) only partially suppresses the MR-rise in Ca<sup>2+</sup>-increase (<span class="html-italic">n</span> = 154). (<b>c</b>) AMTB (20 µM) completely blocks the MR-rise in Ca<sup>2+</sup> (<span class="html-italic">n</span> = 163). (<b>d</b>) Same result with BCTC (20 µM) (<span class="html-italic">n</span> = 109). (<b>e</b>) CPZ (20 µM) temporarily suppresses the MR-rise in Ca<sup>2+</sup>-increase slightly below the baseline (dashed line) (<span class="html-italic">n</span> = 245). (<b>f</b>) Summary of the experiments with MR and the TRP channel blockers in HCjEC. The dashed line is the reference line at 0.1. The asterisks (***) designate significant increases in [Ca<sup>2+</sup>]<sub>i</sub> with MR (<span class="html-italic">t</span> = 400 s and <span class="html-italic">t</span> = 600 s; <span class="html-italic">n</span> = 50–245; *** <span class="html-italic">p</span> &lt; 0.001; paired tested) relative to control (<span class="html-italic">t</span> = 100 s). The hashtags (###) indicate statistically significant differences between MR with and without the aforementioned TRP channel blockers (<span class="html-italic">t</span> = 600 s; <span class="html-italic">n</span> = 50–245; <span class="html-italic">p</span> &lt; 0.001; unpaired tested).</p>
Full article ">Figure 6
<p>Interplay between MR and C5a. Data are means ± SEM. The reagents are added at the time points indicated by arrows. The dashed line is the reference line at 0.1. A control baseline shows a Ca<sup>2+</sup> homeostasis (4 min) in the presence of MR or C5a (<span class="html-italic">n</span> = 197–313). (<b>a</b>) Same experiment as shown in <a href="#cells-13-01329-f001" class="html-fig">Figure 1</a>a, but in the presence of 1 mg/mL MR. 0.05 ng/mL C5a increases the intracellular Ca<sup>2+</sup> (<span class="html-italic">n</span> = 197). (<b>b</b>) Summary of the experiments with C5a and MR in HCjEC. The asterisks (***) designate significant increases in [Ca<sup>2+</sup>]<sub>i</sub> with C5a (<span class="html-italic">t</span> = 600 s; <span class="html-italic">n</span> = 197; *** <span class="html-italic">p</span> &lt; 0.001; paired tested) compared to control (<span class="html-italic">t</span> = 100 s). (<b>c</b>) 0.05 ng/mL C5a completely blocked the MR-induced Ca<sup>2+</sup>-increase to just below the baseline (dashed line; <span class="html-italic">n</span> = 215). (<b>d</b>) Similar analysis as shown in (<b>b</b>), but with C5a preincubation instead of MR. The asterisks (***) designate significant increases in [Ca<sup>2+</sup>]<sub>i</sub> with C5a (<span class="html-italic">t</span> = 600 s; <span class="html-italic">n</span> = 145; *** <span class="html-italic">p</span> &lt; 0.001; paired tested) compared to control (<span class="html-italic">t</span> = 100 s). The hashtags (###) indicate statistically significant differences between MR with and without C5a (<span class="html-italic">t</span> = 600 s; <span class="html-italic">n</span> = 50–245; <span class="html-italic">p</span> &lt; 0.001; unpaired tested).</p>
Full article ">Figure 7
<p>C5a increases whole-cell currents through increases in TRPV1 activity. (<b>a</b>) Time-course recording of the increases in the current induced by C5a (50 ng/mL) that are suppressed by AMG 9810 (20 µM). The dashed line is the reference line at 0 pA/pF. (<b>b</b>) Original traces of current changes induced by voltage ramps in the presence of C5a. Current densities are shown before application as control (labeled as A), during application of 50 ng/mL C5a (labeled as B), and after addition of 20 µM AMG 9810 (labeled as C). (<b>c</b>) Summary of the experiments with C5a and AMG 9810. The asterisks (*) indicate statistically significant differences of whole-cell currents with and without C5a (<span class="html-italic">n</span> = 6; * <span class="html-italic">p</span> &lt; 0.05, ** <span class="html-italic">p</span> &lt; 0.01; paired tested) and significant differences of C5a-induced rises with and without AMG 9810 (<span class="html-italic">n</span> = 6; ** <span class="html-italic">p</span> &lt; 0.01, *** <span class="html-italic">p</span> &lt; 0.005; paired tested). (<b>d</b>) Maximum negative current amplitudes that a voltage step from 0 to −60 mV induces are shown as percent of control values before application of 50 ng/mL C5a (control set to 100%; dashed line). C5a-induces rises in inward currents that 20 µM AMG 9810 suppresses. (<b>e</b>) Same diagram as shown in (<b>d</b>), but related to the maximum positive current amplitudes that a voltage step from 0 to +130 mV induces. AMG 9810. (20 µM) suppresses C5a-rises in outward currents.</p>
Full article ">Figure 8
<p>MR increases whole-cell currents. (<b>a</b>) Time course recording of the current increases that MR induces (1 mg/mL) and declines that occur after replacement with fresh solution. The dashed line is the reference line at 0 pA/pF. (<b>b</b>) Original traces of MR-current rises that occur due to voltage ramps. Current densities are shown before application as a control (labeled as A), during application of 1 mg/mL MR (labeled as B), and after washout (labeled as C). (<b>c</b>) Summary of planar patch-clamp experiments with MR. The asterisks (**) indicate statistically significant differences in whole-cell currents with and without MR (<span class="html-italic">n</span> = 11; <span class="html-italic">p</span> &lt; 0.01; paired tested). (<b>d</b>) Maximum negative current amplitudes that a voltage step from 0 to −60 mV creates are shown as percent of control values before application of 1 mg/mL MR (control set to 100%; dashed line). MR induces rises in inward currents that decline to a control level after washout. The asterisks (***) indicate statistically significant differences in whole-cell currents with and without MR (<span class="html-italic">n</span> = 11; <span class="html-italic">p</span> &lt; 0.001; paired tested). (<b>e</b>) Same diagram as shown in (<b>d</b>), but show the maximum rises in positive current amplitudes that a voltage step from 0 to +130 mV induces. MR also induces a large increase in outward currents.</p>
Full article ">Figure 9
<p>MR increases whole-cell currents through TRPM8 activation. (<b>a</b>) Time course recording of the current increases that MR (1 mg/mL) induces and AMTB (20 µM) inhibits. The dashed line is the reference line at 0 pA/pF. (<b>b</b>) Original traces of rises in currents that MR augments of rises initially induced by voltage ramps. Current densities are shown before application as control (labeled as A), during application of 1 mg/mL MR (labeled as B), and after addition of 20 µM AMTB (labeled as C). (<b>c</b>) Summary of planar patch-clamp experiments with MR + AMTB. The asterisks (***) indicate statistically significant differences in whole-cell currents with and without MR + AMTB (<span class="html-italic">n</span> = 11; *** <span class="html-italic">p</span> &lt; 0.001; paired tested). (<b>d</b>) Maximum negative current amplitudes that a voltage step from 0 to −60 mV induces are shown as percent of control values before application of 1 mg/mL MR + 20 µM AMTB (control set to 100%; dashed line). MR + 20 µM AMTB reduces inward currents below control and washout levels. The asterisks (**) and (***) indicate statistically significant differences in whole-cell currents with and without MR + AMTB (<span class="html-italic">n</span> = 11; ** <span class="html-italic">p</span> &lt; 0.01, *** <span class="html-italic">p</span> &lt; 0.001; paired tested). (<b>e</b>) Same diagram as shown in (<b>d</b>), but shown as maximum positive current amplitudes induced by a voltage step from 0 to +130 mV.</p>
Full article ">Figure 10
<p>C5a suppresses MR-induced rises in currents. (<b>a</b>) Time course recording of the current increases that C5a (50 ng/mL) induces and decreases that result from the application of MR (1 mg/mL). The dashed line is the reference line at 0 pA/pF. (<b>b</b>) Original traces of C5a and rises in currents that MR-induces exceed those that a voltage creates. Current densities are shown before application as control (labelled as A), during application of 50 ng/mL C5a (labeled as B), and after addition of 1 mg/mL MR (labeled as C). (<b>c</b>) Summary of the experiments with C5a + MR. The asterisks (*), (**) and (***) indicate statistically significant differences in whole-cell currents with and without C5a and following the addition of MR (<span class="html-italic">n</span> = 7; * <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; paired tested). (<b>d</b>) Maximum negative current amplitudes at −60 mV are shown as percent of control values before application of 50 ng/mL C5a (control set to 100%; dashed line). C5a increases inward currents, which decline in the presence of MR (<span class="html-italic">n</span> = 7; <span class="html-italic">p</span> &lt; 0.001). The asterisks (***) indicate statistically significant differences in whole-cell currents with and without C5a and following the addition of MR (<span class="html-italic">n</span> = 7; *** <span class="html-italic">p</span> &lt; 0.001; paired tested. (<b>e</b>) Same diagram as shown in (<b>d</b>), but related to maximum outward current amplitudes at +130 mV. The outward currents undergo a similar effect (<span class="html-italic">n</span> = 7; <span class="html-italic">p</span> &lt; 0.001), but the changes are smaller than those in the inward currents.</p>
Full article ">Figure 11
<p>This simplified scheme accounts for how interactions between C5a, MR, and TRPV1 or TRPM8 modulate rises in Ca<sup>2+</sup> influx (black arrows). Notably, C5a and MR elicit TRPV1 and TRPM8 activation in HCjEC (<a href="#cells-13-01329-f001" class="html-fig">Figure 1</a>, <a href="#cells-13-01329-f004" class="html-fig">Figure 4</a>, <a href="#cells-13-01329-f009" class="html-fig">Figure 9</a> and <a href="#app1-cells-13-01329" class="html-app">Figure S3</a>), whereas the interaction between C5a and MR is not fully elucidated. Ca<sup>2+</sup> permeable channels such as TRPV1 can be selectively blocked by AMG 9810 [<a href="#B74-cells-13-01329" class="html-bibr">74</a>] (red marking), and TRPM8 can be blocked by AMTB [<a href="#B58-cells-13-01329" class="html-bibr">58</a>,<a href="#B75-cells-13-01329" class="html-bibr">75</a>] (green marking). Two different physiological assays, such as fluorescence calcium imaging and the planar patch-clamp technique, were used to validate the effect of C5a or MR on intracellular Ca<sup>2+</sup> concentration and whole-cell currents. The C5L2p used contains only the N-terminal part of the C5aR2 and serves to neutralize C5a in the extracellular compartment (orange markings). Therefore, C5L2p does not enable signal transduction [<a href="#B61-cells-13-01329" class="html-bibr">61</a>]. The cells contain 3 GPCRs that can bind C5a (black arrows): These GPCRs are C5aR1, C5aR2 (C5L2), and C3aR. Notably, C5L2 is structurally homologous but deficient in G protein coupling [<a href="#B98-cells-13-01329" class="html-bibr">98</a>]. Finally, C5a has no effect on Ca<sup>2+</sup> regulation and no signal transduction in the presence of C5L2p neutralizing C5a (50 ng/mL) (highlighted in orange) (<a href="#cells-13-01329-f003" class="html-fig">Figure 3</a>).</p>
Full article ">
17 pages, 22700 KiB  
Article
Identification of Schizophrenia Susceptibility Loci in the Urban Taiwanese Population
by Chih-Chung Huang, Yi-Guang Wang, Chun-Lun Hsu, Ta-Chuan Yeh, Wei-Chou Chang, Ajeet B. Singh, Chin-Bin Yeh, Yi-Jen Hung, Kuo-Sheng Hung and Hsin-An Chang
Medicina 2024, 60(8), 1271; https://doi.org/10.3390/medicina60081271 - 6 Aug 2024
Viewed by 629
Abstract
Background and Objectives: Genomic studies have identified several SNP loci associated with schizophrenia in East Asian populations. Environmental factors, particularly urbanization, play a significant role in schizophrenia development. This study aimed to identify schizophrenia susceptibility loci and characterize their biological functions and [...] Read more.
Background and Objectives: Genomic studies have identified several SNP loci associated with schizophrenia in East Asian populations. Environmental factors, particularly urbanization, play a significant role in schizophrenia development. This study aimed to identify schizophrenia susceptibility loci and characterize their biological functions and molecular pathways in Taiwanese urban Han individuals. Materials and Methods: Participants with schizophrenia were recruited from the Taiwan Precision Medicine Initiative at Tri-Service General Hospital. Genotype–phenotype association analysis was performed, with significant variants annotated and analyzed for functional relevance. Results: A total of 137 schizophrenia patients and 26,129 controls were enrolled. Ten significant variants (p < 1 × 10−5) and 15 expressed genes were identified, including rs1010840 (SOWAHC and RGPD6), rs11083963 (TRPM4), rs11619878 (LINC00355 and LINC01052), rs117010638 (AGBL1 and MIR548AP), rs1170702 (LINC01680 and LINC01720), rs12028521 (KAZN and PRDM2), rs12859097 (DMD), rs1556812 (ATP11A), rs78144262 (LINC00977), and rs9997349 (ENPEP). These variants and associated genes are involved in immune response, blood pressure regulation, muscle function, and the cytoskeleton. Conclusions: Identified variants and associated genes suggest a potential genetic predisposition to schizophrenia in the Taiwanese urban Han population, highlighting the importance of potential comorbidities, considering population-specific genetic and environmental interactions. Full article
(This article belongs to the Section Psychiatry)
Show Figures

Figure 1

Figure 1
<p>The study encompassed 137 schizophrenia patients as case groups and 25,927 participants as the control group. Genotyping was conducted using the Taiwan Precision Medicine (TPM) array chip. A dataset of 493,852 SNPs underwent filtration, with 280,153 SNPs passing through and subsequently subjected to chi−squared testing for the detection of risk factors.</p>
Full article ">Figure 2
<p>Manhattan plot of associated variants results in schizophrenia patients. Following chi−squared testing, 280,153 variants were detected among TSGH TPMI participants. Among these, ten highly significant variants were identified based on a <span class="html-italic">p</span>-value &lt; 10<sup>−5</sup> (green dots above the blue dashed line): rs78144262, rs9997349, rs1010840, rs11083963, rs11619878, rs1170702, rs117010638, rs12028521, rs12859097, and rs1556812.</p>
Full article ">Figure 3
<p>Interpretation associations results in Q-Q (quantile-quantile) plot. The distribution of low <span class="html-italic">p</span>-values (represented by high observed −log10(<span class="html-italic">p</span>) values) indicates true associations with schizophrenia. The genomic inflation factor, lambda, was determined based on the median observed and expected chi-square values.</p>
Full article ">Figure 4
<p>Top 10 significant Gene Ontology biological process (GO BP) functions relative to variants. The BP functions of genes associated with identified risk variants were characterized using the GO database. Selected GO terms exhibited with <span class="html-italic">p</span>-values &lt; 0.05.</p>
Full article ">Figure 5
<p>Top 10 significant pathways relative to variants analysis. Genes associated with identified risk variants were characterized using the Reactome database. Selected pathways exhibited with <span class="html-italic">p</span>-values &lt; 0.05.</p>
Full article ">Figure 6
<p>Gene functional and molecular pathway annotation network of AGBL1, ATP11A, DMD, ENPEP, and TPM. The functional annotation network was constructed using GO biological process (green circles) and Reactome pathway (pink circles).</p>
Full article ">
14 pages, 6854 KiB  
Article
Inhibition of Cutaneous TRPV3 Channels by Natural Caffeic Acid for the Alleviation of Skin Inflammation
by Guoji Zhang, Liqin Wang, Yaxuan Qu, Shilun Mo, Xiaoying Sun and Kewei Wang
Molecules 2024, 29(16), 3728; https://doi.org/10.3390/molecules29163728 - 6 Aug 2024
Viewed by 652
Abstract
Natural caffeic acid (CA) and its analogues have been studied for their potential applications in the treatment of various inflammatory and infectious skin diseases. However, the molecular mechanism underlying the effects of the CA remains largely unknown. Here, we report that CA and [...] Read more.
Natural caffeic acid (CA) and its analogues have been studied for their potential applications in the treatment of various inflammatory and infectious skin diseases. However, the molecular mechanism underlying the effects of the CA remains largely unknown. Here, we report that CA and its two analogues, caffeic acid phenethyl ester (CAPE) and caffeic acid methyl caffeate (CAMC), inhibit TRPV3 currents in their concentration- and structure-dependent manners with IC50 values ranging from 102 to 410 μM. At the single-channel level, CA reduces the channel open probability and open frequency without alteration of unitary conductance. CA selectively inhibits TRPV3 relative to other subtypes of thermo-TRPs, such as TRPA1, TRPV1, TRPV4, and TRPM8. Molecular docking combined with site-specific mutagenesis reveals that a residue T636 in the Pore-loop is critical for CA binding to TRPV3. Further in vivo evaluation shows that CA significantly reverses TRPV3-mediated skin inflammation induced by skin sensitizer carvacrol. Altogether, our findings demonstrate that CA exerts its anti-inflammatory effects by selectively inhibiting TRPV3 through binding to the pocket formed by the Pore-loop and the S6. CA may serve as a lead for further modification and identification of specific TRPV3 channel inhibitors. Full article
(This article belongs to the Special Issue Effect of Natural Products on Skin Diseases)
Show Figures

Graphical abstract

Graphical abstract
Full article ">Figure 1
<p>The chemical structures of forsythiaside B (FB), verbascoside (Ver), isochlorogenic acid A (IAA), isochlorogenic acid B (IAB), caffeic acid (CA), caffeic acid phenethyl ester (CAPE), and caffeic acid methyl caffeate (CAMC).</p>
Full article ">Figure 2
<p>Concentration- and structure-dependent inhibition of TRPV3 currents by caffeoyl analogues. (<b>A</b>–<b>C</b>) Left panel, inhibition of whole-cell currents of human TRPV3 (hTRPV3) channel activated by agonist 2-aminoethyl diphenylborinate (2-APB, 50 μM, red bar) and increasing concentrations of caffeic acid (CA, blue bar), caffeic acid phenethyl ester (CAPE, purple bar), or caffeic acid methyl caffeate (CAMC, green bar) from 1 to 1000 μM. Right panel, current–voltage curves of hTRPV3 in response to voltage ramps from −100 to +100 mV from the left panel after the addition of 50 μM 2-APB and co-addition of CA, CAPE, or CAMC from 1 to 1000 μM. (<b>D</b>) The concentration-dependent inhibition of hTRPV3 by caffeoyl analogues at +80 mV was analysed by Hill equation fitting, with IC<sub>50</sub> value of 102.1 ± 19.7 μM (CA, <span class="html-italic">n</span> = 5), 276.7 ± 41.9 μM (CAPE, <span class="html-italic">n</span> = 5), and 409.8 ± 57.9 μM (CAMC, <span class="html-italic">n</span> = 5). Data are expressed as the mean ± SD.</p>
Full article ">Figure 3
<p>Reduction of hTRPV3 single-channel open probability by caffeic acid. (<b>A</b>) Left panel, representative single-channel current traces recorded at −60 mV in inside-out configurations before and after the addition of different TRPV3 agonists (50 μM 2-APB or 300 μM carvacrol) and co-application with 500 μM caffeic acid (CA). All-point amplitude histograms of single-channel currents in 5 s were shown in their right panels. Dotted lines indicate the closed channel state (C) and the opened channel state (O), respectively. (<b>B</b>) Summary of the average open probability (<span class="html-italic">P</span><sub>OPEN</sub>) values of hTRPV3 single channel in the presence of control (circle), different TRPV3 agonists (square) and co-application with CA (triangles) (<span class="html-italic">n</span> = 5, **** <span class="html-italic">p</span> &lt; 0.0001, by unpaired <span class="html-italic">t</span> test). (<b>C</b>) Summary of hTRPV3 single-channel open frequency (Freq) after exposure to control (circle), different TRPV3 agonists (square) and co-application with CA (triangles) (<span class="html-italic">n</span> = 5, **** <span class="html-italic">p</span> &lt; 0.0001, by unpaired <span class="html-italic">t</span> test). (<b>D</b>) Summary of hTRPV3 single channel conductance after exposure to different TRPV3 agonists (square) and co-application with CA (triangles) (<span class="html-italic">n</span> = 5, ns, no significance, by unpaired <span class="html-italic">t</span> test). Data are expressed as the mean ± SD.</p>
Full article ">Figure 4
<p>Selectivity of caffeic acid for TRPV3 over TRPA1, TRPV1, TRPV4, and TRPM8 channels. (<b>A</b>) Left panel, whole-cell current recordings of human TRPV3 (hTRPV3) channels expressed in HEK293T cells in response to 50 μM 2-APB (red bar) and co-application of 1000 μM caffeic acid (CA, blue bar). Right panel, current–voltage curves of hTRPV3 in response to voltage ramps from −100 to +100 mV under control condition (0) after addition of 50 μM 2-APB (1) and co-addition of 1000 μM CA (2) and washout (3). (<b>B</b>) Left panel, whole-cell current recordings of human TRPA1 (hTRPA1) channels expressed in HEK293T cells in response to 300 μM TRPA1 agonist allyl isothiocyanate (AITC, red bar) and co-application of 1000 μM CA (blue bar) and inhibited by TRPA1 antagonist A96 (10 μM, brown bar). Right panel, current–voltage curves of hTRPA1 in response to voltage ramps from −100 to +100 mV under control condition (0) after addition of 300 μM AITC (1) and co-addition of 1000 μM CA (2) and inhibited by 10 μM A96 (3). (<b>C</b>) Left panel, whole-cell current recordings of human TRPV1 (hTRPV1) channels expressed in HEK293T cells in response to 1 μM TRPV1 agonist capsaicin (red bar) or co-application of 1000 μM CA (blue bar) and washout. Right panel, current–voltage curves of hTRPV1 in response to voltage ramps from −100 to +100 mV under control condition (0) after addition of 1 μM capsaicin (1) and co-addition of 1000 μM CA (2) and washout (3). (<b>D</b>) Left panel, whole-cell current recordings of human TRPV4 (hTRPV4) channels expressed in HEK293T cells in response to 0.1 μM TRPV4 agonist GSK1016790A (GSK101, red bar) and co-application of 1000 μM CA (blue bar) and inhibited by 130 mM BaCl<sub>2</sub> (grey bar). Right panel, current–voltage curves of hTRPV4 in response to voltage ramps from −100 to +100 mV under control condition (0) after addition of 0.1 μM GSK101 (1) and co-addition of 1000 μM CA (2) and inhibited by 130 mM BaCl<sub>2</sub> (3). (<b>E</b>) Left panel, whole-cell current recordings of human TRPM8 (hTRPM8) channels expressed in HEK293T cells in response to 500 μM TRPM8 agonist menthol (red bar) or co-application of 1000 μM CA (blue bar) and washout. Right panel, current–voltage curves of hTRPM8 in response to voltage ramps from −100 to +100 mV under control condition (0) after addition of 500 μM menthol (1) and co-addition of 1000 μM CA (2) and washout (3). (<b>F</b>) Summary of average current inhibition of hTRPV3 (red), hTRPA1 (brown), hTRPV1 (purple), hTRPV4 (blue), and hTRPM8 (pink) channels by 1000 μM CA, (<span class="html-italic">n</span> = 5, **** <span class="html-italic">p</span> &lt; 0.0001, by unpaired <span class="html-italic">t</span> test). Data are expressed as the mean ± SD.</p>
Full article ">Figure 5
<p>Binding site and key residues of caffeic acid and its analogues binding to TRPV3 channels. (<b>A</b>) The top-down view of a putative pocket for caffeic acid (CA, blue), caffeic acid phenethyl ester (CAPE, purple), and caffeic acid methyl caffeate (CAMC, green) binding to mouse TRPV3 structure (PDB ID code: 6DVY) from docking. The four subunits of the tetramer are distinguished in four different colours. (<b>B</b>) The side view of CA (blue), CAPE (purple), and CAMC (green) in the pocket formed by Pore-loop (<span class="html-italic">p</span>-loop) and the S6 segment with the key residues T636 or I637 through the hydrogen bonds (yellow dotted line). (<b>C</b>,<b>D</b>) Left panel, representative whole-cell recordings of wild-type (WT) hTRPV3 (<b>C</b>) and T636A (<b>D</b>) mutant expressed in HEK293T cells in the presence of 50 μM 2-APB alone (red bar) and co-application of 500 μM caffeic acid (CA, blue bar). Right panel, current–voltage curves of WT hTRPV3 and T636A mutant in response to voltage ramps from −100 to +100 mV under control condition (0) after addition of 50 μM 2-APB (1) and co-addition of 500 μM CA (2) and washout (3). (<b>E</b>) Summary for WT hTRPV3 (black) or mutants (T636A, red; I637A, blue; F666A, green; L669A, purple; L670A, orange) channel currents inhibition by 500 μM CA (<span class="html-italic">n</span> = 5; ns, no significance; ** <span class="html-italic">p</span> &lt; 0.01; **** <span class="html-italic">p</span> &lt; 0.0001; by unpaired <span class="html-italic">t</span> test). (<b>F</b>) The concentration-dependent inhibition of WT hTRPV3 and mutations outward currents by CA at +80 mV was analysed by Hill equation fitting, with IC<sub>50</sub> value of 102.1 ± 19.7 μM (WT from <a href="#molecules-29-03728-f002" class="html-fig">Figure 2</a>D, <span class="html-italic">n</span> = 5), 104.5 ± 4.8 μM (F666A, <span class="html-italic">n</span> = 5), 179.7 ± 59.9 μM (L669A, <span class="html-italic">n</span> = 5), 305.6 ± 61.1 μM (I637A, <span class="html-italic">n</span> = 5), &gt;1000 μM (T636A, <span class="html-italic">n</span> = 4). (<b>G</b>) The concentration-dependent inhibition of WT hTRPV3 and mutations outward currents by CAPE and CAMC at +80 mV was analysed by Hill equation fitting, with IC<sub>50</sub> value of 276.7 ± 41.9 μM (CAPE-WT from <a href="#molecules-29-03728-f002" class="html-fig">Figure 2</a>D, <span class="html-italic">n</span> = 5), 409.8 ± 57.9 μM (CAMC-WT from <a href="#molecules-29-03728-f002" class="html-fig">Figure 2</a>D, <span class="html-italic">n</span> = 5), 506.3 ± 57.3 μM (CAPE-T636A, <span class="html-italic">n</span> = 5), 416.7 ± 45.1 μM (CAMC-T636A, <span class="html-italic">n</span> = 4), and 522.0 ± 43.8 μM (CAMC-I637A, <span class="html-italic">n</span> = 5). Data are expressed as the mean ± SD.</p>
Full article ">Figure 6
<p>Attenuation of dermatitis induced by skin sensitizer carvacrol by caffeic acid. (<b>A</b>) Schematic drawing of experimental procedures for generation of the mouse dermatitis model of dorsal skin by topical applications of skin sensitizer carvacrol (2%) and subcutaneous injections of caffeic acid (CA) at different concentrations (0.1, 1 mM) for 5 consecutive days. (<b>B</b>) Phenotypic features and histologic images of H&amp;E staining of dorsal skin tissue sections before and after topical applications of carvacrol (2%) for 5 consecutive days with and without subcutaneous injections of different concentrations CA. Scale bar = 100 μm for histologic images. (<b>C</b>) Dermatitis scores of mice in different groups treated with or without CA at different concentrations for 5 consecutive days from B (<span class="html-italic">n</span> = 5, * <span class="html-italic">p</span> &lt; 0.05, ** <span class="html-italic">p</span> &lt; 0.01, *** <span class="html-italic">p</span> &lt; 0.001, **** <span class="html-italic">p</span> &lt; 0.0001, by two-way ANOVA, followed by the Bonferroni’s test). (<b>D</b>) Summary of epidermal thickness of mouse dorsal skin sections (<span class="html-italic">n</span> = 3, **** <span class="html-italic">p</span> &lt; 0.0001, by one-way ANOVA, followed by the Dunnet’s test). Data are expressed as the mean ± SD.</p>
Full article ">
14 pages, 2371 KiB  
Article
Transcription Factor CcFoxO Mediated the Transition from Summer Form to Winter Form in Cacopsylla chinensis
by Chuchu Wei and Songdou Zhang
Int. J. Mol. Sci. 2024, 25(15), 8545; https://doi.org/10.3390/ijms25158545 - 5 Aug 2024
Viewed by 592
Abstract
Amid global climate change featuring erratic temperature fluctuations, insects adapt via seasonal polyphenism, essential for population sustainability and reproductive success. Cacopsylla chinensis, influenced by environment variations, displays a distinct summer form and winter form distinguished by significant morphological variations. Previous studies have [...] Read more.
Amid global climate change featuring erratic temperature fluctuations, insects adapt via seasonal polyphenism, essential for population sustainability and reproductive success. Cacopsylla chinensis, influenced by environment variations, displays a distinct summer form and winter form distinguished by significant morphological variations. Previous studies have highlighted the role of temperature receptor CcTPRM in orchestrating the transition in response to 10 °C temperature. Nevertheless, the contribution of the transcription factor FoxO in this process has remained ambiguous. Here, we aimed to explore the correlation between C. chinensis FoxO (CcFoxO) and cold stress responses, while identifying potential energetic substances for monitoring physiological shifts during this transition from summer to winter form under cold stress by using RNAi. Initially, CcFoxO emerges as responsive to low temperatures (10 °C) and is regulated by CcTRPM. Subsequent investigations reveal that CcFoxO facilitates the accumulation of triglycerides and glycogen, thereby influencing the transition from summer form to winter form by affecting cuticle pigment content, cuticle chitin levels, and cuticle thickness. Thus, the knockdown of CcFoxO led to high mortality and failed transition. Overall, our findings demonstrate that CcFoxO governs seasonal polyphenism by regulating energy storage. These insights not only enhance our comprehension of FoxO functionality but also offer avenues for environmentally friendly management strategies for C. chinensis. Full article
Show Figures

Figure 1

Figure 1
<p>Sequence analysis of FoxO proteins. (<b>A</b>) Schematic of <span class="html-italic">CcFoxO</span> domains and sequence alignment. <span class="html-italic">CcFoxO</span> contains a forkhead (FH) domain. The <span class="html-italic">CcFoxO</span> (<span class="html-italic">C. chinensis</span>, PP931994) FH domain was aligned with its orthologs from <span class="html-italic">DcFoxO</span> (<span class="html-italic">Diaphorina citri</span>, KAI5702794.1), <span class="html-italic">NlFoxO</span> (<span class="html-italic">Nilaparvata lugens</span>, XP_039275858.1), <span class="html-italic">BmFoxO</span> (<span class="html-italic">Bombyx mori</span>, AFD99125.1), and <span class="html-italic">DmFoxO</span> (<span class="html-italic">Drosophila melanogaster</span>, NP_996205.1). The same gene name and GenBank accession number are as below. (<b>B</b>) Phylogenetic analysis of FoxO orthologs from 12 insect species based on the amino acid sequences. The phylogenetic tree (bootstraps with 1000 replicates) was constructed by MEGA-X using maximum-likelihood methods. The <span class="html-italic">CcFoxO</span> protein was set as an outgroup control. <span class="html-italic">PxFoxO</span>: (<span class="html-italic">Plutella xylostella</span>, XP_037964391.1); <span class="html-italic">GmFoxO</span>: (<span class="html-italic">Grapholita molesta</span>, QIM56595.1); <span class="html-italic">PaFoxO</span>: (<span class="html-italic">Pyrrhocoris apterus</span>, XBC28246.1); <span class="html-italic">ClFoxO</span>: (<span class="html-italic">Cimex lectularius</span>, XP_014254467.1); <span class="html-italic">ArFoxO</span>: (<span class="html-italic">Athalia rosae</span>, XP_048514911.1); <span class="html-italic">TcFoxO</span>: (<span class="html-italic">Tribolium castaneum</span>, XP_975200.2); <span class="html-italic">AgFoxO</span>: (<span class="html-italic">Anoplophora glabripennis</span>, XP_018574886.1). (<b>C</b>) The predicted protein tertiary structure of <span class="html-italic">CcFoxO</span>. The conserved forkhead (FH) domain was indicated in the center of the three-dimensional structure.</p>
Full article ">Figure 2
<p>Expression pattern analysis of <span class="html-italic">CcFoxO</span>. (<b>A</b>) The mRNA expression of <span class="html-italic">CcFoxO</span> in SF first instar nymphs in response to different temperatures at day 3, 6 and 10 of 25 °C and 10 °C by qRT-PCR (<span class="html-italic">n</span> = 3). Expression level of <span class="html-italic">CcFoxO</span> at day 3 under 25 °C is used as 1. (<b>B</b>) RNAi efficiency of dsCcTRPM after treatment at 48 h compared to dsEGPF feeding (<span class="html-italic">n</span> = 3). Expression level of <span class="html-italic">CcTRPM</span> under dsCcTRPM treatment is used as 1. (<b>C</b>) Effect of <span class="html-italic">CcTRPM</span> knockdown on the mRNA expression of <span class="html-italic">CcFoxO</span> (<span class="html-italic">n</span> = 3). Expression level of <span class="html-italic">CcFoxO</span> under dsCcTRPM treatment is used as 1. Newly hatched summer-form nymphs were worked on as the experimental subjects. The results are indicated as the mean ± Standard Deviation (SD) and obtained from three independent biological replicates, with per replicate consisting of at least 30 nymphs. Statistical significance was determined using the pairwise Student’s <span class="html-italic">t</span>-test, and significance levels are shown as ** (<span class="html-italic">p</span> &lt; 0.01) or *** (<span class="html-italic">p</span> &lt; 0.001).</p>
Full article ">Figure 3
<p>Effect of <span class="html-italic">CcFoxO</span> knockdown on the abundance of triglycerides, glycogen, and lipid droplet. (<b>A</b>,<b>B</b>) Comparison of the nymph triglycerides and glycogen content after treatment with dsEGFP and dsCcFoxO at 48 h (<span class="html-italic">n</span> = 9). (<b>C</b>) Lipid droplets stained with Nile red in fat bodies of SF first instar nymphs after being treated with dsEGFP and dsCcFoxO at 48 h. Representative images were captured, where lipid droplets are shown in red. The scale bar represents 50 μm. Statistical significance was determined using the pairwise Student’s <span class="html-italic">t</span>-test, and significance levels are shown as *** (<span class="html-italic">p</span> &lt; 0.001).</p>
Full article ">Figure 4
<p>Comparison of the total cuticle pigment content (<b>A</b>), cuticle chitin content (<b>B</b>), and cuticle thickness (<b>C</b>,<b>D</b>) of SF first instar nymphs treated with dsCcFoxO and dsEGFP at 15 days. Data in (<b>B</b>,<b>C</b>) were presented as means ± SD with nine independent biological replicates, and each circle indicated one biological replicate. Scale bar in (<b>D</b>) is 1 μm. Student’s <span class="html-italic">t</span>-test of the pairwise was used to do the statistically analysis (*** <span class="html-italic">p</span> &lt; 0.001).</p>
Full article ">Figure 5
<p>Effect of <span class="html-italic">CcFoxO</span> knockdown on the transition from summer form to winter form and mortality. (<b>A</b>,<b>B</b>) The transition percent and mortality of SF first instar nymphs treated with dsEGFP or dsCcFoxO at 15 days under 10 °C (<span class="html-italic">n</span> = 9). (<b>C</b>) The phenotypes of SF first instar nymphs treated with dsEGFP and dsFoxO at 15 day under 10 °C. Scale bar is 0.5 mm. Data in (<b>A</b>,<b>B</b>) are presented as mean ± SD with nine biological replications. Statistically significant differences were determined with the pair-wise Student’s <span class="html-italic">t</span>-test, and significance levels were denoted by *** (<span class="html-italic">p</span> &lt; 0.001).</p>
Full article ">Figure 6
<p>A model showing the key roles of transcription factor <span class="html-italic">CcFoxO</span> in the transition from summer form to winter form. Under 10 °C condition, low temperatures induced the activation of the temperature receptor <span class="html-italic">CcTRPM</span> in SF first nymphs. Then, <span class="html-italic">CcTRPM</span> significantly increased the expression of transcription factor <span class="html-italic">CcFoxO</span>, promoting the accumulation of energy substances, rising cuticle chitin biosynthesis, as well as cuticle thickness. As a result, the first instar nymphs of SF developed to third instar nymphs of WF in order to better adapt to low temperatures.</p>
Full article ">
14 pages, 603 KiB  
Review
The Role and Function of TRPM8 in the Digestive System
by Zunan Wu, Shuai Peng, Wensha Huang, Yuling Zhang, Yashi Liu, Xiaoyun Yu and Lei Shen
Biomolecules 2024, 14(7), 877; https://doi.org/10.3390/biom14070877 - 21 Jul 2024
Viewed by 798
Abstract
Transient receptor potential (TRP) melastatin member 8 (TRPM8) is a non-selective cation channel that can be activated by low temperatures (8–26 °C), cooling agents (including menthol analogs such as menthol, icilin, and WS-12), voltage, and extracellular osmotic pressure changes. TRPM8 expression has been [...] Read more.
Transient receptor potential (TRP) melastatin member 8 (TRPM8) is a non-selective cation channel that can be activated by low temperatures (8–26 °C), cooling agents (including menthol analogs such as menthol, icilin, and WS-12), voltage, and extracellular osmotic pressure changes. TRPM8 expression has been identified in the digestive system by several research teams, demonstrating its significant involvement in tissue function and pathologies of the digestive system. Specifically, studies have implicated TRPM8 in various physiological and pathological processes of the esophagus, stomach, colorectal region, liver, and pancreas. This paper aims to comprehensively outline the distinct role of TRPM8 in different organs of the digestive system, offering insights for future mechanistic investigations of TRPM8. Additionally, it presents potential therapeutic targets for treating conditions such as digestive tract inflammation, tumors, sensory and functional disorders, and other related diseases. Furthermore, this paper addresses the limitations of existing studies and highlights the research prospects associated with TRPM8. Full article
Show Figures

Figure 1

Figure 1
<p>TRPM8 plays an anti-inflammatory role in colitis. Its mechanisms are as follows: (a) TRPM8 disrupts the PKAca and GSK-3β interaction, thereby impeding the SP apoptotic impact on colonic epithelial cells and mitigating colonic inflammation. (b) TRPM8 upregulates CGRP to suppress the production of pro-inflammatory cytokines by innate immune cells. (c) TRPM8 inhibits TRPV1 and attenuates the pro-inflammatory action of CGRP. (d) TRPM8 modulates the expression of IL-10 and TNF-α to reduce colonic inflammation. The blue lines represent “promote”, and the red lines represent “inhibit” in <a href="#biomolecules-14-00877-f001" class="html-fig">Figure 1</a>.</p>
Full article ">
20 pages, 1712 KiB  
Review
Mechanisms and Functions of Sweet Reception in Oral and Extraoral Organs
by Ryusuke Yoshida and Yuzo Ninomiya
Int. J. Mol. Sci. 2024, 25(13), 7398; https://doi.org/10.3390/ijms25137398 - 5 Jul 2024
Viewed by 683
Abstract
The oral detection of sugars relies on two types of receptor systems. The first is the G-protein-coupled receptor TAS1R2/TAS1R3. When activated, this receptor triggers a downstream signaling cascade involving gustducin, phospholipase Cβ2 (PLCβ2), and transient receptor potential channel M5 (TRPM5). The second type [...] Read more.
The oral detection of sugars relies on two types of receptor systems. The first is the G-protein-coupled receptor TAS1R2/TAS1R3. When activated, this receptor triggers a downstream signaling cascade involving gustducin, phospholipase Cβ2 (PLCβ2), and transient receptor potential channel M5 (TRPM5). The second type of receptor is the glucose transporter. When glucose enters the cell via this transporter, it is metabolized to produce ATP. This ATP inhibits the opening of KATP channels, leading to cell depolarization. Beside these receptor systems, sweet-sensitive taste cells have mechanisms to regulate their sensitivity to sweet substances based on internal and external states of the body. Sweet taste receptors are not limited to the oral cavity; they are also present in extraoral organs such as the gastrointestinal tract, pancreas, and brain. These extraoral sweet receptors are involved in various functions, including glucose absorption, insulin release, sugar preference, and food intake, contributing to the maintenance of energy homeostasis. Additionally, sweet receptors may have unique roles in certain organs like the trachea and bone. This review summarizes past and recent studies on sweet receptor systems, exploring the molecular mechanisms and physiological functions of sweet (sugar) detection in both oral and extraoral organs. Full article
(This article belongs to the Special Issue Molecular Mechanisms Subserving Taste and Olfaction Systems)
Show Figures

Figure 1

Figure 1
<p>Schematic diagram showing molecular mechanisms for sweet detection via TAS1Rs in taste cells. The binding of sweeteners to TAS1R2/TAS1R3 activates a trimeric G-protein (Gα-gustducin, Gβ1 or Gβ3, and Gγ13) and phospholipase Cβ2 (PLCβ2). Then, inositol-1,4,5-trisphosphate (IP<sub>3</sub>) is produced, and [Ca<sup>2+</sup>]<sub>i</sub> is increased by Ca<sup>2+</sup> release from Ca<sup>2+</sup> store. Ca<sup>2+</sup> activates the transient receptor potential channel M5 (TRPM5), leading to cell depolarization and the firing of action potentials (APs) via voltage-gated sodium channels (VGSCs). Then, ATP-permeable CALHM1/3 opens to secrete ATP from the taste cell.</p>
Full article ">Figure 2
<p>Schematic diagram showing molecular mechanisms for sweet detection via glucose transporters in taste cells. Glucose entering via glucose transporters (GLUTs) and/or sodium–glucose transporters (SGLTs) is metabolized to produce ATP. The activity of K<sub>ATP</sub> channels is inhibited (indicated by X) by an increase in [ATP]<sub>i</sub> (indicated by red up arrow), leading to cell depolarization. Na<sup>+</sup> entry through SGLTs also induces cell depolarization. Such depolarization activates voltage-gated channels such as voltage-gated sodium channels (VGSCs).</p>
Full article ">Figure 3
<p>Schematic diagram showing molecular mechanisms for leptin signaling in sweet-sensitive taste cells. Leptin binding to the leptin receptor (Ob-Rb) induces the activation of phosphoinositide 3-kinase (PI3K), leading to the production of phosphatidylinositol (3,4,5)-trisphosphate (PIP<sub>3</sub>) and phosphorylation of AKT. These signaling components might activate K<sub>ATP</sub> channels to induce cell hyperpolarization, leading to the suppression of sweet responses (indicated by blue down arrow).</p>
Full article ">Figure 4
<p>Schematic diagram showing possible molecular mechanisms for cannabinoid signaling in sweet-sensitive taste cells. Endocannabinoids binding to the cannabinoid receptor (CB<sub>1</sub>) activate phospholipase Cβ2 (PLCβ2) via the Gαq pathway. The synergistic or additive activation of PLCβ2 results in enhanced sweet responses in sweet-sensitive taste cells (indicated by red up arrow). Alternatively, the activation of CB<sub>1</sub> suppresses adenylyl cyclase (AC) activity via the Gαi/o pathway, leading to a decrease in [cAMP]<sub>i</sub> (indicated by blue down arrow). The reduction in cAMP levels decreases the activity of protein kinase A (PKA), leading to the disinhibition of PLC signaling (indicated by X) and enhanced sweet responses in sweet-sensitive taste cells.</p>
Full article ">
18 pages, 4475 KiB  
Article
In Silico Electrophysiological Investigation of Transient Receptor Potential Melastatin-4 Ion Channel Biophysics to Study Detrusor Overactivity
by Chitaranjan Mahapatra and Ravindra Thakkar
Int. J. Mol. Sci. 2024, 25(13), 6875; https://doi.org/10.3390/ijms25136875 - 22 Jun 2024
Viewed by 643
Abstract
Enhanced electrical activity in detrusor smooth muscle (DSM) cells is a key factor in detrusor overactivity which causes overactive bladder pathological disorders. Transient receptor potential melastatin-4 (TRPM4) channels, which are calcium-activated cation channels, play a role in regulating DSM electrical activities. These channels [...] Read more.
Enhanced electrical activity in detrusor smooth muscle (DSM) cells is a key factor in detrusor overactivity which causes overactive bladder pathological disorders. Transient receptor potential melastatin-4 (TRPM4) channels, which are calcium-activated cation channels, play a role in regulating DSM electrical activities. These channels likely contribute to depolarizing the DSM cell membrane, leading to bladder overactivity. Our research focuses on understanding TRPM4 channel function in the DSM cells of mice, using computational modeling. We aimed to create a detailed computational model of the TRPM4 channel based on existing electrophysiological data. We employed a modified Hodgkin-Huxley model with an incorporated TRP-like current to simulate action potential firing in response to current and synaptic stimulus inputs. Validation against experimental data showed close agreement with our simulations. Our model is the first to analyze the TRPM4 channel’s role in DSM electrical activity, potentially revealing insights into bladder overactivity. In conclusion, TRPM4 channels are pivotal in regulating human DSM function, and TRPM4 channel inhibitors could be promising targets for treating overactive bladder. Full article
(This article belongs to the Special Issue TRP Channels in Physiology and Pathophysiology 2.0)
Show Figures

Figure 1

Figure 1
<p>Schematic illustration of the proposed physiological role for TRPM4 channels in DSM cells. According to the postulated mechanism, the TRPM4 channels via sarcoplasmic reticulum Ca<sup>2+</sup>-dependent activation participate in a positive feedback loop to maximize DSM contractility by providing Na<sup>+</sup>-depolarizing conductance.</p>
Full article ">Figure 2
<p>Simulated steady-state activation curve showing log (Cai). The solid line represents the result from simulation, where the solid filled triangle shows the adapted experimental data from Demion et al., 2007 [<a href="#B89-ijms-25-06875" class="html-bibr">89</a>].</p>
Full article ">Figure 3
<p>DSM model showing initial fluctuation (<b>a</b>) and constant resting membrane potential maintained at −52 mV (<b>b</b>).</p>
Full article ">Figure 4
<p>(<b>a</b>) The model generated AP (solid red line) and depolarization (dashed red line) with the current stimulus. (<b>b</b>) The model generated AP (solid red line), experimental AP (solid blue line), and simulated depolarization (dashed red line) with synaptic input stimulus.</p>
Full article ">Figure 5
<p>DSM action potential after a 10% (solid red line) and 20% (solid black line) increase in the maximum conductance of the TRPM4 ion channel.</p>
Full article ">Figure 6
<p>Sensitivity analysis of the TRPM4 channel conductance for DSM resting membrane potential and action potential duration.</p>
Full article ">Figure 7
<p>DSM cell generates action potential (solid red line) with mutation of T-type Ca<sup>2+</sup> channel. The dashed line depicts the absence of action potential.</p>
Full article ">Figure 8
<p>Schematic overview of parallel conductance model for ionic current. Further elucidation is provided in the subsequent paragraph.</p>
Full article ">Figure 9
<p>Schematic diagram of 10-state Markov model for BK channel. A detailed explanation is provided in the text.</p>
Full article ">Figure 10
<p>Schematic diagram illustrating all ionic components within a DSM cell. The accompanying paragraph provides descriptions for each component.</p>
Full article ">
18 pages, 325 KiB  
Review
Latest Insights into the In Vivo Studies in Murine Regarding the Role of TRP Channels in Wound Healing—A Review
by Alexandra Grigore, Oana Andreia Coman, Horia Păunescu, Mihnea Costescu and Ion Fulga
Int. J. Mol. Sci. 2024, 25(12), 6753; https://doi.org/10.3390/ijms25126753 - 19 Jun 2024
Cited by 1 | Viewed by 765
Abstract
Wound healing involves physical, chemical and immunological processes. Transient receptor potential (TRP) and other ion channels are implicated in epidermal re-epithelization. Ion movement across ion channels can induce transmembrane potential that leads to transepithelial potential (TEP) changes. TEP is present in epidermis surrounding [...] Read more.
Wound healing involves physical, chemical and immunological processes. Transient receptor potential (TRP) and other ion channels are implicated in epidermal re-epithelization. Ion movement across ion channels can induce transmembrane potential that leads to transepithelial potential (TEP) changes. TEP is present in epidermis surrounding the lesion decreases and induces an endogenous direct current generating an epithelial electric field (EF) that could be implicated in wound re-epithelialization. TRP channels are involved in the activation of immune cells during mainly the inflammatory phase of wound healing. The aim of the study was to review the mechanisms of ion channel involvement in wound healing in in vivo experiments in murine (mice, rats) and how can this process be influenced. This review used the latest results published in scientific journals over the last year and this year to date (1 January 2023–31 December 3000) in order to include the in-press articles. Some types of TRP channels, such as TRPV1, TRPV3 and TRPA1, are expressed in immune cells and can be activated by inflammatory mediators. The most beneficial effects in wound healing are produced using agonists of TRPV1, TRPV4 and TRPA1 channels or by inhibiting with antagonists, antisense oligonucleotides or knocking down TRPV3 and TRPM8 channels. Full article
19 pages, 3533 KiB  
Article
Chemical Composition and Neuroprotective Properties of Indonesian Stingless Bee (Geniotrigona thoracica) Propolis Extract in an In-Vivo Model of Intracerebral Hemorrhage (ICH)
by Steven Tandean, Iskandar Japardi, Muhammad Rusda, Rr Suzy Indharty, Aznan Lelo, Renindra Ananda Aman, Mustafa Mahmud Amin, Andre Marolop Pangihutan Siahaan, Putri Chairani Eyanoer, Celine Augla D’Prinzessin, Ronny Lesmana, Milena Popova, Boryana Trusheva, Vassya Bankova and Felix Zulhendri
Nutrients 2024, 16(12), 1880; https://doi.org/10.3390/nu16121880 - 14 Jun 2024
Viewed by 1087
Abstract
Stroke is the world’s second-leading cause of death. Current treatments for cerebral edema following intracerebral hemorrhage (ICH) mainly involve hyperosmolar fluids, but this approach is often inadequate. Propolis, known for its various beneficial properties, especially antioxidant and anti-inflammatory properties, could potentially act as [...] Read more.
Stroke is the world’s second-leading cause of death. Current treatments for cerebral edema following intracerebral hemorrhage (ICH) mainly involve hyperosmolar fluids, but this approach is often inadequate. Propolis, known for its various beneficial properties, especially antioxidant and anti-inflammatory properties, could potentially act as an adjunctive therapy and help alleviate stroke-associated injuries. The chemical composition of Geniotrigona thoracica propolis extract was analyzed by GC-MS after derivatization for its total phenolic and total flavonoid content. The total phenolic content and total flavonoid content of the propolis extract were 1037.31 ± 24.10 μg GAE/mL and 374.02 ± 3.36 μg QE/mL, respectively. By GC-MS analysis, its major constituents were found to be triterpenoids (22.4% of TIC). Minor compounds, such as phenolic lipids (6.7% of TIC, GC-MS) and diterpenic acids (2.3% of TIC, GC-MS), were also found. Ninety-six Sprague Dawley rats were divided into six groups; namely, the control group, the ICH group, and four ICH groups that received the following therapies: mannitol, propolis extract (daily oral propolis administration after the ICH induction), propolis-M (propolis and mannitol), and propolis-B+A (daily oral propolis administration 7 days prior to and 72 h after the ICH induction). Neurocognitive functions of the rats were analyzed using the rotarod challenge and Morris water maze. In addition, the expression of NF-κB, SUR1-TRPM4, MMP-9, and Aquaporin-4 was analyzed using immunohistochemical methods. A TUNEL assay was used to assess the percentage of apoptotic cells. Mannitol significantly improved cognitive–motor functions in the ICH group, evidenced by improved rotarod and Morris water maze completion times, and lowered SUR-1 and Aquaporin-4 levels. It also significantly decreased cerebral edema by day 3. Similarly, propolis treatments (propolis-A and propolis-B+A) showed comparable improvements in these tests and reduced edema. Moreover, combining propolis with mannitol (propolis-M) further enhanced these effects, particularly in reducing edema and the Virchow-Robin space. These findings highlight the potential of propolis from the Indonesian stingless bee, Geniotrigona thoracica, from the Central Tapanuli region as a neuroprotective, adjunctive therapy. Full article
(This article belongs to the Special Issue Bee Products in Human Health—2nd Edition)
Show Figures

Figure 1

Figure 1
<p>Completion time (in s) for the Rotarod challenge on days 3, 7, and 14. Different letters indicate a statistically significant difference (<span class="html-italic">p</span> &lt; 0.05). For example, groups labeled with ‘a’ and ‘b’ are significantly different from each other, while groups labeled with ‘ab’ and ‘b’ are not significantly different from each other.</p>
Full article ">Figure 2
<p>Morris water maze completion time (s) on days 3, 7, and 14. Different letters indicate a statistically significant difference (<span class="html-italic">p</span> &lt; 0.05). For example, groups labeled with ‘a’ and ‘b’ are significantly different from each other, while groups labeled with ‘ab’ and ‘b’ are not significantly different from each other.</p>
Full article ">Figure 3
<p>Percentage of cells expressing SUR-1 on days 3, 7, and 14. Different letters indicate a statistically significant difference (<span class="html-italic">p</span> &lt; 0.05). For example, groups labeled with ‘a’ and ‘b’ are significantly different from each other, while groups labeled with ‘ab’ and ‘b’ are not significantly different from each other.</p>
Full article ">Figure 4
<p>Percentage of cells expressing NF-κB on days 3, 7, and 14. Different letters indicate a statistically significant difference (<span class="html-italic">p</span> &lt; 0.05). For example, groups labeled with ‘a’ and ‘b’ are significantly different from each other, while groups labeled with ‘ab’ and ‘b’ are not significantly different from each other.</p>
Full article ">Figure 5
<p>Percentage of cells expressing Aquaporin-4 on days 3, 7, and 14. Different letters indicate a statistically significant difference (<span class="html-italic">p</span> &lt; 0.05). For example, groups labeled with ‘a’ and ‘b’ are significantly different from each other, while groups labeled with ‘ab’ and ‘b’ are not significantly different from each other.</p>
Full article ">Figure 6
<p>Percentage of cells expressing MMP-9 on days 3, 7, and 14. Different letters indicate a statistically. For example, groups labeled with ‘a’ and ‘b’ are significantly different from each other, while groups labeled with ‘ab’ and ‘b’ are not significantly different from each other.</p>
Full article ">Figure 7
<p>Edema index (%). Different letters indicate a statistically significant difference (<span class="html-italic">p</span> &lt; 0.05). For example, groups labeled with ‘a’ and ‘b’ are significantly different from each other, while groups labeled with ‘ab’ and ‘b’ are not significantly different from each other.</p>
Full article ">Figure 8
<p>Virchow-Robin space (μm). Different letters indicate a statistically significant difference (<span class="html-italic">p</span> &lt; 0.05). For example, groups labeled with ‘a’ and ‘b’ are significantly different from each other, while groups labeled with ‘ab’ and ‘b’ are not significantly different from each other.</p>
Full article ">Figure 9
<p>Percentage of apoptotic cells based on the TUNEL assay. Different letters indicate a statistically significant difference (<span class="html-italic">p</span> &lt; 0.05). For example, groups labeled with ‘a’ and ‘b’ are significantly different from each other, while groups labeled with ‘ab’ and ‘b’ are not significantly different from each other.</p>
Full article ">
18 pages, 2495 KiB  
Article
Signal Transduction of Transient Receptor Potential TRPM8 Channels: Role of PIP5K, Gq-Proteins, and c-Jun
by Gerald Thiel and Oliver G. Rössler
Molecules 2024, 29(11), 2602; https://doi.org/10.3390/molecules29112602 - 1 Jun 2024
Viewed by 785
Abstract
Transient receptor potential melastatin-8 (TRPM8) is a cation channel that is activated by cold and “cooling agents” such as menthol and icilin, which induce a cold sensation. The stimulation of TRPM8 activates an intracellular signaling cascade that ultimately leads to a change in [...] Read more.
Transient receptor potential melastatin-8 (TRPM8) is a cation channel that is activated by cold and “cooling agents” such as menthol and icilin, which induce a cold sensation. The stimulation of TRPM8 activates an intracellular signaling cascade that ultimately leads to a change in the gene expression pattern of the cells. Here, we investigate the TRPM8-induced signaling pathway that links TRPM8 channel activation to gene transcription. Using a pharmacological approach, we show that the inhibition of phosphatidylinositol 4-phosphate 5 kinase α (PIP5K), an enzyme essential for the biosynthesis of phosphatidylinositol 4,5-bisphosphate, attenuates TRPM8-induced gene transcription. Analyzing the link between TRPM8 and Gq proteins, we show that the pharmacological inhibition of the βγ subunits impairs TRPM8 signaling. In addition, genetic studies show that TRPM8 requires an activated Gα subunit for signaling. In the nucleus, the TRPM8-induced signaling cascade triggers the activation of the transcription factor AP-1, a complex consisting of a dimer of basic region leucine zipper (bZIP) transcription factors. Here, we identify the bZIP protein c-Jun as an essential component of AP-1 within the TRPM8-induced signaling cascade. In summary, with PIP5K, Gq subunits, and c-Jun, we identified key molecules in TRPM8-induced signaling from the plasma membrane to the nucleus. Full article
(This article belongs to the Section Bioorganic Chemistry)
Show Figures

Graphical abstract

Graphical abstract
Full article ">Figure 1
<p>Biosynthesis, metabolism and hydrolysis of phosphatidylinositol 4,5-bisphosphate. Phosphoinositides are phosphorylated metabolites of phosphatidylinositol. Catalyzed by PIP5K, phosphatidylinositol 4,5-bisphosphate (PI(4,5)P<sub>2</sub>) is synthesisized from phosphatidylinositol 4-phosphate (PI 4-phosphate). Phosphatidylinositol 4,5-bisphosphate can be phosphorylated by phosphatidylinositol 3-kinases to generate phosphatidylinositol 3,4,5-trisphosphate (PI(3,4,5)P<sub>3</sub>), a lipid mediator essential to activate the serin/threonine protein kinase AKT. Phosphatidylinositol 4,5-bisphosphate can also be metabolized by phospholipase C (PLC), which generates IP<sub>3</sub> and diacylgycerol.</p>
Full article ">Figure 2
<p>Pharmacological inhibition of PIP5Kα attenuates TRPM8 intracellular signaling. (<b>a</b>) Modular structure of TRPM8. The interaction sites of phosphatidylinositol 4,5-bisphosphate with the channel are shown (preS1 segment, S4-S5 linker, TRP domain). (<b>b</b>) Chemical structure of the PIP5Kα inhibitor ISA-2011B. (<b>c</b>) Luciferase reporter gene under control of the collagenase promoter (Coll.luc), which serves as a sensor for measuring AP-1 activity. Shown is the chromatin-embedded provirus, which also contains the Woodchuck hepatitis virus posttranscriptional regulatory element (WPRE) and the HIV flap element. The U3 region of the 5′ LTR has been deleted. (<b>d</b>) HEK293-M8 cells were infected with a recombinant lentivirus containing the Coll.luc reporter gene. Cells were serum-starved for 24 h, preincubated for 3 h with ISA-2011B (10 μM), and then stimulated with icilin (1 μM) in serum-reduced medium in the presence of the PIP5Kα inhibitor. Cell extracts were prepared, and luciferase activities and protein concentrations were determined. The luciferase activity was normalized to the protein concentration. Data shown are means +/− SD of three experiments performed in quadruplicate (*** <span class="html-italic">p</span> &lt; 0.001).</p>
Full article ">Figure 3
<p>Pharmacological inhibition of PIP5Kα attenuates intracellular signaling following stimulation of TRPM3 and Ca<sub>v</sub>1.2 Ca<sup>2+</sup> channels. (<b>a</b>) Modular structure of TRPM3 showing the interaction sites of phosphatidylinositol 4,5-bisphosphate with the channel (preS1 segment, S4-S5 linker, and TRP domain). (<b>b</b>) T-REx-TRPM3 cells containing a Coll.luc reporter gene integrated into the chromatin were serum-starved for 24 h in the presence of tetracycline (1 μg/mL) to induce TRPM3 expression. The serum-starved cells were preincubated with ISA-2011B (10 μM) for 3 h, and then stimulated with pregnenolone sulfate (20 μM) for 24 h in the presence of the inhibitor. Cells were harvested and analyzed as described in the legend to <a href="#molecules-29-02602-f002" class="html-fig">Figure 2</a> (n = 4; *** <span class="html-italic">p</span> &lt; 0.001). (<b>c</b>) Modular structure of Ca<sub>v</sub>1.2 voltage-gated Ca<sup>2+</sup> channels, consisting of the α1 subunit, which forms the pore, and the auxiliary subunits α2δ, β, and γ. (<b>d</b>) INS-1 832/13 insulinoma cells were infected with a recombinant lentivirus containing the Coll.luc reporter gene. Cells were serum-starved in medium containing 0.5% serum and 2 mM glucose for 24 h. Cells were preincubated in the same medium with ISA-2011B (10 μM) for three hours. Stimulation of the cells was performed with KCl (25 mM) and the voltage-gated Ca<sup>2+</sup> channel activator FPL64176 (2.5 μM) in the presence of the inhibitor for 24 h. Cells were harvested and analyzed as described in the legend to <a href="#molecules-29-02602-f002" class="html-fig">Figure 2</a> (n = 3; *** <span class="html-italic">p</span> &lt; 0.001).</p>
Full article ">Figure 4
<p>RGS2 expression reduces signal transduction after stimulation of TRPM8 channels or Gαq-coupled designer receptors. (<b>a</b>) Modular structure of RGS2. The RGS domain is responsible for binding to Gαq, while the N-terminal domain is required for targeting the protein to the plasma membrane. (<b>b</b>) HEK293-M8 cells containing the Coll.luc reporter gene were infected with a recombinant lentivirus encoding either RGS2 or β-galactosidase (mock). Cells were serum-starved for 24 h, and then stimulated with icilin (1 μM) in serum-reduced medium for 24 h. Cells were harvested and analyzed as described in the legend to <a href="#molecules-29-02602-f002" class="html-fig">Figure 2</a> (n = 3, *** <span class="html-italic">p</span> &lt; 0.001). (<b>c</b>) HEK293 cells were infected with a lentivirus encoding the Gαq-coupled designer receptor Rαq. Cells were additionally infected with a lentivirus containing the Coll.luc reporter gene. Furthermore, the cells were infected with a lentivirus encoding either RGS2 or β-galactosidase (mock). We incubated the cells in medium containing 0.05% serum for 24 h. Stimulation of the cells was performed with CNO (1 μM) for 24 h in serum-reduced medium. Cells were harvested and analyzed as described in the legend to <a href="#molecules-29-02602-f002" class="html-fig">Figure 2</a> (n = 5, *** <span class="html-italic">p</span> &lt; 0.001). (<b>d</b>) T-REx-TRPM3 cells containing a chromatin-integrated Coll.luc reporter gene were serum-starved for 24 h in the presence of tetracycline (1 μg/mL). The serum-starved cells were infected with a lentivirus encoding either RGS2 or β-galactosidase (mock). Cells were stimulated with pregnenolone sulfate (20 μM) for 24 h. Cells were harvested and analyzed as described in the legend to <a href="#molecules-29-02602-f002" class="html-fig">Figure 2</a> (n = 3; n.s., not significant). (<b>e</b>) INS-1 832/13 insulinoma cells were infected with a lentivirus containing the reporter gene Coll.luc. Additionally, we infected the cells with a lentivirus encoding either RGS2 or β-galactosidase (mock). Cells were incubated in medium containing 0.5% serum and 2 mM glucose for 24 h, and then stimulated with KCl (25 mM) and the FPL64176 (2.5 μM) for 24 h. Cells were harvested and analyzed as described in the legend to <a href="#molecules-29-02602-f002" class="html-fig">Figure 2</a> (n = 3; n.s., not significant).</p>
Full article ">Figure 5
<p>The Gβγ inhibitor gallein attenuates signal transduction mediated by TRPM8 and TRPM3 channels. (<b>a</b>) Chemical structure of gallein. (<b>b</b>) T-REx-TRPM3 cells containing a chromatin-embedded Coll.luc reporter gene were serum-starved for 24 h in the presence of tetracycline (1 μg/mL) to induced TRPM3 expression. Serum-starved cells were preincubated with gallein (20 μM) for three hours, and then stimulated with pregnenolone sulfate (20 μM) for 24 h in the presence of the compound. Cells were harvested and analyzed as described in the legend to <a href="#molecules-29-02602-f002" class="html-fig">Figure 2</a> (n = 3; *** <span class="html-italic">p</span> &lt; 0.001). (<b>c</b>) HEK293-M8 cells containing the Coll.luc reporter gene integrated into the chromatin were serum-starved for 24 h, preincubated for 3 h with gallein (20 μM), and then stimulated with icilin (1 μM) in serum-reduced medium in the presence of gallein. Cells were harvested and analyzed as described in the legend to <a href="#molecules-29-02602-f002" class="html-fig">Figure 2</a> (n = 3; *** <span class="html-italic">p</span> &lt; 0.001).</p>
Full article ">Figure 6
<p>The transcription factor c-Jun is required to link the stimulation of TRPM8 with the activation of AP-1. (<b>a</b>) Modular structure of the transcription factor c-Jun and its dominant-negative c-Jun mutant c-JunΔN. c-JunΔN contains the C-terminal amino acids 188 to 331 of c-Jun, which comprise the bZIP domain. The mutant lacks the transcriptional activation domain. (<b>b</b>) Expression of c-JunΔN attenuates icilin-induced activation of AP-1 in HEK293-M8 cells. Cells were infected with a lentivirus containing a Coll.luc reporter gene. Cells were additionally infected with a lentivirus encoding either c-JunΔN or β-galactosidase (mock). Serum-starved cells were stimulated with icilin (1 μM) for 24 h. Cells were harvested and analyzed as described in the legend to <a href="#molecules-29-02602-f002" class="html-fig">Figure 2</a> (n = 3; *** <span class="html-italic">p</span> &lt; 0.001).</p>
Full article ">Figure 7
<p>Signal transduction of TRPM8 channels. The TRPM8 channel is a tetramer embedded into the plasma membrane. The channel interacts with phosphatidylinositol 4,5-bisphosphate, which is essential for its activation. Reducing phosphatidylinositol 4,5-bisphosphate levels by blocking its biosynthesis with a PIP5K inhibitor impairs TRPM8 activation. PIP5K is, therefore, an important regulator of TRPM8 signaling. The stimulation of Gαq-coupled receptors triggers the dissociation of the trimeric G proteins into a GTP-bound α-subunit and the βγ subunits. The Gα subunit interacts with TRPM8 channels and is essential for TRPM8 activation. The α-subunit and the βγ subunits bind to PLCβ and activate the enzyme by increasing <span class="html-italic">k<sub>cat</sub></span> and changing the orientation of the enzyme in the membrane to its substrate. This conformational change may remove the blockade of PLCβ to TRPM8 channels. After the activation of TRPM8, Ca<sup>2+</sup> ions flow into the cells through the channel and trigger the activation of ERK1/2, which acts as signal transducer. The kinase translocates into the cell nucleus and activates AP-1, which is composed of the bZIP proteins c-Jun and c-Fos.</p>
Full article ">
19 pages, 3969 KiB  
Article
Roles of Thermosensitive Transient Receptor Channels TRPV1 and TRPM8 in Paclitaxel-Induced Peripheral Neuropathic Pain
by Wen-Wen Li, Yan Zhao, Huai-Cun Liu, Jiao Liu, Sun-On Chan, Yi-Fei Zhong, Tang-Yu Zhang, Yu Liu, Wei Zhang, Yu-Qi Xia, Xiao-Chun Chi, Jian Xu, Yun Wang and Jun Wang
Int. J. Mol. Sci. 2024, 25(11), 5813; https://doi.org/10.3390/ijms25115813 - 27 May 2024
Cited by 1 | Viewed by 1488
Abstract
Paclitaxel, a microtubule-stabilizing chemotherapy drug, can cause severe paclitaxel-induced peripheral neuropathic pain (PIPNP). The roles of transient receptor potential (TRP) ion channel vanilloid 1 (TRPV1, a nociceptor and heat sensor) and melastatin 8 (TRPM8, a cold sensor) in PIPNP remain controversial. In this [...] Read more.
Paclitaxel, a microtubule-stabilizing chemotherapy drug, can cause severe paclitaxel-induced peripheral neuropathic pain (PIPNP). The roles of transient receptor potential (TRP) ion channel vanilloid 1 (TRPV1, a nociceptor and heat sensor) and melastatin 8 (TRPM8, a cold sensor) in PIPNP remain controversial. In this study, Western blotting, immunofluorescence staining, and calcium imaging revealed that the expression and functional activity of TRPV1 were upregulated in rat dorsal root ganglion (DRG) neurons in PIPNP. Behavioral assessments using the von Frey and brush tests demonstrated that mechanical hyperalgesia in PIPNP was significantly inhibited by intraperitoneal or intrathecal administration of the TRPV1 antagonist capsazepine, indicating that TRPV1 played a key role in PIPNP. Conversely, the expression of TRPM8 protein decreased and its channel activity was reduced in DRG neurons. Furthermore, activation of TRPM8 via topical application of menthol or intrathecal injection of WS-12 attenuated the mechanical pain. Mechanistically, the TRPV1 activity triggered by capsaicin (a TRPV1 agonist) was reduced after menthol application in cultured DRG neurons, especially in the paclitaxel-treated group. These findings showed that upregulation of TRPV1 and inhibition of TRPM8 are involved in the generation of PIPNP, and they suggested that inhibition of TRPV1 function in DRG neurons via activation of TRPM8 might underlie the analgesic effects of menthol. Full article
(This article belongs to the Special Issue New Drugs Regulating Cytoskeletons in Human Health and Diseases)
Show Figures

Figure 1

Figure 1
<p>Mechanical hypersensitivity in a rat model of paclitaxel-induced peripheral neuropathic pain (PIPNP). (<b>A</b>) Experimental protocol for establishing a rat model of PIPNP. Red arrows indicate the administration time of paclitaxel (PTX). (<b>B</b>) Mechanical hypersensitivity indicated by von Frey test in PIPNP. Left: Time course of paw withdrawal threshold (PWT) after PTX or vehicle (Veh) injection. Right: Statistical analysis of the area under the curve (AUC). (<b>C</b>) Percentage of rats with obvious mechanical pain on day 14 after paclitaxel treatment (PWT &lt; 10 g, <span class="html-italic">n</span> = 48). (<b>D</b>) Time course effect and AUC analysis of dynamic score indicated by brush test after PTX or Veh treatment. (<b>E</b>) Time course effect and AUC analysis of paw withdrawal latency (PWL) indicated by Hargreaves test after PTX or Veh injection. (<b>F</b>) Time course effect of cold score indicated by acetone test and its AUC analysis. The data were analyzed using a two-way ANOVA followed by Bonferroni post hoc tests for the time course effect. The AUCs were analyzed using an unpaired <span class="html-italic">t</span>-test. * <span class="html-italic">p</span> &lt; 0.05, ** <span class="html-italic">p</span> &lt; 0.01, *** <span class="html-italic">p</span> &lt; 0.001, <span class="html-italic">n</span> = 6 in (<b>B</b>,<b>D</b>,<b>E</b>) <span class="html-italic">n</span> = 5 in (<b>F</b>). All data are means ± SEM.</p>
Full article ">Figure 2
<p>Upregulation of TRPV1 expression and function following paclitaxel treatment. (<b>A</b>) Representative immunostaining images of TRPV1 on day 14 post-Veh and PTX injection. The scale bars represent 100 μm (left panel) and 50 μm (right panel). (<b>B</b>) Quantitative analysis of TRPV1-positive neurons within the fixed area (1 mm<sup>2</sup>). Data were analyzed using an unpaired <span class="html-italic">t</span>-test. *** <span class="html-italic">p</span> &lt; 0.001, <span class="html-italic">n</span> = 6. (<b>C</b>). Western blot analysis of TRPV1 protein levels in DRG on day 14 following Veh and PTX treatment. TRPV1 expression levels were normalized to β-actin. The histograms display the statistical analysis of the relative TRPV1 protein levels in Veh and PTX groups. Data were analyzed using an unpaired <span class="html-italic">t</span>-test. *** <span class="html-italic">p</span> &lt; 0.001, <span class="html-italic">n</span> = 4. (<b>D</b>) Images of dissociated L4–L5 DRG neurons from Veh and PTX groups before and after capsaicin (Cap) stimulation. Neurons responsive to capsaicin are indicated by arrows. The scale bars represent 100 μm. (<b>E</b>) Changes in the F340/F380 ratio in Cap-responsive neurons in Veh and PTX groups. (<b>F</b>) Statistic analysis of the mean F340/F380 ratio in Cap-responsive neurons in Veh and PTX groups. Data were analyzed using a two-way ANOVA followed by the Bonferroni post hoc test (3 independent experiments). *** <span class="html-italic">p</span> &lt; 0.001. <span class="html-italic">n</span> = 5, 7 in panels (<b>E</b>,<b>F</b>). Veh: vehicle, BL: baseline, PTX: paclitaxel.</p>
Full article ">Figure 3
<p>Attenuation of paclitaxel-induced mechanical pain by intraperitoneal injection of the TRPV1 antagonist capsazepine (CPZ). (<b>A</b>) Static mechanical hypersensitivity in PIPNP was reduced following intraperitoneal injection of CPZ. Left: Time course of PWT after intraperitoneal injection of CPZ (30 mg/kg) on days 4, 8, 12, and 16 after paclitaxel treatment. <span class="html-italic">n</span> = 11. Right: Statistical analysis of the AUC of PWT from the left panel. (<b>B</b>) Dynamic mechanical allodynia in PIPNP was alleviated by intraperitoneal injection of CPZ. Left: Time course of the dynamic score after intraperitoneal injection of CPZ (30 mg/kg) on days 4, 8, 12, and 16 post-PTX and Veh treatment. <span class="html-italic">n</span> = 6. Right: Statistic analysis of the AUC of the dynamic score from the left panel. The pink areas in left panels of (<b>A</b>,<b>B</b>) indicate the period of making the PIPNP model. Data were analyzed using a two-way ANOVA followed by Bonferroni post hoc tests and an unpaired <span class="html-italic">t</span>-test. ** <span class="html-italic">p</span> &lt; 0.01, *** <span class="html-italic">p</span> &lt; 0.001. <sup>##</sup> <span class="html-italic">p</span> &lt; 0.01, <sup>###</sup> <span class="html-italic">p</span> &lt; 0.001.</p>
Full article ">Figure 4
<p>Intrathecal injection of CPZ significantly alleviates mechanical pain in PIPNP. (<b>A</b>) Time course (left) and AUC analysis of the PWT (right) following intrathecal injection of 10 μg of CPZ on days 4, 8, 12, and 16 post-PTX treatment. <span class="html-italic">n</span> = 7, 5. (<b>B</b>) Time course (left) and AUC analysis (right) of the dynamic score following intrathecal injection of 10 μg CPZ on days 4, 8, 12, and 16 post-PTX treatment. <span class="html-italic">n</span> = 8, 6. The pink areas in left panels of (<b>A</b>,<b>B</b>) indicate the period of making the PIPNP model.Data were analyzed using a two-way ANOVA followed by Bonferroni post hoc tests and an unpaired <span class="html-italic">t</span>-test. * <span class="html-italic">p</span> &lt; 0.05, ** <span class="html-italic">p</span>&lt; 0.01.</p>
Full article ">Figure 5
<p>Reduced expression and function of TRPM8 following paclitaxel treatment. (<b>A</b>) Representative photomicrographs of TRPM8 immunofluorescence staining on day 14 after Veh and PTX treatment. (<b>B</b>) Statistical analysis of the number of TRPM8-positive neurons in the fixed area (1 μm<sup>2</sup>). Data were analyzed using a <span class="html-italic">t</span>-test. * <span class="html-italic">p</span> &lt; 0.05, <span class="html-italic">n</span> = 5, 4. (<b>C</b>) Western blot analysis of TRPM8 protein levels in DRGs on day 14 after Veh and PTX treatment. TRPM8 expression was normalized to β-actin expression. Histograms show the relative amount of TRPM8 protein in Veh and PTX-treated rats. The data were analyzed using an unpaired <span class="html-italic">t</span>-test. *** <span class="html-italic">p</span> &lt; 0.001, <span class="html-italic">n</span> = 5. (<b>D</b>,<b>F</b>) The response of DRG neurons to menthol was significantly reduced in the PTX-treated group compared with the Veh-treated group. (<b>D</b>) Representative images of the calcium signals in DRG neurons isolated from Veh and PTX-treated rats after menthol (200 μM) stimulation. White arrows indicate the DRG neurons responsive to menthol. (<b>F</b>) Statistical analysis of the F340/F380 ratio in the Veh and PTX groups following menthol stimulation. Data were analyzed using a two-way ANOVA followed by Bonferroni post hoc tests (3 independent experiments). *** <span class="html-italic">p</span> &lt; 0.001, <span class="html-italic">n</span> = 26, 22. (<b>E</b>,<b>G</b>) The response of DRG neurons to WS-12 (a selective agonist of TRPM8) was significantly decreased in the PTX-treated group compared with the Veh-treated group. (<b>E</b>) Representative calcium images of DRG neurons isolated from Veh- and PTX-treated rats after WS-12 (10 μM) stimulation. White arrows indicate the DRG neurons responsive to WS-12. (<b>G</b>) Statistical analysis of the F340/F380 ratio following WS-12 stimulation. Data were analyzed using a two-way ANOVA followed by Bonferroni post hoc tests, 3 independent experiments, *** <span class="html-italic">p</span> &lt; 0.001, <span class="html-italic">n</span> = 7, 16. Scale bars in (<b>D</b>,<b>E</b>) represent 100 μm. Veh: vehicle, PTX: paclitaxel, BL: baseline.</p>
Full article ">Figure 6
<p>Activation of TRPM8 significantly attenuated mechanical hypersensitivity in PIPNP. (<b>A</b>,<b>B</b>) Intrathecal injection of 10 μg of WS-12 significantly alleviated both static and dynamic mechanical pain. (<b>A</b>) Time course and statistical analysis of PWT in PTX-treated rats on days 8, 10, 12, and 14 after intrathecal injection of WS-12 or the control. (<b>B</b>) Time course and AUC analysis of the dynamic score in the PTX-treated rats following administration of WS-12 or the control. (<b>C</b>,<b>D</b>) Topical application of 1% menthol demonstrated a significant analgesic effect on paclitaxel-induced mechanical pain. (<b>C</b>) Time course of PWT and AUC analysis in the vehicle-treated, PTX-treated, and PTX treated with menthol application groups on days 7, 14, 21, and 28. (<b>D</b>) Time course and AUC statistical analysis of the dynamic score in the vehicle-treated, PTX-treated, and PTX treated with menthol application groups on days 7, 14, 21, and 28. The pink areas in the left panels of (<b>A</b>–<b>D</b>) indicate the period of making the PIPNP model. The data were analyzed using a two-way ANOVA followed by Bonferroni post hoc tests and an unpaired <span class="html-italic">t</span>-test. * <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, <sup>##</sup> <span class="html-italic">p</span> &lt; 0.01, <sup>###</sup> <span class="html-italic">p</span> &lt; 0.001. <span class="html-italic">n</span> = 8, 5 in (<b>A</b>); <span class="html-italic">n</span> = 11, 9 in (<b>B</b>); <span class="html-italic">n</span> = 6 in (<b>C</b>) and <span class="html-italic">n</span> = 5 in (<b>D</b>).</p>
Full article ">Figure 7
<p>Activation of TRPM8 by menthol inhibited the function of TRPV1 in vitro and in vivo. (<b>A</b>–<b>D</b>) The calcium imaging experiment demonstrated that the DRG neurons pre-activated by 200 μM menthol showed a significantly decreased response to 1 μM capsaicin. (<b>A</b>) Representative calcium images of DRG neurons from PTX-treated and vehicle-treated rats. Neurons responsive to menthol are indicated by red arrows (menthol<sup>+</sup>), while those unresponsive to menthol are indicated by blue arrows (menthol<sup>−</sup>). (<b>B</b>) Statistical analysis of the calcium signal changes after capsaicin treatment (ΔRatio after capsaicin) in the menthol-responsive and unresponsive neurons from the vehicle-treated rats. (<b>C</b>) Data analysis of calcium signal changes in DRG neurons in response to capsaicin between the menthol-responsive and unresponsive groups from the PTX-treated rats. (<b>D</b>) Comparison of the overall (including both menthol<sup>+</sup> and menthol<sup>−</sup> neurons) calcium signal changes to capsaicin in the vehicle and PTX groups after activation of TRPM8 by menthol. Data were analyzed using a two-way ANOVA followed by Bonferroni post hoc tests, 3 independent experiments, * <span class="html-italic">p</span> &lt; 0.05, *** <span class="html-italic">p</span> &lt; 0.001, <span class="html-italic">n</span> = 15, 9 in (<b>B</b>); <span class="html-italic">n</span> = 12, 11 in (<b>C</b>); <span class="html-italic">n</span> = 8, 5 in (<b>D</b>). (<b>E</b>) The topical application of 1% menthol significantly attenuated CFA-induced inflammatory heat hyperalgesia. Left: Paw withdrawal latency (PWL) in Hargreaves test for control, CFA, and CFA/menthol groups. Right: Statistical analysis of PWL from the left figure. Black arrows in the left panels of (<b>B</b>–<b>D</b>) indicate the time point of administration of menthol or capsaicin. All data are presented as means ± SEM. Data were analyzed using a two-way ANOVA followed by Bonferroni post hoc tests and an unpaired <span class="html-italic">t</span>-test. * <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, <sup>#</sup> <span class="html-italic">p</span> &lt; 0.05, <sup>###</sup> <span class="html-italic">p</span> &lt; 0.001, <span class="html-italic">n</span> = 5, 5, 6.</p>
Full article ">
14 pages, 2788 KiB  
Article
Identification of Differentially Expressed Genes and microRNAs in the Gray and White Feather Follicles of Shitou Geese
by Pengyun Guo, Junpeng Chen, Lei Luo, Xumeng Zhang, Xiujin Li, Yunmao Huang, Zhongping Wu and Yunbo Tian
Animals 2024, 14(10), 1508; https://doi.org/10.3390/ani14101508 - 20 May 2024
Viewed by 911
Abstract
The Shitou goose, a highly recognized indigenous breed with gray plumage originating from Chaozhou Raoping in Guangdong Province, China, is renowned for being the largest goose species in the country. Notably, during the pure breeding process of Shitou geese, approximately 2% of the [...] Read more.
The Shitou goose, a highly recognized indigenous breed with gray plumage originating from Chaozhou Raoping in Guangdong Province, China, is renowned for being the largest goose species in the country. Notably, during the pure breeding process of Shitou geese, approximately 2% of the offspring in each generation unexpectedly exhibited white plumage. To better understand the mechanisms underlying white plumage color formation in Shitou geese, we conducted a comparative transcriptome analysis between white and gray feather follicles, aiming to identify key genes and microRNAs that potentially regulate white plumage coloration in this unique goose breed. Our results revealed a number of pigmentation genes, encompassing TYR, TYRP1, EDNRB2, MLANA, SOX10, SLC45A2, GPR143, TRPM1, OCA2, ASIP, KIT, and SLC24A5, which were significantly down-regulated in the white feather follicles of Shitou geese. Among these genes, EDNRB2 and KIT emerged as the most promising candidate genes for white plumage coloration in Shitou geese. Additionally, our analysis also uncovered 46 differentially expressed miRNAs. Of these, miR-144-y may play crucial roles in the regulation of feather pigmentation. Furthermore, the expression of novel-m0086-5p, miR-489-y, miR-223-x, miR-7565-z, and miR-3535-z exhibits a significant negative correlation with the expression of pigmentation genes including TYRP1, EDNRB2, MLANA, SOX10, TRPM1, and KIT, suggesting these miRNAs may indirectly regulate the expression of these genes, thereby influencing feather color. Our findings provide valuable insights into the genetic mechanisms underlying white plumage coloration in Shitou geese and contribute to the broader understanding of avian genetics and coloration research. Full article
(This article belongs to the Special Issue Genetic Analysis of Important Traits in Domestic Animals)
Show Figures

Figure 1

Figure 1
<p>Phenotypic comparison between gray and white Shitou geese. (<b>A</b>) Male (<b>left</b>) and female (<b>right</b>) gray-feathered Shitou geese. (<b>B</b>) Male (<b>left</b>) and female (<b>right</b>) white-feathered Shitou geese.</p>
Full article ">Figure 2
<p>Volcano plot of differentially expressed genes between gray and white feather follicles in Shitou geese. Tyrosinase (<span class="html-italic">TYR</span>), Tyrosinase-related protein 1 (<span class="html-italic">TYRP1</span>), Endothelin receptor type B-like (<span class="html-italic">EDNRB2</span>), Melan-A (<span class="html-italic">MLANA</span>), SRY-box transcription factor 10 (SOX10), Solute Carrier Family 45 Member 2 (<span class="html-italic">SLC45A2</span>), G Protein-coupled Receptor 143 (<span class="html-italic">GPR143</span>), Transient Receptor Potential Cation Channel Subfamily M Member 1 (<span class="html-italic">TRPM1</span>), Oculocutaneous Albinism II (OCA2), Agouti Signaling Protein (<span class="html-italic">ASIP</span>), Kit Proto-Oncogene Receptor Tyrosine Kinase (<span class="html-italic">KIT</span>), Solute Carrier Family 24 Member 5 (<span class="html-italic">SLC24A5</span>).</p>
Full article ">Figure 3
<p>Analysis of differentially expressed miRNAs in gray and white feather follicles of Shitou geese. (<b>A</b>) The ratio of different types of small RNA in each sample; all: all samples; G1: Gray feather1; G2: Gray feather2; G3: Gray feather3; W1: White feather1; W2: White feather2; W3: White feather3. (<b>B</b>) Volcano map of differentially expressed miRNAs from gray and white feathers.</p>
Full article ">Figure 4
<p>Construction of the miRNA-mRNA interaction network based on expression level correlations. The pink ellipses represent differentially expressed genes, the yellow ellipses represent pigment genes, and the blue diamonds represent differentially expressed miRNAs. The darker the color of the connecting lines, the greater the correlation coefficient.</p>
Full article ">Figure 5
<p>Verification of the sequencing results using qRT-PCR analysis. (<b>A</b>) Relative expression of Tyrosinase (<span class="html-italic">TYR</span>), Endothelin receptor type B-like (<span class="html-italic">EDNRB2</span>), SRY-box transcription factor 10 (SOX10) and Kit Proto-Oncogene Receptor Tyrosine Kinase (<span class="html-italic">KIT</span>) gene in gray and white feather follicles of Shitou geese. (<b>B</b>) Relative expression of <span class="html-italic">miR-451-x</span>, <span class="html-italic">miR-144-y</span>, <span class="html-italic">miR-204-x</span>, and <span class="html-italic">miR-196-x</span> in gray and white feather follicles of Shitou geese. * Indicates a significant difference (<span class="html-italic">p</span> &lt; 0.05). ** Indicates an extremely significant difference (<span class="html-italic">p</span> &lt; 0.01).</p>
Full article ">
15 pages, 18790 KiB  
Article
Whole-Transcriptome Analysis Sheds Light on the Biological Contexts of Intramuscular Fat Deposition in Ningxiang Pigs
by Zhao Jin, Hu Gao, Yawei Fu, Ruimin Ren, Xiaoxiao Deng, Yue Chen, Xiaohong Hou, Qian Wang, Gang Song, Ningyu Fan, Haiming Ma, Yulong Yin and Kang Xu
Genes 2024, 15(5), 642; https://doi.org/10.3390/genes15050642 - 19 May 2024
Cited by 1 | Viewed by 1004
Abstract
The quality of pork is significantly impacted by intramuscular fat (IMF). However, the regulatory mechanism of IMF depositions remains unclear. We performed whole-transcriptome sequencing of the longissimus dorsi muscle (IMF) from the high (5.1 ± 0.08) and low (2.9 ± 0.51) IMF groups [...] Read more.
The quality of pork is significantly impacted by intramuscular fat (IMF). However, the regulatory mechanism of IMF depositions remains unclear. We performed whole-transcriptome sequencing of the longissimus dorsi muscle (IMF) from the high (5.1 ± 0.08) and low (2.9 ± 0.51) IMF groups (%) to elucidate potential mechanisms. In summary, 285 differentially expressed genes (DEGs), 14 differentially expressed miRNAs (DEMIs), 83 differentially expressed lncRNAs (DELs), and 79 differentially expressed circRNAs (DECs) were identified. DEGs were widely associated with IMF deposition and liposome differentiation. Furthermore, competing endogenous RNA (ceRNA) regulatory networks were constructed through co-differential expression analyses, which included circRNA-miRNA-mRNA (containing 6 DEMIs, 6 DEGs, 47 DECs) and lncRNA-miRNA-mRNA (containing 6 DEMIs, 6 DEGs, 36 DELs) regulatory networks. The circRNAs sus-TRPM7_0005, sus-MTUS1_0004, the lncRNAs SMSTRG.4269.1, and MSTRG.7983.2 regulate the expression of six lipid metabolism-related target genes, including PLCB1, BAD, and GADD45G, through the binding sites of 2-4068, miR-7134-3p, and miR-190a. For instance, MSTRG.4269.1 regulates its targets PLCB1 and BAD via miRNA 2_4068. Meanwhile, sus-TRPM7_0005 controls its target LRP5 through ssc-miR-7134-3P. These findings indicate molecular regulatory networks that could potentially be applied for the marker-assisted selection of IMF to enhance pork quality. Full article
(This article belongs to the Special Issue Advances in Pig Genetic and Genomic Breeding of 2024)
Show Figures

Figure 1

Figure 1
<p>The fat content of the LDM samples and volcano plot. (<b>A</b>) The carcass weight, the content of intramuscular fat (IMF), and the monounsaturated fatty acid (MUFA) and polyunsaturated fat (PUFA) levels between the low- and high-IMF groups and (<b>B</b>) correlation analysis. The volcano plot displaying the differentially expressed genes (DEGs) (<b>C</b>), differentially expressed miRNAs (DEMIs) (<b>D</b>), differentially expressed lncRNAs (DELs) (<b>E</b>), and differentially expressed circRNAs (DECs) (<b>F</b>), including upregulated and downregulated genes in the two groups (low-IMF vs. high-IMF groups).</p>
Full article ">Figure 2
<p>GO annotation and KEGG analysis for DEGs. (<b>A</b>) GO annotation of DEGs; (<b>B</b>) GO annotation in BP and MF of DEGs; and (<b>C</b>,<b>D</b>) KEGG analyses of DEGs (The red boxes are genes/entries related to lipid metabolism).</p>
Full article ">Figure 3
<p>For GO and KEGG annotations, the host genes of DECs and target genes of DELs were analyzed. (<b>A</b>) GO functional classification of cis-lncRNA target genes; (<b>B</b>) KEGG pathway annotation of trans-lncRNA target genes; (<b>C</b>) target genes of trans-lncRNAs involved in differential expression of lipid metabolic pathways; (<b>D</b>) KEGG pathway enrichment annotation of DEC host genes; and (<b>E</b>) lipid-associated pathway-enriched host genes for DECs (The red boxes are entries related to lipid metabolism).</p>
Full article ">Figure 4
<p>The ceRNA networks. (<b>A</b>) Regulatory networks of CircRNA-miRNA-mRNA. (<b>B</b>) Regulatory networks of lncRNA-miRNA-mRNA. qPCR analysis to verify the RNA-sequencing results. (<b>C</b>) The miRNA, lncRNA, and circRNA expression levels. (<b>D</b>) The mRNA expression levels (The main CE network diagrams are shown in red).</p>
Full article ">
13 pages, 1768 KiB  
Article
Anti-Nociceptive Effects of Sphingomyelinase and Methyl-Beta-Cyclodextrin in the Icilin-Induced Mouse Pain Model
by Ádám Horváth, Anita Steib, Andrea Nehr-Majoros, Boglárka Kántás, Ágnes Király, Márk Racskó, Balázs István Tóth, Eszter Szánti-Pintér, Eva Kudová, Rita Skoda-Földes, Zsuzsanna Helyes and Éva Szőke
Int. J. Mol. Sci. 2024, 25(9), 4637; https://doi.org/10.3390/ijms25094637 - 24 Apr 2024
Viewed by 989
Abstract
The thermo- and pain-sensitive Transient Receptor Potential Melastatin 3 and 8 (TRPM3 and TRPM8) ion channels are functionally associated in the lipid rafts of the plasma membrane. We have already described that cholesterol and sphingomyelin depletion, or inhibition of sphingolipid biosynthesis decreased the [...] Read more.
The thermo- and pain-sensitive Transient Receptor Potential Melastatin 3 and 8 (TRPM3 and TRPM8) ion channels are functionally associated in the lipid rafts of the plasma membrane. We have already described that cholesterol and sphingomyelin depletion, or inhibition of sphingolipid biosynthesis decreased the TRPM8 but not the TRPM3 channel opening on cultured sensory neurons. We aimed to test the effects of lipid raft disruptors on channel activation on TRPM3- and TRPM8-expressing HEK293T cells in vitro, as well as their potential analgesic actions in TRPM3 and TRPM8 channel activation involving acute pain models in mice. CHO cell viability was examined after lipid raft disruptor treatments and their effects on channel activation on channel expressing HEK293T cells by measurement of cytoplasmic Ca2+ concentration were monitored. The effects of treatments were investigated in Pregnenolone-Sulphate-CIM-0216-evoked and icilin-induced acute nocifensive pain models in mice. Cholesterol depletion decreased CHO cell viability. Sphingomyelinase and methyl-beta-cyclodextrin reduced the duration of icilin-evoked nocifensive behavior, while lipid raft disruptors did not inhibit the activity of recombinant TRPM3 and TRPM8. We conclude that depletion of sphingomyelin or cholesterol from rafts can modulate the function of native TRPM8 receptors. Furthermore, sphingolipid cleavage provided superiority over cholesterol depletion, and this method can open novel possibilities in the management of different pain conditions. Full article
(This article belongs to the Special Issue Molecular Links between Sensory Nerves, Inflammation, and Pain 3.0)
Show Figures

Figure 1

Figure 1
<p>Effect of lipid raft disruptors on cell viability. The effect of SMase (<b>A</b>), Myr (<b>B</b>), MCD (<b>C</b>), and C1 (<b>D</b>) on viability is presented. Data are presented as means ± standard error of mean (SEM), n = 9; three independent experiments in triplicates. Statistical analysis was performed using one-way Analysis of Variance (ANOVA) followed by Dunnett’s post hoc test. (** <span class="html-italic">p</span> &lt; 0.01; *** <span class="html-italic">p</span> &lt; 0.001 MCD/C1 samples vs. saline control/DMSO control).</p>
Full article ">Figure 2
<p>Effect of lipid raft disruptors on icilin-induced acute nocifensive pain. The effect of 50 mU SMase (<b>A</b>), 1 mM Myr (<b>B</b>), 15 mM MCD (<b>C</b>), and 100 µM C1 compound (<b>D</b>) on nocifensive behavior is presented. Data are presented as means ± SEM, n = 6–27 mice/group. Statistical analysis was performed using one-way ANOVA followed by Bonferroni’s post hoc test. (* <span class="html-italic">p</span> &lt; 0.05; ** <span class="html-italic">p</span> &lt; 0.01; SMase pretreated group vs. saline pretreated group).</p>
Full article ">Figure 3
<p>Effect of 100 µM C1 on PS and CIM-0216-induced acute nocifensive pain. Data are presented as means ± SEM, n = 5–8 mice/group. Statistical analysis was performed using a one-way ANOVA followed by Bonferroni’s post hoc test.</p>
Full article ">Figure 4
<p>Effect of lipid raft disruptors on the agonist-induced activity of TRPM8 and TRPM3. HEK293T cells overexpressing either human TRPM8 (hTRPM8) or TRPM3α2 variant of mouse TRPM3 (mTRPM3α2) were pretreated with (<b>A</b>) SMase, (<b>B</b>) Myriocin, (<b>C</b>) MCD, and (<b>D</b>) C1 compound and subjected to [Ca<sup>2+</sup>]<sub>IC</sub> measurements upon applying various concentrations of agonists, as indicated in the figure. Each data point is presented as the mean ± SEM of three independent measurements (n = 3). Concentration–response curves were fitted as described in the <a href="#sec4-ijms-25-04637" class="html-sec">Section 4</a>.</p>
Full article ">
Back to TopTop