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

Journals

Article Types

Countries / Regions

Search Results (201)

Search Parameters:
Keywords = dorsal root ganglia

Order results
Result details
Results per page
Select all
Export citation of selected articles as:
22 pages, 13437 KiB  
Article
The Intrinsic Neuronal Activation of the CXCR4 Signaling Axis Is Associated with a Pro-Regenerative State in Cervical Primary Sensory Neurons Conditioned by a Sciatic Nerve Lesion
by Petr Dubový, Ivana Hradilová-Svíženská, Václav Brázda, Anna Jambrichová, Viktorie Svobodová and Marek Joukal
Int. J. Mol. Sci. 2025, 26(1), 193; https://doi.org/10.3390/ijms26010193 - 29 Dec 2024
Viewed by 546
Abstract
CXCL12 and CXCR4 proteins and mRNAs were monitored in the dorsal root ganglia (DRGs) of lumbar (L4–L5) and cervical (C7–C8) spinal segments of naïve rats, rats subjected to sham operation, and those undergoing unilateral complete sciatic nerve transection (CSNT) on post-operation day 7 [...] Read more.
CXCL12 and CXCR4 proteins and mRNAs were monitored in the dorsal root ganglia (DRGs) of lumbar (L4–L5) and cervical (C7–C8) spinal segments of naïve rats, rats subjected to sham operation, and those undergoing unilateral complete sciatic nerve transection (CSNT) on post-operation day 7 (POD7). Immunohistochemical, Western blot, and RT-PCR analyses revealed bilaterally increased levels of CXCR4 protein and mRNA in both lumbar and cervical DRG neurons after CSNT. Similarly, CXCL12 protein levels increased, and CXCL12 mRNA was upregulated primarily in lumbar DRGs ipsilateral to the nerve lesion. Intrathecal application of the CXCR4 inhibitor AMD3100 following CSNT reduced CXCL12 and CXCR4 protein levels in cervical DRG neurons, as well as the length of afferent axons regenerated distal to the ulnar nerve crush. Furthermore, treatment with the CXCR4 inhibitor decreased levels of activated Signal Transducer and Activator of Transcription 3 (STAT3), a critical transforming factor in the neuronal regeneration program. Administration of IL-6 increased CXCR4 levels, whereas the JAK2-dependent STAT3 phosphorylation inhibitor (AG490) conversely decreased CXCR4 levels. This indicates a link between the CXCL12/CXCR4 signaling axis and IL-6-induced activation of STAT3 in the sciatic nerve injury-induced pro-regenerative state of cervical DRG neurons. The role of CXCR4 signaling in the axon-promoting state of DRG neurons was confirmed through in vitro cultivation of primary sensory neurons in a medium supplemented with CXCL12, with or without AMD3100. The potential involvement of conditioned cervical DRG neurons in the induction of neuropathic pain is discussed. Full article
(This article belongs to the Special Issue Advances in Peripheral Nerve Regeneration)
Show Figures

Figure 1

Figure 1
<p>(<b>a</b>) Representative sections of DRGs from naïve rats (Naive) and rats undergoing sterile sham (Sham) and CSNT operations on POD7 (CSNT7D). DRGs from CSNT-operated rats were taken from lumbar (L4-DRG) and cervical (C7-DRG) segments on both ipsilateral (i) and contralateral (c) sides. All sections were immunostained for CXCL12 detection under identical conditions. Increased intensities of CXCL12 immunofluorescence were observed in the neuronal bodies (arrows) and satellite glial cells (arrowheads) of both lumbar and cervical DRGs compared with those from naïve or sham-operated rats. Scale bars = 40 µm. (<b>b</b>) The results of CXCL12-IF intensities measured in the neuronal cytoplasm of large (L), medium (M), and small (S) cervical DRG neurons from naïve rats (Naïve), as well as both ipsilateral and contralateral DRGs of Sham- and CSNT-operated rats for POD7. *** Significant differences (<span class="html-italic">p</span> &lt; 0.001) compared with sham-operated rats, as was determined using a Mann–Whitney U-test.</p>
Full article ">Figure 2
<p>(<b>a</b>) Representative sections of DRGs from naïve rat (Naive) and rats after sterile sham (Sham) and CSNT operations on POD7 (CSNT7D). DRGs from CSNT-operated rats were taken from lumbar (L4-DRG) and cervical (C7-DRG) segments on both ipsilateral (i) and contralateral (c) sides. All sections were immunostained for CXCR4 detection under identical conditions. Sciatic nerve lesions induced an increased intensity of CXCR4 immunofluorescence in neuronal nuclei (arrowheads) and a diffuse pattern in the bodies of some DRG neurons (arrows). The heightened intensities of CXCR4 immunofluorescence predominantly loaded nuclei of both DRG neurons and their satellite glial cells. Scale bars = 40 µm. (<b>b</b>) The results of CXCR4-IF intensities measured in the neuronal nuclei of the cervical (C) and lumbar (L) DRGs from both ipsilateral (i) and contralateral (c) sides removed from naïve rats (Naïve) as well as from Sham- and CSNT-operated rats on POD7. *** Significant differences (<span class="html-italic">p</span> &lt; 0.001) compared with sham-operated rats, as was determined using a Mann–Whitney U-test.</p>
Full article ">Figure 3
<p>Results of real-time PCR (RT-PCR) of relative CXCL12 and CXCR4 mRNA levels in DRGs of lumbar (L) and the cervical (C) spinal segments (L4–L5) and (C7–C8), respectively, from the ipsilateral (i) and contralateral (c) sides. Tissue samples were collected from naïve rats as well as from sham- and CSNT-operated rats at POD7 (n = 9 for each group). Relative expressions were calculated using Actin as the housekeeping gene and normalized to naïve controls. * Significant differences (<span class="html-italic">p</span> &lt; 0.05) compared with sham-operated rats were determined using a Mann–Whitney U-test.</p>
Full article ">Figure 4
<p>(<b>a</b>) The representative sections through the cervical (C7-DRG) and lumbar (L4-DRG) DRGs harvested from the ipsilateral (i) side of rats subjected to CSNT on POD7, and after intrathecal application of artificial cerebrospinal fluid (ACSF) or AMD3100. Sections were incubated under the same conditions to demonstrate CXCR4 and CXCL12 immunofluorescence. The results showed reduced intensities of both CXCL12 and CXCR4 immunofluorescence in DRG sections from rats treated with AMD3100 compared with those treated with ACSF. Changes in CXCR4 immunofluorescence intensities were observed in neuronal nuclei (arrowheads), while changes in CXCL12 immunofluorescence intensities were noted in the neuronal bodies (arrows) and satellite glial cells (arrowheads). Scale bars = 50 µm. (<b>b</b>) Representative blots of CXCR4 and CXCL12 proteins with equal protein loading confirmed by actin levels (Actin). Samples of DRGs from both lumbar (L) and cervical (C) segments were collected from naïve rats and from the ipsilateral (i) and contralateral (c) sides of rats undergoing CSNT on POD7, followed by intrathecal application of either artificial cerebrospinal fluid (ACSF) or AMD3100. (<b>c</b>) The densitometry results of CXCR4 and CXCL12 protein bands, normalized to Actin levels. Densities of CXCR4 and CXCL12 bands from the naïve DRGs were set as 1 for reference. *, ** denote significant differences (<span class="html-italic">p</span> &lt; 0.05, <span class="html-italic">p</span> &lt; 0.01, respectively) compared with naïve controls, while †† indicates a significant difference (<span class="html-italic">p</span> &lt; 0.01) compared with ACSF-treated counterparts, as determined by the Mann–Whitney U-test.</p>
Full article ">Figure 5
<p>In vivo assay of the pro-regenerative state of cervical DRG neurons conditioned by unilateral CSNT on POD7. The results revealed that intrathecal application of AMD3100 significantly reduced the lengths of SCG10-immunopositive sensory axons regenerated distal to the crush point compared with rats subjected to CSNT and treated with ACSF. (<b>a</b>) Representative longitudinal sections through the ulnar nerve distal to the crush site following prior CSNT and intrathecal application of either ACSF or AMD3100. Arrows indicate the maximal length of regenerated axons. Scale bars = 250 µm. (<b>b</b>) The lengths of SCG10-immunopositive axons measured distal to the ulnar nerve crush (n = 4 for each group). ** indicates a significant difference (<span class="html-italic">p</span> &lt; 0.01) compared with rats treated with ACSF, as determined by the Mann–Whitney U-test.</p>
Full article ">Figure 6
<p>The primary sensory neurons isolated from cervical DRGs were cultivated in vitro (<b>a</b>) in control medium (Control), in medium with the addition of CXCL12 (CXCL12), and in medium supplemented with both CXCL12 and AMD3100 (CXCL12+AMD3100). Scale bars = 75 µm. (<b>b</b>) The measurement results showed that the addition of CXCL12 into the medium significantly increased the number and length of neurites per neuron. Conversely, the medium supplemented with CXCL12 and AMD3100 reduced the outgrowth of neurites. * denotes a significant difference (<span class="html-italic">p</span> &lt; 0.05) compared with DRG neurons cultivated in control medium; † indicates a significant difference (<span class="html-italic">p</span> &lt; 0.05) between DRG neurons treated with CXCL12 and CXCL12 + AMD3100 media.</p>
Full article ">Figure 7
<p>(<b>a</b>) The results of STAT3(Y705) immunofluorescence staining are present in representative sections of DRGs. Sections through the cervical (C7-DRG) harvested from a naïve rat (Naïve) and a rat subjected to CSNT on POD7 (CSNT7D). In addition, sections of the cervical (C7-DRGi) and lumbar (L4-DRGi) ganglia ipsilateral (i) to CSNT on POD7 were obtained from rats after intrathecal application of artificial cerebrospinal fluid (ACSF) or AMD3100. The sections were immunostained under identical conditions to detect STAT3(Y705) in the neuronal nuclei (arrowheads). Scale bars = 25 µm. (<b>b</b>) Intensities of STAT3(Y705) immunofluorescence measured in the nuclei of DRG neurons (n = 3 for each group) were significantly decreased in DRGs of L4–L5 and C7–C8 segments (L4–L5 DRG, C7–C8 DRG) from both ipsilateral (i) and contralateral (c) sides of rats undergoing CSNT followed intrathecal application of AMD3100, compared with those from ACSF-treated controls. **, *** indicate significant differences (<span class="html-italic">p</span> &lt; 0.01 or <span class="html-italic">p</span> &lt; 0.001, respectively) compared with DRGs of ACSF-treated rats as determined by the Mann–Whitney U-test.</p>
Full article ">Figure 8
<p>(<b>a</b>) The representative sections of cervical DRGs (C7) harvested from rats subjected to CSNT on POD7 followed by intrathecal application of ACSF, IL-6 or JAK2 inhibitor AG490. Sections were immunostained for CXCR4 under identical conditions. Scale bars = 25 µm. (<b>b</b>) Measurement of immunofluorescence intensities in neuronal nuclei demonstrated that intrathecal administration of IL-6 increased, while JAK2 inhibitor significantly decreased, the intensities of CXCR4-IF compared to those of control rats treated with ACSF (n = 3 for each group). ***, * denote significant differences (<span class="html-italic">p</span> &lt; 0.001 or <span class="html-italic">p</span> &lt; 0.05, respectively) compared with DRGs of ACSF-treated rats, as determined by the Mann–Whitney U-test.</p>
Full article ">Figure 9
<p>The representative longitudinal sections distal to the UN crush from rats with a prior sciatic nerve lesion for 7 days. The sections were double-immunostained for CXCR4 and GAP43 (<b>a</b>) as well as CXCL12 and GAP43 (<b>b</b>) or GFAP (<b>c</b>). Merged pictures detected immunopositivity for CXCR4 in the growth cones (long arrows) as well as CXCL12 in non-neuronal cells that displayed GFAP immunopositivity indicating their Schwann cell origin (arrowheads). Scale bars = 30 µm.</p>
Full article ">Figure 10
<p>A diagram illustrating the rat experimental groups, timelines, and samples collected from cervical (C7–C8) and lumbar (L4–L5) DRGs, as well as the segments distal to UN crush, for analyses. (<b>A</b>) Samples were used for immunohistochemical analysis of CXCL12 and CXCR4 proteins and their mRNAs using RT-PCR. (<b>B</b>) Samples were analyzed by immunohistochemistry and western blot for CXCL12 and CXCR4 proteins, as well as immunohistochemical analysis of STAT3. (<b>C</b>) Nerve segments distal to the UN crush were used for SCG10 immunohistochemical analysis. (<b>D</b>) Cervical DRGs were analyzed for CXCR4 in neuronal nuclei by immunohistochemistry.</p>
Full article ">
11 pages, 1460 KiB  
Article
Charactering Neural Spiking Activity Evoked by Acupuncture Through Coupling Generalized Linear Model
by Qing Qin, Kaiyue Zhang, Yanqiu Che, Chunxiao Han, Yingmei Qin and Shanshan Li
Entropy 2024, 26(12), 1088; https://doi.org/10.3390/e26121088 - 13 Dec 2024
Viewed by 588
Abstract
Acupuncturing the ST36 acupoint can evoke a responding activity in the spinal dorsal root ganglia and generate spikes. In order to identify the responding mechanism of different acupuncture manipulations, in this paper the spike history of neurons is taken as the starting point [...] Read more.
Acupuncturing the ST36 acupoint can evoke a responding activity in the spinal dorsal root ganglia and generate spikes. In order to identify the responding mechanism of different acupuncture manipulations, in this paper the spike history of neurons is taken as the starting point and the coupling generalized linear model is adopted to encode the neuronal spiking activity evoked by different acupuncture manipulations. Then, maximum likelihood estimation is used to fit the model parameters and estimate the coupling parameters of stimulus, the self-coupling parameters of the neuron’s own spike history and the cross-coupling parameters of other neurons’ spike history. We use simulation data to test the estimation algorithm’s effectiveness and analyze the main factors that evoke neuronal responding activity. Finally, we use the coupling generalized linear model to encode neuronal spiking activity evoked by two acupuncture manipulations, and estimate the coupling parameters of stimulus, the self-coupling parameters and the cross-coupling parameters. The results show that in acupuncture experiments, acupuncture stimulus is the inducing factor of neuronal spiking activity, and the cross-coupling of other neurons’ spike history is the main factor of neuronal spiking activity. Additionally, the higher the amplitude of the neuronal spiking waveform, the greater the cross-coupling parameter. This lays a theoretical foundation for the scientific application of acupuncture therapy. Full article
(This article belongs to the Section Information Theory, Probability and Statistics)
Show Figures

Figure 1

Figure 1
<p>The effect of spike history of three neurons on the spiking activity of neuron 1.</p>
Full article ">Figure 2
<p>Numbers of neuronal spiking events by varying the coupling parameter of stimulus.</p>
Full article ">Figure 3
<p>Numbers of neuronal spiking events by varying the coupling parameters of spike history.</p>
Full article ">Figure 4
<p>Spiking classification. Raw data evoked by the two acupuncture manipulations: (<b>a</b>) “twirling manipulation”; (<b>b</b>) “lifting−thrusting manipulation”; (<b>c</b>) the two−dimensional map of wavelet coefficients for all spiking waveforms; (<b>d</b>) four average spiking waveforms in the classification results: neuron 1 (red), neuron 2 (blue), neuron 3 (green) and neuron 4 (cyan).</p>
Full article ">
16 pages, 4203 KiB  
Article
HC-HA/PTX3 from Human Amniotic Membrane Induced Differential Gene Expressions in DRG Neurons: Insights into the Modulation of Pain
by Shao-Qiu He, Chi Zhang, Xue-Wei Wang, Qian Huang, Jing Liu, Qing Lin, Hua He, Da-Zhi Yang, Scheffer C. Tseng and Yun Guan
Cells 2024, 13(22), 1887; https://doi.org/10.3390/cells13221887 - 15 Nov 2024
Viewed by 889
Abstract
Background: The biologics derived from human amniotic membranes (AMs) demonstrate potential pain-inhibitory effects in clinical settings. However, the molecular basis underlying this therapeutic effect remains elusive. HC-HA/PTX3 is a unique water-soluble regenerative matrix that is purified from human AMs. We examined whether HC-HA/PTX3 [...] Read more.
Background: The biologics derived from human amniotic membranes (AMs) demonstrate potential pain-inhibitory effects in clinical settings. However, the molecular basis underlying this therapeutic effect remains elusive. HC-HA/PTX3 is a unique water-soluble regenerative matrix that is purified from human AMs. We examined whether HC-HA/PTX3 can modulate the gene networks and transcriptional signatures in the dorsal root ganglia (DRG) neurons transmitting peripheral sensory inputs to the spinal cord. Methods: We conducted bulk RNA-sequencing (RNA-seq) of mouse DRG neurons after treating them with HC-HA/PTX3 (15 µg/mL) for 10 min and 24 h in culture. Differential gene expression analysis was performed using the limma package, and Gene Ontology (GO) and protein–protein interaction (PPI) analyses were conducted to identify the networks of pain-related genes. Western blotting and in vitro calcium imaging were used to examine the protein levels and signaling of pro-opiomelanocortin (POMC) in DRG neurons. Results: Compared to the vehicle-treated group, 24 h treatment with HC-HA/PTX3 induced 2047 differentially expressed genes (DEGs), which were centered on the ATPase activity, receptor–ligand activity, and extracellular matrix pathways. Importantly, PPI analysis revealed that over 50 of these DEGs are closely related to pain and analgesia. Notably, HC-HA/PTX3 increased the expression and signaling pathway of POMC, which may affect opioid analgesia. Conclusions: HC-HA/PTX3 induced profound changes in the gene expression in DRG neurons, centered around various neurochemical mechanisms associated with pain modulation. Our findings suggest that HC-HA/PTX3 may be an important biological active component in human AMs that partly underlies its pain inhibitory effect, presenting a new strategy for pain treatment. Full article
(This article belongs to the Section Cells of the Nervous System)
Show Figures

Figure 1

Figure 1
<p><b>HC-HA/PTX3 (HHP) treatment of dorsal root ganglia (DRG) neurons for RNA-sequencing.</b> (<b>A</b>) Schematic diagram of experimental procedure. RNA-seq was performed on cultured wildtype (WT) mouse DRG neurons treated with vehicle or HC-HA/PTX3 (15 µg/mL) for 10 min or 24 h. (<b>B</b>) Neuronal markers, such as <span class="html-italic">Calca</span> (CGRP) or <span class="html-italic">Tubb3</span>, had high-level expression in the RNA samples from cultured DRG neurons, whereas those for glial (<span class="html-italic">Apoe</span>, <span class="html-italic">Fabb7</span>) and endothelial (<span class="html-italic">Cldn5</span>, <span class="html-italic">Fit1</span>) markers were much lower. N = 9. Data are mean ± SEM. (<b>C</b>) Principal component analysis (PCA) of the samples treated with vehicle or HC-HA/PTX3 (10 min, 24 h). PCAs were generated after batch effect correction by applying ComBat-seq.</p>
Full article ">Figure 2
<p><b>HC-HA/PTX3 (HHP) treatment for 24 h broadly changed gene expression in dorsal root ganglia (DRG) neurons, which may affect molecular functions and cellular components.</b> (<b>A</b>) Volcano plot of differentially expressed genes (DEGs) in cultured wildtype (WT) DRG neurons after vehicle or 24 h of HC-HA/PTX3 treatment (15 µg/mL). DEGs were identified by a [log2 fold change (FC)] &gt; 0.5 and an FDR &lt; 0.05. Significant downregulated and upregulated genes are designated in blue and red colors. (<b>B</b>) Venn diagram representing the number of DEGs identified after batch effect correction by ComBat-seq (red), edgeR (blue), and in both (overlap). (<b>C</b>) Heatmap shows the log2FC of the 200 most variable genes in 24 h HC-HA/PTX3-group as compared to the vehicle group. (<b>D</b>,<b>E</b>) Gene Ontology (GO) enrichment analysis of DEGs induced by HC-HA/PTX3. ClusterProfiler analysis of significantly enriched GO terms within molecular function categories (<b>D</b>) and cellular component categories (<b>E</b>). FDR-adjusted <span class="html-italic">p</span> &lt; 0.05.</p>
Full article ">Figure 3
<p><b>Long-term HC-HA/PTX3 treatment altered the expression of pain-related genes in dorsal root ganglia (DRG) neurons.</b> (<b>A</b>) The pain-related PPI network of the DEGs in cultured wildtype (WT) mouse DRG neurons after 24 h of HC-HA/PTX3 treatment (15 µg/mL) as compared to the vehicle. The network was built based on the pain interactome using DEGs, together with neighbors from the pain interactome network (DEGs and non-DEGs). (<b>B</b>) Heatmap shows the fold change of DEGs in each biological replicate of vehicle- and HC-HA/PTX3-treated samples (15 µg/mL, 24 h). The genes shown on the heatmap that encode neuropeptides, cytokines, GPCRs, extracellular matrix (ECMs) were significantly regulated by HC-HA/PTX3. (<b>C</b>) The pro-opiomelanocortin (POMC) network of DEGs induced by 24 h treatment with HC-HA/PTX3 (15 µg/mL) in WT DRG neurons. Colored edges mark the type of interaction. Colored nodes mark the expression changes (up/down/no change) after HC-HA/PTX3 treatment. Node size indicates the number of interactions against pain interactome, as explained in the legend. (<b>D</b>) Western immunoblotting (cropped blots, full-length blots are presented in <a href="#app1-cells-13-01887" class="html-app">Supplementary File S1</a>) shows an upregulation of POMC expression after 24 h of HC-HA/PTX3 treatment (15 µg/mL). The quantification of POMC protein levels (28 kDa), which were normalized to GAPDH (37 kDa). The mean POMC level in the HC-HA/PTX3 group was considered to be 100%. N = 6 mice/group. Data are mean ± SEM. Unpaired Student’s <span class="html-italic">t</span>-test. t = 3.69, df = 12, ** <span class="html-italic">p</span> &lt; 0.01.</p>
Full article ">Figure 4
<p><b>Long-term HC-HA/PTX3 treatment attenuated capsaicin-evoked responses in dorsal root ganglia (DRG) neurons, which involved increased endogenous pro-opiomelanocortin (POMC) signaling.</b> (<b>A</b>) Left: Images showing calcium response of cultured wildtype (WT) mouse DRG neurons to capsaicin (CAP, 0.3 μM, 20 s), which are marked in red. Scale bars, 50 μm. Right: The quantification shows that the percentage of capsaicin-responsive neurons was significantly decreased by HC-HA/PTX3 (15 µg/mL, 24 h). This effect of HC-HA/PTX3 was blocked by co-treatment with CTOP [a mu-opioid receptor (MOR) antagonist,10 nM, 24 h] with HC-HA/PTX3. N = 4–8 mice/group. Data are mean ± SEM. One-way ANOVA followed by Bonferroni post hoc test. F<sub>(2,15)</sub> = 19.22, *** <span class="html-italic">p</span> &lt; 0.001 versus vehicle. <sup>##</sup> <span class="html-italic">p</span> &lt; 0.01 versus HC-HA/PTX3. (<b>B</b>) Left: Representative traces show that capsaicin (0.3 μM, 20 s) evoked an increase in [Ca<sup>2+</sup>]i in WT mouse DRG neurons, which was significantly reduced by HC-HA/PTX3 (15 µg/mL, 24 h) but not in those co-treated with CTOP. Right: Quantification of responses of individual neurons in each group, as shown in A. Data are mean ± SEM. One-way ANOVA followed by Bonferroni post hoc test. F<sub>(2, 326)</sub> = 15.18, *** <span class="html-italic">p</span> &lt; 0.001 versus vehicle. <sup>##</sup> <span class="html-italic">p</span> &lt; 0.01 versus HC-HA/PTX3. (<b>C</b>) Left: Images showing calcium response of DRG neurons from MOR conditioning knockout (cKO) mice to capsaicin (0.3 μM, 20 s). Scale bars, 50 μm. Right: The percentage of capsaicin-responsive neurons following HC-HA/PTX3 (15 µg/mL, 24 h) or vehicle treatment. N = 5 mice/group. Data are mean ± SEM. Unpaired Student’s <span class="html-italic">t</span>-test. t = 0.34, df = 8, <span class="html-italic">p</span> &gt; 0.05. (<b>D</b>) Representative traces show that capsaicin (0.3 μM, 20 s) evoked an increase in [Ca<sup>2+</sup>]i in DRG neurons from MOR cKO mice. Right: Quantification of calcium responses to capsaicin of individual MOR cKO DRG neurons from the vehicle-treated and HC-HA/PTX3-treated groups, as shown in C. Data are mean ± SEM. Unpaired Student’s <span class="html-italic">t</span>-test. t = 0.82, df = 260, <span class="html-italic">p</span> &gt; 0.05.</p>
Full article ">
14 pages, 2122 KiB  
Article
The Fundamental Neurobiological Mechanism of Oxidative Stress-Related 4E-BP2 Protein Deamidation
by Davis Joseph
Int. J. Mol. Sci. 2024, 25(22), 12268; https://doi.org/10.3390/ijms252212268 - 15 Nov 2024
Viewed by 4492
Abstract
Memory impairment is caused by the absence of the 4E-BP2 protein in the brain. This protein undergoes deamidation spontaneously in the neurons. 4E-BP2 deamidation significantly alters protein synthesis in the neurons and affects the balance of protein production required for a healthy nervous [...] Read more.
Memory impairment is caused by the absence of the 4E-BP2 protein in the brain. This protein undergoes deamidation spontaneously in the neurons. 4E-BP2 deamidation significantly alters protein synthesis in the neurons and affects the balance of protein production required for a healthy nervous system. Any imbalance in protein production in the nervous system causes neurodegenerative diseases. Discovering what causes 4E-BP2 deamidation will make it possible to control this balance of protein production and develop effective treatments against neurodegenerative diseases such as Alzheimer’s and Parkinson’s. The purpose of this work is to discover the neurobiological mechanism that causes the deamidation reaction in the 4E-BP2 protein by performing immunoblotting in the retinal ganglia, the optic nerve, the dorsal root ganglia, the sciatic nerve, and the whole brain, extracted via dissection from 2-month-old, Wild-type male mice. The results show that axons and their unique properties cause neuron-specific 4E-BP2 deamidation in the nervous system, confirming conclusively that axons are the critical factors behind the fundamental neurobiological mechanism of 4E-BP2 protein deamidation. Full article
(This article belongs to the Special Issue Antioxidants in Health and Diseases)
Show Figures

Figure 1

Figure 1
<p>4E-BP2 western blot of the whole brain, the optic nerve, and the retinal ganglia from 2-month-old WT mice, using GAPDH as the control: (<b>A</b>) Immunoblotting data. (<b>B</b>) Bonferroni multiple comparisons test of the deamidation ratios of the three organs studied. The star (“*”) between columns symbolizes a significant difference with a <span class="html-italic">p</span>-value of less than 0.05 between results, whereas “ns” stands for “not significant”. Three stars (“***”) between columns symbolize a significant difference with a <span class="html-italic">p</span>-value of less than 0.001 between results. More stars mean a more substantial difference between results.</p>
Full article ">Figure 2
<p>4E-BP2 western blot of the sciatic nerve and the dorsal root ganglia (DRG) from 2-month-old WT mice, using GAPDH as the control: (<b>A</b>) Immunoblotting data. (<b>B</b>) <span class="html-italic">T</span>-test comparing the two organs’ deamidation ratios. Three stars (“***”) between columns symbolize a significant difference between results with a <span class="html-italic">p</span>-value of less than 0.001.</p>
Full article ">Figure 3
<p><span class="html-italic">T</span>-test comparison between the deamidation ratios of the whole brain and the sciatic nerve. Four stars (“****”) between columns symbolize a significant difference between results with a <span class="html-italic">p</span>-value of less than 0.0001.</p>
Full article ">Figure 4
<p>Six-step chemical reaction of deamidation occurring in 4E-BP2 at positions N99 and N102.</p>
Full article ">Figure 5
<p>Biochemical flow sheet of the impact of the axon on protein production in the mammalian organism (all images used are royalty-free, except for the 5′ cap structure image, which the author made, and the 4E-BP and eIF4E images obtained using AlphaFold [<a href="#B47-ijms-25-12268" class="html-bibr">47</a>,<a href="#B48-ijms-25-12268" class="html-bibr">48</a>,<a href="#B49-ijms-25-12268" class="html-bibr">49</a>]).</p>
Full article ">
21 pages, 4570 KiB  
Article
Levodopa Impairs Lysosomal Function in Sensory Neurons In Vitro
by Oyedele J. Olaoye, Asya Esin Aksoy, Santeri V. Hyytiäinen, Aia A. Narits and Miriam A. Hickey
Biology 2024, 13(11), 893; https://doi.org/10.3390/biology13110893 - 2 Nov 2024
Cited by 1 | Viewed by 1135
Abstract
Parkinson’s disease (PD) is the second-most common neurodegenerative disease worldwide. Patients are diagnosed based upon movement disorders, including bradykinesia, tremor and stiffness of movement. However, non-motor signs, including constipation, rapid eye movement sleep behavior disorder, smell deficits and pain are well recognized. Peripheral [...] Read more.
Parkinson’s disease (PD) is the second-most common neurodegenerative disease worldwide. Patients are diagnosed based upon movement disorders, including bradykinesia, tremor and stiffness of movement. However, non-motor signs, including constipation, rapid eye movement sleep behavior disorder, smell deficits and pain are well recognized. Peripheral neuropathy is also increasingly recognized, as the vast majority of patients show reduced intraepidermal nerve fibers, and sensory nerve conduction and sensory function is also impaired. Many case studies in the literature show that high-dose levodopa may induce or exacerbate neuropathy in PD, which is thought to involve levodopa’s metabolism to homocysteine. Here, we treated primary cultures of dorsal root ganglia and a sensory neuronal cell line with levodopa to examine effects on cell morphology, mitochondrial content and physiology, and lysosomal function. High-dose levodopa reduced mitochondrial membrane potential. At concentrations observed in the patient, levodopa enhanced immunoreactivity to beta III tubulin. Critically, levodopa reduced lysosomal content and also reduced the proportion of lysosomes that were acidic, thereby impairing their function, whereas homocysteine tended to increase lysosome content. Levodopa is a critically important drug for the treatment of PD. However, our data suggest that at concentrations observed in the patient, it has deleterious effects on sensory neurons that are not related to homocysteine. Full article
(This article belongs to the Special Issue Lysosomes and Diseases Associated with Its Dysfunction)
Show Figures

Figure 1

Figure 1
<p>Primary cultures of DRGs were prepared and treated for 24 h or 7 days in hypoxia (hypoxia to mimic endogenous conditions and prevent levodopa auto-oxidation). Cultures were then examined for mitochondrial membrane potential (tetramethylrhodamine, methyl ester (TMRM)), reactive oxygen species (ROS) using dihydroethidium, beta III tubulin and lysosome content using Lysotracker red (red dots in the green DRG soma) and lysosome acidity (Lysotracker red + Lysosensor green, red + green = yellow dots in green DRG soma) as detailed in the text. Cells from the 50B11 cell line were also treated, then cultured in hypoxia for 24 h, and lysosome content was examined, as detailed in the text, using Lysotracker (red dots in the cell soma at top right). Figure made using BioRender.</p>
Full article ">Figure 2
<p>Time-dependent cytotoxicity of rotenone in DRGs. Cells were fixed and stained for MAP2, a pan-neuronal marker. (<b>A</b>–<b>F</b>) Example photomicrographs following 7 days (168 h) of treatment; (<b>A</b>) = control, (<b>B</b>) = 1 nM, (<b>C</b>) = 10 nM, (<b>D</b>) = 100 nM, (<b>E</b>) = 500 nM, (<b>F</b>) = 1 µM. Scale bar in (<b>F</b>) = 200 µm and is for all photomicrographs. Photomicrographs are not edited except for contrast, and they depict the MAP2 channel. The graph shows the time-dependent effect of rotenone on DRGs. Symbols in white and shades of grey to black show data from individual technical replicates. Symbols in shades of orange to brown show experimental means. * <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; 0.0001.</p>
Full article ">Figure 3
<p>Effect of levodopa on mitochondrial membrane potential (ΔΨ<sub>M</sub>) in primary sensory neurons (DRGs) cultured in hypoxia. Data are normalized to percent of control cells treated with 0 µM levodopa and 0 nM rotenone in hypoxia. (<b>A</b>) A dose of 30 µM levodopa increases ΔΨ<sub>M</sub> at 24 h in hypoxia; however, this effect is lost at 7 days and instead 300 µM levodopa inhibits ΔΨ<sub>M</sub>. Photomicrographs show TMRM staining in DRG soma treated with 0 µM levodopa (left) or 300 µM levodopa (LD, right) in hypoxia. Scale bar = 10 µm, for both images. (<b>B</b>) (left) Beta III tubulin immunocytochemistry shows no effect of high-dose (300 µM) levodopa on soma size; however, mean percent fluorescence for ATP5b (right), a mitochondrial marker, was reduced. (<b>C</b>) No impact of levodopa is observed in the context of parkinsonism (rotenone) at 24 h. (<b>D</b>) At 7 days, the deleterious effect of 300 µM levodopa is maintained in mild ΔΨ<sub>M</sub> inhibition (1 nM rotenone). Further, both 30 µM and 300 µM levodopa reduce ΔΨ<sub>M</sub> caused by 10 nM rotenone. No additive effects of levodopa are observed at stronger ΔΨ<sub>M</sub> inhibition caused by 500 nM rotenone. Data in (<b>A</b>–<b>D</b>) are shown as box plots of technical replicates with whiskers depicting 5–95% percentiles, black circles depicting remaining data points, lines depicting medians and “+” symbols depicting means. Light and dark orange circles show experiment means. (<b>A</b>,<b>C</b>,<b>D</b>): dashed lines show 100% (mean of control cells not treated with levodopa or rotenone) for reference. * <span class="html-italic">p</span> &lt; 0.05, *** <span class="html-italic">p</span> &lt; 0.001, **** <span class="html-italic">p</span> &lt; 0.0001.</p>
Full article ">Figure 4
<p>Effect of levodopa on oxidative stress in primary sensory neurons (DRGs) cultured in hypoxia. (<b>A</b>) Alone, only the highest concentration of levodopa (300 µM) induces mild oxidative stress. (<b>B</b>) In the context of parkinsonism (rotenone), levodopa increases oxidative stress at 1 nM rotenone. At moderate mitochondrial inhibition (10 nM, 100 nM), levodopa tended to reduce oxidative stress, possibly due to its ability to accept electrons [<a href="#B33-biology-13-00893" class="html-bibr">33</a>] and this effect is lost at 500 nM rotenone. Data are technical replicates and are shown as box plots with whiskers depicting 5–95% percentiles, black circles depicting remaining data points, lines depicting medians and “+” symbols depicting means. Light and dark orange circles show means for each experiment. *** <span class="html-italic">p</span> &lt; 0.001, **** <span class="html-italic">p</span> &lt; 0.0001 post hoc tests, following ANOVA, as discussed in text. Photomicrograph for 0 µM levodopa shows example original image of cells treated with 0 µM levodopa for 7 days and stained using dihydroethidium, with outlined area shown zoomed in and its corresponding segmentation; no cells were positive for oxidative stress. Photomicrograph for 300 µM levodopa shows example original image of cells treated with 300 µM levodopa for 7 days, then stained with dihydroethidium, with outlined area shown zoomed in and its corresponding segmentation; two cells were positive for oxidative stress in the zoomed in area. Arrowheads in photomicrographs point to DRG soma negative for oxidative stress (e.g., 0 µM levodopa). Arrows in photomicrographs point to DRG soma positive for oxidative stress. Photomicrographs are not modified, except for cropping for the zoomed images. Image processing picked out only DRG soma that contained reactive oxygen species. Note that the segmentation was for the entire image; here, we have used zoomed in areas to better demonstrate the segmentation. Scale bars = 200 µm.</p>
Full article ">Figure 5
<p>Effect of levodopa on beta III tubulin in primary sensory neurons (DRGs) cultured in hypoxia. (<b>A</b>) Alone, 3µM and 30 µM levodopa increased percent area immunoreactive for beta III tubulin. (<b>B</b>) In the context of parkinsonian sensory neurons (treated with rotenone), levodopa tended to increase percent area positive for beta III tubulin at 30 µM and 300 µM. (<b>C</b>) We examined fluorescence intensity for beta III tubulin, at the level of individual neurites in cells. Cells treated with levodopa only again showed increased fluorescence at 30 µM. (<b>D</b>) In the context of parkinsonian sensory neurons (treated with rotenone), 30 µM and 300 µM levodopa tended to increase fluorescence for beta III tubulin. Data are technical replicates shown as box plots with whiskers depicting 5–95% percentiles, black circles depicting remaining data points, lines depicting medians and “+” symbols depicting means. Light orange circles depict experimental means. * <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 post hoc tests, following ANOVA, as discussed in text. Photomicrographs show example original images of cells treated for 24 h in hypoxia with 0 µM, 3 µM, 30 µM, 300 µM levodopa, only, then stained for beta III tubulin. Photomicrographs are not modified. Scale bar bottom right = 200 µm, for all images.</p>
Full article ">Figure 6
<p>Effect of levodopa alone on lysosome content and acidity in primary DRGs cultured in hypoxia for 7 days. (<b>A</b>) A dose of 300 µM levodopa reduced lysosome content in DRG soma. (<b>B</b>) Acidity of lysosomes was impaired by 300 µM levodopa, because less lysosomes labelled with Lysotracker were co-labelled with Lysosensor. Data are shown as box plots with whiskers depicting 5–95% percentiles, black circles depicting remaining data points, lines depicting medians and “+” symbols depicting means. *** <span class="html-italic">p</span> &lt; 0.001. Orange circles show experiment means. Photomicrographs depict control (left) and levodopa-treated (right) soma. Lysosomes were identified as Lysotracker (red)-positive puncta, and Lysosensor (green) was used to show appropriate acidity; thus, properly acidified lysosomes show co-localised (yellow) puncta. Scale bar = 10 µm, for both images.</p>
Full article ">Figure 7
<p>Effect of levodopa on lysosomes in the 50B11 immortalised sensory cell line, treated and incubated in hypoxia for 24 h. (<b>A</b>) A dose of 300 µM levodopa reduced lysosome content in comparison to control-treated cells and cells treated with 30 µM levodopa. (<b>B</b>) Both 30 µM and 300 µM levodopa reduced lysosome content when cells are treated with entacapone. (<b>C</b>) Photomicrographs show example images of cells treated with 0 µM or 300 µM levodopa for 24 h in hypoxia and labelled with Lysotracker (red) and Hoechst (nuclei). Photomicrographs are not modified except for brightness and contrast. Scale bar = 20 µm, for both photomicrographs. Outlines show examples of CellProfiler segmentations of lysosomes (red) at 20 pixels and 60 pixels from individual nuclei (blue). Note: For segmentation, the entire nucleus was required to be within the field of view, and not to touch the image border. The green lines show the appropriate boundaries. (<b>D</b>) Individual lysosome size per boundary, showing that in the context of entacapone, 300 µM levodopa in particular reduces lysosome size. (<b>E</b>) Number of lysosomes per boundary. In the context of entacapone, both 30 µM and 300 µM levodopa reduce lysosome number. Data in graphs are shown as box plots with whiskers depicting 5–95% percentiles, black circles depicting remaining data points, lines depicting medians and “+” symbols depicting means. * <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. Light orange circles show experiment means. Entacapone = 1 µM [<a href="#B38-biology-13-00893" class="html-bibr">38</a>].</p>
Full article ">
17 pages, 2545 KiB  
Article
miR-30c-5p Gain and Loss of Function Modulate Sciatic Nerve Injury-Induced Nucleolar Stress Response in Dorsal Root Ganglia Neurons
by Raquel Francés, Jorge Mata-Garrido, Miguel Lafarga, María A. Hurlé and Mónica Tramullas
Int. J. Mol. Sci. 2024, 25(21), 11427; https://doi.org/10.3390/ijms252111427 - 24 Oct 2024
Viewed by 2931
Abstract
Neuropathic pain is a prevalent and debilitating chronic syndrome that is often resistant to treatment. It frequently arises as a consequence of damage to first-order nociceptive neurons in the lumbar dorsal root ganglia (DRG), with chromatolysis being the primary neuropathological response following sciatic [...] Read more.
Neuropathic pain is a prevalent and debilitating chronic syndrome that is often resistant to treatment. It frequently arises as a consequence of damage to first-order nociceptive neurons in the lumbar dorsal root ganglia (DRG), with chromatolysis being the primary neuropathological response following sciatic nerve injury (SNI). Nevertheless, the function of miRNAs in modulating this chromatolytic response in the context of neuropathic pain remains unexplored. Our previous research demonstrated that the intracisternal administration of a miR-30c mimic accelerates the development of neuropathic pain, whereas the inhibition of miR-30c prevents pain onset and reverses established allodynia. In the present study, we sought to elucidate the role of miR-30c-5p in the pathogenesis of neuropathic pain, with a particular focus on its impact on DRG neurons following SNI. The organisation and ultrastructural changes in DRG neurons, particularly in the protein synthesis machinery, nucleolus, and Cajal bodies (CBs), were analysed. The results demonstrated that the administration of a miR-30c-5p mimic exacerbates chromatolytic damage and nucleolar stress and induces CB depletion in DRG neurons following SNI, whereas the administration of a miR-30c-5p inhibitor alleviates these effects. We proposed that three essential cellular responses—nucleolar stress, CB depletion, and chromatolysis—are the pathological mechanisms in stressed DRG neurons underlying neuropathic pain. Moreover, miR-30c-5p inhibition has a neuroprotective effect by reducing the stress response in DRG neurons, which supports its potential as a therapeutic target for neuropathic pain management. This study emphasises the importance of miR-30c-5p in neuropathic pain pathogenesis and supports further exploration of miRNA-based treatments. Full article
(This article belongs to the Special Issue Molecular Mechanisms of mRNA Transcriptional Regulation: 2nd Edition)
Show Figures

Figure 1

Figure 1
<p>miR-30c-5p modulation effects on the chromatolysis developed by dorsal root ganglia neurons after spared nerve injury. (<b>A</b>–<b>H</b>) Dissociated dorsal root ganglia (DRG) neurons double stained with propidium iodide (PI, red) and Lamin B1 (green). Note the prominent NBs and round nuclei in sham rats treated with vehicle (<b>A</b>,<b>E</b>), miR-30c-5p mimic (<b>B</b>), or miR-30c-5p inhibitor (<b>F</b>), reflecting a normal distribution of the protein synthesis machinery and nuclear location. DRG neurons from day-5 (<b>C</b>) or day-10 SNI rats (<b>G</b>) exhibited central chromatolysis with dispersion and severe loss of NBs in the centre of the neuronal body, accumulations of Nissl substance at the marginal cytoplasm, and peripheral displacement of the nucleus, which were aggravated by treatment with miR-30c-5p mimic (<b>D</b>). Administration of miR-30c-5p inhibitor reduced the chromatolytic response observed after SNI (<b>H</b>). (<b>I</b>,<b>J</b>) Percentage of neurons showing chromatolysis. (<b>K</b>,<b>L</b>) Percentage of neurons showing eccentricity of the nucleus. The percentage of damaged neurons and eccentric nuclei was determined in 1000 neurons per rat (n = 3 rats per group). ** <span class="html-italic">p</span> &lt; 0.01, *** <span class="html-italic">p</span> &lt; 0.001 vs. Sham; # <span class="html-italic">p</span> &lt; 0.05, ## <span class="html-italic">p</span> &lt; 0.01, ### <span class="html-italic">p</span> &lt; 0.001 vs. SNI (Two-way ANOVA followed by the Bonferroni post hoc test). Scale bar: 5 µm.</p>
Full article ">Figure 2
<p>Electron micrographs illustrating the ultrastructural characteristics of dorsal root ganglia neurons after administration of miR-30c-5p mimic or inhibitor to SNI rats. In dorsal root ganglia (DRG) neurons from sham (<b>A</b>) and SNI rats treated with miR-30c-5p inhibitor (<b>C</b>), the most prominent organelles are the NBs, composed of RER cisterns (<b>C</b>, arrow) and rosettes of free polyribosomes (<b>A</b>, arrow). Bundles of neurofilaments (NF) interspersed between NBs, profiles of Golgi complexes, and mitochondria are also apparent. In DRGs from SNI rats treated with vehicle (<b>B</b>) or miR-30c-5p mimic (<b>D</b>), the NBs disaggregated, leaving an extensive cleared chromatolytic area in the centre of the cell body, free of NBs. The increased number of NFs and the abundance of mitochondria (M)—some of which are very small (&lt;0.5 µm)—in chromatolytic areas are also noteworthy. Scale bar: 5 µm.</p>
Full article ">Figure 3
<p>miR-30c-5p modulation effects on the nucleolar organisation of dorsal root ganglion neurons after spared nerve injury. (<b>A</b>–<b>H</b>) Dissociated dorsal root ganglia (DRG) neurons double immunostained for upstream binding factor (UBF, green) and Lamin B1 (red). DRG neurons from sham rats treated with vehicle (<b>A</b>,<b>E</b>), miR-30c-5p mimic (<b>B</b>), or miR-30c-5p inhibitor (<b>F</b>), and day-10 SNI rats treated with miR-30c-5p inhibitor (<b>H</b>) presented a normal UBF distribution as small dots corresponding to FCs. In contrast, DRG neurons from day-5 SNI rats treated with vehicle (<b>C</b>) or miR-30c-5p mimic (<b>D</b>) and day-10 SNI rats treated with vehicle (<b>G</b>) showed segregation of UBF nucleolar staining into one or a few giant FCs. (<b>I</b>,<b>J</b>) The percentage of neurons showing UBF-positive giant FCs was determined in 1000 neurons per rat (n = 3 rats per group); ** <span class="html-italic">p</span> &lt; 0.01, *** <span class="html-italic">p</span> &lt; 0.001 vs. Sham; ### <span class="html-italic">p</span> &lt; 0.001 vs. SNI (Two-way ANOVA followed by the Bonferroni post hoc test). Scale bar: 5 µm.</p>
Full article ">Figure 4
<p>Representative electron micrographs illustrating the ultrastructural nucleolar characteristics of dorsal root ganglia neurons after administration of miR-30c-5p mimic or inhibitor to SNI rats. Sham (<b>A</b>) and SNI rats treated with miR-30c-5p inhibitor (<b>C</b>) exhibit the typical nucleolar organisation of DRG neurons, characterised by the presence of numerous small-sized fibrillar centres (*, FCs), surrounded by a ring of dense fibrillar component (DFC), and areas of granular component (GC), preferentially at the nucleolar periphery. SNI rats treated with vehicle (<b>B</b>) or with miR-30c-5p mimic (<b>D</b>) present severe nucleolar alterations, including the formation of enlarged FCs and segregation of large masses of GC and DFC at the nucleolar periphery. Scale bar: 2 µm.</p>
Full article ">Figure 5
<p>miR-30c-5p modulation effects on the number of Cajal bodies in dorsal root ganglion neurons after spared nerve injury. Representative images of dissociated DRG neurons immunolabeled for coilin (green) and counterstained with propidium iodide ((PI), red). Example of neurons showing 0 (<b>A</b>), 1 (<b>B</b>), and 2 (<b>C</b>) CBs. (<b>D</b>,<b>E</b>) Quantitative analysis of the percentage of neurons carrying 0, 1, or more than 2 CBs in each of our experimental groups. The number of CBs per neuron was determined in 1000 neurons per rat, in 3 rats of each group (sham; SNI + vehicle; SNI + miR-30c-5p inhibitor; SNI + miR-30c-5p mimic). The quantification analysis indicates that, regardless of the experimental condition, most neurons present 1 CB. There is a significant increase in the percentage of neurons showing more than 2 CBs in SNI rats treated with miR-30c-5p inhibitor. The proportion of neurons without CBs is significantly increased in SNI rats treated with vehicle or miR-30c-5p mimic. (<b>F</b>,<b>G</b>) Electron microscopy of CBs in DRG neurons from SNI rats treated with miR-30c-5p inhibitor showing 3 CBs (<b>F</b>) and a hypertrophic CB physically close to the nucleolus (<b>G</b>). * <span class="html-italic">p</span> &lt; 0.05, *** <span class="html-italic">p</span> &lt; 0.01 vs. Sham; ### <span class="html-italic">p</span> &lt; 0.001 vs. SNI). (Two-way ANOVA followed by the Bonferroni post hoc test). Scale bar: 5 µm. Scale bar: 2 µm.</p>
Full article ">
31 pages, 9469 KiB  
Article
Elucidation of Medusozoan (Jellyfish) Venom Constituent Activities Using Constellation Pharmacology
by Angel A. Yanagihara, Matías L. Giglio, Kikiana Hurwitz, Raechel Kadler, Samuel S. Espino, Shrinivasan Raghuraman and Baldomero M. Olivera
Toxins 2024, 16(10), 447; https://doi.org/10.3390/toxins16100447 - 17 Oct 2024
Viewed by 1361
Abstract
Within the phylum Cnidaria, sea anemones (class Anthozoa) express a rich diversity of ion-channel peptide modulators with biomedical applications, but corollary discoveries from jellyfish (subphylum Medusozoa) are lacking. To bridge this gap, bioactivities of previously unexplored proteinaceous and small molecular weight (~15 kDa [...] Read more.
Within the phylum Cnidaria, sea anemones (class Anthozoa) express a rich diversity of ion-channel peptide modulators with biomedical applications, but corollary discoveries from jellyfish (subphylum Medusozoa) are lacking. To bridge this gap, bioactivities of previously unexplored proteinaceous and small molecular weight (~15 kDa to 5 kDa) venom components were assessed in a mouse dorsal root ganglia (DRG) high-content calcium-imaging assay, known as constellation pharmacology. While the addition of crude venom led to nonspecific cell death and Fura-2 signal leakage due to pore-forming activity, purified small molecular weight fractions of venom demonstrated three main, concentration-dependent and reversible effects on defined heterogeneous cell types found in the primary cultures of mouse DRG. These three phenotypic responses are herein referred to as phenotype A, B and C: excitatory amplification (A) or inhibition (B) of KCl-induced calcium signals, and test compound-induced disturbances to baseline calcium levels (C). Most notably, certain Alatina alata venom fractions showed phenotype A effects in all DRG neurons; Physalia physalis and Chironex fleckeri fractions predominantly showed phenotype B effects in small- and medium-diameter neurons. Finally, specific Physalia physalis and Alatina alata venom components induced direct excitatory responses (phenotype C) in glial cells. These findings demonstrate a diversity of neuroactive compounds in jellyfish venom potentially targeting a constellation of ion channels and ligand-gated receptors with broad physiological implications. Full article
(This article belongs to the Section Animal Venoms)
Show Figures

Figure 1

Figure 1
<p>Biochemical signaling pathways that result in increased intracellular calcium concentration ([Ca<sup>2+</sup>]<sub>i</sub>). Different cell processes can lead to the influx of extracellular Ca<sup>2+</sup> or the release of Ca<sup>2+</sup> from the intracellular stores, both leading to increased [Ca<sup>2+</sup>]<sub>i</sub>. When assessing the effects of venom component toxins by constellation pharmacology using high extracellular potassium [K<sup>+</sup>] pulses (e.g., 25 mM KCl) as a depolarizing stimulus, three different effects could be observed according to the interaction of the venom components to one or more of the mechanistic pathways depicted here: amplification (phenotype A), which augments or prolongs subsequent membrane depolarization duration (e.g., potential mechanisms include blocking voltage-gated potassium channels or delayed inactivation of sodium channels); inhibition (phenotype B), which decreases or shortens the membrane depolarization duration (e.g., potential mechanisms include blocking voltage-gated sodium or calcium channels or activation of hyperpolarizing ion channels including GABA or glycine receptors); and direct effects (phenotype C) on voltage-gated ion channels and ligand-gated ion channels, including G-protein coupled receptors (GPCRs), causing a depolarizing elevation of cytosolic calcium levels. Finally, as with other calcium imaging-based assays, calcium ionophores and cytolytic events involving the loss of cell membrane integrity can result in a transient increase in the intracellular Ca<sup>2+</sup>, which is most often irreversible and leads to cell death with leakage of the Fura-2.</p>
Full article ">Figure 2
<p>Output of the constellation pharmacology assay. <b>(A</b>) Representative tracing of two pulses of KCl with the application and incubation with phosphate-buffered saline (PBS) as a control. The application of extracellular KCl causes membrane depolarization, leading to an elevation of intracellular calcium and thus an increase in the Fura-2 signal. Note that the signal peak heights are reproducible and reversible; (<b>B</b>) Representative images of dorsal root ganglia (DRG) cells indicating Fura-2 intensity at the time of PBS incubation (<span class="html-italic">t</span><sub>1</sub>) and during the application of KCl (<span class="html-italic">t</span><sub>2</sub>) from A are shown.</p>
Full article ">Figure 3
<p>Representation of the different phenotypes evoked by venom fractions. <span class="html-italic">X</span>-axis is time in minutes and <span class="html-italic">Y</span>-axis represents the ratiometric emission of the Fura-2 excited at 340 nm and 380 nm. (<b>A</b>) Phenotype A: amplification of the KCl-induced Fura-2 signal. This phenotype could be elicited by voltage-gated potassium channel (VGKC) inhibitors, blockers of the voltage-gated sodium channel (VGSC) inhibition, or modulation of ligand-gated receptors; (<b>B</b>) phenotype B: inhibition of the KCl-induced Fura-2 signal. VGSC inhibitors or modulators of ligand-gated receptors are examples that could lead to this phenotype; (<b>C</b>) phenotype C: direct effect consisting of a spontaneous (independent from the depolarization stimulus) increase in [Ca<sup>2+</sup>]<sub>i</sub> upon incubation with the venom fraction. VGSC activators or ligand-gated receptor modulators could directly induce an increase in the influx of Ca<sup>2+</sup>; (<b>D</b>) example of dual effects. Complex venom fractions or isolated compounds could elicit a combination of phenotypes (typically A and C, or B and C) on the same cell. Some VGKC blockers, such as the conotoxin κM-RIIIJ, for instance, could cause both a direct effect and amplification [<a href="#B21-toxins-16-00447" class="html-bibr">21</a>].</p>
Full article ">Figure 4
<p>Porin activity of the <span class="html-italic">Alatina alata</span> crude venom using constellation pharmacology: (<b>A</b>) Representative traces of DRG cells exposed to crude venom for 3.5 min (blue and red boxes). The <span class="html-italic">X</span>-axis represents time. The <span class="html-italic">Y</span>-axis represents the ratiometric measurement of Fura-2 at 340 nm and 380 nm. K40 = 10 s pulse with 40 mM KCl; K = 10 s pulse of 25 mM KCl. Untreated crude venom (left) causes an increase in [Ca<sup>2+</sup>]i and sustained high ratiometric values. Crude venom denatured by heating at 100 °C for 10 min showed no effect; (<b>B</b>) comparison of the ratiometric signal (top row) with the deconvoluted signal of Fura-2 excited at 340 nm (middle row) and 380 nm (lower row); (<b>C</b>) bright field images of the DRG cells following 3.5 min treatment with crude or heat-denatured venom before and after the exposure with the crude venom. Scale bar = 100 μm.</p>
Full article ">Figure 5
<p>Size fractionation chromatograms of cnidarian venom preparations using size exclusion column high-performance liquid chromatography (SEC-HPLC). Major peaks are labeled: Peak I, molecular weight (MW) of ~250 kDa to 60 kDa comprising large proteins; Peak II MW ~50 kDa to 30 kDa comprising proteins including porins; Peak III MW ~15 kDa to 5 kDa comprising peptides and molecules investigated in this study; Peak IV MW ~3 kDa to 1 kDa. (<b>A</b>) <span class="html-italic">P. physalis</span> venom; (<b>B</b>) <span class="html-italic">C. fleckeri</span> venom; (<b>C</b>) <span class="html-italic">A. alata</span> venom. The fractions highlighted in Peak III were analyzed as subfractions (Peak III A–D) as well as pooled together.</p>
Full article ">Figure 6
<p>Overall comparison of the effect of venom Peak III from three medusozoan species using constellation pharmacology: (<b>A</b>) proportion of different neuron responses; (<b>B</b>) profile of representative cells comprising panel A. Two profiles of each neuron class are shown. From top to bottom: large-diameter unlabeled neurons (large gray circle), medium-diameter GFP-positive neurons (peptidergic nociceptors) (green square), medium-diameter IB4-positive neurons (non-peptidergic nociceptors) (orange square), and small-unlabeled neurons (gray triangle). The <span class="html-italic">x</span> axis represents time. The <span class="html-italic">y</span> axis represents the ratiometric measurement of Fura-2 at 340 nm and 380 nm. K = 10-s pulse with 25 mM KCl. Vehicle control: white box = 3.5 min incubation with PBS. The incubation with the venom fractions is represented by the blue (<span class="html-italic">P. physalis</span>), orange (<span class="html-italic">C. fleckeri</span>), and green (<span class="html-italic">A. alata</span>) boxes.</p>
Full article ">Figure 7
<p>Effect of the Peak III pooled venom fractions from <span class="html-italic">Physalia physalis</span> on DRG cells: (<b>A</b>) proportion of different cell types responding to the peptide fraction of <span class="html-italic">P. physalis</span> using two different concentrations; (<b>B</b>) representative calcium traces from panel A exposed to the venom fraction for 3.5 min (blue box). Three profiles of each neuron class are shown. From top to bottom: glial cells, large-unlabeled neurons (L1–L4), large-GFP positive neurons (L5–L6), medium-GFP positive neurons (peptidergic nociceptors), medium-size IB4 positive neurons (non-peptidergic nociceptors), and small-unlabeled neurons. The <span class="html-italic">X</span> axis represents time. The <span class="html-italic">Y</span> axis represents the ratiometric measurement of Fura-2 at 340 nm and 380 nm. K = 10 s pulse with 25 mM KCl; A = 10 s pulse of AITC; M = 10 s pulse of menthol; C = 10 s pulse of capsaicin; K40 = 10 sec pulse of 40 mM KCl. Vehicle control: white box = 2.5 min incubation with PBS. Grey box = 3.5 min incubation with the conotoxin κM-RIIIJ.</p>
Full article ">Figure 8
<p>Effect of the Peak III pooled venom fraction from <span class="html-italic">Chironex fleckeri</span> on DRG cells: (<b>A</b>) proportion of different cell classes responding to the peptide fraction of <span class="html-italic">C. fleckeri</span> using two different concentrations; (<b>B</b>) profile of representative cells from panel A exposed to the venom fraction for 3.5 min (orange box). Three profiles of each neuron class are shown. From top to bottom: glial cells, large-unlabeled neurons (L1–L4), large-GFP positive neurons (L5–L6), medium-GFP positive neurons (peptidergic nociceptors), medium-size IB4 positive neurons (non-peptidergic nociceptors), and small-unlabeled neurons. The <span class="html-italic">X</span> axis represents time. The <span class="html-italic">Y</span> axis represents the ratiometric measurement of Fura-2 at 340 nm and 380 nm. K = 10 s pulse with 25 mM KCl; A = 10 s pulse of AITC; M = 10 s pulse of menthol; C = 10 s pulse of capsaicin; K40 = 10 s pulse of 40 mM KCl. Vehicle control:white box = 3.5 min incubation with PBS. Grey box = 3.5 min incubation with the conotoxin κM-RIIIJ.</p>
Full article ">Figure 9
<p>Effects of Peak III pooled venom fractions from <span class="html-italic">Alatina alata</span> on DRG cells: (<b>A</b>) proportions of responses in different cell classes at two different concentrations; (<b>B</b>) time-course Fura-2 signal intensity response profile of representative cells from panel A exposed to Peak III for 3.5 min (green box). Three profiles of each neuron class are shown. From top to bottom: glial cells, large-unlabeled neurons (L1–L4), large-GFP positive neurons (L5–L6), medium-GFP positive neurons (peptidergic nociceptors), medium-size IB4 positive neurons (non-peptidergic nociceptors), and small-unlabeled neurons. The <span class="html-italic">X</span> axis represents time. The <span class="html-italic">Y</span> axis represents the ratiometric measurement of Fura-2 at 340 nm and 380 nm. K = 10 s pulse with 25 mM KCl; A = 10 s pulse of AITC; M = 10 s pulse of menthol; C = 10 s pulse of capsaicin; K40 = 10 s pulse of 40 mM KCl. Vehicle control: white box = 3.5 min incubation with PBS. Grey box = 3.5 min incubation with the conotoxin κM-RIIIJ.</p>
Full article ">Figure 10
<p>Bioactivity-resolution within Peak III of <span class="html-italic">A. alata</span> using constellation pharmacology: (<b>A</b>) chromatogram of individual and pooled fractions comprising Peak III of <span class="html-italic">Alatina alata</span> venom shown in <a href="#toxins-16-00447-f004" class="html-fig">Figure 4</a>. Each shade of green (A–D) represents the different subfractions used for the constellation pharmacology assay; (<b>B</b>) proportion of neurons responding to the different subfractions classified by type of response.</p>
Full article ">Figure 11
<p>Venom-derived toxins in paralysis and pain: (<b>A</b>) Motor neuron–muscle synapse and toxin targets. Flaccid paralysis (left): the block of voltage-gated sodium channels (VGSC) inhibits the transmission of the electric signal to the neuronal terminus; the block of the voltage-gated calcium channels (VGCC) prevents the fusion of the synaptic vesicles for the release of the neurotransmitter acetylcholine (ACh). The block of the muscarinic (mAChR) and nicotinic (nAChR) acetylcholine receptors in the postsynaptic membrane inhibits the transduction of the signal into the myocyte. Tetanic paralysis: the block of voltage-gated potassium channels (VGKC) and the activation or activity prolongation of the VGSC result in more neurotransmitter release and overstimulation of the muscle. Modified from Trim and Trim 2013 [<a href="#B75-toxins-16-00447" class="html-bibr">75</a>]; (<b>B</b>) scheme of the pain-related receptors and the pain-signaling transmission to the central nervous system (CNS). Activation of the pain transducer receptors acid-sensing ion channel (ASIC) and transient receptor potential channel ankyrin 1 (TRPA1) and vanilloid 1 (TRPV1) lead to the activation of nociceptive neurons. The inhibition of either of these receptors has been related to analgesia. Additionally, the block of the VGCC at the nociceptor terminal abolishes the signal transduction to the CNS. Modified from Bohlen 2012 [<a href="#B76-toxins-16-00447" class="html-bibr">76</a>].</p>
Full article ">Figure 12
<p>Jellyfish collection, intact cnidae recovery and venom preparation, as described in 5.1. From left to right (arrows denote steps in this process), schematic images of <span class="html-italic">Chironex fleckeri</span>, <span class="html-italic">Alatina alata</span> and <span class="html-italic">Physalia physalis</span>; schematic representation of freshly excised tentacles (pink) in chilled 1 M trisodium citrate; spontaneously shed intact cnidae are represented as pink ovals; light micrograph of <span class="html-italic">Alatina alata</span> eurytele [<a href="#B86-toxins-16-00447" class="html-bibr">86</a>] in live replete tentacle (scale bar 65 micron, inset 15 micron); 0.5 mm mesh sieving of citrate solution of tentacles to remove shed intact cnidae from depleted tentacles; pelleted cnidae are then ruptured in a chilled pressure cell disruptor (French Press); complete ruptured slurry is quickly centrifuged to pellet collagenous structural components (capsule walls and tubules), total cnidae content venom (supernatant) is then aliquoted and snap frozen in liquid nitrogen [<a href="#B1-toxins-16-00447" class="html-bibr">1</a>].</p>
Full article ">Figure 13
<p>DRG dissection and dissociation protocol. Lumbar DRGs (L1–L6) were removed from the mice and transferred to a petri dish where the nerves were trimmed for cleaning. Individual DRGs were excised to expose the cells. Enzymatic tissue dissociation was performed using a trypsin solution for 20 min at 37 °C. DRGs were then mechanically dissociated by pipetting using Pasteur pipettes with decreasing diameter tips and sieved through a 45-µm cell strainer. Cell suspension was centrifuged and pelleted cells were resuspended in 100 μL of the culture medium. Aliquots of the resuspended solution were plated in 24-well plates within a silicone ring and incubated overnight in a cell culture chamber at 37 °C.</p>
Full article ">Figure 14
<p>Workflow to screen for effects of cnidarian venom fractions on dorsal root ganglia (DRG) cells using constellation pharmacology. Primary cultures were plated, as described in the <a href="#sec5-toxins-16-00447" class="html-sec">Section 5</a>. Eighteen hours later, DRG cells were incubated with Fura-2 for 1 h (30 min in culture chamber and 30 min at room temperature). Continuous Fura-2 calcium imaging was conducted during pulse-chase exposure of the cells in the flow cell well to control solutions, depolarizing KCl pulses and the specific pharmacological agents comprising constellation pharmacology. Venom fractions were applied in between successive KCl pulses (blue box). Cell responses (Fura-2 ratiometric intensity over time) were transformed into profiles of individual cells to be analyzed. KCl, potassium chloride 25 mM; AITC, allyl isothiocyanate (agonist of the transient receptor potential ankyrin 1 channel or TRPA1); Men, menthol (agonist of the transient receptor potential melastatin 8 channel or TRPM8); Cap, capsaicin (agonist of the transient receptor potential vanilloid 1 channel or TRPV1).</p>
Full article ">
22 pages, 2170 KiB  
Article
The Endocannabinoid System of the Nervous and Gastrointestinal Systems Changes after a Subnoxious Cisplatin Dose in Male Rats
by Yolanda López-Tofiño, Mary A. Hopkins, Ana Bagues, Laura Boullon, Raquel Abalo and Álvaro Llorente-Berzal
Pharmaceuticals 2024, 17(10), 1256; https://doi.org/10.3390/ph17101256 - 24 Sep 2024
Viewed by 840
Abstract
Background/Objectives: Cisplatin, a common chemotherapy agent, is well known to cause severe side effects in the gastrointestinal and nervous systems due to its toxic and pro-inflammatory effects. Although pharmacological manipulation of the endocannabinoid system (ECS) can alleviate these side effects, how chemotherapy affects [...] Read more.
Background/Objectives: Cisplatin, a common chemotherapy agent, is well known to cause severe side effects in the gastrointestinal and nervous systems due to its toxic and pro-inflammatory effects. Although pharmacological manipulation of the endocannabinoid system (ECS) can alleviate these side effects, how chemotherapy affects the ECS components in these systems remains poorly understood. Our aim was to evaluate these changes. Methods: Male Wistar rats received cisplatin (5 mg/kg, i.p.) or saline on day 0 (D0). Immediately after, serial X-rays were taken for 24 h (D0). Body weight was recorded (D0, D1, D2 and D7) and behavioural tests were performed on D4. On D7, animals were euthanized, and gastrointestinal tissue, dorsal root ganglia (DRGs) and brain areas were collected. Expression of genes related to the ECS was assessed via Rt-PCR, while LC-MS/MS was used to analyse endocannabinoid and related N-acylethanolamine levels in tissue and plasma. Results: Animals treated with cisplatin showed a reduction in body weight. Cisplatin reduced gastric emptying during D0 and decreased MAGL gene expression in the antrum at D7. Despite cisplatin not causing mechanical or heat sensitivity, we observed ECS alterations in the prefrontal cortex (PFC) and DRGs similar to those seen in other chronic pain conditions, including an increased CB1 gene expression in L4/L5 DRGs and a decreased MAGL expression in PFC. Conclusions: A single dose of cisplatin (5 mg/kg, i.p.), subnoxious, but capable of inducing acute gastrointestinal effects, caused ECS changes in both gastrointestinal and nervous systems. Modulating the ECS could alleviate or potentially prevent chemotherapy-induced toxicity. Full article
Show Figures

Graphical abstract

Graphical abstract
Full article ">Figure 1
<p>Body weight in grams (<b>A</b>), food intake per animal in grams (<b>B</b>), and water intake per animal in millilitres (<b>C</b>) on day 0, 1, 2, and 7 after treatment with saline (n = 6) or a single dose of cisplatin (i.p., 5 mg/kg; n = 6) in male Wistar rats. Data are expressed as Mean ± SEM of saline- vs. cisplatin-treated animals. Statistical significance set at <span class="html-italic">p</span> &lt; 0.05. Repeated measures ANOVA followed by a <span class="html-italic">t</span>-test or Kruskal–Wallis in each time-point: * significant difference.</p>
Full article ">Figure 2
<p>Radiographic analysis of gastrointestinal motility 1, 2, 3, 4, 6, 8, and 24 h after treatment with saline (n = 6) or a single dose of cisplatin (i.p., 5 mg/kg; n = 6) in the stomach (<b>A</b>), small intestine (<b>B</b>), caecum (<b>C</b>), and colorectum (<b>D</b>) in male Wistar rats. Contents, in arbitrary units, are expressed as Mean ± SEM of saline- vs. cisplatin-treated animals. Statistical significance set at <span class="html-italic">p</span> &lt; 0.05. Repeated measures ANOVA followed by a <span class="html-italic">t</span>-test or Kruskal–Wallis in each time-point: * significant difference. Representative X-rays of rats treated either with saline or a single dose of cisplatin (i.p., 5 mg/kg) at time-point 2 h (<b>E</b>), 8 h (<b>F</b>), and 24 h (<b>G</b>); scale bar: 3 cm. S: stomach; SI: small intestine; C: caecum; FP: faecal pellet (in colorectum).</p>
Full article ">Figure 3
<p>Behavioural analysis 4 days after treatment with saline (n = 6) or a single dose of cisplatin (i.p., 5 mg/kg; n = 6) in male Wistar rats. Locomotor activity in beam counts/30 min (<b>A</b>), mechanical tactile sensitivity measured by von Frey filaments as mechanical threshold (grams) (<b>B</b>), and heat tactile sensitivity measured by Hargreaves’ test as thermal threshold (seconds) (<b>C</b>). Data are expressed as Mean ± SEM of saline- vs. cisplatin-treated animals. Statistical significance set at <span class="html-italic">p</span> &lt; 0.05.</p>
Full article ">Figure 4
<p>Gene expression of the <span class="html-italic">mgll</span> gene in the antrum of male rats after treatment with saline (n = 6) or a single dose of cisplatin (i.p., 5 mg/kg; n = 5). The mean percentage of the saline-treated group (2<sup>−ΔCT</sup>) for saline is expressed as Mean ± SEM of saline- vs. cisplatin-treated animals. Statistical significance set at <span class="html-italic">p</span> &lt; 0.05. <span class="html-italic">t</span>-test: * significant difference.</p>
Full article ">Figure 5
<p>Gene expression in the nervous system of male rats after treatment with saline (n = 5–6) or a single dose of cisplatin (i.p., 5 mg/kg; n = 4–6). Levels of <span class="html-italic">cnr1</span> in the dorsal root ganglia L4 (<b>A</b>) and L5 (<b>B</b>); levels of <span class="html-italic">cnr2</span> (<b>C</b>), <span class="html-italic">mgll</span> (<b>D</b>), and <span class="html-italic">ppara</span> (<b>E</b>) in the prefrontal cortex (PFC); and levels of <span class="html-italic">cnr1</span> (<b>F</b>) and <span class="html-italic">mgll</span> (<b>G</b>) in the amygdala (AMY). The mean percentage of the saline-treated group (2<sup>−ΔCT</sup>) for saline is expressed as Mean ± SEM of saline- vs. cisplatin-treated animals. Statistical significance set at <span class="html-italic">p</span> &lt; 0.05. <span class="html-italic">t</span>-test: * significant difference.</p>
Full article ">Figure 6
<p>Hypothetical mechanism of the effect of cisplatin on the pain corticoamygdalar pathway. In normal conditions, activation of basolateral amygdala (BLA) pyramidal cells induces the secretion of 2-AG by pyramidal cells in the medial prefrontal cortex (mPFC) thanks to the activation of mGluR5. Administration of 5 mg/kg of cisplatin induced a significant increase in the expression levels of MAGL in the prefrontal cortex, which in turn would deplete the 2AG secreted by prefrontal pyramidal cells and, hence, reduce the activation of CB1 in prefrontal GABAergic interneurons. Prefrontal GABAergic interneurons would then increase their inhibitory influx on mPFC pyramidal cells projecting to GABAergic interneurons in the central nucleus of the amygdala (CeA), which would lead to increased firing of the amygdalar signalling outputs, an effect frequently observed in pain-related emotional responses in animal models of arthritis, colitis, formalin-induced inflammation, and neuropathic pain. The yellow-filled arrow represents the increase in MAGL gene expression observed in the PFC, while yellow-striped arrows represent the hypothetical changes that would follow it in the corticoamygdalar pathway. Schematic representation adapted from [<a href="#B50-pharmaceuticals-17-01256" class="html-bibr">50</a>].</p>
Full article ">Figure 7
<p>Experimental outline. Male Wistar rats were treated either with cisplatin (i.p., 5 mg/kg; n = 6) or saline (i.p., 0.9% NaCl; n = 6) on day 0. Immediately after, animals were administered barium contrast by gavage and submitted to a radiographic evaluation of gastrointestinal transit during the first 24 h after treatment. Food intake and body weight were recorded on days 1, 2, and 7 and locomotor activity (actimeter) and pain-related behaviours (von Frey and Hargreave’s test) were measured on day 4. On day 7, gastrointestinal and nervous tissues were extracted for further analyses.</p>
Full article ">
17 pages, 5153 KiB  
Article
Axonal Growth and Fasciculation of Spinal Neurons Promoted by Aldynoglia in Alkaline Fibrin Hydrogel: Influence of Tol-51 Sulfoglycolipid
by Vinnitsa Buzoianu-Anguiano, Alejandro Arriero-Cabañero, Alfonso Fernández-Mayoralas, Mabel Torres-Llacsa and Ernesto Doncel-Pérez
Int. J. Mol. Sci. 2024, 25(17), 9173; https://doi.org/10.3390/ijms25179173 - 23 Aug 2024
Viewed by 882
Abstract
Traumatic spinal cord injury (tSCI) has complex pathophysiological events that begin after the initial trauma. One such event is fibroglial scar formation by fibroblasts and reactive astrocytes. A strong inhibition of axonal growth is caused by the activated astroglial cells as a component [...] Read more.
Traumatic spinal cord injury (tSCI) has complex pathophysiological events that begin after the initial trauma. One such event is fibroglial scar formation by fibroblasts and reactive astrocytes. A strong inhibition of axonal growth is caused by the activated astroglial cells as a component of fibroglial scarring through the production of inhibitory molecules, such as chondroitin sulfate proteoglycans or myelin-associated proteins. Here, we used neural precursor cells (aldynoglia) as promoters of axonal growth and a fibrin hydrogel gelled under alkaline conditions to support and guide neuronal cell growth, respectively. We added Tol-51 sulfoglycolipid as a synthetic inhibitor of astrocyte and microglia in order to test its effect on the axonal growth-promoting function of aldynoglia precursor cells. We obtained an increase in GFAP expression corresponding to the expected glial phenotype for aldynoglia cells cultured in alkaline fibrin. In co-cultures of dorsal root ganglia (DRG) and aldynoglia, the axonal growth promotion of DRG neurons by aldynoglia was not affected. We observed that the neural precursor cells first clustered together and then formed niches from which aldynoglia cells grew and connected to groups of adjacent cells. We conclude that the combination of alkaline fibrin with synthetic sulfoglycolipid Tol-51 increased cell adhesion, cell migration, fasciculation, and axonal growth capacity, promoted by aldynoglia cells. There was no negative effect on the behavior of aldynoglia cells after the addition of sulfoglycolipid Tol-51, suggesting that a combination of aldynoglia plus alkaline fibrin and Tol-51 compound could be useful as a therapeutic strategy for tSCI repair. Full article
(This article belongs to the Special Issue The Function of Glial Cells in the Nervous System)
Show Figures

Figure 1

Figure 1
<p>Subacute and chronic lesions of the spinal cord in human patients observed by magnetic resonance imaging. Syringomyelia cyst in a child due to idiopathic causes (<b>a</b>) shows the enlargement of the cavity within the spinal cord (<b>b</b>). The dotted areas (<b>c</b>,<b>d</b>) indicate a decade of evolution in a patient from subacute injury (<b>c</b>) to chronic spinal injury (<b>d</b>) produced by traumatic disk herniation at the second to third vertebral thoracic levels, T2–T3 (see arrows). Note the formation of syringomyelia below the lesion at the beginning (<b>c</b>) and how ten years later, the spinal cord becomes necrotic (<b>d</b>).</p>
Full article ">Figure 2
<p>Alkaline conditions for fibrin hydrogel gelation and Tol-51 sulfoglycolipid mass spectrum with molecular structures. (<b>a</b>) Fibrin hydrogels 1 h after gelation in physiological and alkaline conditions, with normal or double (2×) concentration of TBS ingredients. (<b>b</b>) Signal at 550.33 in the negative ion mode mass spectrum for the sulfoglycolipid Tol-51. In the box, the carbon structure and the ball-and-stick model of the Tol-51 molecule are depicted at top and bottom, respectively. Note the greater transparency in the fibrin hydrogel under alkaline conditions (<b>a</b>) and the folding possibility in aliphatic moiety for Tol-51 compound (<b>b</b>).</p>
Full article ">Figure 3
<p>Increasing expression of GFAP after differentiation of neural precursor cells to aldynoglia phenotype in alkaline fibrin hydrogels. After one week in cell culture, aldynoglia–GFP cells in fibrin hydrogels made clusters (<b>e</b>,<b>i</b>,<b>m</b>) in the GFP signaling areas, but not in control cells (<b>a</b>). No significant differences in constitutive GFP expression were obtained in areas for cell differentiation analysis (<b>q</b>). However, in these GFP+ areas, the alkaline fibrin variants had significantly higher GFAP expression (<b>j</b>,<b>n</b>) than controls (<b>b</b>,<b>f</b>), as shown in (<b>r</b>). The white arrows point to GFAP bridges between clusters of aldynoglia. Nuclei revealed by Hoechst agent (<b>c</b>,<b>g</b>,<b>k</b>,<b>o</b>) and merging zones of GFP, GFAP signals (<b>d</b>,<b>h</b>,<b>l</b>,<b>p</b>) are shown One-way ANOVA followed by Tukey’s test were used in statistical analysis (* <span class="html-italic">p</span> = 0.01, ** <span class="html-italic">p</span> = 0.001 and *** <span class="html-italic">p</span> = 0.0001); magnification bars at 500 µm.</p>
Full article ">Figure 4
<p>GFAP expression for aldynoglia cultured in alkaline fibrins in the presence of sulfoglycolipid Tol-51. Significant differences in the areas of GFAP expression were obtained in the presence of alkaline fibrins and sulfoglycolipid Tol-51 versus control fibrin, but no additive effect was observed (<b>j</b>). Less aggregation of aldynoglia–GFP cells (<b>g</b>–<b>i</b>) was observed in alkaline (2×) fibrin than in other fibrin hydrogels (<b>a</b>–<b>f</b>). White arrows point to some of the GFAP glial process that linked aldynoglia cell clusters inside hydrogels (<b>d</b>,<b>f</b>). Saturation zones (white) were not considered for GFAP expression evaluation. One-way ANOVA followed by Kruskal–Wallis test: * <span class="html-italic">p</span> = 0.05, ** <span class="html-italic">p</span> = 0.005 and *** <span class="html-italic">p</span> = 0.0005; magnification bars at 100 µm.</p>
Full article ">Figure 5
<p>Dynamic cellular interaction of aldynoglia–GFP cells and DRG explants in alkaline fibrin hydrogel. Cell cultures was recorded by time-lapse video up to 96 h, and profuse migration of aldynoglia–GFP cells to DRGs was observed in this period (<b>a</b>–<b>d</b>). Differentiation of neurospheres into aldynoglia inside alkaline fibrin began at approximately 24 h in cell culture (<b>a</b>). Aldynoglia–GFP interaction with DRG explant, fibrin degradation, and cell migration are indicated (<b>a’</b>–<b>d’</b>). Alkaline fibrin degradation and massive cellular invasion of DRGs by aldynoglia–GFP cells are denoted by white dotted lines and arrows, respectively (<b>a’</b>–<b>d’</b>). On the third and fourth day (<b>c</b>,<b>d</b>), large areas of cell migration with aldynoglia–GFP and ganglion cells were also observed throughout the alkaline fibrin matrix—see green dotted area (<b>c’</b>,<b>d’</b>). A, aldynoglia; DRG, dorsal root ganglion; N, neurosphere; arrowheads, aldynoglia–GFP and DRG cell contact site; white dotted line, zones of alkaline fibrin hydrogel degradation by aldynoglia–GFP cells; green dotted area, cell migration zone in alkaline fibrin.</p>
Full article ">Figure 6
<p>Alkaline fibrin and Tol-51 sulfoglycolipid during axonal growth promotion of spinal neurons by aldynoglia. DRGs were plated on alkaline fibrin containing aldynoglia cells and incubated in the presence of Tol-51 for ten days. Axonal outgrowth of spinal neurons was obtained after subtracting the central body of ganglia (dotted circles) from the area of total axonal outgrowth (dotted lines) revealed by monoclonal anti-BIII tubulin antibody, in red. Some promotion of DRG axonal growth by Tol-51 was observed in the absence of aldynoglia cells (<b>a</b>,<b>d</b>,<b>g</b>,<b>j</b> and bar graph <b>m</b>) and in their presence (<b>b</b>,<b>e</b>,<b>h</b>,<b>k</b> and bar graph <b>n</b>), but differences were not significant. The presence of aldynoglia is revealed by GFP expression in merged images (<b>c</b>,<b>f</b>,<b>i</b>,<b>l</b>). Magnification bars at 500 µm. DRGs, dorsal root ganglia.</p>
Full article ">Figure 7
<p>Alkaline fibrin and Tol-51 sulfoglycolipid promoted axonal fasciculation of spinal neurons by aldynoglia. DRGs were plated on alkaline fibrin containing aldynoglia cells and incubated in the presence of Tol-51 for ten days. Neurons and axonal bundles were revealed by monoclonal anti-βIII tubulin antibody, in red. The fasciculation of spinal neurons was obtained by measuring the axonal bundle width at ≥8.0 µm, dotted circles; and 20 µm away from DRG body, solid line. Axonal bundle fasciculation by Tol-51 similar to control was observed in the absence of aldynoglia cells (<b>a</b>,<b>c</b>,<b>e</b>,<b>g</b>; and bar graph in <b>i</b>). The aldynoglia presence increased axonal bundle width (<b>b</b>,<b>d</b>,<b>f</b>,<b>h</b>,<b>j</b>) and a significant dose response for assayed Tol-51 concentrations were obtained (###; <span class="html-italic">p</span> = 0.0006). Values of (<b>i</b>) were compared with counterparts in (<b>j</b>); ** <span class="html-italic">p</span> = 0.01; **** <span class="html-italic">p</span> = 0.0001; one-way ANOVA. DRGs, dorsal root ganglia; magnification bars at 500 µm.</p>
Full article ">
14 pages, 4045 KiB  
Article
Dipeptidyl Peptidase (DPP)-4 Inhibitors and Pituitary Adenylate Cyclase-Activating Polypeptide, a DPP-4 Substrate, Extend Neurite Outgrowth of Mouse Dorsal Root Ganglia Neurons: A Promising Approach in Diabetic Polyneuropathy Treatment
by Masahiro Yamaguchi, Saeko Noda-Asano, Rieko Inoue, Tatsuhito Himeno, Mikio Motegi, Tomohide Hayami, Hiromi Nakai-Shimoda, Ayumi Kono, Sachiko Sasajima, Emiri Miura-Yura, Yoshiaki Morishita, Masaki Kondo, Shin Tsunekawa, Yoshiro Kato, Koichi Kato, Keiko Naruse, Jiro Nakamura and Hideki Kamiya
Int. J. Mol. Sci. 2024, 25(16), 8881; https://doi.org/10.3390/ijms25168881 - 15 Aug 2024
Cited by 1 | Viewed by 1225
Abstract
Individuals suffering from diabetic polyneuropathy (DPN) experience debilitating symptoms such as pain, paranesthesia, and sensory disturbances, prompting a quest for effective treatments. Dipeptidyl-peptidase (DPP)-4 inhibitors, recognized for their potential in ameliorating DPN, have sparked interest, yet the precise mechanism underlying their neurotrophic impact [...] Read more.
Individuals suffering from diabetic polyneuropathy (DPN) experience debilitating symptoms such as pain, paranesthesia, and sensory disturbances, prompting a quest for effective treatments. Dipeptidyl-peptidase (DPP)-4 inhibitors, recognized for their potential in ameliorating DPN, have sparked interest, yet the precise mechanism underlying their neurotrophic impact on the peripheral nerve system (PNS) remains elusive. Our study delves into the neurotrophic effects of DPP-4 inhibitors, including Diprotin A, linagliptin, and sitagliptin, alongside pituitary adenylate cyclase-activating polypeptide (PACAP), Neuropeptide Y (NPY), and Stromal cell-derived factor (SDF)-1a—known DPP-4 substrates with neurotrophic properties. Utilizing primary culture dorsal root ganglia (DRG) neurons, we meticulously evaluated neurite outgrowth in response to these agents. Remarkably, all DPP-4 inhibitors and PACAP demonstrated a significant elongation of neurite length in DRG neurons (PACAP 0.1 μM: 2221 ± 466 μm, control: 1379 ± 420, p < 0.0001), underscoring their potential in nerve regeneration. Conversely, NPY and SDF-1a failed to induce neurite elongation, accentuating the unique neurotrophic properties of DPP-4 inhibition and PACAP. Our findings suggest that the upregulation of PACAP, facilitated by DPP-4 inhibition, plays a pivotal role in promoting neurite elongation within the PNS, presenting a promising avenue for the development of novel DPN therapies with enhanced neurodegenerative capabilities. Full article
(This article belongs to the Special Issue Peripheral Neuropathies: Molecular Research and Novel Therapy)
Show Figures

Figure 1

Figure 1
<p>Expression and activity of DPP-4 in the peripheral nervous system. (<b>A</b>) RT-PCR of DPP-4 mRNA on several tissues: liver, dorsal root ganglia (DRG), sciatic nerve (SCN), heart, and skeletal muscle (SM). (<b>B</b>–<b>D</b>) Immunostaining of DPP-4 protein on DRG. (<b>B</b>): DPP-4. (<b>C</b>): DAPI. (<b>D</b>): Merge. Scale bar: 50 μm. (<b>E</b>) Quantified DPP-4 activity of DRG and liver, μU: microunits; mgPro: mg Protein; ns: not significant between DRG and liver.</p>
Full article ">Figure 2
<p>Neurite outgrowth of DRG neurons by Diprotin A. (<b>A</b>–<b>D</b>) Representative fluorescence figures of DRG neurons cultured in the absence (<b>A</b>) or presence (<b>B</b>–<b>D</b>) of Diprotin A ((<b>B</b>): 1 μM, (<b>C</b>): 10 μM, (<b>D</b>): 50 μM). Scale bar: 50 μm. (<b>E</b>) Quantified neurite length in each DRG neuron. *: <span class="html-italic">p</span> &lt; 0.05 versus control; ns: not significant versus control.</p>
Full article ">Figure 3
<p>Quantified neurite length in each DRG neuron with or without sitagliptin. ***: <span class="html-italic">p</span> &lt; 0.001, ****: <span class="html-italic">p</span> &lt; 0.0001 versus control.</p>
Full article ">Figure 4
<p>Neurite outgrowth of DRG neurons by linagliptin. (<b>A</b>–<b>C</b>) Representative fluorescence figures of DRG neurons cultured in the absence (<b>A</b>) or presence (<b>B</b>,<b>C</b>) of linagliptin ((<b>B</b>): 10 nM, (<b>C</b>): 100 nM). Scale bar: 50 μm. (<b>D</b>) Quantified neurite length in each DRG neuron. *: <span class="html-italic">p</span> &lt; 0.05; ns: not significant versus control.</p>
Full article ">Figure 5
<p>Neurite outgrowth of DRG neurons by PACAP. (<b>A</b>–<b>C</b>) Representative fluorescence figures of DRG neurons cultured in the absence (<b>A</b>) or presence (<b>B</b>,<b>C</b>) of PACAP ((<b>B</b>): 0.1 μM, (<b>C</b>): 1 μM). Scale bar: 50 μm. (<b>D</b>) Quantified neurite length in each DRG neuron. ***: <span class="html-italic">p</span> &lt; 0.001, ****: <span class="html-italic">p</span> &lt; 0.0001 versus control.</p>
Full article ">Figure 6
<p>Neurite outgrowth of DRG neurons by NPY and SDF-1α. (<b>A</b>–<b>C</b>) Representative fluorescence figures of DRG neurons cultured in the absence (<b>A</b>) or presence of (<b>B</b>) NPY 0.1 μM and (<b>C</b>) SDF-1α 0.1 μM. Scale bar: 50 μm. (<b>D</b>) Quantified neurite length in each DRG neuron. ns: not significant versus control.</p>
Full article ">Figure 7
<p>Cell viability in 50B11 cells treated with DPP-4 inhibitors and PACAP. H<sub>2</sub>O<sub>2</sub> (−): no supplementation of H<sub>2</sub>O<sub>2</sub>. 0.1 mM H<sub>2</sub>O<sub>2</sub> (+): supplementation of 0.1 mM H<sub>2</sub>O<sub>2</sub>. 100% cell viability: control without H<sub>2</sub>O<sub>2</sub>. H<sub>2</sub>O<sub>2</sub>: hydrogen peroxide. ***: <span class="html-italic">p</span> &lt; 0.001, ****: <span class="html-italic">p</span> &lt; 0.0001 versus control with 0.1 mM H<sub>2</sub>O<sub>2</sub>.</p>
Full article ">
31 pages, 17050 KiB  
Article
SARS-CoV-2 Rapidly Infects Peripheral Sensory and Autonomic Neurons, Contributing to Central Nervous System Neuroinvasion before Viremia
by Jonathan D. Joyce, Greyson A. Moore, Poorna Goswami, Telvin L. Harrell, Tina M. Taylor, Seth A. Hawks, Jillian C. Green, Mo Jia, Matthew D. Irwin, Emma Leslie, Nisha K. Duggal, Christopher K. Thompson and Andrea S. Bertke
Int. J. Mol. Sci. 2024, 25(15), 8245; https://doi.org/10.3390/ijms25158245 - 28 Jul 2024
Cited by 3 | Viewed by 17268
Abstract
Neurological symptoms associated with COVID-19, acute and long term, suggest SARS-CoV-2 affects both the peripheral and central nervous systems (PNS/CNS). Although studies have shown olfactory and hematogenous invasion into the CNS, coinciding with neuroinflammation, little attention has been paid to susceptibility of the [...] Read more.
Neurological symptoms associated with COVID-19, acute and long term, suggest SARS-CoV-2 affects both the peripheral and central nervous systems (PNS/CNS). Although studies have shown olfactory and hematogenous invasion into the CNS, coinciding with neuroinflammation, little attention has been paid to susceptibility of the PNS to infection or to its contribution to CNS invasion. Here we show that sensory and autonomic neurons in the PNS are susceptible to productive infection with SARS-CoV-2 and outline physiological and molecular mechanisms mediating neuroinvasion. Our infection of K18-hACE2 mice, wild-type mice, and golden Syrian hamsters, as well as primary peripheral sensory and autonomic neuronal cultures, show viral RNA, proteins, and infectious virus in PNS neurons, satellite glial cells, and functionally connected CNS tissues. Additionally, we demonstrate, in vitro, that neuropilin-1 facilitates SARS-CoV-2 neuronal entry. SARS-CoV-2 rapidly invades the PNS prior to viremia, establishes a productive infection in peripheral neurons, and results in sensory symptoms often reported by COVID-19 patients. Full article
(This article belongs to the Special Issue Coronavirus Disease (COVID-19): Pathophysiology 5.0)
Show Figures

Graphical abstract

Graphical abstract
Full article ">Figure 1
<p><b>SARS-CoV-2 infection of TG and SCG in hACE2 and WT mice.</b> (<b>a</b>) SARS-CoV-2 RNA was detected at increasing concentration in TGs of hACE2 and WT mice in both inoculum groups from 3 to 6 dpi. The TG provides sensory innervation to the face, including the nasal septum, and sends projections to the brain stem, thereby providing an alternative entry point for SARS-CoV-2. No statistically significant differences were detected between the groups (F(7, 40) = 1.855, <span class="html-italic">p</span> = 0.1032). (<b>b</b>) SARS-CoV-2 RNA was detected at increasing concentrations in SCGs of hACE2 and WT mice in both inoculum groups from 3 to 6 dpi. The SCG provides sympathetic innervation to the salivary glands, lacrimal glands, and blood vessels of the head, neck, and brain. Three-way ANOVA detected a significant difference (F(7, 40) = 3.118, <span class="html-italic">p</span> = 0.0101) in RNA genome copy number. Tukey’s honestly significant difference (HSD) post hoc tests detected significant differences between the WT groups inoculated with 10<sup>3</sup> PFU assessed at 3 and 6 dpi (<span class="html-italic">p</span> = 0.048). (<b>c</b>) Percentage of SCG and TG neurons infected by 6 dpi in tissue sections from 10<sup>5</sup> PFU-inoculated hACE2 and WT mice. SCGs showed high levels of infection in each mouse type (93–96%) as did TGs (37–41%). (<b>d</b>) Immunofluorescence for SARS-CoV-2 nucleocapsid (SARS-N, grey)- and NeuN (red)-labeled neurons in TG and SCG sections at 6 dpi in 10<sup>5</sup> PFU-inoculated mice. SARS-N was more prevalent in hACE2 than in WT but observable in both. No SARS-N was detected in ganglia from uninfected animals. Neurons in SCG were particularly sensitive to infection; in the magnified single z-plane of the area shown in the yellow box on hACE2-infected SCGs, significant vacuolization can be observed in infected hACE2 SCG cells. Contrast for NeuN was increased in the z-plane to better illustrate residual NeuN immunoreactivity inside SARS-N-negative vacuoles in the SCGs. This cytopathology was common across numerous SCGs in hACE2 mice. (<b>e</b>) Immunofluorescence for SARS-CoV-2 spike (SARS-S, grey)- and NeuN (red)-labeled neurons in TG and SCG sections at 6 dpi in 10<sup>5</sup> PFU-inoculated mice. Immunostaining was similar to that for SARS-N, with greater SARS-S in hACE2 neurons, but present in both. Vacuolization was again observed in SCG neurons. SARS-S was absent in uninfected neurons. (<b>f</b>) Immunofluorescence for double stranded RNA (dsRNA, grey)- and NeuN (red)-labeled neurons in TG and SCG sections at 6 dpi in 10<sup>5</sup> PFU-inoculated mice. Immunostaining for dsRNA, a marker of viral replication, was similar to SARS-S and SARS-N immunostaining in TGs and SCGs. dsRNA was present in greater concentrations in SCGs than TGs and in hACE2 mice than WT mice but was present in both ganglia and both mouse types. Positive dsRNA immunostaining indicates SARS-CoV-2 genome replication in TGs and SCGs. Some nonspecific extracellular binding was observed in uninfected mice, but no intracellular immunofluorescence was observed. Data are the mean ± s.e.m. Log-transformed RNA genome copy numbers were statistically compared by three-way ANOVA (independent variables: inocula, days post infection, genotype). Pairwise comparisons were conducted using Tukey’s HSD post hoc tests. * <span class="html-italic">p</span> &lt; 0.05. <span class="html-italic">n</span> = 6 for all animals/timepoints/tissues (except 10<sup>5</sup> hACE2 6 dpi TG: <span class="html-italic">n</span> = 7; 10<sup>3</sup> hACE2 6 dpi TG: <span class="html-italic">n</span> = 5). Scale bars are 25 μm. See <a href="#app1-ijms-25-08245" class="html-app">Figure S2</a> for unmerged images. See <a href="#app1-ijms-25-08245" class="html-app">Figure S6</a> for additional antibody validation via Western blot, hACE2 genotyping, hACE2 protein expression, and RT-qPCR controls.</p>
Full article ">Figure 2
<p><b>SARS-CoV-2 infection of LS-DRG, including satellite glial cells, and LS spinal cord in hACE2 and WT mice.</b> (<b>a</b>) SARS-CoV-2 RNA was detected in increasing concentrations in LS-DRGs of hACE2 and WT mice in both inoculum groups from 3 to 6 dpi. The LS-DRG conveys sensory information (pain, pressure, position) from the periphery and organs to the spinal cord. Three-way ANOVA detected a significant difference (F(7, 41) = 4.590, <span class="html-italic">p</span> = 0.0007) in RNA genome copy number. Tukey’s HSD detected differences between the hACE2 and WT groups inoculated with 10<sup>5</sup> PFU at 3 dpi (<span class="html-italic">p</span> = 0.018). <span class="html-italic">n</span> = 6 for all animals/timepoints (except 10<sup>5</sup> hACE2 6 dpi: <span class="html-italic">n</span> = 7). (<b>b</b>) SARS-CoV-2 RNA was detected in lumbosacral spinal cords of hACE2 and WT mice in both inoculum groups at both time points. No statistically significant differences were detected between the groups (F(7, 25) = 1.3054, <span class="html-italic">p</span> = 0.2885). Samples sizes were four for 10<sup>5</sup> hACE2 mice at 3 dpi and five at 6 dpi; five for all timepoints for 10<sup>3</sup> hACE2 mice; four for all timepoints for 10<sup>5</sup> WT mice; and three for all timepoints for 10<sup>3</sup> WT mice. (<b>c</b>) Percentage of LS-DRG neurons infected by 6 dpi in tissue sections from 10<sup>5</sup> PFU-inoculated hACE2 and WT mice. A total of 42% of neurons were infected in hACE2 mice, and 24% were infected in WT mice. (<b>d</b>) Immunofluorescence for SARS-N (grey) and NeuN (red) in LS-DRG sections from 10<sup>5</sup> PFU-inoculated and uninfected hACE2 and WT mice at 6 dpi. SARS-N was more prevalent in hACE2 than in WT but observable in both. No SARS-N was detected in uninfected mice. Detection of RNA and SARS-N in peripheral neurons with no direct connection to the oronasopharynx suggests spread via hematogenous dissemination or via axonal transport. (<b>e</b>) Immunofluorescence for SARS-S (grey) and NeuN (red) in LS-DRG sections from 10<sup>5</sup> PFU-inoculated hACE2 and WT mice at 6 dpi. SARS-S immunostaining was similar to that of SARS-N, with neuronal staining in hACE2 mice and satellite glial cell (SGC) staining in WT mice. (<b>f</b>) Immunofluorescence for dsRNA (grey) and NeuN (red) in LS-DRG sections from 10<sup>5</sup> PFU-inoculated hACE2 and WT mice at 6 dpi. dsRNA immunostaining was similar to that of SARS-N and SARS-S with neuronal staining in hACE2 mice and SGC staining in WT mice. Presence of dsRNA indicated viral genome replication. (<b>g</b>) SARS-N (grey) was detected in numerous satellite glial cells (SGCs, glutamine synthetase (GS, red), as denoted by the yellow arrows, surrounding infected LS-DRG neurons at 6 dpi in 10<sup>5</sup> PFU-inoculated mice. See <a href="#app1-ijms-25-08245" class="html-app">Video S1</a> for 3D rendering of this image. (<b>h</b>) Representative image of immunofluorescence for SARS-N (grey) and NeuN (red) in spinal cord cross-sections from a 10<sup>5</sup> PFU-inoculated hACE2 mouse at 6 dpi. SARS-N was observed as discrete puncta in the neuronal cytoplasm, reminiscent of viral replication complexes (arrowheads in h1 and h2, which are magnified areas in the yellow boxes). See <a href="#app1-ijms-25-08245" class="html-app">Video S2</a> for 3D rendering of this image (<b>i</b>) Representative image of immunofluorescence for SARS-N (grey) and microglial marker Iba1 (red) in spinal cord cross-sections from a 10<sup>5</sup> PFU-inoculated hACE2 mouse at 6 dpi. Microglia processes are present throughout the cord, as are discrete SARS-N puncta. (<b>j</b>) Representative image of immunofluorescence for SARS-N (grey) and astrocyte marker S100B (red) in spinal cord cross-sections from a 10<sup>5</sup> PFU-inoculated hACE2 mouse at 6 dpi. (Independent variables: inocula, days post infection, genotype). Pairwise comparisons were conducted using Tukey’s HSD post hoc tests. * <span class="html-italic">p</span> &lt; 0.05. See <a href="#app1-ijms-25-08245" class="html-app">Figure S2</a> for unmerged LS-DRGs. See <a href="#app1-ijms-25-08245" class="html-app">Figure S3</a> for unmerged spinal cord sections and controls. See <a href="#app1-ijms-25-08245" class="html-app">Figure S6</a> for additional antibody validation via Western blot, hACE2 genotyping, hACE2 protein expression, and RT-qPCR controls.</p>
Full article ">Figure 3
<p><b>SARS-CoV-2 infection of the olfactory bulb and various brain regions in hACE2 and WT mice.</b> SARS-CoV-2 RNA was detected in increasing concentrations from 3 to 6 dpi in the olfactory bulb (<b>a</b>), hippocampus (<b>b</b>), cortex (<b>c</b>), brainstem (<b>d</b>), and cerebellum (<b>e</b>) of hACE2 and WT mice in both inoculum groups. Three-way ANOVA detected a significant difference (F(7, 41) = 11.825, <span class="html-italic">p</span> ≤ 0.0001) in RNA genome copy number in the olfactory bulb. Tukey’s HSD detected differences in the olfactory bulb between hACE2 and WT groups inoculated with 10<sup>5</sup> PFU assessed at 3 dpi (<span class="html-italic">p</span> &lt; 0.0001) as well as between those groups assessed at 6 dpi (<span class="html-italic">p =</span> 0.004). A significant difference (F(7, 41) = 5.433, <span class="html-italic">p</span> = 0.0002) was also detected in the hippocampi by three-way ANOVA; however, Tukey’s HSD revealed it occurred between non-biologically relevant comparisons. A significant difference (F(7, 41) = 7.217, <span class="html-italic">p</span> ≤ 0.0001) was also detected in the brainstem of the hACE2 and WT groups inoculated with 10<sup>5</sup> PFU assessed at 3 dpi (<span class="html-italic">p</span> = 0.0107). While differences were detected in the cortex (F(7, 41) = 6.302, <span class="html-italic">p</span> = &lt;0.0001) none were between relevant groups. A significant difference was detected in the cerebellum (F(7, 41) = 6.996, <span class="html-italic">p</span> ≤ 0.0001) of the WT groups inoculated with 10<sup>3</sup> PFU when compared between 3 and 6 dpi. (<b>f</b>) Heatmap showing recovery of infectious SARS-CoV-2 from homogenates of the olfactory bulb and specific brain regions assessed for viral RNA. Recovery of infectious virus varied across individual animals, with some having no regions with the recoverable virus and some with the virus in all regions. Of note, infectious virus was recovered from the hippocampi (5 PFU/mg homogenate) and brainstems (3 PFU/mg homogenate) of some WT mice, which are regions functionally impacted in COVID-19 disease. Data are the mean ± s.e.m. Log-transformed RNA genome copy numbers were statistically compared by three-way ANOVA (independent variables: inocula, days post infection, genotype). Pairwise comparisons were conducted using Tukey’s HSD post hoc tests. * <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 for all animals/timepoints/tissues (except 10<sup>5</sup> hACE2 6 dpi brain regions: <span class="html-italic">n</span> = 7). See <a href="#app1-ijms-25-08245" class="html-app">Figure S3</a> for control hACE2 brain sections and <a href="#app1-ijms-25-08245" class="html-app">Figure S4</a> for infected and control WT brain sections. See <a href="#app1-ijms-25-08245" class="html-app">Figure S6</a> for additional antibody validation via Western blot, hACE2 genotyping, hACE2 protein expression, and RT-qPCR controls.</p>
Full article ">Figure 4
<p><b>SARS-CoV-2 infection of primary PNS neuronal cultures from SCG, TG, and LS-DRG of hACE2 and WT mice.</b> Primary neuronal cultures were generated from SCG, TG, and LS-DRG from 8–10 wk old hACE2 and WT mice. (<b>a</b>) SARS-CoV-2 RNA was quantified by RT-qPCR separately in neuronal cells and media to generate a 4–5-day viral genome replication profile. Intracellular replication patterns were similar between hACE2 and WT neurons, although at reduced levels in WT neurons, with increasing viral RNA detected in media over the course of infection. hACE2 LS-DRGs had peaks in genome replication at ~48 hpi and ~96 hpi, indicating successive rounds of replication. Results are from three separate neuronal cultures, each with duplicate technical replicates per ganglion/timepoint. (<b>b</b>) Infectious virus was quantified by plaque assay on Vero E6 cells from cellular and media fractions of SCG, TG, and LS-DRG neuronal cultures to generate growth curves in primary neurons from hACE2 and WT mice. Infectious virus was not recovered from SCG neurons, indicating abortive infection, likely mediated by cytotoxicity. Infectious virus was recovered from TG and LS-DRG neurons, indicating productive infection of these neurons, although sustained production of viral progeny is dependent on hACE2. Results are based on sample sizes as reported for above for panel a. (<b>c</b>) hACE2 neuronal cultures: immunofluorescence for SARS-N (grey) and either tyrosine hydroxylase (TH, red) or Isolectin-B4 (IB4, red) to counterstain SCG and LS-DRG neurons, respectively, or glutamine synthetase (GS, red) to stain satellite glial cells. SARS-N was observed in neurons from each of the ganglia. Infected neurons were largely free of neurites by 1 dpi. At 3 dpi, many infected neurons exhibited cytopathologies such as degraded neurites, enlarged multi-nucleated cell bodies (arrow) compared to uninfected neurons (arrowhead), and SARS-N+ puncta reminiscent of viral replication compartments. See <a href="#app1-ijms-25-08245" class="html-app">Supplementary Video S3</a> for 3D rendering of LS-DRG at 3 dpi. See <a href="#app1-ijms-25-08245" class="html-app">Video S4</a> for 3D rendering of TG at 2 dpi. (<b>d</b>) WT neuronal cultures: immunofluorescence for SARS-N (grey) and either TH or IB4 (red) to counterstain neurons or GS (red) to stain satellite glial cells. Immunostaining revealed a similarly heterogenous infection of neurons and satellite glial cells as observed in hACE2 neurons. (<b>e</b>) Immunofluorescence for SARS-N (grey) and IB4 (red) to counterstain hACE2 LS-DRG neurons at 3 dpi shows a variety of phenotypes of infected cells, including neurons with a loss of membrane integrity (inset 1), SARS-N+ puncta within and surrounding neurons (inset 2), and seemingly healthy neurons with extensive neurites with strong SARS-N+ staining (arrow in inset 3). Infected satellite glial cells were also observed (arrowheads in inset 3); many appeared to be activated, noted by the presence of extended cellular processes. These findings are similar to immunostaining of LS-DRGs of hACE2 and WT mice in vivo, which also contained numerous infected satellite glial cells. (<b>f</b>) Percentage of hACE2 autonomic (SCG) and sensory (LS-DRG) cultured neurons positive for SARS-N were counted from 1 to 3 dpi. A small percentage of autonomic (SCG) neurons were visibly infected, with significant observable cell death, similar to in vivo observations. Infection in sensory (LS-DRG) neurons was consistent from 1 to 3 dpi, with ~5% infected. Infection of neurons ex vivo is less efficient than infection in vivo. Scale bar = 20 μm. Data are the mean ± s.e.m. See <a href="#app1-ijms-25-08245" class="html-app">Figure S5</a> for unmerged and control images. See <a href="#app1-ijms-25-08245" class="html-app">Figure S6</a> for additional antibody validation via Western blot, hACE2 genotyping, hACE2 protein expression, and RT-qPCR controls.</p>
Full article ">Figure 5
<p><b>Pre-viremic neuroinvasion of SARS-CoV-2 into PNS and functionally connected CNS tissues of hACE2 and WT mice as early as 18 h post-infection.</b> (<b>a</b>) Although no viral RNA was detected in blood, low levels of SARS-CoV-2 RNA were detected in PNS and CNS of both hACE2 and WT mice as early as 18 hpi. CNS invasion in hACE2 and WT mice was common in the olfactory bulb, hippocampus, cortex, and brainstem. Viral RNA was detected in PNS ganglia and the tissues they innervate (SCG-salivary gland, TG-brainstem) in both hACE2 and WT mice. Viral RNA was detected separately in the LS-DRGs and spinal cords of some mice. By 42 hpi, viral RNA was detected in blood in only one hACE2 mouse but had increased in the brainstem, TG, and olfactory bulb, indicating replication in these tissues. Sample sizes were hACE2 mice <span class="html-italic">n</span> = 10 (5 per timepoint) and WT mice <span class="html-italic">n</span> = 10 (5 per timepoint). (<b>b</b>) Heatmaps visually displaying RT-qPCR values from panel a. Neuroinvasion in both PNS and CNS occurs rapidly before detectable viremia, thereby indicating direct neural entry and trans-synaptic spread of SARS-CoV-2. Detection of viral RNA in the LS-DRGs but not the spinal cord in some mice and vice versa in others indicates that separate entry routes exist for the LS-DRG and spinal cord. Invasion of the cord likely occurs from the brainstem, as all mice with early spinal cord infection also had brainstem infection. Invasion of the LS-DRG may occur from the periphery. Infectious virus was not detected by plaque assay, indicating the virus had not yet started replicating at these early time points. See <a href="#app1-ijms-25-08245" class="html-app">Figure S6</a> for hACE2 genotyping, hACE2 protein expression, and RT-qPCR controls.</p>
Full article ">Figure 6
<p><b>Pre-viremic neuroinvasion of SARS-CoV-2 into PNS and functionally connected CNS tissues in Syrian golden hamsters as early as 18 h post-infection.</b> (<b>a</b>) Similar to hACE2 and WT mice, no viral RNA was detected in hamster blood at 18 hpi; however, low levels of SARS-CoV-2 RNA were detected in PNS and CNS tissues. Viral RNA was detected in PNS ganglia and the tissues they innervate (SCG-lacrimal gland, TG-brainstem, LS-DRG-spinal cord) as early as 18 hpi. By 42 hpi, viral RNA was increased in all PNS tissues (excluding SCGs), indicating replication in these tissues. Lacrimal glands harbored notably high viral RNA concentrations. Viral RNA copy number was stable in SCGs at 3 dpi (4 log) and had returned to levels observed at 18 hpi in TG and LS-DRGs (4 log and 3 log, respectively), indicating a burst of viral replication. Sample size was <span class="html-italic">n</span> = 9 (3 hamsters per timepoint). (<b>b</b>) Heatmaps visually displaying RT-qPCR values from panel a. Neuroinvasion in both PNS and CNS occurs rapidly before detectable viremia, thereby indicating direct neural entry and trans-synaptic spread of SARS-CoV-2, as observed in mice. (<b>c</b>) Immunofluorescence for SARS-N (grey) and NeuN (red) in LS-DRG, TG, and SCG sections from infected and uninfected hamsters at 18 hpi, 42 hpi, and 3 dpi. SARS-N appears in all ganglia by 18 hpi and continues to be present to 3 dpi. Significant pathology/vacuolization is observed in SCGs at 3 dpi, similar to the pathology observed in mice. (<b>d</b>) Results of von Frey threshold test showing a decrease in the amount of force required to elicit a withdrawal reflex, indicating allodynia as a consequence of infection. A paired samples two-tailed t-test found this decrease to be significant (t(8) = 7.606, <span class="html-italic">p =</span> 0.0001). Presence of SARS-CoV-2 neuroinvasion of LS-DRGs likely mediates this allodynia, which was noted in some animals (five of nine) as early as 18 hpi and all remaining animals (three of three) by 3 dpi. Scale bar = 20 μm. Data are the mean ± s.e.m. See <a href="#app1-ijms-25-08245" class="html-app">Figure S4</a> for immunostaining of infected hamster brain. See <a href="#app1-ijms-25-08245" class="html-app">Figure S6</a> for additional antibody validation via Western blot and RT-qPCR controls. **** <span class="html-italic">p</span> &lt; 0.0001.</p>
Full article ">Figure 7
<p><b>The role of NRP-1 during entry of SARS-CoV-2 in PNS neurons.</b> (<b>a</b>) Western blot of ganglia (SCG, TG, LS-DRG) from uninfected hACE2 and WT mice showing the presence of the transmembrane glycoprotein neuropilin-1 (NRP-1) on primary neurons and satellite glial cells in culture. NRP-1 has been shown to increase SARS-CoV-2 entry into non-neuronal cells. (<b>b</b>) Immunofluorescence for NRP-1 (red) and DAPI (blue) in primary neuronal cultures of LS-DRG neurons from uninfected hACE2 and WT mice, showing that NRP-1 is present across the soma and processes of neurons (arrows) as well as satellite glial cells (arrowheads). Immunostaining for NRP-1 was used to demonstrate spatial distribution of NRP-1 across the neurons in culture as a complement to our detection of NRP-1 in whole neuron homogenate by Western blot. (<b>c</b>) Treatment of hACE2 and WT LS-DRG neurons with the NRP-1 antagonist EG00229 prior to infection with SARS-CoV-2 significantly reduced viral RNA concentrations at 2 dpi, the initial peak of viral replication in hACE2 LS-DRG neurons as determined through our replication/growth curves, by 99.8% (t(4) = 4.896, <span class="html-italic">p</span> = 0.0081) in hACE2 neurons (DMSO: 1,739,333.3 SARS-CoV-2 RNA copies/200 ng RNA; EG00229: 3,362 SARS-CoV-2 RNA copies/200 ng RNA) and 86.7% (t(4) = 4.165, <span class="html-italic">p</span> = 0.0141) in WT neurons (DMSO: 158,933.3 SARS-CoV-2 RNA copies/200 ng RNA; EG00229: 21,233.3 SARS-CoV-2 RNA copies/200 ng RNA). Thus, NRP-1 is an entry factor mediating viral entry into neurons expressing hACE2 and also enhances viral entry into WT neurons. These data also indicate that additional host proteins are involved in neuronal entry. Scale bar = 20 μm. Data are the mean ± s.e.m. Results are from two separate neuronal cultures, each with duplicate technical replicates per ganglion. See <a href="#app1-ijms-25-08245" class="html-app">Figure S6</a> for additional antibody validation via Western blot, hACE2 genotyping, hACE2 protein expression, NRP-1 protein expression, and RT-qPCR controls. * <span class="html-italic">p</span> &lt; 0.05, ** <span class="html-italic">p</span> &lt; 0.01.</p>
Full article ">
23 pages, 17872 KiB  
Article
Lysophosphatidic Acid Receptors LPAR5 and LPAR2 Inversely Control Hydroxychloroquine-Evoked Itch and Scratching in Mice
by Caroline Fischer, Yannick Schreiber, Robert Nitsch, Johannes Vogt, Dominique Thomas, Gerd Geisslinger and Irmgard Tegeder
Int. J. Mol. Sci. 2024, 25(15), 8177; https://doi.org/10.3390/ijms25158177 - 26 Jul 2024
Viewed by 1345
Abstract
Lysophosphatidic acids (LPAs) evoke nociception and itch in mice and humans. In this study, we assessed the signaling paths. Hydroxychloroquine was injected intradermally to evoke itch in mice, which evoked an increase of LPAs in the skin and in the thalamus, suggesting that [...] Read more.
Lysophosphatidic acids (LPAs) evoke nociception and itch in mice and humans. In this study, we assessed the signaling paths. Hydroxychloroquine was injected intradermally to evoke itch in mice, which evoked an increase of LPAs in the skin and in the thalamus, suggesting that peripheral and central LPA receptors (LPARs) were involved in HCQ-evoked pruriception. To unravel the signaling paths, we assessed the localization of candidate genes and itching behavior in knockout models addressing LPAR5, LPAR2, autotaxin/ENPP2 and the lysophospholipid phosphatases, as well as the plasticity-related genes Prg1/LPPR4 and Prg2/LPPR3. LacZ reporter studies and RNAscope revealed LPAR5 in neurons of the dorsal root ganglia (DRGs) and in skin keratinocytes, LPAR2 in cortical and thalamic neurons, and Prg1 in neuronal structures of the dorsal horn, thalamus and SSC. HCQ-evoked scratching behavior was reduced in sensory neuron-specific Advillin-LPAR5−/− mice (peripheral) but increased in LPAR2−/− and Prg1−/− mice (central), and it was not affected by deficiency of glial autotaxin (GFAP-ENPP2−/−) or Prg2 (PRG2−/−). Heat and mechanical nociception were not affected by any of the genotypes. The behavior suggested that HCQ-mediated itch involves the activation of peripheral LPAR5, which was supported by reduced itch upon treatment with an LPAR5 antagonist and autotaxin inhibitor. Further, HCQ-evoked calcium fluxes were reduced in primary sensory neurons of Advillin-LPAR5−/− mice. The results suggest that LPA-mediated itch is primarily mediated via peripheral LPAR5, suggesting that a topical LPAR5 blocker might suppress “non-histaminergic” itch. Full article
(This article belongs to the Collection Feature Papers in “Molecular Biology”)
Show Figures

Graphical abstract

Graphical abstract
Full article ">Figure 1
<p>LPA release in the skin and thalamus evoked by itch stimulation. (<b>A</b>) Concentrations of LPA species in the thalamic region after intradermal injection of saline (vehicle), Cp 48/80 or HCQ. Tissue was collected 30–45 min after injection. The box is the interquartile range, whiskers show minimum and maximum, and scatters represent the mice. Each mouse is depicted as two scatters of each two samples. (<b>B</b>) Plasma LPAs of the mice shown in A, n = 8–10 per group. (<b>C</b>) Concentrations of LPA species in the skin on the side of the HCQ injection (ipsilateral) and the opposite side (contralateral). The scatters represent individual mice. Data were compared with a two-way ANOVA and subsequent post hoc <span class="html-italic">t</span>-test using an adjustment of alpha according to Šidák. * <span class="html-italic">p</span> &lt; 0.05, *** <span class="html-italic">p</span> &lt; 0.001. (<b>D</b>) Heatmap of mRNA expression of candidate genes involved in itch signaling in the CNS and periphery. The color lines represent the mice (n = 7). Abbreviations: ScN, sciatic nerve; DRG, dorsal root ganglia; SC, spinal cord; vThal, ventral thalamus; dThal, dorsal thalamus; SSC, somatosensory cortex. MRGPR, mas-related G protein-coupled receptor; AGMO, alkylglycerolmonooxygenase.</p>
Full article ">Figure 2
<p>LPAR5 reporter gene expression and localization in the skin. (<b>A</b>) X-Gal (blue) histology of skin sections in naïve LPAR5-β-Galactosidase reporter mice. The numbered rectangles show the regions used for zoom-in images. (<b>B</b>) β-Galactosidase immunofluorescence in the skin of naïve LPAR5-β-Galactosidase reporter mice. LPAR5-flfl mice were used as negative control. Cytokeratin 14 (CK14) was used as marker for keratinocytes, mast cell tryptase (MCT) for mast cells and cluster of differentiation CD11b for macrophages. DAPI is a nuclear counterstain. Eight LPAR5-LacZ mice were used to assess LPAR5 localization in the skin. White arrows point to double labeled cells.</p>
Full article ">Figure 3
<p>LPAR5 reporter gene expression and localization in DRGs. (<b>A</b>) X-Gal histology of the DRGs in naïve LPAR5-β-Galactosidase reporter mice. Ten LPAR5-LacZ mice were used to assess LPAR5 localization in the DRGs and other tissues. Numbered rectangles show the regions used for zoom-in images. (<b>B</b>) β-Galactosidase immunofluorescence in the DRGs of naïve LPAR5-β-Galactosidase reporter mice. LPAR5-flfl mice were used as negative control. Neuron subtypes showing LPAR5 reporter expression were assessed by co-immunofluorescence studies with markers for DRG subpopulations. CGRP, calcitonin gene-related peptide (peptidergic); SP, substance P (peptidergic); IB4, isolectin B4 (C-fiber glutamatergic); P2X3, purine receptor (C-fiber, purinergic); TRPV1, transient receptor potential channel (heat sensitive, nonmyelinated); TrkA, tyrosine kinase A (NGF-responsive). The rectangles show the area used for zoom-in and the arrow heads in the zoom-in images point to double positive immunoreactive spots.</p>
Full article ">Figure 4
<p>LPAR2 mRNA expression in brain regions by RNA scope technology. For each site, two exemplary images of two mice are shown for LPAR2 in comparison with the housekeeping positive control gene PPIA (Peptidylprolyl Isomerase A) and the negative control bacterial gene, dapB (dipicolinate reductase). Slides were developed with a colorimetric red HRP substrate, and positive reactions occur as red dots. Slides were counterstained with hematoxylin (blue). RNAscope studies were performed on three wildtype mice. Red rectangles indicate the region from where the images were taken.</p>
Full article ">Figure 5
<p>Prg1/LPPR4 reporter gene expression in the CNS, spinal cord and DRGs. (<b>A</b>,<b>B</b>) X-Gal histology and β-Galactosidase immunofluorescence of the somatosensory cortex (<b>A</b>) and thalamus (<b>B</b>) in naïve Prg1-β-Galactosidase reporter mice. Neurons were counterstained with the nuclear neuronal marker NeuN (SSC, thalamus). (<b>C</b>,<b>D</b>) X-Gal histology and β-Galactosidase immunofluorescence in the spinal cord dorsal horn in naïve Prg1-β-Galactosidase reporter mice. Subpanel C shows overviews of the dorsal horn. FDG (Fluorescein di-β-D-galactopyranoside) was used for detection of β-Gal. Isolectin B4 (IB4) was used to highlight Lamina II nerve fiber terminals. Subpanel D shows higher magnifications of the dorsal horn with counterstaining of IB4, NeuN (neurons), Gad67 (GABAergic synapses) and Homer (glutamatergic synapses). Prg1-β-Gal dots reveal synaptic localization. (<b>E</b>) X-Gal histology and β-Galactosidase immunofluorescence in the DRGs. No specific signal was observed. DRGs were negative. Eleven Prg1-LacZ mice were used to assess the localization of Prg1 along the itch-signaling axis.</p>
Full article ">Figure 6
<p>Itch behavior in LPAR2<sup>−/−</sup> and Advillin-LPAR5<sup>−/−</sup> mice. Scratching behavior was evoked by intradermal injection of Cp 48/80 (histamine-evoked itch), HCQ (non-histaminergic itch) and Bam 8–22 (peptidergic itch) and observed for 30 min in 5 min intervals. Column headlines indicate the mouse line, and row headlines show the stimulus. Time courses are shown in the left panels (mean ± sem), and box plots show the total scratching time. The box is the interquartile range, the line the median, whiskers minimum to maximum. Scatters show individual mice (n = 11 for LPAR2 experiments; n = 18–22 for LPAR5 experiments). Total scratching times were compared by unpaired, 2-tailed Student’s <span class="html-italic">t</span>-test; * <span class="html-italic">p</span> &lt; 0.05.</p>
Full article ">Figure 7
<p>Itch behavior in Prg1/LPPR4<sup>−/−</sup> and Prg2/LPPR3<sup>−/−</sup> mice. Scratching behavior was evoked by intradermal injection of Cp 48/80 (histamine-evoked itch), HCQ (non-histaminergic itch) and Bam 8–22 (peptidergic itch) and observed for 30 min at 5 min intervals. Column headlines indicate the mouse line, and row headlines show the stimulus. Time courses are shown in the left panels (mean ± sem), and box plots show the total scratching time. The box is the interquartile range, the line the median, whiskers minimum to maximum. Scatters show individual mice (n = 16–26 for Prg1 experiments; n = 12 for Prg2 experiments). Total scratching times were compared by unpaired, 2-tailed Student’s <span class="html-italic">t</span>-test; * <span class="html-italic">p</span> &lt; 0.05.</p>
Full article ">Figure 8
<p>Itch behavior in GFAP-ENPP2<sup>−/−</sup> mice and effects of autotaxin, LPAR2 and LPAR5 inhibitors. Scratching behavior was evoked by intradermal injection of Cp 48/80 (histamine-evoked itch) or HCQ (non-histaminergic itch) and observed for 30 min at 5 min intervals. Time courses are shown in the left panels (mean ± sem), and box plots show the total scratching time. The box is the interquartile range, the line the median, whiskers minimum to maximum. Scatters show individual mice. Total scratching times were compared by unpaired, 2-tailed Student’s <span class="html-italic">t</span>-test; * <span class="html-italic">p</span> &lt; 0.05. Drugs were injected 20 min before itch stimulation. Control mice received an equal volume of vehicle (5% DMSO in saline). (<b>A</b>) Scratching behavior of GFAP-ENPP2<sup>−/−</sup> versus floxed control mice (ENPP2-flfl), n = 6 per group. GFAP-ENPP2<sup>−/−</sup> mice carry a deletion of autotaxin/ENPP2 specifically in astrocytes. (<b>B</b>) Scratching behavior in mice treated with the autotaxin inhibitors, PF-8380 or vehicle, i.p., n = 10 per group. (<b>C</b>) Scratching behavior in mice treated with the LPAR2 antagonist H2L5186303 or vehicle, i.p., n = 10 or 12 per group. (<b>D</b>) Scratching behavior in mice treated with the LPAR5 antagonist TC LPA5 4 or vehicle, i.p., n = 10 per group. * <span class="html-italic">p</span> &lt; 0.05, *** <span class="html-italic">p</span> &lt; 0.001.</p>
Full article ">Figure 9
<p>Histamine and HCQ-evoked calcium fluxes in primary sensory neurons of Advillin-LPAR5<sup>−/−</sup> mice. Calcium imaging was performed by analysis of absorbance ratios of Fura-2 upon excitation at 340 and 380 nm. Primary DRG neurons of Advillin-LPAR5<sup>−/−</sup> and LPAR5-flfl control mice were stimulated sequentially with 1 mM histamine, 0.5 mM HCQ and final 75 mM K<sup>+</sup> (in <b>A</b>) or final 0.1 µM capsaicin (in <b>B</b>). The left panels show the time courses of the fold change in [Ca<sup>2+</sup>]<sub>i</sub> versus baseline. The Chi-square contingency analysis shows the total numbers of neurons and stimulus-responsive neurons. A response was defined as a fold increase in [Ca<sup>2+</sup>]<sub>i</sub> &gt; 1.3. Time courses show responsive neurons. “Not” means not responsive. (<b>A</b>) Calcium imaging using a final stimulus of high potassium (75 mM K<sup>+</sup>) to evoke depolarization-dependent calcium influx. (<b>B</b>) Calcium imaging using capsaicin as final stimulus to evoke TRPV1-dependent calcium influx.</p>
Full article ">Figure 10
<p>Putative LPA-pruriceptive signaling paths. LPAs are released from keratinocytes upon pruriceptive stimulation and activate fiber terminals of LPAR5-positive pruriceptive neurons, causing glutamate or neuropeptide release from central terminals and activation of postsynaptic neurons. Calcium influx stimulates LPA generation and release from secondary neurons. Extracellularly, LPA is generated from LPC via autotaxin (ENPP2). Synaptic LPA can be recycled via Prg1/LPPR4.</p>
Full article ">
13 pages, 2643 KiB  
Article
The Influence of Sex Steroid Hormone Fluctuations on Capsaicin-Induced Pain and TRPV1 Expression
by Edgardo Mota-Carrillo, Rebeca Juárez-Contreras, Ricardo González-Ramírez, Enoch Luis and Sara Luz Morales-Lázaro
Int. J. Mol. Sci. 2024, 25(15), 8040; https://doi.org/10.3390/ijms25158040 - 24 Jul 2024
Viewed by 1509
Abstract
Sexual dimorphism among mammals includes variations in the pain threshold. These differences are influenced by hormonal fluctuations in females during the estrous and menstrual cycles of rodents and humans, respectively. These physiological conditions display various phases, including proestrus and diestrus in rodents and [...] Read more.
Sexual dimorphism among mammals includes variations in the pain threshold. These differences are influenced by hormonal fluctuations in females during the estrous and menstrual cycles of rodents and humans, respectively. These physiological conditions display various phases, including proestrus and diestrus in rodents and follicular and luteal phases in humans, distinctly characterized by varying estrogen levels. In this study, we evaluated the capsaicin responses in male and female mice at different estrous cycle phases, using two murine acute pain models. Our findings indicate that the capsaicin-induced pain threshold was lower in the proestrus phase than in the other three phases in both pain assays. We also found that male mice exhibited a higher pain threshold than females in the proestrus phase, although it was similar to females in the other cycle phases. We also assessed the mRNA and protein levels of TRPV1 in the dorsal root and trigeminal ganglia of mice. Our results showed higher TRPV1 protein levels during proestrus compared to diestrus and male mice. Unexpectedly, we observed that the diestrus phase was associated with higher TRPV1 mRNA levels than those in both proestrus and male mice. These results underscore the hormonal influence on TRPV1 expression regulation and highlight the role of sex steroids in capsaicin-induced pain. Full article
(This article belongs to the Special Issue TRP Channel, 2nd Edition)
Show Figures

Figure 1

Figure 1
<p>Evaluation of capsaicin-induced acute pain in male and female mice. (<b>A</b>) Paw licking assay: vehicle solution injection (bars with empty symbols) induced minimal paw licking response in all experimental groups (males: <span class="html-italic">n</span> = 6, P: <span class="html-italic">n</span> = 14, E: <span class="html-italic">n</span> = 6, M: <span class="html-italic">n</span> = 3, and D: <span class="html-italic">n</span> = 12). The nociceptive response was observed following capsaicin injection in all experimental groups (bars with filled symbols, males: <span class="html-italic">n</span> = 11, P: <span class="html-italic">n</span> = 17, E: <span class="html-italic">n</span> = 9, M: <span class="html-italic">n</span> = 3, and D: <span class="html-italic">n</span> = 14). (<b>B</b>) Cheek injection assay: The graph presents data generated from counting the number of wiping bouts following injection of either capsaicin or vehicle solution. Nociceptive response to capsaicin application was observed in all experimental groups (bars with filled symbols, males: <span class="html-italic">n</span> = 18, P: <span class="html-italic">n</span> = 12, E: <span class="html-italic">n</span> = 6, M: <span class="html-italic">n</span> = 4, and D: <span class="html-italic">n</span> = 8). The vehicle induced minimal wiping bouts in all experimental groups (bars with empty symbols, males: <span class="html-italic">n</span> = 6, P: <span class="html-italic">n</span> = 4, E: <span class="html-italic">n</span> = 3, M: <span class="html-italic">n</span> = 6, and D: <span class="html-italic">n</span> = 7). Proestrus (P), estrus (E), metestrus (M), and diestrus (D). Data are the mean ± SEM. One-way ANOVA followed by Tukey’s post hoc test, **, ***, and **** for <span class="html-italic">p</span> &lt; 0.05, <span class="html-italic">p</span> &lt; 0.001, and <span class="html-italic">p</span> &lt; 0.0001, respectively; ns: non-statistically significant.</p>
Full article ">Figure 2
<p>Evaluation of TRPV1 gene expression in the dorsal root ganglia (DRGs) of male and female mice in the proestrus (P) and diestrus (D) phases. (<b>A</b>) TRPV1 mRNA levels were assessed using qPCR. The data obtained showed similar TRPV1 mRNA levels across the three experimental groups. (<b>B</b>) Western Blot experiments demonstrated that TRPV1 protein levels were lower in the diestrus phase than in the proestrus phase (<span class="html-italic">n</span> = 9). Data are the mean ± SEM. One-way ANOVA followed by Tukey’s post hoc test. *** <span class="html-italic">p</span> &lt; 0.005; ns: non-statically significant. ADU: arbitrary densitometric unit.</p>
Full article ">Figure 3
<p>Evaluation of TRPV1 gene expression in the trigeminal ganglia (TGs) of male and female mice in the proestrus (P) and diestrus phases (D). (<b>A</b>) TRPV1 mRNA levels were assessed using qPCR. The data showed that TRPV1 mRNA levels are higher in diestrus than in proestrus and male mice (<span class="html-italic">n</span> = 3). (<b>B</b>) Western Blot experiments demonstrated that TRPV1 protein levels are lower in the diestrus phase compared to females in the proestrus phase (<span class="html-italic">n</span> = 13). Data are the mean ± SEM. One-way ANOVA test followed by Tukey’s post hoc test. * <span class="html-italic">p</span> &lt; 0.05; *** <span class="html-italic">p</span> &lt; 0.005. ADU: arbitrary densitometric unit.</p>
Full article ">Figure 4
<p>Calcium imaging recordings of TG neurons treated with low and high 17β-estradiol concentrations. (<b>A</b>) Summary graphs of the calcium imaging recordings performed on TG neurons in response to 500 nM capsaicin (left and middle) and high potassium (right) stimuli. The left graph shows the percentage of capsaicin-responding neurons under different experimental conditions: control condition (Ctr, <span class="html-italic">n</span> = 86), low E2 concentration (LE2, <span class="html-italic">n</span> = 125), and high E2 concentration (HE2, <span class="html-italic">n</span> = 117). The middle graph in A shows the amplitude responses to capsaicin, indicating that neurons from all groups respond similarly. The right graph in A displays the amplitude of the responses to the high potassium stimulus. (<b>B</b>) Representative recording of TG neuron groups, the capsaicin-positive neurons treated with HE2 show faster responses to capsaicin compared to those treated with LE2. The arrows represent the time capsaicin is added, and the arrowhead is when the high K<sup>+</sup> solution is added. (<b>C</b>) Representative recording of the capsaicin response latency under each treatment (records taken from B) and the summary graph of the time to peak of the capsaicin responses. The bar graph represents the mean ± SEM. One-way ANOVA followed by Tukey’s post hoc test. *, **, and **** for <span class="html-italic">p</span> &lt; 0.05, <span class="html-italic">p</span> &lt; 0.01, and <span class="html-italic">p</span> &lt; 0.005, respectively.</p>
Full article ">Figure 5
<p>Evaluation of ERα total protein in TGs from male and female mice in the proestrus and diestrus phases. (<b>A</b>) Representative WB for ERα (upper panel) and GAPDH (lower panel) immunodetection. The membrane was stripped for TRPV1 immunodetection (middle panel). (<b>B</b>) The graph shows the normalized data of the mean values for ERα (<span class="html-italic">n</span> = 3) or TRPV1 (<span class="html-italic">n</span> = 13; these data are the same as those used in <a href="#ijms-25-08040-f003" class="html-fig">Figure 3</a>) with respect to load control (GAPDH). Two-way ANOVA with Šídák’s multiple comparison test. ** <span class="html-italic">p</span> &lt; 0.005. ADU: arbitrary densitometric unit.</p>
Full article ">
19 pages, 7891 KiB  
Article
The Combination of Molecular Hydrogen and Heme Oxygenase 1 Effectively Inhibits Neuropathy Caused by Paclitaxel in Mice
by Ignacio Martínez-Martel, Xue Bai, Rebecca Kordikowski, Christie R. A. Leite-Panissi and Olga Pol
Antioxidants 2024, 13(7), 856; https://doi.org/10.3390/antiox13070856 - 17 Jul 2024
Viewed by 1483
Abstract
Chemotherapy-provoked peripheral neuropathy and its associated affective disorders are important adverse effects in cancer patients, and its treatment is not completely resolved. A recent study reveals a positive interaction between molecular hydrogen (H2) and a heme oxygenase (HO-1) enzyme inducer, cobalt [...] Read more.
Chemotherapy-provoked peripheral neuropathy and its associated affective disorders are important adverse effects in cancer patients, and its treatment is not completely resolved. A recent study reveals a positive interaction between molecular hydrogen (H2) and a heme oxygenase (HO-1) enzyme inducer, cobalt protoporphyrin IX (CoPP), in the inhibition of neuropathic pain provoked by nerve injury. Nevertheless, the efficacy of CoPP co-administered with hydrogen-rich water (HRW) on the allodynia and emotional disorders related to paclitaxel (PTX) administration has not yet been assessed. Using male C57BL/6 mice injected with PTX, we examined the effects of the co-administration of low doses of CoPP and HRW on mechanical and thermal allodynia and anxiodepressive-like behaviors triggered by PTX. Moreover, the impact of this combined treatment on the oxidative stress and inflammation caused by PTX in the amygdala (AMG) and dorsal root ganglia (DRG) were studied. Our results indicated that the antiallodynic actions of the co-administration of CoPP plus HRW are more rapid and higher than those given by each of them when independently administered. This combination inhibited anxiodepressive-like behaviors, the up-regulation of the inflammasome NLRP3 and 4-hydroxynonenal, as well as the high mRNA levels of some inflammatory mediators. This combination also increased the expression of NRF2, HO-1, superoxide dismutase 1, glutathione S-transferase mu 1, and/or the glutamate-cysteine ligase modifier subunit and decreased the protein levels of BACH1 in the DRG and/or AMG. Thus, it shows a positive interaction among HO-1 and H2 systems in controlling PTX-induced neuropathy by modulating inflammation and activating the antioxidant system. This study recommends the co-administration of CoPP plus HRW as an effective treatment for PTX-provoked neuropathy and its linked emotive deficits. Full article
(This article belongs to the Special Issue Experimental and Therapeutic Targeting of Heme Oxygenase)
Show Figures

Figure 1

Figure 1
<p>Schematic design of the experiments performed to evaluate whether the administration of CoPP (2.5 mg/kg), HRW (0.15 mM), or CoPP (2.5 mg/kg) plus HRW (0.15 mM) given intraperitoneally, twice a day, over three consecutive days, can inhibit the nociceptive responses (<b>A</b>) and/or the emotive disorders (<b>B</b>) caused by PTX. CP: cold plate test; EPM: elevated plus maze; FST: forced swimming test; PTX: paclitaxel; TST: tail suspension test; VEH; vehicle; VF: von Frey filaments.</p>
Full article ">Figure 2
<p>The inhibition of mechanical allodynia produced by the intraperitoneal administration of CoPP (2.5 mg/kg) and HRW (0.15 mM), alone and combined, 2 times per day, from days 19 to 21 after PTX injection, are represented. Data are presented as the von Frey filaments strength (g) on the left (<b>A</b>) and right (<b>C</b>) paws and their respective AUC to examine the global effect of these treatments (<b>B</b>,<b>D</b>). In all figures, symbols show significant differences vs. subjects given * VEH-VEH-VEH, VEH-CoPP-VEH, VEH-VEH-HRW or VEH-CoPP-HRW, + vs. PTX-VEH-VEH, # vs. PTX-CoPP-VEH and <span>$</span> vs. PTX-HRW-VEH (<span class="html-italic">p</span> &lt; 0.05, one-way ANOVA and Tukey test). Mean values ± SEM of 6 animals per group.</p>
Full article ">Figure 3
<p>The inhibition of thermal allodynia produced by the intraperitoneal administration of CoPP (2.5 mg/kg) and HRW (0.15 mM), alone and combined, 2 times per day, from days 19 to 21 after PTX injection, are represented. Data are presented as the paw lifts (number) in the left (<b>A</b>) and right (<b>C</b>) paws and their respective AUC to examine the global effect of these treatments (<b>B</b>,<b>D</b>). In all figures, symbols show significant differences vs. subjects given * VEH-VEH-VEH, VEH-CoPP-VEH, VEH-VEH-HRW or VEH-CoPP-HRW, + vs. PTX-VEH-VEH, # vs. PTX-CoPP-VEH and <span>$</span> vs. PTX-HRW-VEH (<span class="html-italic">p</span> &lt; 0.05, one-way ANOVA and Tukey test). Mean values ± SEM of 6 animals per group.</p>
Full article ">Figure 4
<p>The inhibition of anxious-like behaviors associated with PIPN induced by the intraperitoneal administration of CoPP (2.5 mg/kg) or HRW (0.15 mM), alone and combined, given 2 times per day over three consecutive days, are represented. The effects of CoPP, HRW, CoPP plus HRW, or VEH in animals given the VEH are also shown. The number of entrances to the open arms (<b>A</b>) and closed arms (<b>B</b>) and the proportion of time passed in the open arms (<b>C</b>) of the EPM are represented. In all graphs, * signifies significant differences vs. animals treated with VEH-VEH-VEH and + vs. animals treated with PTX-VEH-VEH (<span class="html-italic">p</span> &lt; 0.05, one-way ANOVA and Tuckey test). Data are expressed as mean values ± SEM of 8 animals per group.</p>
Full article ">Figure 5
<p>The inhibition of depressant-like behaviors associated with PIPN induced by the intraperitoneal administration of CoPP (2.5 mg/kg) or HRW (0.15 mM), alone and combined, given 2 times per day over three consecutive days, are represented. The effects of CoPP, HRW, CoPP plus HRW, or VEH in animals given the VEH are also shown. In the TST (<b>A</b>) and FST (<b>B</b>), the time that the animals remain immobile (s) is represented. In both graphs, * signifies significant differences vs. animals treated with VEH-VEH-VEH and + vs. animals treated with PTX-VEH-VEH (<span class="html-italic">p</span> &lt; 0.05, one-way ANOVA and Tuckey test). Data are expressed as mean values ± SEM of 8 animals per group.</p>
Full article ">Figure 6
<p>The effects of CoPP combined with HRW on the protein levels of NLRP3, 4-HNE, NRF2, HO-1, GSTM1, SOD-1, and BACH1 in the DRG of PTX-injected mice. This combined treatment reversed the up-regulation of NLRP3 (<b>A</b>) and 4-HNE (<b>B</b>), increased the protein levels of NRF2 (<b>C</b>), HO-1 (<b>D</b>), and SOD-1 (<b>F</b>), and decreased those of BACH1 (<b>G</b>) in the DRG of PTX-injected mice. No changes in GSTM1 levels (<b>E</b>) were observed. VEH-injected mice treated with VEH plus VEH were used as controls. In all graphs, symbols denote significant changes vs., * VEH-VEH-VEH treated mice, + vs. PTX-injected animals treated with VEH-VEH and # vs. PTX-injected mice treated whit CoPP-HRW (<span class="html-italic">p</span> &lt; 0.05; one-way ANOVA and Tukey test). Data are presented as mean values ± SEM of 3 samples/groups.</p>
Full article ">Figure 7
<p>The effects of CoPP combined with HRW on the protein levels of NLRP3, 4-HNE, NRF2, HO-1, GSTM1, SOD-1, and BACH1 in the AMG of PTX-injected mice. This combined treatment reversed the up-regulation of NLRP3 (<b>A</b>) and 4-HNE (<b>B</b>) and increased the protein levels of GSTM1 (<b>E</b>) in the AMG of PTX-injected mice. No changes in NRF2 (<b>C</b>), HO-1 (<b>D</b>), SOD-1 (<b>F</b>), and BACH1 (<b>G</b>) were identified. VEH-injected mice treated with VEH plus VEH were used as controls. In all graphs, symbols denote significant changes vs., * VEH-VEH-VEH treated mice, + vs. PTX-injected animals treated with VEH-VEH and # vs. PTX-injected mice treated with CoPP-HRW (<span class="html-italic">p</span> &lt; 0.05; one-way ANOVA, followed by the Tukey test). Data are presented as mean values ± SEM of 3 samples/group.</p>
Full article ">Figure 8
<p>The effects of CoPP combined with HRW on the mRNA levels of IL-1β, IL6, TNFα, GCLC, and GCLM in the DRG of PTX-injected mice. This combined treatment reversed the up-regulation of IL-1B (<b>A</b>). No changes in the IL-6 (<b>B</b>), TNFα (<b>C</b>), GCLC (<b>D</b>), or GCLM (<b>E</b>) were identified. VEH-injected mice treated with VEH plus VEH were used as controls. In all graphs, symbols denote significant changes vs., * VEH-VEH-VEH treated mice (<span class="html-italic">p</span> &lt; 0.05; one-way ANOVA and # vs. PTX-injected mice treated with CoPP-HRW, followed by the Tukey test). Data are presented as mean values ± SEM of 3 samples/group.</p>
Full article ">Figure 9
<p>The effects of CoPP combined with HRW on the mRNA levels of IL-1β, IL6, TNFα, GCLC, and GCLM in the AMG of PTX-injected mice. This combined treatment reversed the up-regulation of IL-6 (<b>B</b>) and TNFα (<b>C</b>) and maintained the increased expression of GCLM (<b>E</b>) provoked by PTX. No changes in IL-1β (<b>A</b>) and GCLC (<b>D</b>) were identified. VEH-injected mice treated with VEH plus VEH were used as controls. In all graphs, symbols denote significant changes vs., * VEH-VEH-VEH treated mice and # vs. PTX-injected mice treated with CoPP-HRW, followed by the Tukey test. Data are presented as mean values ± SEM of 3 samples/groups.</p>
Full article ">
22 pages, 4456 KiB  
Article
GRT-X Stimulates Dorsal Root Ganglia Axonal Growth in Culture via TSPO and Kv7.2/3 Potassium Channel Activation
by Léa El Chemali, Suzan Boutary, Song Liu, Guo-Jun Liu, Ryan J. Middleton, Richard B. Banati, Gregor Bahrenberg, Rainer Rupprecht, Michael Schumacher and Liliane Massaad-Massade
Int. J. Mol. Sci. 2024, 25(13), 7327; https://doi.org/10.3390/ijms25137327 - 3 Jul 2024
Cited by 1 | Viewed by 1397
Abstract
GRT-X, which targets both the mitochondrial translocator protein (TSPO) and the Kv7.2/3 (KCNQ2/3) potassium channels, has been shown to efficiently promote recovery from cervical spine injury. In the present work, we investigate the role of GRT-X and its two targets in the axonal [...] Read more.
GRT-X, which targets both the mitochondrial translocator protein (TSPO) and the Kv7.2/3 (KCNQ2/3) potassium channels, has been shown to efficiently promote recovery from cervical spine injury. In the present work, we investigate the role of GRT-X and its two targets in the axonal growth of dorsal root ganglion (DRG) neurons. Neurite outgrowth was quantified in DRG explant cultures prepared from wild-type C57BL6/J and TSPO-KO mice. TSPO was pharmacologically targeted with the agonist XBD173 and the Kv7 channels with the activator ICA-27243 and the inhibitor XE991. GRT-X efficiently stimulated DRG axonal growth at 4 and 8 days after its single administration. XBD173 also promoted axonal elongation, but only after 8 days and its repeated administration. In contrast, both ICA27243 and XE991 tended to decrease axonal elongation. In dissociated DRG neuron/Schwann cell co-cultures, GRT-X upregulated the expression of genes associated with axonal growth and myelination. In the TSPO-KO DRG cultures, the stimulatory effect of GRT-X on axonal growth was completely lost. However, GRT-X and XBD173 activated neuronal and Schwann cell gene expression after TSPO knockout, indicating the presence of additional targets warranting further investigation. These findings uncover a key role of the dual mode of action of GRT-X in the axonal elongation of DRG neurons. Full article
(This article belongs to the Collection Feature Papers in “Molecular Biology”)
Show Figures

Figure 1

Figure 1
<p>Impact of various compounds on axonal growth in embryonic C57BL/6 DRG explants at DIV4. (<b>A</b>) depicts the experimental protocol; (<b>B</b>) displays confocal microscopy images of DRG explants at DIV4 following treatments initiated at DIV1. Scale bar = 100 µm; (<b>C</b>) shows the mean of 3–4 DRGs per treatment across 4 independent experiments after quantification using Neurite-J; (<b>D</b>) represents the mean and the SD of the calculation of the areas under the curve (AUCs) for the different treatment groups after normalizing to non-treated (NT). The data are presented as mean ± standard deviation, with statistical significance indicated as *** <span class="html-italic">p</span> &lt; 0.001 and **** <span class="html-italic">p</span> &lt; 0.0001 (ANOVA followed by Tukey’s Multiple Comparison Test). The double-sided arrow represents a significant difference between GRT-X and all other treatments.</p>
Full article ">Figure 2
<p>GRT-X maintains a positive effect on axonal growth in C57BL/6 DRG explants at DIV8 after a single dose of treatment. (<b>A</b>) represents the experimental protocol; (<b>B</b>) represents confocal microscopy images of DRGs at DIV8 following their treatment with the different molecules at DIV1. Scale bar = 100 µm; (<b>C</b>) represents the mean of 3–4 DRGs per treatment across 3 independent experiments after quantification using Neurite-J; (<b>D</b>) represents the calculation of the areas under the curve of the different groups after normalizing to the non-treated group (NT). Data represent mean ± SD; * <span class="html-italic">p</span> &lt; 0.05, ** <span class="html-italic">p</span> &lt; 0.01 (ANOVA followed by Tukey’s Multiple Comparison Test). The double-sided arrow represents a significant difference between GRT-X and XBD173, XE991, and ICA27243.</p>
Full article ">Figure 3
<p>Effect of different treatments on axonal growth in C57BL/6 embryonic DRGs at DIV8 after two treatment sessions. (<b>A</b>) illustrates the experimental protocol; (<b>B</b>) displays confocal microscopy images of DRGs at DIV8 following treatment with various molecules at DIV1 and DIV4. Scale bar = 100 µm; (<b>C</b>) presents the graph resulting from Neurite-J quantification of neurites from 3–4 DRGs per treatment across 3 independent experiments; (<b>D</b>) shows the calculation of the areas under the curve (AUCs) for the different treatment groups after normalizing to the non-treated (NT) group. The data are presented as mean ± standard deviation, and statistical analysis was conducted using ANOVA followed by Tukey’s Multiple Comparison Test. * <span class="html-italic">p</span> &lt; 0.05, and *** <span class="html-italic">p</span> &lt; 0.001. The double-sided arrow represents a significant difference between GRT-X and all other treatments.</p>
Full article ">Figure 4
<p>mRNA expression of GRT-X target genes (<b>B</b>) and genes involved in myelination (<b>C</b>), Schwann cell differentiation (<b>D</b>), and neuronal structure (<b>E</b>,<b>F</b>) in C57Bl/6. (<b>A</b>): protocol used; (<b>B</b>–<b>F</b>): relative quantification of mRNA levels of TSPO, Kv7.2, Kv7.3, Mpz, Mbp, Plp, Dhh, Krox20, CNPase, Tfap2α, Peripherin, Cad19, and Nfh from dissociated cultures of C57BL/6 embryonic DRGs at DIV4 following the different treatments. Data represent mean ± SD of 3 independently performed manipulations. * <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 (ANOVA followed by Tukey’s Multiple Comparison Test). Only comparisons between the non-treated group and all the other treatments are represented in this figure. For further information, please refer to <a href="#app1-ijms-25-07327" class="html-app">Supplementary Table S1</a>.</p>
Full article ">Figure 5
<p>GRT-X has no effect on axonal growth of TSPO-KO embryonic DRG explants at DIV4. (<b>A</b>) represents the experimental protocol; (<b>B</b>) represents confocal microscopy images of DRG explants at DIV4 following their treatment with the different molecules at DIV1. Scale bar = 100 µm; (<b>C</b>) represents the graph plotted from 3 independent experiments following Neurite-J quantification of neurites from 3 or 4 DRGs per treatment; (<b>D</b>) represents the calculation of the areas under the curve of the different groups after normalizing to the non-treated (NT) group. Data represent mean ± SD of 3 independent experiments. *** <span class="html-italic">p</span> &lt; 0.001 (ANOVA followed by Tukey’s Multiple Comparison Test). The double-sided arrow represents a significant difference between NT, GRT-X, and XBD173 and both treatments XE991 and ICA27243.</p>
Full article ">Figure 6
<p>mRNA expression of GRT-X target genes (<b>B</b>) and genes involved in myelination (<b>C</b>), Schwann cell differentiation (<b>D</b>), and neuronal structure (<b>E</b>,<b>F</b>) in TSPO-KO DRG. (<b>A</b>): protocol used; (<b>B</b>–<b>F</b>): relative quantification of mRNA levels of Kv7.2, Kv7.3, Mpz, Mbp, Plp, Dhh, Krox20, CNPase, Tfap2α, Peripherin, Cad19, and Nfh from dissociated cultures of TSPO-KO embryonic DRGs at DIV4 following the different treatments. Data represent the mean ± SD of 3 independently performed experiments. * <span class="html-italic">p</span> &lt; 0.05, ** <span class="html-italic">p</span> &lt; 0.01, *** <span class="html-italic">p</span> &lt; 0.001, **** <span class="html-italic">p</span> &lt; 0.0001 (ANOVA followed by Tukey’s Multiple Comparison Test). Only comparisons with the non-treated group are represented in this figure. For further information, please refer to <a href="#app1-ijms-25-07327" class="html-app">Supplementary Table S2</a>.</p>
Full article ">Figure 7
<p>(<b>A</b>): Chemical structure of GRT-X (N-[(3-fluorophenyl)-methyl]-1-(2- methoxyethyl)-4-methyl-2-oxo-(7-trifluoromethyl)-1H-quinoline-3-caboxylic acid amide); (<b>B</b>): 2D representation of GRT-X. Black represents carbon atoms, blue represents nitrogen atoms, and red represents oxygen atoms.</p>
Full article ">Figure 8
<p>Experimental design (<b>A</b>) for DRG explant cultures: Embryos were surgically removed from pregnant females at E13.5 and spinal cords subsequently extracted. Each treatment (NT, GRT-X, XBD173, XE991, ICA27243, and EtOH) was allocated to a separate 4-well plate, with each well containing a DRG explant from a distinct embryo. Confocal microscopy was conducted at the designated time points after mounting the DRG explants on slides. (<b>B</b>) Dissociated DRG cultures: Spinal cords were obtained following the same procedure as in (<b>A</b>). A total of 40 DRGs per spinal cord were harvested, dissociated, and cultured in individual PETRI dishes, each representing a distinct treatment group (NT, GRT-X, XBD173, XE991, ICA27243, and EtOH). Real-time quantitative polymerase chain reaction (RT-qPCR) analysis was performed at the specified time points.</p>
Full article ">Scheme 1
<p>Representation of the effects of GRT-X according to our results [<a href="#B22-ijms-25-07327" class="html-bibr">22</a>].</p>
Full article ">
Back to TopTop